[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (6,046)

Search Parameters:
Keywords = probiotics

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 2795 KiB  
Article
Effects of Bifidobacterium-Fermented Milk on Obesity: Improved Lipid Metabolism through Suppression of Lipogenesis and Enhanced Muscle Metabolism
by Hitomi Maruta, Yusuke Fujii, Naoki Toyokawa, Shoji Nakamura and Hiromi Yamashita
Int. J. Mol. Sci. 2024, 25(18), 9934; https://doi.org/10.3390/ijms25189934 (registering DOI) - 14 Sep 2024
Viewed by 217
Abstract
Obesity is a major global health concern. Studies suggest that the gut microflora may play a role in protecting against obesity. Probiotics, including lactic acid bacteria and Bifidobacterium, have garnered attention for their potential in obesity prevention. However, the effects of Bifidobacterium [...] Read more.
Obesity is a major global health concern. Studies suggest that the gut microflora may play a role in protecting against obesity. Probiotics, including lactic acid bacteria and Bifidobacterium, have garnered attention for their potential in obesity prevention. However, the effects of Bifidobacterium-fermented products on obesity have not been thoroughly elucidated. Bifidobacterium, which exists in the gut of animals, is known to enhance lipid metabolism. During fermentation, it produces acetic acid, which has been reported to improve glucose tolerance and insulin resistance, and exhibit anti-obesity and anti-diabetic effects. Functional foods have been very popular around the world, and fermented milk is a good candidate for enrichment with probiotics. In this study, we aim to evaluate the beneficial effects of milks fermented with Bifidobacterium strains on energy metabolism and obesity prevention. Three Bifidobacterium strains (Bif-15, Bif-30, and Bif-39), isolated from newborn human feces, were assessed for their acetic acid production and viability in milk. These strains were used to ferment milk. Otsuka–Long–Evans Tokushima Fatty (OLETF) rats administered Bif-15-fermented milk showed significantly lower weight gain compared to those in the water group. The phosphorylation of AMPK was increased and the expression of lipogenic genes was suppressed in the liver of rats given Bif-15-fermented milk. Additionally, gene expression related to respiratory metabolism was significantly increased in the soleus muscle of rats given Bif-15-fermented milk. These findings suggest that milk fermented with the Bifidobacterium strain Bif-15 can improve lipid metabolism and suppress obesity. Full article
Show Figures

Figure 1

Figure 1
<p>Total body weight gain and total food intake. (<b>A</b>) Changes in body weight during the intervention period, (<b>B</b>) total body weight gain, (<b>C</b>) abdominal white adipose tissue weight, and (<b>D</b>) total food intake. Body weight changes in rats administered distilled water (water), acetic acid (ace), skim milk (milk), or <span class="html-italic">Bifidobacterium</span>-fermented milk (Bif-15, Bif-30, and Bif-39) starting from 15 weeks of age. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 2
<p>Effect of <span class="html-italic">Bifidobacterium</span>-fermented milks on the phosphorylation of AMPK in the liver. Total protein was isolated from the livers of OLETF rats in the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups, as described in the <a href="#sec4-ijms-25-09934" class="html-sec">Section 4</a>. Western blotting was performed to determine AMP-activated protein kinase (AMPK) phosphorylation in the liver. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">Bifidobacterium</span>-fermented milks on mRNA levels in the liver. Total RNA was isolated from the liver at 24 weeks of age. Real-time PCR analysis was performed to determine the mRNA levels of <span class="html-italic">Mixipl</span> (<b>A</b>), <span class="html-italic">LPK</span> (<b>B</b>), <span class="html-italic">ACC</span> (<b>C</b>), and <span class="html-italic">Fas</span> (<b>D</b>) in the liver of rats from the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 4
<p>Histological sections of the liver. (<b>A</b>) Representative images of Oil Red O staining (×100 magnification, scale bar = 500 μm) in the livers of OLETF rats from the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups. (<b>B</b>) Red areas (lipid droplets) were measured using ImageJ software. ** <span class="html-italic">p</span> &lt; 0.01, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">Bifidobacterium</span>-fermented milks on the expression of the GPR43 (<b>A</b>), MEF2A (<b>B</b>), PGC-1α (<b>C</b>), and SDH (<b>D</b>) genes in the soleus muscle. Total RNA was isolated from the soleus muscle of OLETF rats in the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups at 24 weeks of age. Real-time PCR analysis was performed to determine the mRNA levels of <span class="html-italic">GPR43</span>, <span class="html-italic">Mef2a</span>, <span class="html-italic">Ppargc1a</span>, and <span class="html-italic">Sdha</span> in the soleus muscle. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). ** <span class="html-italic">p</span> &lt; 0.01, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 6
<p>Effects of <span class="html-italic">Bifibacterium</span> fermented milks on the expression of the GPR43 (<b>A</b>), MEF2A (<b>B</b>), PGC-1α (<b>C</b>), and SDH (<b>D</b>) genes in the gastrocnemius muscle of rats. Total RNA was isolated from the gastrocnemius muscle of OLETF rats at 24 weeks of age. Real-time PCR analysis was performed to determine the mRNA levels of <span class="html-italic">GPR43</span>, <span class="html-italic">Mef2a</span>, <span class="html-italic">Ppargc1a</span>, and <span class="html-italic">Sdha</span> in the gastrocnemius muscle. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 7
<p>Effects of <span class="html-italic">Bifibacterium</span>-fermented milks on the phosphorylated AMPK, PGC-1α, and MEF2A protein levels in the soleus muscle of rats. Total protein was isolated from the soleus muscle of OLETF rats in the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups at 24 weeks of age. Western blotting was carried out to determine the levels of pAMPK (<b>A</b>), PGC-1α (<b>B</b>), and MEF2A (<b>C</b>), as described in the <a href="#sec4-ijms-25-09934" class="html-sec">Section 4</a>. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 8
<p>Effects of <span class="html-italic">Bifibacterium</span>-fermented milks on the phosphorylated AMPK, PGC-1α, and MEF2A protein levels in the gastrocnemius muscle of rats. Total protein was isolated from the gastrocnemius muscles of OLETF rats in the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups at 24 weeks of age. Western blotting was carried out to determine the levels of pAMPK (<b>A</b>), PGC-1α (<b>B</b>), and MEF2A (<b>C</b>), as described in the <a href="#sec4-ijms-25-09934" class="html-sec">Section 4</a>. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, according to Dunnett’s test, compared to the water group.</p>
Full article ">Figure 9
<p>Effects of <span class="html-italic">Bifibacterium</span>-fermented milks on the mtDNA levels in the skeletal muscles. Genomic DNA was isolated from the soleus and gastrocnemius muscles of OLETF rats in the water, ace, milk, Bif-15, Bif-30, and Bif-39 groups at 24 weeks of age. Real-time PCR analysis was performed to determine <span class="html-italic">ND1</span> levels in the soleus (<b>A</b>) and gastrocnemius (<b>B</b>) muscles. Each value represents the mean ± SE (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001, according to Dunnett’s test, compared to the water group.</p>
Full article ">
17 pages, 3031 KiB  
Article
Functional Muffins Exert Bifidogenic Effects along with Highly Product-Specific Effects on the Human Gut Microbiota Ex Vivo
by Stef Deyaert, Jonas Poppe, Lam Dai Vu, Aurélien Baudot, Sarah Bubeck, Thomas Bayne, Kiran Krishnan, Morgan Giusto, Samuel Moltz and Pieter Van den Abbeele
Metabolites 2024, 14(9), 497; https://doi.org/10.3390/metabo14090497 (registering DOI) - 14 Sep 2024
Viewed by 191
Abstract
GoodBiome™ Foods are functional foods containing a probiotic (Bacillus subtilis HU58™) and prebiotics (mainly inulin). Their effects on the human gut microbiota were assessed using ex vivo SIFR® technology, which has been validated to provide clinically predictive insights. GoodBiome™ Foods (BBM/LCM/OSM) [...] Read more.
GoodBiome™ Foods are functional foods containing a probiotic (Bacillus subtilis HU58™) and prebiotics (mainly inulin). Their effects on the human gut microbiota were assessed using ex vivo SIFR® technology, which has been validated to provide clinically predictive insights. GoodBiome™ Foods (BBM/LCM/OSM) were subjected to oral, gastric, and small intestinal digestion/absorption, after which their impact on the gut microbiome of four adults was assessed (n = 3). All GoodBiome™ Foods boosted health-related SCFA acetate (+13.1/14.1/13.8 mM for BBM/LCM/OSM), propionate (particularly OSM; +7.4/7.5/8.9 mM for BBM/LCM/OSM) and butyrate (particularly BBM; +2.6/2.1/1.4 mM for BBM/LCM/OSM). This is related to the increase in Bifidobacterium species (B. catenulatum, B. adolescentis, B. pseudocatenulatum), Coprococcus catus and Bacteroidetes members (Bacteroides caccae, Phocaeicola dorei, P. massiliensis), likely mediated via inulin. Further, the potent propionogenic potential of OSM related to increased Bacteroidetes members known to ferment oats (s key ingredient of OSM), while the butyrogenic potential of BBM related to a specific increase in Anaerobutyricum hallii, a butyrate producer specialized in the fermentation of erythritol (key ingredient of BBM). In addition, OSM/BBM suppressed the pathogen Clostridioides difficile, potentially due to inclusion of HU58™ in GoodBiome™ Foods. Finally, all products enhanced a spectrum of metabolites well beyond SCFA, including vitamins (B3/B6), essential amino acids, and health-related metabolites such as indole-3-propionic acid. Overall, the addition of specific ingredients to complex foods was shown to specifically modulate the gut microbiome, potentially contributing to health benefits. Noticeably, our findings contradict a recent in vitro study, underscoring the critical role of employing a physiologically relevant digestion/absorption procedure for a more accurate evaluation of the microbiome-modulating potential of complex foods. Full article
(This article belongs to the Special Issue Natural Metabolites on Gut Microbiome Modulation)
Show Figures

Figure 1

Figure 1
<p><b>Schematic overview of the study design using ex vivo SIFR<sup>®</sup> technology.</b> (<b>a</b>) Reactor design using the ex vivo SIFR<sup>®</sup> technology to evaluate the impact of GoodBiome<sup>TM</sup> Foods against an unsupplemented parallel control (NSC = no substrate control). (<b>b</b>) Timeline and analysis at different timepoints.</p>
Full article ">Figure 2
<p><b>The fecal microbiota covered clinically relevant interpersonal differences.</b> Abundances (%) of the key families (top 15), as quantified via shallow shotgun sequencing, in the fecal microbiota of each of the four human adults that provided a fecal donation for the current SIFR<sup>®</sup> study.</p>
Full article ">Figure 3
<p><b>GoodBiome™ Foods exerted marked effects on microbial metabolic activity over time.</b> The effects on (<b>A</b>) pH, (<b>B</b>) gas production, (<b>C</b>) total SCFA, (<b>D</b>) acetate, (<b>E</b>) propionate, (<b>F</b>) butyrate, (<b>G</b>) valerate, and (<b>H</b>) bCFA were compared for GoodBiome™ Foods versus an unsupplemented control (NSC) at 6 h, 24 h, 30 h, and 48 h after the initiation of colonic incubation. Data were presented as means across simulations for four individual donors (n = 3 per donor). The statistical significance of the treatment effects for the test products vs. NSC within each timepoint can be found in <a href="#app1-metabolites-14-00497" class="html-app">Figures S2 and S3</a>.</p>
Full article ">Figure 4
<p><b>GoodBiome™ Foods exerted significant impact on microbial composition at phylum level.</b> Samples were collected 30 h after the colonic incubations were initiated. Data were expressed as average absolute levels (cells/mL) of each phylum across simulations for four individual donors (n = 3 per donor). The statistical significance of the potential treatment effects within each comparison was determined via Benjamani–Hochberg post hoc testing. Significant changes (<span class="html-italic">p</span><sub>adjusted</sub> &lt; 0.05) were indicated with asterisks.</p>
Full article ">Figure 5
<p><b>GoodBiome™ Foods exerted significant impact on microbial composition at species level.</b> The bar charts were generated for species that were significantly (FDR = 0.05) affected by any of the treatments at 30 h, expressed as log2fold change (treatment/NSC), averaged across four human adults (n = 3 per donor). Purple and red bars indicated significant/consistent decreases and increases, respectively. Notable health- or disease-related taxa are highlighted in a gray box.</p>
Full article ">Figure 6
<p><b>The GoodBiome™ Foods exerted significant impact on taxa that are potentially relevant for human health.</b> Violin plots, expressed as log2fold change (treatment/NSC), were presented for four individual human adults (n = 3). The data were presented for (<b>A</b>) <span class="html-italic">Clostridiodes difficile</span> (<b>B</b>) <span class="html-italic">Bifidobacteriaceae</span>, (<b>C</b>) <span class="html-italic">Anaerobutyricum hallii</span>, (<b>D</b>) <span class="html-italic">Bacteroidaceae</span>, <span class="html-italic">Bacteroidales_u_f</span>, and <span class="html-italic">Tannerellaceae</span>. For (<b>B</b>–<b>D</b>), Pearson correlation analysis demonstrated significant positive correlations (<span class="html-italic">p</span> &lt; 0.05) between the absolute levels of these taxa (cells/mL) and the concentration (mM) of the most relevant SCFA related to these taxa, i.e., (<b>A</b>) acetate, (<b>B</b>) butyrate, and (<b>C</b>) propionate.</p>
Full article ">Figure 7
<p><b>The GoodBiome™ Foods exerted significant impact on the production of microbial metabolites, well beyond SCFA.</b> The bars were generated for metabolites that were significantly (FDR = 0.05) affected by any of the treatments, expressed as log2fold change (treatment/NSC), averaged across four human adults (n = 3 per test subject). Purple and red bars indicated significant decreases and increases, respectively.</p>
Full article ">
16 pages, 2637 KiB  
Proceeding Paper
Optimizing the Formulation of Homemade Milk Kefir Drink from India: Comprehensive Microbial, Physicochemical, Nutritional, and Bioactivity Profiling
by Muskan Chadha, Ratnakar Shukla, Rohit Kumar Tiwari, Shalini Choudhary, Anisha Adya and Karuna Singh
Eng. Proc. 2024, 67(1), 44; https://doi.org/10.3390/engproc2024067044 (registering DOI) - 14 Sep 2024
Viewed by 51
Abstract
Kefir is a naturally fermented milk drink with rich probiotic content. This study aimed to develop and optimize homemade cow milk kefir (HCMK) and evaluate its microbial, chemical, nutritional, and antioxidant properties. HCMK was optimized using response surface methodology, where the independent variables [...] Read more.
Kefir is a naturally fermented milk drink with rich probiotic content. This study aimed to develop and optimize homemade cow milk kefir (HCMK) and evaluate its microbial, chemical, nutritional, and antioxidant properties. HCMK was optimized using response surface methodology, where the independent variables were kefir grain inoculum (2–4%) and fermentation time (20 h–28 h), and the dependent variables were total bacterial count, pH, and overall acceptability. HCMK was prepared using 3% w/v Indian kefir grains inoculated into cow milk and fermented at 25 °C for 24 h. Optimized dependent variables were 2.08 × 108 ± 0.34 CFU/mL, pH 4.52 ± 0.05, and overall acceptability of 6.55 ± 0.21. Nutritional analysis revealed protein 3.6 g/100 mL, carbohydrates 2.66 g/100mL, fat 3.4 g/100 mL, iron 2.99 mg/100 mL, and calcium 29.3 mg/100 mL. Antioxidant profiling elucidated 54% radical scavenging activity and 18 mgGAE/100 mL total phenolic content. GC-MS revealed the presence of bioactive compounds with documented antioxidant, anti-inflammatory, and antimicrobial activities. This study highlights HCMK as a healthy probiotic functional food with significant antioxidant potential. Full article
(This article belongs to the Proceedings of The 3rd International Electronic Conference on Processes)
Show Figures

Figure 1

Figure 1
<p>Formulation of milk kefir drink using household conditions. (<b>i</b>) Boiling milk at 100 °C, (<b>ii</b>) kefir grain inoculum, (<b>iii</b>) addition of milk, (<b>iv</b>) fermentation at room temperature, (<b>v</b>) sieving kefir grains for reuse, and (<b>vi</b>) HCMK drink stored at 4 °C.</p>
Full article ">Figure 2
<p>Three-dimensional response graphs showing the effects of kefir grains and fermentation time on (<b>a</b>) total bacterial count, (<b>b</b>) pH, and (<b>c</b>) overall acceptability.</p>
Full article ">Figure 3
<p>Sensory evaluation of HCMK compared to control.</p>
Full article ">Figure 4
<p>Microbial analysis of HCMK compared to control; *** <span class="html-italic">p</span> &lt; 0.001 indicates a significant difference.</p>
Full article ">Figure 5
<p>GCMS chromatogram of HCMK.</p>
Full article ">Figure 6
<p>Antioxidant efficacy of HCMK compared to control; *** <span class="html-italic">p</span> &lt; 0.001 indicates a significant difference.</p>
Full article ">Figure 7
<p>DPPH radical scavenging activity % of the HCMK and control; % radical scavenging activity (RSA).</p>
Full article ">
22 pages, 3939 KiB  
Article
Investigating the Effect of Bacilli and Lactic Acid Bacteria on Water Quality, Growth, Survival, Immune Response, and Intestinal Microbiota of Cultured Litopenaeus vannamei
by Ana Sofía Vega-Carranza, Ruth Escamilla-Montes, Jesús Arturo Fierro-Coronado, Genaro Diarte-Plata, Xianwu Guo, Cipriano García-Gutiérrez and Antonio Luna-González
Animals 2024, 14(18), 2676; https://doi.org/10.3390/ani14182676 (registering DOI) - 14 Sep 2024
Viewed by 207
Abstract
Shrimp is one of the most important aquaculture industries. Therefore, we determined the effect of nitrifying-probiotic bacteria on water quality, growth, survival, immune response, and intestinal microbiota of Litopenaeus vannamei cultured without water exchange. In vitro, only Bacillus licheniformis used total ammonia nitrogen [...] Read more.
Shrimp is one of the most important aquaculture industries. Therefore, we determined the effect of nitrifying-probiotic bacteria on water quality, growth, survival, immune response, and intestinal microbiota of Litopenaeus vannamei cultured without water exchange. In vitro, only Bacillus licheniformis used total ammonia nitrogen (TAN), nitrites, and nitrates since nitrogen bubbles were produced. TAN decreased significantly in the treatments with B. licheniformis and Pediococcus pentosaceus and Leuconostoc mesenteroides, but no differences were observed in nitrites. Nitrates were significantly higher in the treatments with bacteria. The final weight was higher only with bacilli and bacilli and LAB treatments. The survival of shrimp in the bacterial treatments increased significantly, and superoxide anion increased significantly only in lactic acid bacteria (LAB) treatment. The activity of phenoloxidase decreased significantly in the treatments with bacteria compared to the control. Shrimp treated with bacilli in the water showed lower species richness. The gut bacterial community after treatments was significantly different from that of the control. Linoleic acid metabolism was positively correlated with final weight and superoxide anion, whereas quorum sensing was correlated with survival. Thus, bacilli and LAB in the water of hyperintensive culture systems act as heterotrophic nitrifers, modulate the intestinal microbiota and immune response, and improve the growth and survival of shrimp. This is the first report on P. pentosaceus and L. mesenteroides identified as nitrifying bacteria. Full article
(This article belongs to the Special Issue The Application of Probiotics for Sustainable Aquaculture)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Concentration of TAN, nitrites, and nitrates on Days 7 (<b>A</b>), 15 (<b>B</b>), and 30 (<b>C</b>) in the shrimp culture system without water exchange and treated with bacteria. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Data are mean ± SD. Different letters indicate significant differences.</p>
Full article ">Figure 2
<p>Survival of shrimp cultured without water exchange and treated with bacteria. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Data are mean ± SD. Different letters indicate significant differences.</p>
Full article ">Figure 3
<p>Total hemocyte count in <span class="html-italic">L. vannamei</span> cultured without water exchange and treated with bacteria. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Data are mean ± SD.</p>
Full article ">Figure 4
<p>Superoxide anion in hemocytes of <span class="html-italic">L. vannamei</span> cultured without water exchange and treated with bacteria. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Data are mean ± SD. Different letters indicate significant differences.</p>
Full article ">Figure 5
<p>Phenoloxidase activity (absorbance) in hemolymph of <span class="html-italic">L. vannamei</span> cultured without water exchange and treated with bacteria. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Data are mean ± SD. Different letters indicate significant differences.</p>
Full article ">Figure 6
<p>Venn analysis of the bacterial communities in the shrimp intestine at the OTUs level. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water.</p>
Full article ">Figure 7
<p>Most abundant bacterial phyla (%) in the shrimp intestine. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Patescibacteria phylum (* no significant differences [<span class="html-italic">p</span> &gt; 0.05], ** significant differences [<span class="html-italic">p</span> &lt; 0.05]). The analysis was done with Shaman.</p>
Full article ">Figure 8
<p>Most abundant bacterial genera (%) in the shrimp intestines. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). The analysis was done with Shaman.</p>
Full article ">Figure 9
<p>Beta diversity of intestinal microbiota of <span class="html-italic">L. vannamei</span> at the genus level using non-metric multidimensional scaling based on Jaccard distances in MicrobiomeAnalyst. Treatments: (I) Control without bacteria in the water; (II) bacilli in the water; (III) LAB in the water; (IV) bacilli + LAB in the water. ANOSIM test, <span class="html-italic">p</span> &lt; 0.008.</p>
Full article ">Figure 10
<p>Correlation among linoleic acid metabolism of intestinal bacteria of <span class="html-italic">L. vannamei</span> and immune and productive variables. Spearman correlation analysis.</p>
Full article ">Figure 11
<p>Correlation among quorum sensing of intestinal bacteria of <span class="html-italic">L. vannamei</span> and immune and productive variables. Spearman correlation analysis.</p>
Full article ">
14 pages, 2997 KiB  
Article
Lactic Acid Bacterial Fermentation of Esterified Agave Fructans in Simulated Physicochemical Colon Conditions for Local Delivery of Encapsulated Drugs
by Carmen Miramontes-Corona, Abraham Cetina-Corona, María Esther Macías-Rodríguez, Alfredo Escalante, Rosa Isela Corona-González and Guillermo Toriz
Fermentation 2024, 10(9), 478; https://doi.org/10.3390/fermentation10090478 (registering DOI) - 14 Sep 2024
Viewed by 179
Abstract
Understanding drug release in the colon is fundamental to developing efficient treatments for colon-related diseases, while unraveling the relationship between the colonic microbiota and excipients is crucial to unveiling the effect of biomaterials on the release of drugs. In this contribution, the bio-release [...] Read more.
Understanding drug release in the colon is fundamental to developing efficient treatments for colon-related diseases, while unraveling the relationship between the colonic microbiota and excipients is crucial to unveiling the effect of biomaterials on the release of drugs. In this contribution, the bio-release of ibuprofen (encapsulated in acetylated and palmitoylated agave fructans) was evaluated by fermentation with lactic acid bacteria in simulated physicochemical (pH and temperature) colon conditions. It was observed that the size of the acyl chain (1 in acetyl and 15 in palmitoyl) was critical both in the growth of the microorganisms and in the release of the drug. For example, both the bacterial growth and the release of ibuprofen were more favored with acetylated fructan microspheres. Among the microorganisms evaluated, Bifidobacterium adolescentis and Lactobacillus brevis showed great potential as probiotics useful to release drugs from modified fructans. The production of short-chain fatty acids (lactic, acetic, and propionic acids) in the course of fermentations was also determined, since such molecules have a positive effect both on colon-related diseases and on the regulation of the intestinal microbiota. It was found that a higher concentration of acetate is related to a lower growth of bacteria and less release of ibuprofen. Full article
(This article belongs to the Special Issue Fermentation: 10th Anniversary)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) A model of agave fructan according to [<a href="#B33-fermentation-10-00478" class="html-bibr">33</a>]; in native fructan, R is OH; in palmitoylated fructan, 4 Rs are substituted with palmitoyl moieties; for acetylated fructan, about 62 Rs should be acetyl groups. (<b>B</b>) <sup>1</sup>H NMR spectra: (<b>a</b>) native fructan analyzed in D<sub>2</sub>O; H-C* denotes the proton at the anomeric carbon in glucose (<b>b</b>) palmitoylated fructan obtained in CDCl<sub>3</sub>; (<b>c</b>) acetylated fructans analyzed in <sup>d6</sup>DMSO.</p>
Full article ">Figure 2
<p>Scanning Electron Micrographs of (<b>a</b>) acetylated (7920X) and (<b>b</b>) palmitoylated fructan (1750X).</p>
Full article ">Figure 3
<p>Growth profile of strains and their consortium obtained by optical density of (<b>a</b>) acetylated fructan microspheres and (<b>b</b>) palmitoylated fructan microspheres: <span class="html-italic">B. adolescentis</span> (●); <span class="html-italic">Weisella paramesenteroides</span> (◆); <span class="html-italic">Enterococcus mundtii</span> (◼); <span class="html-italic">Lactobacillus brevis</span> (▲); <span class="html-italic">consortium</span> (∗).</p>
Full article ">Figure 4
<p>Percentage of ibuprofen release as function of fermentation time from (<b>a</b>) acetylated and (<b>b</b>) palmitoylated fructan microspheres with <span class="html-italic">B. adolescentis</span> (●); <span class="html-italic">Weisella paramesenteroides</span> (◆); <span class="html-italic">Enterococcus mundtii</span> (◼); <span class="html-italic">Lactobacillus brevis</span> (▲); and <span class="html-italic">consortium</span> (∗).</p>
Full article ">Figure 5
<p>Production of SCFA (g/L) by lactic acid bacteria at 48 h of fermentation: (<b>a</b>) acetylated fructan microspheres and (<b>b</b>) palmitoylated fructan microspheres, (■) lactic acid, (□) acetic acid, and (<span class="html-fig-inline" id="fermentation-10-00478-i001"><img alt="Fermentation 10 00478 i001" src="/fermentation/fermentation-10-00478/article_deploy/html/images/fermentation-10-00478-i001.png"/></span>) propionic acid. BA: <span class="html-italic">B. adolescentis</span>; WP: <span class="html-italic">Weisella paramesenteroides</span>; EL: <span class="html-italic">Enterococcus mundtii</span>; LB: <span class="html-italic">Lactobacillus brevis</span>; C: <span class="html-italic">consortium</span>.</p>
Full article ">Scheme 1
<p>Phylogenetic tree for identification of <span class="html-italic">Weissella paramesenteroides</span> (Jal1), <span class="html-italic">Enterococcus mundtii</span> (BT5inv), and <span class="html-italic">Lactobacillus brevis</span> (Col18) isolated from carposphere of tomato. The red boxes indicate the identified strains in the phylogenetic tree.</p>
Full article ">
31 pages, 3407 KiB  
Review
Glucose Metabolism-Modifying Natural Materials for Potential Feed Additive Development
by Wei-Chih Lin, Boon-Chin Hoe, Xianming Li, Daizheng Lian and Xiaowei Zeng
Pharmaceutics 2024, 16(9), 1208; https://doi.org/10.3390/pharmaceutics16091208 - 13 Sep 2024
Viewed by 269
Abstract
Glucose, a primary energy source derived from animals’ feed ration, is crucial for their growth, production performance, and health. However, challenges such as metabolic stress, oxidative stress, inflammation, and gut microbiota disruption during animal production practices can potentially impair animal glucose metabolism pathways. [...] Read more.
Glucose, a primary energy source derived from animals’ feed ration, is crucial for their growth, production performance, and health. However, challenges such as metabolic stress, oxidative stress, inflammation, and gut microbiota disruption during animal production practices can potentially impair animal glucose metabolism pathways. Phytochemicals, probiotics, prebiotics, and trace minerals are known to change the molecular pathway of insulin-dependent glucose metabolism and improve glucose uptake in rodent and cell models. These compounds, commonly used as animal feed additives, have been well studied for their ability to promote various aspects of growth and health. However, their specific effects on glucose uptake modulation have not been thoroughly explored. This article focuses on glucose metabolism is on discovering alternative non-pharmacological treatments for diabetes in humans, which could have significant implications for developing feed additives that enhance animal performance by promoting insulin-dependent glucose metabolism. This article also aims to provide information about natural materials that impact glucose uptake and to explore their potential use as non-antibiotic feed additives to promote animal health and production. Further exploration of this topic and the materials involved could provide a basis for new product development and innovation in animal nutrition. Full article
Show Figures

Figure 1

Figure 1
<p>Pathway of insulin-dependent glucose uptake and AMPK activation-induced translocation of GLUT4 vesicles. Pointed arrows represent activation or translocation in the signaling pathways. The “P” symbol with a circle represents a phosphorylation event. After insulin binds to the IR and triggers tyrosine autophosphorylation, the IRSs are activated to promote the PI3K/AKT pathway, eventually inducing the translocation of GLUT4 vesicles to the cell membranes and facilitating glucose uptake.</p>
Full article ">Figure 2
<p>Putative mechanisms by which ROS interferes with insulin-dependent glucose uptake. Pointed arrows represent activation or translocation in the signaling pathways. Line-headed arrows indicate inhibition of the signaling process. The “P” symbol with a circle represents a phosphorylation event. ROS directly inhibits the expression of PI3K/AKB and the translocation of GLUT4. ROS also causes protein misfolding, which can directly inflict IR. The cellular damage can induce the production of TNF-α, which inhibits the phosphorylation of IR. ER stress caused by ROS facilitates the ubiquitination of IRS via the JNK pathway, inhibiting the downstream signaling.</p>
Full article ">Figure 3
<p>Potential crosstalk between oxidative stress and inflammation through phytochemicals in poultry [<a href="#B75-pharmaceutics-16-01208" class="html-bibr">75</a>]. Copyright@2019, Animal Bioscience, Seoul, Republic of Korea.</p>
Full article ">Figure 4
<p>Mechanisms of phytochemicals to modulate glucose metabolism. Upward arrows (↑) and downward arrows (↓) represent the upregulation and downregulation of specific molecules, respectively. The antioxidant and anti-inflammatory effects of phytochemicals prevent the potential inhibition of insulin-dependent metabolism by oxidative stresses. Also, phytochemicals can directly stimulate the expressions of IRSs, PI3K/AKT, and AMPK, eventually facilitating GLUT4 translocation.</p>
Full article ">Figure 5
<p>Mechanisms of probiotics and prebiotics on the modulation of insulin-dependent glucose metabolism. Upward arrows (↑) and downward arrows (↓) represent the upregulation and downregulation of specific molecules, respectively. Probiotics and prebiotics can work independently or synergistically to modulate the intestinal microbiome and facilitate the production of SCFAs. The SCFAs then stimulate the GPR 41 and 43, promote pro-insulin GLP-1, and enhance host insulin secretion. Individually, prebiotics can directly promote the IRS/PI3K/AKT pathway and stimulate GLUT4 translocation. On the other hand, probiotics directly stimulate PI3K/AKT/GLUT4 with their metabolites, such as surfactin. Furthermore, some probiotics, after successful colonization in the intestinal environment, can exert their benefits through SCFA production.</p>
Full article ">Figure 6
<p>Mechanisms of Se and Cr<sup>3+</sup> on the modulation of insulin-dependent glucose metabolism. These two trace elements can reduce cellular ER stress via their antioxidant effects. Upward arrows (↑) and downward arrows (↓) represent the upregulation and downregulation of specific molecules, respectively. The reduced ER stress prevents IRSs from ubiquitination, which preserves the downstream signaling cascade of insulin-dependent glucose uptake. Se and Cr<sup>3+</sup> have been reported to promote the activation of serine and threonine kinases with subsequent IRS phosphorylation. Se and Cr<sup>3+</sup> have the ability to respectively stimulate GLP-1 for increased insulin production or activate AMPK-facilitated GLUT4 translocation.</p>
Full article ">Figure 7
<p>The importance of gut microbial metabolism in regulating insulin sensitivity in humans and mice [<a href="#B210-pharmaceutics-16-01208" class="html-bibr">210</a>]. Upward arrows (↑) and downward arrows (↓) represent the upregulation and downregulation of specific molecules, respectively. Question mark (?) indicates the hypothesized modulation effects. Copyright@2024, Nature, London, UK.</p>
Full article ">Figure 8
<p>Flow chart showing how functional non-antibiotic feed additives support animal health and preserve good product quality.</p>
Full article ">
15 pages, 3636 KiB  
Article
Effects of Kimchi Intake on the Gut Microbiota and Metabolite Profiles of High-Fat-Induced Obese Rats
by Dong-Wook Kim, Quynh-An Nguyen, Saoraya Chanmuang, Sang-Bong Lee, Bo-Min Kim, Hyeon-Jeong Lee, Gwang-Ju Jang and Hyun-Jin Kim
Nutrients 2024, 16(18), 3095; https://doi.org/10.3390/nu16183095 - 13 Sep 2024
Viewed by 290
Abstract
With rising global obesity rates, the demand for effective dietary strategies for obesity management has intensified. This study evaluated the potential of kimchi with various probiotics and bioactive compounds as a dietary intervention for high-fat diet (HFD)-induced obesity in rats. Through a comprehensive [...] Read more.
With rising global obesity rates, the demand for effective dietary strategies for obesity management has intensified. This study evaluated the potential of kimchi with various probiotics and bioactive compounds as a dietary intervention for high-fat diet (HFD)-induced obesity in rats. Through a comprehensive analysis incorporating global and targeted metabolomics, gut microbiota profiling, and biochemical markers, we investigated the effects of the 12-week kimchi intake on HFD-induced obesity. Kimchi intake modestly mitigated HFD-induced weight gain and remarkably altered gut microbiota composition, steroid hormones, bile acids, and metabolic profiles, but did not reduce adipose tissue accumulation. It also caused significant shifts in metabolomic pathways, including steroid hormone metabolism, and we found substantial interactions between dietary interventions and gut microbiota composition. Although more research is required to fully understand the anti-obesity effects of kimchi, our findings support the beneficial role of kimchi in managing obesity and related metabolic disorders. Full article
(This article belongs to the Section Prebiotics and Probiotics)
Show Figures

Figure 1

Figure 1
<p>Comparison of gut microbiota’s relative abundance after 12 weeks of kimchi intervention. (<b>A</b>) Bar charts showing the overall microbial composition at genus levels in feces from rats fed ND, HFD, and KHD diets, with the average relative abundance. (<b>B</b>) Chao1 and Shannon indices calculated after rarefying to an equal number of sequence reads. (<b>C</b>) Principal-coordinate analysis plots of weighted UniFrac distance dissimilarities (PC1 and PC2). (<b>D</b>) Relative abundances of bacteria at the genus level. ND, control; HFD; high-fat diet; KHD, high-fat-kimchi diet. * <span class="html-italic">p</span>-value &lt; 0.05, ** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt;0.001, # <span class="html-italic">p</span>-value &lt; 0.0001. Data represent the relative abundance of microbes analyzed from six samples. Different letters on the bar and box plot indicate significant differences in the <span class="html-italic">t</span>-test at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>(<b>A</b>) Metabolite analysis of rats fed HFD and KHD. Partial least-squares discriminant analysis (PLS-DA) score plot obtained from UPLC-Q-TOF MS data of plasma, urine, large intestine, liver, and kidney (<span class="html-italic">n</span> = 10). (<b>B</b>) Fold change of identified metabolites. Metabolites were analyzed using UPLC-Q-TOF MS via an Acquity BEH C18 column (2.1 mm × 100 mm, 1.7 μm) with a positive ESI mode. The qualification of PLS-DA models was evaluated by R2X, R2Y, Q2, and <span class="html-italic">p</span>-values. R2X and R2Y show the fitting quality of the models, while Q2 shows their prediction quality. ND, control; HFD; high-fat diet; KHD, high-fat-kimchi diet.</p>
Full article ">Figure 3
<p>Proposed metabolomic pathway associated with HFD and kimchi intake and the relative abundance of metabolites. Box plots present the relative abundance of metabolites analyzed by UPLC-Q-TOF MS, with significant differences determined by <span class="html-italic">t</span>-tests at <span class="html-italic">p</span>-values &lt; 0.05 (*), &lt;0.01 (**), &lt;0.001 (***), and &lt;0.0001 (#). N, normal-diet group; H, high-fat-diet group; K, kimchi-high-fat-diet group; P, plasma; L, liver; K, kidney; LI, large intestinal residues.</p>
Full article ">Figure 4
<p>Proposed bile acid and steroid hormone pathway associated with HFD and kimchi intake and their relative abundances. Box plots present the relative abundance of bile acids and steroid hormones analyzed by UPLC-Q-TOF MS, with significant differences determined using <span class="html-italic">t</span>-tests at <span class="html-italic">p</span>-values of &lt;0.05 (*), &lt; 0.01 (**), &lt;0.001 (***), and &lt;0.0001 (#). N, normal-diet group; H, high-fat-diet group; K, kimchi-high-fat-diet group.</p>
Full article ">Figure 5
<p>Analysis of correlations of gut microbiota with bile acid and steroid hormones (<b>A</b>), and gut microbiota with identified metabolites’ data (<b>B</b>). The correlation matrix was analyzed and visualized with a heat map. Positive correlations are shown in blue, and negative correlations are shown in red. A dark color means a stronger correlation.</p>
Full article ">
18 pages, 2294 KiB  
Article
Goat Cheese Produced with Sunflower (Helianthus annuus L.) Seed Extract and a Native Culture of Limosilactobacillus mucosae: Characterization and Probiotic Survival
by Dôrian Cordeiro Lima Júnior, Viviane Maria da Silva Quirino, Alícia Santos de Moura, Joyceana Oliveira Correia, João Ricardo Furtado, Isanna Menezes Florêncio, Márcia Maria Cândido da Silva, Hévila Oliveira Salles, Karina Maria Olbrich dos Santos, Antonio Silvio do Egito and Flávia Carolina Alonso Buriti
Foods 2024, 13(18), 2905; https://doi.org/10.3390/foods13182905 - 13 Sep 2024
Viewed by 273
Abstract
The microbiological and biochemical properties of a goat cheese produced using Helianthus annuus (sunflower) seed extract as a coagulant and the potentially probiotic autochthonous culture Limosilactobacillus mucosae CNPC007 were examined in comparison to a control cheese devoid of the autochthonous culture. Throughout a [...] Read more.
The microbiological and biochemical properties of a goat cheese produced using Helianthus annuus (sunflower) seed extract as a coagulant and the potentially probiotic autochthonous culture Limosilactobacillus mucosae CNPC007 were examined in comparison to a control cheese devoid of the autochthonous culture. Throughout a 60-day storage period at 6 ± 1 °C, lactobacilli maintained a count of above 8 log CFU/g. Additionally, its viability in cheeses subjected to the in vitro gastrointestinal conditions demonstrated improvement over this period. Specifically, the recovery of lactobacilli above 6 log CFU/g was observed in 16.66% of the samples in the first day, increasing to 66.66% at both 30 and 60 days. While total coliforms were detected in both cheese trials, this sanitary parameter exhibited a decline in L. mucosae cheeses during storage, falling below the method threshold (<3 MPN/g) at 60 days. This observation suggests a potential biopreservative effect exerted by this microorganism, likely attributed to the higher acidity of L. mucosae cheeses at that point (1.80 g/100 g), which was twice that of the control trial (0.97 g/100 g). Furthermore, distinct relative proportions of >30 kDa, 30–20 kDa, and <20 kDa proteins during storage was verified for L. mucosae and control cheeses. Consequently, either the H. annuus seed extract or the L. mucosae CNPC007 autochthonous culture influenced the biochemical properties of the cheese, particularly in terms of proteolysis. Moreover, L. mucosae CNPC007 acidification property resulted in a biopreservative effect throughout the storage period, indicating the potential as a promising source of probiotics for this product. Full article
(This article belongs to the Special Issue Probiotics: Selection, Cultivation, Evaluation and Application)
Show Figures

Figure 1

Figure 1
<p>Complete cheesemaking process of the cheese trials studied.</p>
Full article ">Figure 2
<p>Sodium dodecyl sulfate polyacrylamide gel electrophoresis of control and <span class="html-italic">L. mucosae</span> cheeses in the day of packing (day 1) and after 30 and 60 days of storage at 6 ± 1 °C. The SDS-PAGE pattern of standard mixture Sigma Marker (Wide Range 6500–200,000 Da, Sigma-Aldrich) is shown in lane 1. The pattern of control cheeses on days 1, 30, and 60 is shown in lanes 2, 4 and 6, respectively, while the pattern of <span class="html-italic">L. mucosae</span> cheeses is shown in lanes 3, 5, and 7 for the same sampling periods, respectively. αs-CN = αs-casein. β-CN = β-casein. para-κ-CN = para-κ-casein, * = highlight for the intense bands verified in <span class="html-italic">L. mucosae</span> cheeses.</p>
Full article ">Figure 3
<p>Densitometric analysis of bands obtained by SDS-PAGE for cheeses at 1, 30, and 60 days of storage of control (lanes 2, 4, and 6, respectively) and <span class="html-italic">L. mucosae</span> (lanes 3, 5, and 7, respectively) trials. The SDS-PAGE pattern of standard mixture Sigma Marker (Wide Range 6500–200,000 Da, Sigma-Aldrich) is labelled as “S”. * = highlight for the prominent peaks marked with a circle in the region below 6.5 kDa in control cheeses at 30 and 60 days.</p>
Full article ">Figure 4
<p>Relative intensity (%) of proteins obtained by SDS-PAGE and quantified by densitometric analysis using GelAnalyzer 23.1.1 in control and <span class="html-italic">L. mucosae</span> cheeses after 1, 30, and 60 days of storage.</p>
Full article ">
22 pages, 9427 KiB  
Article
The Indigenous Probiotic Lactococcus lactis PH3-05 Enhances the Growth, Digestive Physiology, and Gut Microbiota of the Tropical Gar (Atractosteus tropicus) Larvae
by Graciela María Pérez-Jiménez, Carina Shianya Alvarez-Villagomez, Marcel Martínez-Porchas, Estefanía Garibay-Valdez, César Antonio Sepúlveda-Quiroz, Otilio Méndez-Marín, Rafael Martínez-García, Ronald Jesús-Contreras, Carlos Alfonso Alvarez-González and Susana del Carmen De la Rosa-García
Animals 2024, 14(18), 2663; https://doi.org/10.3390/ani14182663 - 13 Sep 2024
Viewed by 221
Abstract
Probiotics in aquaculture hold promise for enhancing fish health and growth. Due to their increased specificity and affinity for their host, indigenous probiotics may offer isolated and potentially amplified benefits. This study investigated the effects of Lactococcus lactis PH3-05, previously isolated from adults [...] Read more.
Probiotics in aquaculture hold promise for enhancing fish health and growth. Due to their increased specificity and affinity for their host, indigenous probiotics may offer isolated and potentially amplified benefits. This study investigated the effects of Lactococcus lactis PH3-05, previously isolated from adults of tropical gar (Atractosteus tropicus), on the growth, survival, digestive enzyme activity, intestinal morphology, expression of barrier and immune genes, and intestinal microbiota composition in the larvae of tropical gar. Larvae were fed with live L. lactis PH3-05 concentrations of 104, 106, and 108 CFU/g for 15 days alongside a control diet without probiotics. Higher concentrations of L. lactis PH3-05 (106 and 108 CFU/g) positively influenced larval growth, increasing hepatocyte area and enterocyte height. The 106 CFU/g dose significantly enhanced survival (46%) and digestive enzyme activity. Notably, the 108 CFU/g dose stimulated increased expression of muc-2 and il-10 genes, suggesting enhanced mucosal barrier function and anti-inflammatory response. Although L. lactis PH3-05 did not significantly change the diversity, structure, or Phylum level composition of intestinal microbiota, which was constituted by Proteobacteria, Bacteroidota, Chloroflexi, and Firmicutes, an increase in Lactobacillus abundance was observed in fish fed with 106 CFU/g, suggesting enhanced probiotic colonization. These results demonstrate that administering L. lactis PH3-05 at 106 CFU/g promotes growth, survival, and digestive health in A. tropicus larvae, establishing it as a promising indigenous probiotic candidate for aquaculture applications. Full article
(This article belongs to the Special Issue Fish Nutrition, Physiology and Management)
Show Figures

Figure 1

Figure 1
<p>Representative images of the liver and digestive system of <span class="html-italic">A. tropicus</span> larvae treated with CD and 10<sup>6</sup> CFU/g of <span class="html-italic">L. lactis</span> PH3-05: (<b>a</b>) CD, (<b>b</b>) 10<sup>6</sup> CFU/g of <span class="html-italic">L. lactis</span> PH3-05. The liver images show melanomacrophagic centers (yellow arrow) and hepatocytes (circle). Images of the intestine display the height of enterocytes (yellow line) of <span class="html-italic">A. tropicus</span> larvae: (<b>c</b>) CD, (<b>d</b>) 10<sup>6</sup> CFU/g of <span class="html-italic">L. lactis</span> PH3-05.</p>
Full article ">Figure 2
<p>Relative expression levels of intestinal barrier function and immune system genes in <span class="html-italic">A. tropicus</span> larvae fed with <span class="html-italic">L. lactis</span> PH3-05 supplementation (10<sup>4</sup>, 10<sup>6,</sup> and 10<sup>8</sup> CFU/g) and the control diet. Values are mean ± SD. Data are presented as fold-change relative to control diet samples (set to 1). Significant differences between treatments are indicated by letters (<span class="html-italic">p</span> &lt; 0.05). (<b>a</b>) Mucus layer protein (<span class="html-italic">muc-2)</span>; (<b>b</b>) Tight junction protein (<span class="html-italic">zo-2</span>); (<b>c</b>) Pro-inflammatory cytokine (<span class="html-italic">il-</span>8); (<b>d</b>) Anti-inflammatory cytokine (<span class="html-italic">il-</span>10).</p>
Full article ">Figure 3
<p>Alpha diversity of gut microbiota in <span class="html-italic">A. tropicus</span> larvae treated with <span class="html-italic">L. lactis</span> PH3-05 supplementations (10<sup>4</sup>, 10<sup>6,</sup> and 10<sup>8</sup> CFU/g) and control diet: (<b>a</b>) Chao 1, (<b>b</b>) ACE, and (<b>c</b>) Shannon–Weaver indexes were calculated from the ASVs.</p>
Full article ">Figure 4
<p>A relative abundance of bacterial phyla is present in the intestinal microbiota of <span class="html-italic">A. tropicus</span> larvae fed with <span class="html-italic">L. lactis</span> PH3-05 (10<sup>4</sup>, 10<sup>6,</sup> and 10<sup>8</sup> CFU/g) and a control diet.</p>
Full article ">Figure 5
<p>Relative abundance of the bacterial genus is present in the intestinal microbiota of <span class="html-italic">A. tropicus</span> larvae fed with <span class="html-italic">L. lactis</span> (10<sup>4</sup>, 10<sup>6</sup>, and 10<sup>8</sup> CFU/g) and a control diet.</p>
Full article ">Figure 6
<p>Principal coordinate analysis (PCoA) based on beta diversity analyses with Bray–Curtis (<b>a</b>), Jaccard (<b>b</b>), Weighted Unifrac (<b>c</b>), and Unweighted Unifrac (<b>d</b>) indexes of gut bacterial profiles of <span class="html-italic">A. tropicus</span> larvae treated fed <span class="html-italic">L. lactis</span> PH3-05 (10<sup>4</sup>, 10<sup>6</sup>, and 10<sup>8</sup> CFU/g) and a control diet.</p>
Full article ">Figure 7
<p>Heat map of microbial functions in the digestive tract of <span class="html-italic">A. tropicus</span> larvae fed with <span class="html-italic">L. lactis</span> PH3-05 (10<sup>4</sup>, 10<sup>6</sup>, and 10<sup>8</sup> CFU/g) and a control diet. Predictions are based on level 3 functional annotations using the KEGG database.</p>
Full article ">
16 pages, 3191 KiB  
Review
Postbiotics: Mapping the Trend
by Veroniki Stelmach, George Stavrou, Ioannis Theodorou, Eleni Semertzidou, Georgios Tzikos, Alexandra-Eleftheria Menni, Anne Shrewsbury, Aris Ioannidis and Katerina Kotzampassi
Nutrients 2024, 16(18), 3077; https://doi.org/10.3390/nu16183077 - 12 Sep 2024
Viewed by 241
Abstract
Background: Since the consensus of ISAPP on the definition of the term “postbiotic” there has been an enthusiasm for publications in review form—their number being disproportionate to the primary research. The aim of this bibliometry is to analyze the bibliometric trends of this [...] Read more.
Background: Since the consensus of ISAPP on the definition of the term “postbiotic” there has been an enthusiasm for publications in review form—their number being disproportionate to the primary research. The aim of this bibliometry is to analyze the bibliometric trends of this newfound interest in the field. Methods: Search of the PubMed database for review articles on postbiotics, published between November 2021 and June 2024. Results: Analysis was performed on 92 review articles, the number corresponding to 2.9 reviews per month. China, Poland, Italy, Iran and India had the maximum productivity among the 32 countries involved; 21 articles were published in 13 journals with the highest impact factor, while 45 were in 16 journals with an IF between 4.0 and 4.9. The authors were mainly affiliated to universities with specialization in both basic research and technology, as well as food science. The top five publications regarding the citations received, published in Foods (2), EBioMedicine, Biomolecules, and Front. Nutr., have collected between 138 and 109 citations. Conclusions: The ever-growing number of reviews regarding postbiotics is perhaps disproportionate to the actual original research in the field. Further clinical trials would extend and deepen the subject and facilitate the drowning of more robust conclusions in relation to their effects. Full article
(This article belongs to the Section Prebiotics and Probiotics)
Show Figures

Figure 1

Figure 1
<p>Flowchart.</p>
Full article ">Figure 2
<p>Monthly publication rate per year.</p>
Full article ">Figure 3
<p>Authors contributing to more than two publications.</p>
Full article ">Figure 4
<p>Scientists contributing with more than one publication, stratified as first and/or last authors.</p>
Full article ">Figure 5
<p>Eleven countries of first authors, having at least two publications each.</p>
Full article ">Figure 6
<p>Number of articles per country, according to affiliation of the first and last authors, presenting only countries having a minimum of two publications.</p>
Full article ">Figure 7
<p>Country collaboration [one arrow represents one paper]: red frame for active countries; the direction of arrows indicates the cooperative country [blue frame].</p>
Full article ">Figure 8
<p>Affiliations of first and last authors.</p>
Full article ">Figure 9
<p>Specialization of first and last authors affiliated with universities.</p>
Full article ">Figure 10
<p>Thirteen journals with two or more publications.</p>
Full article ">Figure 11
<p>Thirteen journals with two or more publications [red open bars] and their Impact Factors [blue bars].</p>
Full article ">Figure 12
<p>Top five journals among the 92 analyzed, according to the Impact Factor.</p>
Full article ">Figure 13
<p>Distribution of journals based on their Impact Factor and the articles hosted.</p>
Full article ">
13 pages, 656 KiB  
Article
Influence of Diet on Bowel Function and Abdominal Symptoms in Children and Adolescents with Hirschsprung Disease—A Multinational Patient-Reported Outcome Survey
by Judith Lindert, Hannah Day, Marta de Andres Crespo, Eva Amerstorfer, Sabine Alexander, Manouk Backes, Carlotta de Filippo, Andrzej Golebiewski, Paola Midrio, Mazeena Mohideen, Anna Modrzyk, Anette Lemli, Roxana Rassouli-Kirchmeier, Marijke Pfaff-Jongman, Karolina Staszkiewicz, Lovisa Telborn, Pernilla Stenström, Karolin Holström, Martina Kohl, Joe Curry, Stavros Loukogeorgakis and Joseph R Davidsonadd Show full author list remove Hide full author list
Children 2024, 11(9), 1118; https://doi.org/10.3390/children11091118 - 12 Sep 2024
Viewed by 423
Abstract
Introduction: This study aimed to understand the influence of diet and nutrition items on gastrointestinal symptoms in patients with Hirschsprung Disease (HD). Method: An online questionnaire was created to obtain patient-reported outcomes using the multinational Holistic Care in Hirschsprung Disease Network. This was [...] Read more.
Introduction: This study aimed to understand the influence of diet and nutrition items on gastrointestinal symptoms in patients with Hirschsprung Disease (HD). Method: An online questionnaire was created to obtain patient-reported outcomes using the multinational Holistic Care in Hirschsprung Disease Network. This was distributed in Dutch, English, German, Italian, Polish, and Swedish via patient associations. Information on demographics, the extension of disease, current diet, and the influence of food ingredients on bowel function were obtained. Results: In total, 563 questionnaires were answered by parents or patients themselves. The length of the aganglionic segment was short in 33%, long in 45%, total colonic aganglionosis (TCA) in 11%, and involved the small intestine in 10%. Overall, 90% reported following a mixed diet, and 31% reported taking probiotics, with twice as many patients taking probiotics in the TCA group compared to standard HD. Mealtimes and behaviours around eating were affected by 61%, while 77% had established food items that worsened symptoms, and of these, 80% stated that they had worked these items out themselves. A high-fibre diet was followed by 24% and 18% a low-fibre diet. Symptoms were reported, particularly from dairy in 30%, fruits in 39%, pulses in 54%, and sugar in 48%. Conclusions: This first multinational survey on diet and bowel function in HD reports an association between certain dietary items with gastrointestinal symptoms. This study can support an improved understanding of the interaction between food items and bowel function in children with HD. We suggest a multidisciplinary approach to balance dietary exclusions and support adequate growth, preventing nutrition deficiencies and enhancing quality of life. Full article
Show Figures

Figure 1

Figure 1
<p>Overall onset of adverse bowel symptoms noted by patients and their families according to food items.</p>
Full article ">Figure 2
<p>Checklist role of dietician.</p>
Full article ">
17 pages, 3182 KiB  
Article
Lacticaseibacillus paracsei HY7207 Alleviates Hepatic Steatosis, Inflammation, and Liver Fibrosis in Mice with Non-Alcoholic Fatty Liver Disease
by Hyeon-Ji Kim, Hye-Jin Jeon, Dong-Gun Kim, Joo-Yun Kim, Jae-Jung Shim and Jae-Hwan Lee
Int. J. Mol. Sci. 2024, 25(18), 9870; https://doi.org/10.3390/ijms25189870 - 12 Sep 2024
Viewed by 320
Abstract
Non-alcoholic fatty acid disease (NAFLD) is caused by a build-up of fat in the liver, inducing local inflammation and fibrosis. We evaluated the effects of probiotic lactic acid-generating bacteria (LAB) derived from a traditional fermented beverage in a mouse model of NAFLD. The [...] Read more.
Non-alcoholic fatty acid disease (NAFLD) is caused by a build-up of fat in the liver, inducing local inflammation and fibrosis. We evaluated the effects of probiotic lactic acid-generating bacteria (LAB) derived from a traditional fermented beverage in a mouse model of NAFLD. The LAB isolated from this traditional Korean beverage were screened using the human hepatic cell line HepG2, and Lactocaseibacillus paracasei HY7207 (HY7207), which was the most effective inhibitor of fat accumulation, was selected for further study. HY7207 showed stable productivity in industrial-scale culture. Whole-genome sequencing of HY7207 revealed that the genome was 2.88 Mbp long, with 46.43% GC contents and 2778 predicted protein-coding DNA sequences (CDSs). HY7207 reduced the expression of lipogenesis and hepatic apoptosis-related genes in HepG2 cells treated with palmitic acid. Furthermore, the administration of 109 CFU/kg/day of HY7207 for 8 weeks to mice fed an NAFLD-inducing diet improved their physiologic and serum biochemical parameters and ameliorated their hepatic steatosis. In addition, HY7207 reduced the hepatic expression of genes important for lipogenesis (Srebp1c, Fasn, C/ebpa, Pparg, and Acaca), inflammation (Tnf, Il1b, and Ccl2), and fibrosis (Col1a1, Tgfb1, and Timp1). Finally, HY7207 affected the expression of the apoptosis-related genes Bax (encoding Bcl2 associated X, an apoptosis regulator) and Bcl2 (encoding B-cell lymphoma protein 2) in the liver. These data suggest that HY7207 consumption ameliorates NAFLD in mice through effects on liver steatosis, inflammation, fibrosis, and hepatic apoptosis. Thus, L. paracasei HY7207 may be suitable for use as a functional food supplement for patients with NAFLD. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Screening of 15 LAB strains for their ability to reduce lipid accumulation in PA-treated HepG2 cells. (<b>B</b>) Growth curve of HY7207 in industrial-scale culture. (<b>C</b>) Genome map of <span class="html-italic">Lactocaseibacilllus paracasei</span> HYY7207. From the outer circle to the inner circle, the genomic features are indicated as follows: genome size, gray color; forward strand and reverse strand CDSs, different colors, according to the COG classification; tRNA, blue line; rRNA, red line; GC skew, green and red peaks; GC ratio, blue and yellow peaks. Data are presented as the mean ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 vs. the control group, * <span class="html-italic">p</span> &lt; 0.05 vs. the PA group; LAB: lactic acid bacteria; PA: palmitic acid; CDS: protein-coding DNA sequence; COG: clusters of orthologous groups; G: guanine, C: cytosine.</p>
Full article ">Figure 2
<p>Effect of HY7207 on the expression of lipogenesis and apoptosis-related genes in PA-treated HepG2 cells. (<b>A</b>) <span class="html-italic">SREBP-1c</span>, (<b>B</b>) <span class="html-italic">FASN</span>, (<b>C</b>) <span class="html-italic">C</span>/<span class="html-italic">EBPα</span>, (<b>D</b>) <span class="html-italic">BAX</span>, (<b>E</b>) <span class="html-italic">BCL-2</span>, (<b>F</b>) <span class="html-italic">BAX</span>/<span class="html-italic">BCL-2</span> ratio (<b>G</b>) <span class="html-italic">CASP3</span>, and (<b>H</b>) <span class="html-italic">CASP9</span>. The results are presented as the mean ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the control group; ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. the PA group. PA: palmitic acid; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207; SREBP-1c: sterol regulatory element-binding protein 1; FASN: fatty acid synthase; C/EBPα: CCAAT/enhancer-binding protein alpha; BAX: BCL2-associated X, apoptosis regulator; BCL-2: BCL2, apoptosis regulator; CASP3: Caspase 3; CASP9: Caspase 9.</p>
Full article ">Figure 3
<p>Effect of HY7207 on mice fed an NAFLD-inducing diet. (<b>A</b>) Body mass changes, (<b>B</b>) body mass gain, (<b>C</b>) food efficiency ratio, (<b>D</b>) liver tissue mass, (<b>E</b>) liver/body mass ratio, (<b>F</b>) epididymal fat mass, and (<b>G</b>) epididymal fat/body mass ratio. The results are presented as the mean ± SD. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the CON group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. the ND group. CON: Untreated group; ND: NAFLD-inducing diet group; MFM: metformin; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207; FER: food efficiency ratio.</p>
Full article ">Figure 4
<p>Effect of HY7207 on the blood biochemistry of NAFLD-inducing diet-fed mice. (<b>A</b>) AST, (<b>B</b>) ALT, (<b>C</b>) ALP, and (<b>D</b>) γ-GPT activities; (<b>E</b>) TG, (<b>F</b>) T-CHO, (<b>G</b>) LDL-C, (<b>H</b>) HDL-C, and (<b>I</b>) GLU concentrations. The results are presented as the mean ± SD. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the CON group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. the ND group. CON: Untreated group; ND: NAFLD-inducing diet group; MFM: Metformin group; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207 group; AST: aspartate aminotransferase; ALT: alanine aminotransferase; ALP: alkaline phosphatase; γ-GTP: gamma-glutamyl transferase; TG: triglyceride; T-CHO: total cholesterol; LDL-C: low-density lipoprotein-cholesterol; HDL-C: high-density lipoprotein-cholesterol; GLU: glucose.</p>
Full article ">Figure 5
<p>Effect of HY7207 on the hepatic histology of NAFLD-inducing diet-fed mice. (<b>A</b>) Liver morphology, (<b>B</b>) histology of the liver (hematoxylin and eosin-stained sections; 100× magnification), and (<b>C</b>) steatosis grade of the mice. (<b>D</b>) NAS grade of the mice. Results are presented as the mean ± SD. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the CON group; * <span class="html-italic">p</span> &lt; 0.05 vs. the ND group. CON: Untreated group; ND: NAFLD-inducing diet-fed group; MFM: Metformin group; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207 group; NAS: NAFLD activity score.</p>
Full article ">Figure 6
<p>Effect of HY7207 on the hepatic gene expression of NAFLD-inducing diet-fed mice. The expression of genes related to (<b>A</b>) lipogenesis, (<b>B</b>) inflammation, (<b>C</b>) fibrosis, and (<b>D</b>) hepatic apoptosis is shown. Results are presented as the mean ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the CON group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. the ND group. CON: Untreated group; ND: NAFLD-inducing diet-fed group; MFM: Metformin group; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207 group.</p>
Full article ">Figure 7
<p>(<b>A</b>) Effects of HY7207 on the histology of the epididymal fat of NAFLD-inducing diet-fed mice. (H&amp;E staining; 100× magnification). Effect of HY7207 on the expression of lipogenesis related genes in NAFLD-inducing diet-fed mice. (<b>B</b>) <span class="html-italic">Srebp-1c</span>, (<b>C</b>) <span class="html-italic">C</span>/<span class="html-italic">ebpα</span>, (<b>D</b>) <span class="html-italic">Pparγ</span> and (<b>E</b>) <span class="html-italic">Acaca</span>. Results are presented as the mean ± SD. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. the CON group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. the ND group. CON: Untreated group; ND: NAFLD-inducing diet-fed group; MFM: Metformin group; HY7207: <span class="html-italic">Lacticaseibacillus paracasei</span> HY7207 group.</p>
Full article ">Figure 8
<p>A flow chart of animal experiments.</p>
Full article ">
18 pages, 2041 KiB  
Review
Revolutionizing Cosmetic Ingredients: Harnessing the Power of Antioxidants, Probiotics, Plant Extracts, and Peptides in Personal and Skin Care Products
by Hye Yung Choi, Yun Jung Lee, Chul Min Kim and Young-Mi Lee
Cosmetics 2024, 11(5), 157; https://doi.org/10.3390/cosmetics11050157 - 12 Sep 2024
Viewed by 638
Abstract
The burgeoning interest in natural components in personal care products has led to significant research and development of ingredients such as plant extracts, antioxidants, peptides, and probiotics. These components have been recognized for their potential to enhance skin health through various mechanisms, addressing [...] Read more.
The burgeoning interest in natural components in personal care products has led to significant research and development of ingredients such as plant extracts, antioxidants, peptides, and probiotics. These components have been recognized for their potential to enhance skin health through various mechanisms, addressing consumer demand for products that are both effective and benign. Plant extracts, known for their rich composition of bioactive compounds, offer a myriad of benefits including antioxidant, anti-inflammatory, and antimicrobial properties, making them invaluable in skin care formulations. Antioxidants, derived from both plants and other natural sources, play a pivotal role in protecting the skin from oxidative damage, thereby preventing premature aging and promoting skin vitality. Bioactive peptides have garnered attention owing to their multifunctional activities that include promoting collagen synthesis, inhibiting enzymes responsible for skin degradation, and reducing inflammation, thereby contributing to skin regeneration and anti-aging. Probiotics have expanded their utility beyond gut health to skin care, where they help in maintaining skin microbiome balance, thus enhancing skin barrier function and potentially mitigating various skin disorders. The purpose of this review is to explore the individual roles of plant extracts, antioxidants, peptides, and probiotics in personal care products, while emphasizing their synergistic effects when combined. By integrating these natural components, this paper aims to highlight the potential for developing innovative skincare formulations that not only address specific skin concerns but also contribute to overall skin health, aligning with the increasing consumer preference for natural and holistic skincare solutions. Full article
Show Figures

Figure 1

Figure 1
<p>Bioactive compounds in personal care products. Skin and hair care products contain a range of bioactive compounds including antioxidants, anti-inflammatories and neuromodulators. Some compounds are plant-specific (e.g., Nimbidin, Aloin and Curcumin) and others are available in many sources.</p>
Full article ">Figure 2
<p>Peptides used in skincare products.</p>
Full article ">Figure 3
<p>Molecular structure of antioxidant compounds used in personal care products.</p>
Full article ">
11 pages, 4485 KiB  
Article
Characterization of Exopolysaccharides Isolated from Donkey Milk and Its Biological Safety for Skincare Applications
by Chiara La Torre, Pierluigi Plastina, Diana Marisol Abrego-Guandique, Paolino Caputo, Cesare Oliviero Rossi, Giorgia Francesca Saraceno, Maria Cristina Caroleo, Erika Cione and Alessia Fazio
Polysaccharides 2024, 5(3), 493-503; https://doi.org/10.3390/polysaccharides5030031 - 12 Sep 2024
Viewed by 179
Abstract
Kefiran is a heteropolysaccharide that is considered a postbiotic and is obtained by kefir grains fermented in cow’s milk, while little is known about the donkey milk (DM) variety. Postbiotics are recognised as having important human health benefits that are very similar to [...] Read more.
Kefiran is a heteropolysaccharide that is considered a postbiotic and is obtained by kefir grains fermented in cow’s milk, while little is known about the donkey milk (DM) variety. Postbiotics are recognised as having important human health benefits that are very similar to probiotics but without the negative effects associated with their ingestion. Donkey is a monogastric animal, as are humans, and when used as an alternative food for infants who suffer from cow milk protein allergies, DM could therefore display more biocompatibility. In this study, the DM kefiran was extracted by ultrasound from kefir grains cultured in donkey milk and fully characterized for its structural and physicochemical properties by Fourier-transform infrared spectroscopy (FT-IR), High-Performance Liquid Chromatography- Refractive Index (HPLC-RI), Scanning electron microscope (SEM), Differential Scanning Calorimeters (DSC) and rheological analyses. In addition, tests were conducted on keratinocytes cell lines and human red blood cells to assess the nontoxicity and haemolysis degree of the polymer. The extraction yield of the DM kefiran was 6.5 ± 0.15%. The FT-IR analysis confirmed the structure of the polysaccharide by showing that the stretching of the C-O-C and C-O bonds in the ring, which formed two bands at 1157 and 1071 cm−1, respectively, and the anomeric band at 896 cm−1 indicates the β configuration and vibrational modes of glucose and galactose. Results were confirmed by HPLC-RI analysis indicating that the ratio glucose/galactose was 1:0.87. Furthermore, the SEM analysis showed a porous and homogeneous structure. The rheological analysis confirmed the pseudoplastic nature of the polymer, while the DSC analysis highlighted excellent thermal resistance (324 °C). Finally, DM kefiran was revealed to have biologically acceptable toxicity, showing a haemolytic activity of less than 2% when using fresh human red blood cells and showing no cytotoxicity on human keratinocytes. Therefore, kefiran obtained by DM shows an excellent biocompatibility, establishing it as a promising polymer for bioengineering human tissue for regenerative applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Fourier-transform infrared spectroscopy spectra of kefiran from DM.</p>
Full article ">Figure 2
<p>Intensity of refraction index signals (µV) versus retention times (min) after HPLC separation of monosaccharide standards (glucose 15.9 min, galactose 16.9 min, black and red lines respectively) and kefiran from DM before and after hydrolyzation (blue and pink line respectively).</p>
Full article ">Figure 3
<p>Differential scanning calorimetry spectra of exopolysaccharide from DM, showing three endothermic peaks at 112.4 ± 9.5, 280.6 ± 0.7 and 324.2 ± 3.19 °C.</p>
Full article ">Figure 4
<p>Rheology results of kefiran from DM at 0.1, 1, 10 and 100 s<sup>−1</sup>.</p>
Full article ">Figure 5
<p>Scanning electron microscopy was used to study the surface morphologies. The kefiran from donkey milk was analysed at three magnifications equal to 100×, 1000× and 10,000×, with scales of 1.0 mm, 100.0 µm and 10.0 µm, respectively.</p>
Full article ">Figure 6
<p>Cytotoxicity of kefiran polymers at three concentrations (50, 100 and 200 µg/mL) on the HaCat cell line.</p>
Full article ">
11 pages, 866 KiB  
Article
Probiotic Feed Additives Mitigate Odor Emission in Cattle Farms through Microbial Community Changes
by Min-Kyu Park, Tae-Kyung Hwang, Wanro Kim, YoungJae Jo, Yeong-Jun Park, Min-Chul Kim, HyunWoo Son, DaeWeon Seo and Jae-Ho Shin
Fermentation 2024, 10(9), 473; https://doi.org/10.3390/fermentation10090473 - 12 Sep 2024
Viewed by 199
Abstract
Odor emissions from animal manure present a significant environmental challenge in livestock farming, impacting air quality and farm sustainability. Traditional methods, such as chemical additives and manure treatment, can be costly, labor-intensive, and less eco-friendly. Therefore, this study investigated the effectiveness of microbial [...] Read more.
Odor emissions from animal manure present a significant environmental challenge in livestock farming, impacting air quality and farm sustainability. Traditional methods, such as chemical additives and manure treatment, can be costly, labor-intensive, and less eco-friendly. Therefore, this study investigated the effectiveness of microbial feed additives in reducing these odors. Conducted over three months in 2022 on a Korean beef cattle farm with 20 cattle, the experiment involved feeding a mixture of four microbial strains—Bacillus subtilis KNU-11, Lactobacillus acidophilus KNU-02, Lactobacillus casei KNU-12, and Saccharomyces cerevisiae KNU-06. Manure samples were collected from an experimental group (n = 9) and a control group (n = 11), with microbial community changes assessed through 16S ribosomal RNA gene amplicon sequencing. The results demonstrated significant reductions in specific odorous compounds in the experimental group compared to the control group: ammonia decreased by 64.1%, dimethyl sulfide by 81.3%, butyric acid by 84.6%, and isovaleric acid by 49.8%. Additionally, there was a notable shift in the microbiome, with an increase in the relative abundance of Ruminococcaceae and Prevotellaceae microbes associated with fiver degradation and fermentation, while the control group had higher levels of Bacteroidota and Spirochaetota, which are linked to pathogenicity. This study demonstrates that probiotics effectively alter intestinal microbiota to enhance microorganisms associated with odor mitigation, offering a promising and more sustainable approach to reducing odor emissions in livestock farming. Full article
(This article belongs to the Special Issue Bioconversion of Agricultural Wastes into High-Nutrition Animal Feed)
Show Figures

Figure 1

Figure 1
<p>Comparison of beta diversity between experimental and control groups. The plots illustrate the microbial community diversity for each group across the experimental period: (<b>A</b>) September, (<b>B</b>) October, and (<b>C</b>) November. <span class="html-italic">p</span>-values, calculated using PERMANOVA and ANOSIM, are displayed in the upper right corner of each plot.</p>
Full article ">Figure 2
<p>Biomarkers identified using LEfSe. Red boxes indicate a high abundance of the microorganisms in the corresponding group, while blue boxes represent a lower abundance or relative absence in that group.</p>
Full article ">
Back to TopTop