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

Previous Issue
Volume 11, November
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Vet. Sci., Volume 11, Issue 12 (December 2024) – 31 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
15 pages, 2380 KiB  
Article
Effect of a Broiler-Specific Light Spectrum on Growth Performance and Adrenocortical Activity in Chickens: A Pilot Study on a Commercial Farm
by Livio Galosi, Luca Todini, Laura Menchetti, Annaïs Carbajal, Rupert Palme, Nicola Ruggiero, Roberto Falconi and Alessandra Roncarati
Vet. Sci. 2024, 11(12), 618; https://doi.org/10.3390/vetsci11120618 (registering DOI) - 2 Dec 2024
Abstract
This study evaluated the effect of a broiler-specific light spectrum on productive performance corticosterone (fCC) and androgen dehydroepiandrosterone (fDHEA) concentrations in feathers, and glucocorticoid (GCMs) and androgen (AMs) metabolites in droppings of broilers. Two groups of female Ross 308 broilers were reared under [...] Read more.
This study evaluated the effect of a broiler-specific light spectrum on productive performance corticosterone (fCC) and androgen dehydroepiandrosterone (fDHEA) concentrations in feathers, and glucocorticoid (GCMs) and androgen (AMs) metabolites in droppings of broilers. Two groups of female Ross 308 broilers were reared under white LED (WL, n = 9000) and broiler-specific LED (BSL, n = 9000) lights. The body weight (BW) of 150 randomly selected animals/groups was measured weekly. Droppings and feathers were collected at the end of the cycle (29 days) from 20 animals/group. The BSL group showed higher final BW than WL (1407 ± 11 vs. 1341 ± 15 g, respectively; p < 0.001) and higher indices of uniformity (76.8% vs. 61.2% animals in the 10% around the mean, respectively; p < 0.001). No difference between groups was found in fCC and fDHEA concentrations or in the fCC–fDHEA, indicating similar long-term HPA axis activity during the cycle. A higher concentration of GCMs was found in the BSL group, indicating higher glucocorticoid secretion before sampling, with neither a difference in AMs nor in GCMs–AMs. Finally, there was a positive correlation between fCC and fDHEA and between GCMs and AMs (p < 0.01). Our findings suggest that the use of broiler-specific light improved the productivity performances of chickens without long-term consequences on HPA activation. However, the results of this pilot study in a commercial farm setting must be interpreted with caution and need confirmation. Full article
(This article belongs to the Section Veterinary Physiology, Pharmacology, and Toxicology)
Show Figures

Figure 1

Figure 1
<p>Poultry houses used in this trial. (<b>A</b>). Poultry house WL, equipped with compact fluorescent white light. (<b>B</b>). Poultry house BSL equipped with broiler-specific light (NatureDynamics System, ONCE by Signify, Philips).</p>
Full article ">Figure 2
<p>Body weight (BW) changes in the white LED (WL) and broiler-specific LED (BSL) groups. Values are means ± standard errors. The asterisks indicate significant differences between the two groups for each time point (*** <span class="html-italic">p</span> &lt; 0.001; n.s. not significant). T1 = 8 days; T2 = 15 days; T3 = 22 days; T4 = 29 days.</p>
Full article ">Figure 3
<p>Corticosterone (<b>A</b>) and DHEA (<b>B</b>) in feathers in the white LED (WL) and broiler-specific LED (BSL) groups. Values are means ± standard errors (n.s. = not significant). DHEA = dehydroepiandrosterone.</p>
Full article ">Figure 4
<p>Pyramid chart and test statistics for the Mann–Whitney test comparing the fCC/fDHEA (expressed as pg/mm) distribution of white LED (WL) and broiler-specific LED (BSL) groups. Numbers on the x-axis indicate the frequency of the number of animals, while numbers on the Y-axis indicate the value of the fCC–fDHEA ratio. fCC = corticosterone concentrations in feathers; fDHEA = dehydroepiandrosterone concentrations in feathers.</p>
Full article ">Figure 5
<p>Glucocorticoid (<b>A</b>) and androgen (<b>B</b>) metabolites in droppings in the white LED (WL) and broiler-specific LED (BSL) groups. Values are means ± standard errors. The asterisk indicates significant differences between the two groups (* <span class="html-italic">p</span> &lt; 0.05; n.s. = not significant).</p>
Full article ">Figure 6
<p>Pyramid chart and test statistics for the Mann–Whitney test comparing the distribution of glucocorticoid/androgen metabolites ratio (GCMs/AMs) in the droppings of white LED (WL) and broiler-specific LED (BSL) groups. Numbers on the x-axis indicate the frequency of the number of animals, while numbers on the Y-axis indicate the value of the ratio.</p>
Full article ">
12 pages, 3229 KiB  
Article
Genomic Differences and Mutations in Epidemic Orf Virus and Vaccine Strains: Implications for Improving Orf Virus Vaccines
by Dengshuai Zhao, Yaoxu Shi, Miaomiao Zhang, Ping Li, Yuanhang Zhang, Tianyu Wang, Dixi Yu and Keshan Zhang
Vet. Sci. 2024, 11(12), 617; https://doi.org/10.3390/vetsci11120617 (registering DOI) - 2 Dec 2024
Viewed by 16
Abstract
Orf (ORF) is an acute disease caused by the Orf virus (ORFV), and poses a certain threat to animal and human health. Live attenuated vaccines play an important role in the prevention and control of ORF. The effectiveness of the live attenuated Orf [...] Read more.
Orf (ORF) is an acute disease caused by the Orf virus (ORFV), and poses a certain threat to animal and human health. Live attenuated vaccines play an important role in the prevention and control of ORF. The effectiveness of the live attenuated Orf virus vaccine is influenced by several factors, including the genomic match between the vaccine strain and circulating epidemic strains. Genomic differences between an ORFV epidemic strain (ORFV-2W) and a vaccine strain (ORFV-1V) were identified in this study via analysis of multiple sequence alignments, phylogenetic trees, and single nucleotide polymorphisms. Phylogenetic analysis revealed that ORFV-2W and ORFV-1V were closely related, with a whole genome homology of 99.8%. Furthermore, a deletion in the non-coding region at the end of the whole genome of ORFV-1V was detected. Such non-essential genes in the terminal regions are usually unnecessary for virus replication but may play important roles in pathogenicity, host and tissue tropism. Single nucleotide polymorphism analysis identified three missense mutations in ORF067, ORF072, and the terminal non-coding region of ORFV-1V. Moreover, a frameshift mutation in ORF102 of ORFV-1V was detected. Mutations in individual genes and deletion of terminal non-coding regions may be related to the attenuation of the vaccine strain. These results provide useful context for improving ORFV vaccines. Full article
Show Figures

Figure 1

Figure 1
<p>Whole genome maps of ORFV-1V and ORFV-2W. (<b>A</b>) Whole genome map of ORFV-1V. (<b>B</b>) Whole genome map of ORFV-2W.</p>
Full article ">Figure 2
<p>Whole genome nucleotide sequence alignment analysis. The deleted region of ORFV-1V is represented by the red box.</p>
Full article ">Figure 3
<p>Alignment of ITR sequences from the ORFV-1V and ORFV-2W genomes with those from other isolates. The left terminal 5’-ITR sequences of eleven ORFV isolates were compared using MEGA version 11. The BamHI terminal region (GGATCC) and the telomere resolution sequence (ATT TTTT-N(8)-TAAAT) are indicated as yellow and blue boxes, respectively. The deletion region of ORFV-1V is indicated as a black box.</p>
Full article ">Figure 4
<p>Phylogenetic analysis based on two genes (i.e., ORF011 and ORF059) and whole genome sequence data. Phylogenetic trees were constructed using the neighbor-joining method in MEGA version 11. Numbers above or below branch points indicate the bootstrap support calculated for 1000 replicates. Shown are trees using sequence data for: (<b>A</b>) B2L (<span class="html-italic">ORF011</span>) and (<b>B</b>) F1L (<span class="html-italic">ORF059</span>). Here, ORFV-1V and ORFV-2W are represented by red triangles and red squares, respectively. Also shown is a tree using (<b>C</b>) whole genome sequence data. ORFV-1V and ORFV-2W are shown in red font.</p>
Full article ">Figure 5
<p>Alignment of amino acid sequences of <span class="html-italic">ORF067</span>, <span class="html-italic">ORF072</span>, and <span class="html-italic">ORF102</span>. The sequences of ORFV-1V and other isolates were aligned using MEGA version 11. (<b>A</b>) <span class="html-italic">ORF067</span>. (<b>B</b>) <span class="html-italic">ORF072</span>. (<b>C</b>) <span class="html-italic">ORF102</span>. ORFV-1V is highlighted with a red box for emphasis. Amino acid mutation sites are represented by yellow boxes.</p>
Full article ">Figure 6
<p>Prediction of the tertiary structures of proteins encoded by <span class="html-italic">ORF067</span>, <span class="html-italic">ORF072</span>, and <span class="html-italic">ORF102</span> of the ORFV-1V and ORFV-2W genomes. Tertiary protein structures were predicted using SWISS-MODEL. The template coverage of each of the above proteins is &gt;50%, and all GMQE &gt; 0.2, which indicates that the prediction results are reliable. (<b>A</b>) ORFV-1V <span class="html-italic">ORF067</span>. (<b>B</b>) ORFV-2W <span class="html-italic">ORF067</span>. (<b>C</b>) ORFV-1V <span class="html-italic">ORF072</span>. (<b>D</b>) ORFV-2W <span class="html-italic">ORF072</span>. (<b>E</b>) ORFV-1V <span class="html-italic">ORF102</span>. (<b>F</b>) ORFV-2W <span class="html-italic">ORF102</span>. Amino acid sites are indicated by black arrows. D196H indicates the mutation of amino acid 196 from aspartic acid (Asp) to histidine (His). I113V shows the mutation of isoleucine (Ile) to valine (Val) at position 113. N148 shows the mutation of amino acid 148 to asparagine (Asn).</p>
Full article ">
36 pages, 927 KiB  
Review
Applications of Next-Generation Sequencing Technologies and Statistical Tools in Identifying Pathways and Biomarkers for Heat Tolerance in Livestock
by Gajendirane Kalaignazhal, Veerasamy Sejian, Silpa Mullakkalparambil Velayudhan, Chinmoy Mishra, Ebenezer Binuni Rebez, Surinder Singh Chauhan, Kristy DiGiacomo, Nicola Lacetera and Frank Rowland Dunshea
Vet. Sci. 2024, 11(12), 616; https://doi.org/10.3390/vetsci11120616 (registering DOI) - 2 Dec 2024
Viewed by 56
Abstract
The climate change-associated abnormal weather patterns negatively influences the productivity and performance of farm animals. Heat stress is the major detrimental factor hampering production, causing substantial economic loss to the livestock industry. Therefore, it is important to identify heat-tolerant breeds that can survive [...] Read more.
The climate change-associated abnormal weather patterns negatively influences the productivity and performance of farm animals. Heat stress is the major detrimental factor hampering production, causing substantial economic loss to the livestock industry. Therefore, it is important to identify heat-tolerant breeds that can survive and produce optimally in any given environment. To achieve this goal, a clearer understanding of the genetic differences and the underlying molecular mechanisms associated with climate change impacts and heat tolerance are a prerequisite. Adopting next-generation biotechnological and statistical tools like whole transcriptome analysis, whole metagenome sequencing, bisulphite sequencing, genome-wide association studies (GWAS), and selection signatures provides an opportunity to achieve this goal. Through these techniques, it is possible to identify permanent genetic markers for heat tolerance, and by incorporating those markers in marker-assisted breeding selection, it is possible to achieve the target of breeding for heat tolerance in livestock. This review gives an overview of the recent advancements in assessing heat tolerance in livestock using such ‘omics’ approaches and statistical models. The salient findings from this research highlighted several candidate biomarkers that have the potential to be incorporated into future heat-tolerance studies. Such approaches could revolutionise livestock production in the changing climate scenario and support the food demands of the growing human population. Full article
Show Figures

Figure 1

Figure 1
<p>Applications of biotechnological and statistical tools to establish climate-resilient livestock production.</p>
Full article ">
15 pages, 6448 KiB  
Article
Hermetia illucens Larvae Meal Enhances Immune Response by Improving Serum Immunoglobulin, Intestinal Barrier and Gut Microbiota of Sichuan White Geese After Avian Influenza Vaccination
by Yufei Xie, Yongfeng Hao, Fuxing Gui, Xifeng Li, Huan Huang, Pingrui Yang, Chonghua Zhong and Liting Cao
Vet. Sci. 2024, 11(12), 615; https://doi.org/10.3390/vetsci11120615 (registering DOI) - 2 Dec 2024
Viewed by 40
Abstract
Hermetia illucens Larvae Meal (HILM) has been observed to enhance growth performance and immune function, yet the effects and mechanisms in geese remain less understood. Experiment I included 64 Sichuan White Geese to investigate the optimal additive amount of HILM in diet, and [...] Read more.
Hermetia illucens Larvae Meal (HILM) has been observed to enhance growth performance and immune function, yet the effects and mechanisms in geese remain less understood. Experiment I included 64 Sichuan White Geese to investigate the optimal additive amount of HILM in diet, and experiment II included 32 Sichuan White Geese to access serum immunoglobulin, spleen immune-related genes, intestinal morphology and gut microbiota at the optimal additive amount of HILM. The results showed that the addition of 1% HILM significantly increased the ADG of Sichuan White Geese (p < 0.05), serum H5-R14 and H7-R4 strain titer at 33 d (p < 0.01) and H5-R13 strain titer (p < 0.05) at 40 d, which is the optimal dose of this trial. Experiment II revealed that the 1% HILM significantly increased serum IgG, IgG1, IgG2a, IgG3 and complement C3 (p < 0.05) and the mRNA expressions of IL-6 (p < 0.05) and CD4 (p < 0.01) in the spleen. The intestinal morphology was improved, and the secretion of SIgA and mRNA expression of Occludin in the jejunum were significantly increased (p < 0.05). Additionally, the abundance of Campilobacterota, Barnesiellaceae and Barnesiella was significantly decreased (p < 0.05), while the abundance of Lactobacillaceae was significantly increased (p < 0.05). This research provides new insights into the use of HILM in geese production. Full article
(This article belongs to the Section Nutritional and Metabolic Diseases in Veterinary Medicine)
Show Figures

Figure 1

Figure 1
<p>Effect of HILM on H5-R13 titer (<b>a</b>), H5-R14 titer (<b>b</b>) and H7-R4 titer (<b>c</b>) of AIV specific antibody in Sichuan White Geese. * indicates that there is a significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.05), and ** indicates an extremely significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Effect of HILM on AIV-specific antibody blocking rate of Sichuan White Geese.</p>
Full article ">Figure 3
<p>Effect of HILM on serum IgG (<b>a</b>), IgG1 (<b>b</b>), IgG2a (<b>c</b>), IgG2b (<b>d</b>), IgG3 (<b>e</b>), complement C3 (<b>f</b>), complement C4 (<b>g</b>), and SIgA in jejunum (<b>h</b>) of Sichuan White Geese. * indicates that there is a significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.05), *** and ** indicates an extremely significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Effect of HILM on intestinal morphology of Sichuan White Geese by HE staining.</p>
Full article ">Figure 5
<p>Effect of HILM on intestinal morphology (VH, CD and VH/CD) of duodenum (<b>a</b>–<b>c</b>), jejunum (<b>d</b>–<b>f</b>) and ileum (<b>g</b>–<b>i</b>) in Sichuan White Geese. * indicates that there is a significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of HILM on jejunum (<b>a</b>) and spleen gene mRNA expression (<b>b</b>) of Sichuan White Geese. * indicates that there is a significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.05), and ** indicates an extremely significant difference compared with the control group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>Diversity analysis of HILM impacts on Sichuan White Geese. (<b>a</b>) Venn diagram; (<b>b</b>–<b>e</b>) alpha diversity index as accessed by Chao1, Simpson and Shannon indices; (<b>f</b>) beta diversity as accessed by principal coordinate analysis (PCoA).</p>
Full article ">Figure 8
<p>Effects of HILM on microflora in cecal contents. (<b>a</b>–<b>c</b>) The relative abundance of top 20 bacteria at the phylum, family and genus levels; (<b>d</b>) LDA score of LEfSe analysis; (<b>e</b>) evolutionary branching diagram of LEfSe analysis (Control group was marked in Red and HILM group in Green).</p>
Full article ">
13 pages, 907 KiB  
Article
Effects of Perch on Productivity, Welfare, and Physiological Indicators of Broiler Chickens Reared in Animal Welfare-Certificated Farms
by Byung-Yeon Kwon, Seong-Taek Kim, Da-Hye Kim, Jina Park, Hyun-Gwan Lee, Yong-Sung Jeon, Ju-Young Song, Sang-Ho Kim, Dong-Wook Kim, Chan-Ho Kim and Kyung-Woo Lee
Vet. Sci. 2024, 11(12), 614; https://doi.org/10.3390/vetsci11120614 (registering DOI) - 2 Dec 2024
Viewed by 124
Abstract
This study explored the impact of perches on the productivity and welfare of broilers raised on two animal welfare-certified farms (designated as Farm A and B) in South Korea. Broiler houses in each farm were provided with or without wooden square-shaped perches (2 [...] Read more.
This study explored the impact of perches on the productivity and welfare of broilers raised on two animal welfare-certified farms (designated as Farm A and B) in South Korea. Broiler houses in each farm were provided with or without wooden square-shaped perches (2 × 2 cm) at a rate of 2 m per 1000 birds. The study aimed to assess whether perches could influence productivity measures, such as weight and uniformity, and animal welfare indicators, including corticosterone levels and physical health markers. The findings showed that the effects on productivity were inconsistent, varying by farm and period. Corticosterone levels, as an indicator of stress, were significantly lower in the perch group on farm B, but not on farm A. There were no significant differences in welfare indicators such as footpad dermatitis or feather cleanliness, although gait scores improved in farm B with perch provision. Litter moisture was higher in the perch group of farm A, but showed no difference in farm B. The study concluded that while perches did not consistently improve productivity or welfare, they did help reduce stress in broilers, as indicated by lower corticosterone levels. Full article
Show Figures

Figure 1

Figure 1
<p>The figure of perch used in the experiment.</p>
Full article ">Figure 2
<p>The sampling locations of parameters and the perch placement in houses. The symbol equal (=) represents the perch, while the black diamonds and the black circles denote sampling locations. The black diamond symbols indicate locations for body weight measurement, and the black circle symbols indicate locations for monitoring feathers and fecal corticosterone, footpad dermatitis, hock burn, feather cleanliness, litter, and body surface temperature.</p>
Full article ">
11 pages, 2671 KiB  
Article
Comparative Analysis of Cryotherapy Modalities Using Muscle Tissue Temperature Measurement: Cold Pack, Cold Compression, and Hyperbaric Gaseous Cryotherapy
by Jinyeob Baek, Jaeeon Cheon, Hyeonseo Lim, Yong Yu and Suyoung Heo
Vet. Sci. 2024, 11(12), 613; https://doi.org/10.3390/vetsci11120613 (registering DOI) - 1 Dec 2024
Viewed by 304
Abstract
Cryotherapy is widely used in veterinary medicine to manage pain and swelling, yet optimal methods for specific tissue depths remain unclear. Cryotherapy modalities, including cold packs (CPs), cold compression (CC), and hyperbaric gaseous cryotherapy (HGC), were applied to nine beagle dogs under general [...] Read more.
Cryotherapy is widely used in veterinary medicine to manage pain and swelling, yet optimal methods for specific tissue depths remain unclear. Cryotherapy modalities, including cold packs (CPs), cold compression (CC), and hyperbaric gaseous cryotherapy (HGC), were applied to nine beagle dogs under general anaesthesia. A custom 3D-printed guide facilitated consistent and accurate measurements of tissue temperatures at depths of 1 and 3 cm. After a 20 min application, all modalities significantly reduced the muscular tissue temperatures at both measured depths. HGC exhibited the most effective rapid temperature reduction, whereas CC displayed the most extended sustained cooling effect. In comparison, CPs showed a lower temperature reduction. The effectiveness of cryotherapy varies with tissue depth, and selecting the appropriate method can improve therapeutic outcome. Full article
(This article belongs to the Section Veterinary Surgery)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Blueprint for 3D-printed guides capable of measuring 1cm depth. (<b>B</b>) Three-dimensional-printed guides for inserting the Type K thermocouple probe temperature sensor.</p>
Full article ">Figure 2
<p>Three-dimensional-printed guides and Type K thermocouple probe temperature sensors placed in each hindlimb.</p>
Full article ">Figure 3
<p>Applying a plastic film to minimise the impact of the modality on the opposite hindlimb. Cold compression on the right hindlimb and hyperbaric gas cryotherapy on the left hindlimb.</p>
Full article ">Figure 4
<p>Mean temperature changes over time at a depth of 1 cm during modality application. The X-axis represents time (minutes) and the Y-axis represents temperature (°C).</p>
Full article ">Figure 5
<p>Mean temperature changes over time at a depth of 3 cm during modality application. The X-axis represents time (minutes) and the Y-axis represents temperature (°C).</p>
Full article ">
13 pages, 256 KiB  
Article
Comparison of Ketamine/Diazepam and Tiletamine/Zolazepam Combinations for Anaesthesia Induction in Horses Undergoing Partial Intravenous Anaesthesia (PIVA): A Retrospective Clinical Study
by Carlotta Lambertini, Elena Boanini, Isabelle Casalini, Francesca Spaccini, Riccardo Rinnovati and Noemi Romagnoli
Vet. Sci. 2024, 11(12), 612; https://doi.org/10.3390/vetsci11120612 (registering DOI) - 30 Nov 2024
Viewed by 366
Abstract
The aim of this retrospective clinical study was to compare the combinations of ketamine/diazepam (KD group) and tiletamine/zolazepam (TZ group) for the induction of general anaesthesia in horses undergoing elective surgery. The data from the clinical and the anaesthetic records of 138 horses [...] Read more.
The aim of this retrospective clinical study was to compare the combinations of ketamine/diazepam (KD group) and tiletamine/zolazepam (TZ group) for the induction of general anaesthesia in horses undergoing elective surgery. The data from the clinical and the anaesthetic records of 138 horses from 2021 to 2023 were evaluated, and the horses were divided in two groups: KD (n = 60) and TZ (n = 72). The horses were premedicated with romifidine and methadone IV; anaesthesia was induced with ketamine/diazepam for the KD group and tiletamine/zolazepam for the TZ group and was maintained with isoflurane and a constant rate infusion of romifidine. The data encompassed sex and neuter status, age, breed, weight, American Society of Anaesthesiologists physical status, type of surgical procedure performed under anaesthesia, induction time, induction score, surgery time, recovery time, and the recovery score using a descriptive scale. Baseline heart rate (HR), intraoperative HR, baseline respiratory rate (fR), intraoperative fR, mean arterial pressure (MAP), oxygen saturation (SpO2), and fraction of expired isoflurane (FE’Iso) were also recorded. The induction time was significantly longer (p = 0.004) in the TZ group (60 (40–120)) as compared to the KD group (50 (30–120)). Recovery time was also significantly longer (p ≤ 0.001) in the TZ group (46.5 (15–125)) as compared to the KD group (30 (5–105)). These findings suggested that, in adult horses undergoing elective surgery, TZ could be considered a valid alternative to KD for the induction of general anaesthesia. Additional experimental studies comparing the two induction regimens and their pharmacokinetic and pharmacodynamic characteristics are needed. Full article
(This article belongs to the Special Issue Anesthesia and Pain Management in Large Animals)
14 pages, 4375 KiB  
Article
Study of the Normal Crested Porcupine (Hystrix cristata) Nasal Cavity and Paranasal Sinuses by Cross-Sectional Anatomy and Computed Tomography
by Daniel Morales Bordon, Francisco Suárez-Cabrera, Gregorio Ramírez, Pablo Paz-Oliva, Alejandro Morales-Espino, Alberto Arencibia, Mario Encinoso, Myriam R. Ventura and José Raduan Jaber
Vet. Sci. 2024, 11(12), 611; https://doi.org/10.3390/vetsci11120611 (registering DOI) - 29 Nov 2024
Viewed by 195
Abstract
This study utilized CT imaging to investigate the rostral part of the head of the crested porcupine’s head. By combining CT images with anatomical cross-sections, we have provided a detailed description of the structures in this area. This information could be useful for [...] Read more.
This study utilized CT imaging to investigate the rostral part of the head of the crested porcupine’s head. By combining CT images with anatomical cross-sections, we have provided a detailed description of the structures in this area. This information could be useful for diagnosing disorders and improving their treatment in the nasal cavity and paranasal sinuses of crested porcupines. Full article
Show Figures

Figure 1

Figure 1
<p>Parasagittal MPR CT image of the crested porcupine head depicting the approximate anatomical levels of the sections of the porcupine head. Sections I–XI correspond to <a href="#vetsci-11-00611-f002" class="html-fig">Figure 2</a>, <a href="#vetsci-11-00611-f003" class="html-fig">Figure 3</a>, <a href="#vetsci-11-00611-f004" class="html-fig">Figure 4</a>, <a href="#vetsci-11-00611-f005" class="html-fig">Figure 5</a>, <a href="#vetsci-11-00611-f006" class="html-fig">Figure 6</a>, <a href="#vetsci-11-00611-f007" class="html-fig">Figure 7</a>, <a href="#vetsci-11-00611-f008" class="html-fig">Figure 8</a>, <a href="#vetsci-11-00611-f009" class="html-fig">Figure 9</a>, <a href="#vetsci-11-00611-f010" class="html-fig">Figure 10</a>, <a href="#vetsci-11-00611-f011" class="html-fig">Figure 11</a> and <a href="#vetsci-11-00611-f012" class="html-fig">Figure 12</a>.</p>
Full article ">Figure 2
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the nose, corresponding to line I in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Af: alar fold. Ag: alar groove. An: tip of the nose. Aw: alar wing. Dlnc: dorsal lateral nasal cartilage. Dn: dorsum of nose. Nos: nostrils. Phi: philtrum. Rn: root of nose. Ui: upper incisive tooth.</p>
Full article ">Figure 3
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the nasal vestibule, corresponding to line II in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Af: alar fold. Ag: alar groove. Aw: alar wing. Bf: basal fold (medial part). Bf’: basal fold (lateral part). Dlnc: dorsal lateral nasal cartilage. Hf: hair follicle. I: incisive bone. Lac: lateral accessory nasal cartilage. Ns: nasal septum. Nv: nasal vestibule. Pf: parallel fold. Sf: straight fold. Ui: upper incisive tooth. UiR: upper incisive tooth root. Vlnc: ventral lateral nasal cartilage.</p>
Full article ">Figure 4
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the basal folds, corresponding to line III in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Af: alar fold. Bf: basal fold (medial part). Bf′: basal fold (lateral part). Dlnc: dorsal lateral nasal cartilage. Hf: hair follicle. I: incisive bone. Li: lower incisive tooth. N: nasal bone. Ns: nasal septum. Nv: nasal vestibule. Oc: oral cavity. Pf: parallel fold. Pp: palatine plexus. UiR: upper incisive tooth root. Vlnc: ventral lateral nasal cartilage.</p>
Full article ">Figure 5
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the dorsal nasal concha, corresponding to line IV in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Dnc: dorsal nasal concha. Hf: hair follicle. I: incisive bone. Li: lower incisive tooth. Ll: lower lip. Mx: maxilla. N: nasal bone. Ns: nasal septum. Oc: oral cavity. Pf: parallel fold. Pp: palatine plexus. T: tongue. Ul: upper lip. UiR: upper incisive tooth root. V: vomer. Vn: vomeronasal organ. Vnc: ventral nasal concha.</p>
Full article ">Figure 6
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the dorsal and ventral nasal conchae, corresponding to line V in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Cnm: common nasal meatus. Dnc: dorsal nasal concha. Dnm: dorsal nasal meatus. Hf: hair follicle. I: incisive bone. Li: lower incisive tooth. Ll: lower lip. Mx: maxilla. N: nasal bone. Ns: nasal septum. Oc: oral cavity. Pp: palatine plexus. UiR: upper incisive tooth root. SVnc: sinus of ventral nasal concha (dorsal part). SVnc’: sinus of ventral nasal concha (ventral part). T: tongue. Ul: upper lip. V: vomer. Vn: vomeronasal organ. Vnm: ventral nasal meatus.</p>
Full article ">Figure 7
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the dorsal and ventral nasal conchae, and rostral maxillary sinus, corresponding to line VI in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Cnm: common nasal meatus. Dnc: dorsal nasal concha. Dnm: dorsal nasal meatus. Li: lower incisive tooth. Mx: maxilla. Ms’: rostral maxillary sinus. N: nasal bone. Ncp: nasal cavernous plexuses. Ns: nasal septum. Oc: oral cavity. Pf: parallel fold. Pp: palatine plexus. SVnc: sinus of ventral nasal concha (dorsal part). SVnc’: sinus of ventral nasal concha (ventral part). UiR: upper incisive tooth root. V: vomer. Vn: vomeronasal organ. Vnm: ventral nasal meatus.</p>
Full article ">Figure 8
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the choanae, corresponding to line VII in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Ch: choanae. Ccm: conchal crest of maxilla. Cnm: common nasal meatus. Dnm: dorsal nasal meatus. Li: lower incisive teeth. Mx: maxilla. Mb: mandible. Mnm: middle nasal meatus. Ms’: rostral maxillary sinus. N: nasal bone. Ns: nasal septum. NMxo: nasomaxillary opening. Oc: oral cavity. PMx: first premolar tooth (maxillar). PMb: first premolar tooth (mandibular). SDnc: sinus of dorsal nasal concha. SVnc: sinus of ventral nasal concha (dorsal part). SVnc’: sinus of ventral nasal concha (ventral part). V: vomer. Vn: vomeronasal organ. Vnm: ventral nasal meatus.</p>
Full article ">Figure 9
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the nasopharynx, corresponding to line VIII in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Ch: choanae. Cnm: common nasal meatus. Dnm: dorsal nasal meatus. Li: lower incisive teeth. Mi: mylohyoid muscle. Mx: maxilla. Mb: mandible. Mbu: buccinator muscle. Ms’’: caudal maxillary sinus. N: nasal bone. Ns: nasal septum. NMxo: nasomaxillary opening. Oc: oral cavity. PMx’: second premolar tooth (maxillar). PMb’: second premolar tooth (mandibular). SFr: sinus of frontal bone (rostral part). SVnc: sinus of ventral nasal concha (dorsal part). SVnc’: sinus of ventral nasal concha (ventral part). T: tongue. V: vomer. Vn: vomeronasal organ. Vnc: ventral nasal concha. Vnm: ventral nasal meatus. Zp: zygomatic process.</p>
Full article ">Figure 10
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the middle nasal concha, corresponding to line IX in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Cnm: common nasal meatus. CMxo: conchomaxillary opening. Dnm: dorsal nasal meatus. Li: lower incisive tooth. Mi: mylohyoid muscle. Mx: maxilla. Mb: mandible. Mbu: buccinator muscle. Ms’’: caudal maxillary sinus. N: nasal bone. Np: nasopharynx. Ns: nasal septum. MMx: first molar tooth (maxillar). MMb: first molar teeth (mandibular). Oc: oral cavity. SFr: sinus of frontal bone (rostral part). Smnc: sinus of midle nasal concha. SphS: sphenopalatine sinus (palatine part). SVnc: sinus of ventral nasal concha. T: tongue. V: vomer.</p>
Full article ">Figure 11
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the eyeball, corresponding to line X in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. Cnm: common nasal meatus. Dnm: dorsal nasal meatus. Dr: dorsal rectus muscle of the eye. E: ethmoid bone (tectorial plate). Ect: ectoturbinate. End: endoturbinate. F: frontal bone. L: lens. LiR: lower incisive root. Lp: perpendicular plate (ethmoid bone). Mb: mandible. Mm: masseter muscle (superficial portion). Mm’: masseter muscle (deeper portion). Mx: maxilla. MMx’: second molar tooth (maxillar). MMb’: second molar tooth (mandibular). Np: nasopharynx. Sc: sclera. SFs: septum of frontal sinuses. SFr: sinus of frontal bone (rostral part). SphS: sphenopalatine sinus (palatine part). T: tongue. V: vomer. Vc: vitreous chamber. Vnm: ventral nasal meatus. Vr: ventral rectus muscle of the eye. Zp: zygomatic process.</p>
Full article ">Figure 12
<p>Anatomical section (<b>A</b>), bone window (<b>B</b>), and pulmonary window (<b>C</b>) CT transverse images of a crested porcupine’s nasal cavity at the level of the olfactory bulb, corresponding to line XI in <a href="#vetsci-11-00611-f001" class="html-fig">Figure 1</a>. F: frontal bone. LiR: lower incisive root. Lp: perpendicular plate (ethmoid bone). Mb: mandible. Mm: masseter muscle. Mt: temporalis muscle. Mpt: medial pterygoid muscle. Mptl: lateral pterygoid muscle. Np: nasopharynx. Ob: olfactory bulb. Or: olfactory recess. SFr: sinus of frontal bone (caudolateral part). SFs: septum of frontal sinuses. Sp: soft palate. SphS: sphenopalatine sinus (sphenoidal part). SSphS: septum of sphenopalatine sinus. T: tongue. Zg: zygomatic gland. Zp: zygomatic process.</p>
Full article ">
16 pages, 463 KiB  
Article
Prognostic Factors in 26 Cats Undergoing Surgery for Extra-Hepatic Biliary Obstruction
by Jonathan P. Speelman, Ki-Lam Hui, Nicolas T. Woodbridge, Susanne Pfeiffer, Julia A. Beatty and Alan H. Taylor
Vet. Sci. 2024, 11(12), 610; https://doi.org/10.3390/vetsci11120610 (registering DOI) - 29 Nov 2024
Viewed by 363
Abstract
Surgical management of feline extra-hepatic biliary obstruction (EHBO) has poor survival rates with few prognostic factors reported in the literature. The etiology and clinical findings of feline EHBO and their influence on short-(2 weeks–6 months) and long-term (>6 months) survival and prognosis were [...] Read more.
Surgical management of feline extra-hepatic biliary obstruction (EHBO) has poor survival rates with few prognostic factors reported in the literature. The etiology and clinical findings of feline EHBO and their influence on short-(2 weeks–6 months) and long-term (>6 months) survival and prognosis were examined in an observational clinical retrospective study of 26 client-owned cats undergoing surgery for biliary obstruction at one institution between 2012 and 2020. The etiology of EHBO was determined in 21/26 cats, which included inflammatory causes (14/21), neoplastic causes (6/21), and a duodenal foreign body (1/21). Pre-operative hyperbilirubinemia and post-extubation hypotension (mean arterial pressure < 60 mmHg) were statistically associated with reduced short- and long-term survival. Short-term survival was documented in 17/26 cats, and long-term survival in 13/26 cats. Median survival time (MST) was 86 days (range, 0–1497). An MST of 17 days (range, 2–520) in cats with malignancies was found compared to an MST of 1165 days (range, 61–2268) in EHBO resulting from inflammatory complexes. Pre-operative hyperbilirubinemia and immediate post-operative hypotension may represent prognostic factors for cats undergoing surgery for EHBO. Cats with inflammatory causes of EHBO carry a more favorable prognosis than neoplastic causes. Further studies are required to evaluate the validity of the observed associations. Full article
(This article belongs to the Section Veterinary Surgery)
Show Figures

Figure 1

Figure 1
<p>Kaplan–Meier survival curve for the survivorship function for the time from surgery to death. Note: Dashed lines represent the 95% confidence intervals. The median survival time for cats was 86 days post-surgery.</p>
Full article ">
16 pages, 1304 KiB  
Article
Epidemiology and Molecular Characterisation of Multidrug-Resistant Escherichia coli Isolated from Cow Milk
by Zarin Tasnim Mim, Chandan Nath, Abdullah Al Sattar, Rijwana Rashid, Mehedy Hasan Abir, Shahneaz Ali Khan, Md Abul Kalam, Shahanaj Shano, Rowland Cobbold, John I. Alawneh and Mohammad Mahmudul Hassan
Vet. Sci. 2024, 11(12), 609; https://doi.org/10.3390/vetsci11120609 - 29 Nov 2024
Viewed by 686
Abstract
Antimicrobial resistance (AMR) is a growing global concern and poses a significant threat to public health. The emergence of multidrug-resistant organisms, including Escherichia coli, also presents a risk of transmission to humans through the food chain, including milk. This study aimed to [...] Read more.
Antimicrobial resistance (AMR) is a growing global concern and poses a significant threat to public health. The emergence of multidrug-resistant organisms, including Escherichia coli, also presents a risk of transmission to humans through the food chain, including milk. This study aimed to investigate the prevalence of E. coli in raw milk in the Chattogram metropolitan area (CMA) of Bangladesh and their phenotypic and genotypic antimicrobial resistance patterns. A total of 450 raw cow milk samples were collected from 18 farms within the CMA. The isolation and identification of E. coli were performed following standard bacteriological methods. Antimicrobial susceptibility testing (AST) was conducted using the Kirby–Bauer disc diffusion method. Molecular detection of E. coli and antimicrobial resistance genes was performed using the Polymerase Chain Reaction (PCR). This study found 134 (29.77%) milk samples that tested positive for E. coli. Antimicrobial susceptibility testing (AST) revealed the highest resistance rates (69.40%) to be for ampicillin, amoxicillin–clavulanic acid, cephalothin, and cephalexin, with the lowest resistance (21.64%) being for norfloxacin. A significant correlation (r = 1) was observed between ciprofloxacin and ceftazidime resistance among the antimicrobials tested. All E. coli isolates were classified as multidrug-resistant (MDR), being resistant to three or more antimicrobial classes, with a multiple resistance index >0.2. PCR amplification showed that the blaTEM gene had the highest prevalence (74.19%) among the ESBL and antimicrobial resistance genes tested. In contrast, the blaCMY-1 gene had a lower prevalence (6.45%) among the ESBL genes, while the tetD gene had the lowest prevalence (2.9%) among the resistance genes tested. Positive correlations were observed between antimicrobial resistance and the presence of these resistance genes. This study emphasises the high prevalence of MDR E. coli in raw cow milk and its significant potential impact on public health. It underscores the urgent need for strategic interventions to effectively manage and mitigate AMR in the Bangladeshi dairy sector, focusing on the prudent use of antimicrobials and implementing enhanced AMR surveillance. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical locations of the farms randomly selected for sampling in this study.</p>
Full article ">Figure 2
<p>Heatmap showing the correlation coefficient among antimicrobials tested in this study.</p>
Full article ">Figure 3
<p>MDR profiles of <span class="html-italic">E. coli</span> isolates from raw cow milk.</p>
Full article ">Figure 4
<p>The correlation coefficient between phenotypic AMR and resistance genes.</p>
Full article ">
21 pages, 3310 KiB  
Article
Changes in Gut Microbiota in Peruvian Cattle Genetic Nucleus by Breed and Correlations with Beef Quality
by Carlos Quilcate, Richard Estrada, Yolanda Romero, Diorman Rojas, Rolando Mamani, Renán Dilton Hañari-Quispe, Mery Aliaga, Walter Galindo, Héctor V. Vásquez, Jorge L. Maicelo and Carlos I. Arbizu
Vet. Sci. 2024, 11(12), 608; https://doi.org/10.3390/vetsci11120608 - 29 Nov 2024
Viewed by 395
Abstract
This study evaluated the gut microbiota and meat quality traits in 11 healthy female cattle from the Huaral region of Peru, including 5 Angus, 3 Braunvieh, and 3 F1 Simmental × Braunvieh. All cattle were 18 months old and maintained on a consistent [...] Read more.
This study evaluated the gut microbiota and meat quality traits in 11 healthy female cattle from the Huaral region of Peru, including 5 Angus, 3 Braunvieh, and 3 F1 Simmental × Braunvieh. All cattle were 18 months old and maintained on a consistent lifelong diet. Meat quality traits, including loin area, fat thickness, muscle depth, and marbling, were assessed in vivo using ultrasonography. Fecal samples were collected for microbiota analysis, and DNA was extracted for 16S and 18S rRNA sequencing to characterize bacterial, fungal, and protist communities. Significant correlations were observed between microbial genera and meat traits: Christensenellaceae R-7 and Alistipes were positively associated with marbling and muscle area, while Rikenellaceae RC9 showed a negative correlation with fat thickness. Among fungi, Candida positively correlated with marbling, while Trichosporon was negatively associated with muscle depth. For protists, Entodinium negatively correlated with fat thickness and marbling. Alpha diversity varied by breed, with Angus showing greater bacterial diversity, and beta diversity analyses indicated a strong breed influence on microbial composition. These findings suggest that microbial composition, shaped by breed and dietary consistency, could serve as an indicator of meat quality, offering insights into gut microbiota’s role in optimizing cattle production. Full article
Show Figures

Figure 1

Figure 1
<p>Alpha diversity metrics of bacteria and fungi were evaluated in cattle in three breeds. Diversity indices exhibited include ACE, Chao1, Observe, Pielou, Shannon, and Fisher. (<b>A</b>–<b>D</b>) Alpha diversity of bacteria (<b>E</b>–<b>H</b>) Alpha diversity of fungi. Different letters indicate statistically significant tests. * <span class="html-italic">p</span> &lt; 0.05. p = <span class="html-italic">p-</span>value.</p>
Full article ">Figure 2
<p>Bray–Curtis and Jaccard analysis of beta diversity in the different breeds of cattle. (<b>A</b>) Bray–Curtis of bacteria (<b>B</b>) Jaccard of bacteria (<b>C</b>) Bray–Curtis of fungi. (<b>D</b>) Fungi Jaccard (<b>E</b>) Protist Bray–Curtis (<b>F</b>) Protist Jaccard.</p>
Full article ">Figure 3
<p>Relative abundances in the gut microbiota at the phylum and genus level in different cattle breeds. (<b>A</b>) Bar graph analysis illustrates the abundance of bacterial phyla in each breed. (<b>B</b>) Heat map with the main bacterial abundances of the 20 genera in each breed. (<b>C</b>) Bar graph analysis illustrates the abundance of fungal phyla in each breed. (<b>D</b>) Heat map with the main abundances of fungi of 20 genera in each breed.</p>
Full article ">Figure 4
<p>Heat map of Spearman correlation analysis between the genera of intestinal microbes and meat variables. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 (<b>A</b>) Heat map of bacterial genera. (<b>B</b>) Heat map of fungi genera. (<b>C</b>) Heat map of protist genera.</p>
Full article ">Figure 5
<p>LEfSe analysis of the intestinal microbiome in different breeds. (<b>A</b>) LEfSe of bacteria. (<b>B</b>) LEfSe of fungi. (<b>C</b>) LEfSe of protist.</p>
Full article ">Figure 6
<p>Correlations between the alpha diversity of the intestinal microbiota and hematological and meat variables. (<b>A</b>) Spearman correlation of alpha diversity of bacteria. (<b>B</b>) Spearman correlation of protist alpha diversity.</p>
Full article ">
14 pages, 4925 KiB  
Article
The Effect of Meloxicam on Inflammatory Response and Oxidative Stress Induced by Klebsiella pneumoniae in Bovine Mammary Epithelial Cells
by Kangjun Liu, Shangfei Qiu, Li Fang, Luying Cui, Junsheng Dong, Long Guo, Xia Meng, Jianji Li and Heng Wang
Vet. Sci. 2024, 11(12), 607; https://doi.org/10.3390/vetsci11120607 - 29 Nov 2024
Viewed by 299
Abstract
Klebsiella pneumoniae (K. pneumoniae) is a significant pathogen associated with clinical mastitis in cattle. Anti-inflammatory drugs are necessary to alleviate pain and inflammation during clinical mastitis. Among many drugs, meloxicam (MEL) has been widely used in clinical mastitis because of its [...] Read more.
Klebsiella pneumoniae (K. pneumoniae) is a significant pathogen associated with clinical mastitis in cattle. Anti-inflammatory drugs are necessary to alleviate pain and inflammation during clinical mastitis. Among many drugs, meloxicam (MEL) has been widely used in clinical mastitis because of its excellent inhibitory effect on the cyclooxygenase-2 (COX-2) enzyme. However, the effectiveness of MEL on the inflammatory response and oxidative stress induced by K. pneumoniae are unclear. In the present study, primary BMECs were infected with K. pneumoniae in the presence or absence of plasma maintenance concentration of MEL (0.5 and 5 μM). Following 1 or 3 h of combined treatment with K. pneumoniae and MEL, BMECs were gathered to assess the related indicators. The results showed that MEL at plasma maintenance concentrations exerted no influence on the viability of uninfected BMECs and also had no impact on bacterial load in BMECs. At these concentrations, MEL was able to inhibit the mRNA expression of COX-2, Interleukin (IL)-1β, Tumor necrosis factor α (TNF-α), and IL-6 while simultaneously elevating the mRNA levels of IL-8 in K. pneumoniae-infected BMECs. MEL had clear effects on relieving oxidative stress by increasing the activity of superoxide dismutase (SOD) and catalase (CAT) and the level of total antioxidant capacity (T-AOC). The mechanisms by which MEL mitigated the inflammatory response and oxidative stress were partially attributed to inhibition of the nuclear transcription factor-kappa B (NF-κB) signaling pathway and improvement of the activation of the nuclear factor erythroid 2-related factors (Nrf2) signaling pathway. To conclude, the results manifested that MEL at plasma maintenance concentrations protected BMECs from inflammatory and oxidative damage induced by K. pneumoniae. Full article
(This article belongs to the Special Issue Ruminant Mastitis: Therapies and Control)
Show Figures

Figure 1

Figure 1
<p>The effects of MEL on cell viability. BMECs were treated with different concentrations (0, 0.5, 5, 10, 20, and 40 μΜ) of MEL for 12 h. The cell viability was evaluated with the CCK-8 method.</p>
Full article ">Figure 2
<p>The effects of MEL on bacterial load. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. Each experiment was repeated 4 times.</p>
Full article ">Figure 3
<p>The effects of MEL on the mRNA expression (<b>A</b>) and protein levels (<b>B</b>,<b>C</b>) of COX-2 in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. <span class="html-italic"># p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>).</p>
Full article ">Figure 4
<p>The effects of MEL on the mRNA expression of IL-1β (<b>A</b>), IL-6 (<b>B</b>), IL-8 (<b>C</b>), and TNF-α (<b>D</b>) in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 5
<p>The effects of MEL on the NF-κB signaling pathway. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 1 h in the presence or absence of MEL. (<b>A</b>) The protein expression of p-p65, p65, p-IκBα, and IκBα in BMECs (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>). (<b>B</b>) Changes in the phosphorylation level of p65. (<b>C</b>) Changes in the phosphorylation level of IκBα. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 6
<p>The effect of MEL on the nuclear accumulation of the p65 protein in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 1 h in the presence or absence of MEL.</p>
Full article ">Figure 7
<p>The effects of MEL on the oxidative state of BMECs. (<b>A</b>,<b>B</b>) Changes in the level of ROS. (<b>C</b>) Changes in the level of MDA. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 8
<p>The effects of MEL on the antioxidant capacity of BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. The activity of SOD (<b>A</b>) and CAT (<b>B</b>) and the level of T-AOC (<b>C</b>) were detected using commercial kits. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 9
<p>The effects of MEL on the Nrf2 signaling pathway in BMECs. (<b>A</b>) The key proteins expressed in the Nrf2 signaling pathway were detected by Western blot (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>). (<b>B</b>) Changes in Nrf2 nuclear accumulation. Nuclear protein was extracted from cells. The protein expression levels of Nrf2 (<b>C</b>), Keap1 (<b>D</b>), HO-1 (<b>E</b>), and NQO1 (<b>F</b>) were detected using total protein. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. <span class="html-italic"># p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">
14 pages, 2546 KiB  
Article
Maternal Transmission of Rotavirus to Calves and Comparison of Colostrum and Fecal Microbiota in Holstein and Hanwoo Cattle
by Seon-Ho Kim, Michelle Miguel, Ye Pyae Naing, Yong-Il Cho and Sang-Suk Lee
Vet. Sci. 2024, 11(12), 606; https://doi.org/10.3390/vetsci11120606 - 28 Nov 2024
Viewed by 400
Abstract
This study aimed to evaluate rotavirus transmission to calves and analyze microbial communities in cow milk and neonatal calf feces within dairy and beef cattle. A total of 20 cattle, Hanwoo (n = 10), and Holstein (n = 10) were allotted [...] Read more.
This study aimed to evaluate rotavirus transmission to calves and analyze microbial communities in cow milk and neonatal calf feces within dairy and beef cattle. A total of 20 cattle, Hanwoo (n = 10), and Holstein (n = 10) were allotted for the study, with each breed comprising five cows and five calves. Colostrum samples were obtained from the dam, while feces were obtained from both the dam and calf. Group A rotavirus was identified in the fecal samples through real-time reverse transcription PCR (RT-qPCR). Bacterial communities present in the colostrum and bovine feces were explored using 16S rRNA metagenomic sequencing. The RT-qPCR results showed that the Cq value of one calf and one cow in the Holstein group was < 35, confirming the presence of rotavirus, whereas the Cq value in the Hanwoo group was > 35, indicating a negative result. For the bacterial communities, significant differences (p < 0.05) were found between the colostrum and fecal samples from the dams and calves, but there were no significant differences between Hanwoo and Holstein cattle. Alpha diversity analysis showed that the Chao1 and Shannon indices revealed significant differences (p < 0.05) among the sample types (cow colostrum, cow feces, and calf feces). The bacterial communities in various sample types from both Hanwoo and Holstein cattle were dominated by the phyla Firmicutes, Proteobacteria, and Bacteroidetes. In addition, the genera shared between the cow colostrum and calf fecal microbiota were higher than those shared between cow and calf feces. Overall, the current study detected rotavirus in Holstein but not in Hanwoo cattle; however, no clear evidence showed the transmission of rotavirus from dam to calf. Moreover, significant variations in bacterial compositions were observed among calf feces, cow feces, and colostrum samples, suggesting the presence of unique microbial profiles. Full article
(This article belongs to the Section Veterinary Microbiology, Parasitology and Immunology)
Show Figures

Figure 1

Figure 1
<p>Amplification of bovine rotavirus Group A from Holstein and Hanwoo cow and calf fecal samples. A Cq cut-off value of 35 was considered positive.</p>
Full article ">Figure 2
<p>Alpha diversity of bacterial communities in cow (colostrum and feces) and calf (feces) samples in Hanwoo and Holstein breeds. Boxplots for Chao1 richness index (<b>A</b>) and Shannon diversity index (<b>B</b>).</p>
Full article ">Figure 3
<p>Principal coordinate analysis (PCoA) plot based on the Bray–Curtis dissimilarity in beta diversity based on sample type and breed (<b>A</b>), sample type (cow colostrum, cow feces, and calf feces) (<b>B</b>), breed (Hanwoo and Holstein) (<b>C</b>), different sample types within the Hanwoo group (<b>D</b>), and different sample types within the Holstein group (<b>E</b>). The <span class="html-italic">p</span>-value corresponds to results from the adonis PERMANOVA.</p>
Full article ">Figure 4
<p>Venn diagrams showing the genera shared between cow colostrum and calf feces (<b>A</b>), cow feces and calf feces (<b>B</b>), different cow and calf samples in Hanwoo (<b>C</b>) and Holstein cattle (<b>D</b>), and (<b>E</b>) UpSet diagram showing genus distribution across sample types in different breeds of cattle. The bars represent the number of shared genera between specific samples or those unique to a single sample.</p>
Full article ">Figure 5
<p>Taxonomic composition of bacterial communities in the colostrum and feces of Hanwoo and Holstein cows and calves. Relative abundance at phylum (<b>A</b>) and genus (<b>B</b>) levels.</p>
Full article ">Figure 6
<p>Differential abundance analysis of the bacterial genera in different sample types (cow colostrum, cow feces, and calf feces) (<b>A</b>) and breed (Hanwoo and Holstein) (<b>B</b>).</p>
Full article ">
26 pages, 7060 KiB  
Article
The Venom of Vipera ammodytes ammodytes: Proteomics, Neurotoxic Effect and Neutralization by Antivenom
by Saša R. Ivanović, Dina Rešetar Maslov, Ivana Rubić, Vladimir Mrljak, Irena Živković, Nevena Borozan, Jelica Grujić-Milanović and Sunčica Borozan
Vet. Sci. 2024, 11(12), 605; https://doi.org/10.3390/vetsci11120605 - 28 Nov 2024
Viewed by 375
Abstract
Deep proteomic analyses identified, in total, 159 master proteins (with 1% FDR and 2 unique peptides) from 26 protein families in the venom of Vipera ammodytes ammodytes (Vaa). Data are available via ProteomeXchange with the identifier PXD056495. The relative abundance of PLA2s is [...] Read more.
Deep proteomic analyses identified, in total, 159 master proteins (with 1% FDR and 2 unique peptides) from 26 protein families in the venom of Vipera ammodytes ammodytes (Vaa). Data are available via ProteomeXchange with the identifier PXD056495. The relative abundance of PLA2s is 11.60% of the crude venom, of which 4.35% are neurotoxic Ammodytoxins (Atxs). The neurotoxicity of the venom of Vaa and the neutralizing effect of the antivenom were tested on the neuromuscular preparation of the diaphragm (NPD) of rats. The activity of PLA2 in the venom of Vaa and its neutralization by the antivenom were determined under in vitro conditions. The Vaa venom leads to a progressive decrease in NPD contractions. We administered pre-incubated venom/antivenom mixtures at various ratios of 1:2, 1:10 and 1:20 (w/w) and observed the effects of these mixtures on NPD contractions. The results show that the mean effective time (ET50) for NPD contractions with the 1:20 mixture is highly significantly different (p < 0.001) from the ET50 for the venom and the ET50 for the 1:2 and 1:10 mixture ratios. We also found a highly significant (p < 0.001) reduction in Na+/K+-ATPase activity in the NPD under the influence of the venom. The reduction in the activity of this enzyme was reversible by the antivenom. Under in vitro conditions, we have achieved the complete neutralization of PLA2 by the antivenom. In conclusion, the antivenom abolished the venom-induced progressive decrease in NPD contractions in a concentration-dependent manner. Antivenom with approximately the same mass proportion almost completely restores Na+/K+-ATPase activity in the NPD and completely neutralizes the PLA2 activity of the venom in vitro. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Vipera ammodytes ammodytes</span>. Original photo: Institute of Virology, Vaccines and Sera “Torlak”, Belgrade, Serbia.</p>
Full article ">Figure 2
<p>(<b>A</b>) Distribution of identified proteins for fraction 0 when different protein FASTA databases were used in the analysis; (<b>B</b>) Distribution of identified proteins for all fractions (0, 3A, 5A, 8A, 9A and 10A) when <span class="html-italic">Serpentes</span> protein FASTA databases (DB) were used in the analysis; (<b>C</b>) Distribution of identified proteins for all fractions (0, 3A, 5A, 8A, 9A and 10A) when <span class="html-italic">Vipera</span> protein FASTA databases (DB) were used in the analysis; (<b>D</b>) Distribution of identified proteins for all fractions (0, 3A, 5A, 8A, 9A and 10A) when <span class="html-italic">V. ammodytes</span> protein FASTA databases (DB) were used in the analysis; (<b>E</b>) Distribution of identified proteins for all fractions (0, 3A, 5A, 8A, 9A and 10A) when different protein FASTA databases were used in the analysis. Next to each fraction, the number of proteins identified in this fraction for all three databases used is given in brackets.</p>
Full article ">Figure 3
<p>Relative distribution of protein groups (%) in the <span class="html-italic">Vaa</span> venom determined by nano-liquid chromatography–tandem mass spectrometry-based proteomics: (<b>A</b>) DB <span class="html-italic">V. ammodytes</span>; (<b>B</b>) DB <span class="html-italic">Vipera</span>.</p>
Full article ">Figure 4
<p>Representative recording of contractions of a neuromuscular preparation of the diaphragm (NPD) induced by indirect EFS (·····) in the absence of venom. C<sub>1</sub> and C<sub>2</sub>—control contractions; panc 1 μM—contractions under the influence of 1 μM pancuronium; W<sub>1</sub> and W<sub>2</sub>—contractions after the washout of pancuronium; 10 “packages” of contractions in the function of time.</p>
Full article ">Figure 5
<p>Representative recording of contractions of a neuromuscular preparation of the diaphragm (NPD) induced by indirect EFS (<b>·····</b>) and direct EFS (<b>-----</b>) under the influence of venom. C<sub>1</sub> and C<sub>2</sub>—control contractions; panc 3 μM—contractions under the influence of 3 μM pancuronium; W<sub>1</sub> and W<sub>2</sub>—contractions after the washout of pancuronium; 12 “packages” of contractions induced by indirect EFS; 2 “packages” of contractions induced by direct EFS.</p>
Full article ">Figure 6
<p>Sigmoidal curves of the reduction in contractions of the neuromuscular preparation of the diaphragm (NPD) in a logarithmic function of time under the influence of venom and venom/antivenom mixtures at ratios of 1:2; 1:10 and 1:20 (<span class="html-italic">w</span>/<span class="html-italic">w</span>).</p>
Full article ">Figure 7
<p>Comparison of ET<sub>50</sub> (minutes) after the administration of venom and a venom/antivenom mixture at the ratios of 1:2; 1:10 and 1:20 (<span class="html-italic">w</span>/<span class="html-italic">w</span>) (mean ± SD, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. venom; <sup>+++</sup> <span class="html-italic">p</span> &lt; 0.001 between different mass ratios of venom/antivenom).</p>
Full article ">Figure 8
<p>Representative recording of contraction peaks of the neuromuscular preparations of the diaphragm (NPD) induced by indirect EFS: (<b>A</b>) Control contractions; (<b>B</b>) Contractions under the influence of pancuronium <span class="html-italic">(tetanic fade)</span>; (<b>C</b>) Contractions under the influence of venom; (<b>D</b>) Contractions under the influence of a mixture of venom/antivenom at a ratio of 1:20 (<span class="html-italic">w</span>/<span class="html-italic">w</span>) (white arrows show a facilitated release of neurotransmitters; black arrows show a reduced release of neurotransmitters—<span class="html-italic">tetanic fade</span>).</p>
Full article ">Figure 9
<p>AChE activity (U/mg P) in the neuromuscular preparations of the diaphragm (NPD) without the presence of venom (control), under the influence of venom and for the mixture of venom/antivenom at a ratio of 1:2, 1:10 and 1:20 (<span class="html-italic">w</span>/<span class="html-italic">w</span>) (mean ± SD, <span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 10
<p>Na<sup>+</sup>/K<sup>+</sup>-ATPase activity (U/mg P) in the neuromuscular preparations of the diaphragm (NPD) without the presence of venom (control), under the influence of venom and under the influence of a mixture of venom and antivenom at the ratios of 1:2; 1:10 and 1:20 (<span class="html-italic">w</span>/<span class="html-italic">w</span>) (mean ± SD, *** <span class="html-italic">p</span> &lt; 0.001 vs. control; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. venom; <sup>++</sup> <span class="html-italic">p</span>&lt;0.01 between different mass ratios of venom/antivenom).</p>
Full article ">Figure 11
<p>(<b>A</b>) Activity of the PLA2 in increasing concentrations of the <span class="html-italic">Vaa</span> venom (mg/mL); (<b>B</b>) Inhibition of the PLA2 activity in 1 mg/mL of the <span class="html-italic">Vaa</span> venom by increasing concentrations of the antivenom (mg/mL).</p>
Full article ">
12 pages, 303 KiB  
Article
Growth Performance, Rumen Fermentation, and Meat Quality of Finishing Lambs Supplemented with Calcium Propionate or Sodium Propionate
by Lucero Abigail Velázquez-Cruz, Pedro Abel Hernández-García, Germán David Mendoza-Martínez, Enrique Espinosa-Ayala, Héctor Aarón Lee-Rangel, Gabriela Vázquez-Silva, Pablo Benjamín Razo-Ortíz, Cesar Díaz-Galván, José Felipe Orzuna-Orzuna and María Eugenia de la Torre-Hernández
Vet. Sci. 2024, 11(12), 604; https://doi.org/10.3390/vetsci11120604 - 28 Nov 2024
Viewed by 375
Abstract
This study aimed to evaluate the effect of dietary supplementation with calcium propionate (CaPr) or sodium propionate (NaPr) on growth performance, ruminal fermentation, and meat quality of finishing lambs. Twenty-seven non-castrated Creole male lambs (24.95 ± 2.15 kg body weight (BW); 4.5 ± [...] Read more.
This study aimed to evaluate the effect of dietary supplementation with calcium propionate (CaPr) or sodium propionate (NaPr) on growth performance, ruminal fermentation, and meat quality of finishing lambs. Twenty-seven non-castrated Creole male lambs (24.95 ± 2.15 kg body weight (BW); 4.5 ± 0.5 months old) were randomly assigned to three treatments: (a) CON: basal diet without the addition of CaPr or Na Pr; (b) basal diet + CaPr (10 g/kg DM); and (c) basal diet + NaPr (10 g/kg DM). The data were analyzed using a completely randomized experimental design, with each lamb considered an experimental unit (nine replicates/treatment). Dietary supplementation with CaPr or NaPr did not affect (p > 0.05) growth performance or dietary energetics. However, greater (p = 0.05) apparent dry matter digestibility was observed in the lambs that consumed the diet with NaPr10. Dietary supplementation with CaPr or NaPr did not affect (p > 0.05) ruminal pH or ruminal concentrations of ammonia nitrogen, acetate, propionate, butyrate, and total volatile fatty acids. However, ruminal lactate concentration increased (p = 0.01) in lambs consuming the NaPr diet. Hot carcass weight and yield, backfat thickness, meat pH, meat color (L*, a*, and b*), cooking loss, and water holding capacity were not affected by dietary supplementation with CaPr or NaPr. In conclusion, dietary supplementation with 10 g/kg DM of calcium propionate or sodium propionate does not affect growth performance, dietary energetics, ruminal fermentation, and the meat quality of finishing lambs. Full article
(This article belongs to the Section Nutritional and Metabolic Diseases in Veterinary Medicine)
9 pages, 1354 KiB  
Brief Report
Mutation in Wzz(fepE) Linked to Altered O-Antigen Biosynthesis and Attenuated Virulence in Rough Salmonella Infantis Variant
by Nneka Vivian Iduu, Steven Kitchens, Stuart B. Price and Chengming Wang
Vet. Sci. 2024, 11(12), 603; https://doi.org/10.3390/vetsci11120603 - 28 Nov 2024
Viewed by 372
Abstract
Salmonella enterica serovar Infantis has emerged as a prevalent foodborne pathogen in poultry with significant global health implications. This study investigates the molecular characteristics influencing virulence in a S. Infantis rough variant collected from a poultry farm in the USA. In this study, whole [...] Read more.
Salmonella enterica serovar Infantis has emerged as a prevalent foodborne pathogen in poultry with significant global health implications. This study investigates the molecular characteristics influencing virulence in a S. Infantis rough variant collected from a poultry farm in the USA. In this study, whole genome sequencing and comparative genomics were performed on smooth and rough poultry S. Infantis isolates, while chicken embryo lethality assay was conducted to assess their virulence. Comparative genomics between isolates was analyzed using Mauve pairwise Locally Collinear Blocks to measure the genetic conservation. Embryo survival rates between the isolates were compared using the Kaplan–Meier curves. High genomic conservation was observed between the two isolates, but a frameshift mutation was detected in the Wzz(fepE) gene of the rough variant, resulting in early protein truncation. The chicken embryo lethality assay showed that the lethality rate of the smooth strain was higher than that of the rough strain (p < 0.05). This study identifies a frameshift mutation in the Wzz(fepE) gene, leading to protein truncation, which may reduce bacterial virulence by impacting O-antigen biosynthesis in the rough Salmonella Infantis variant. These findings deepen our understanding of S. Infantis pathogenesis and suggest that targeting the Wzz(fepE) gene or related pathways could be a promising strategy for developing effective vaccines and therapeutic interventions. Full article
Show Figures

Figure 1

Figure 1
<p>High genomic similarities between the whole genome sequences of <span class="html-italic">Sal</span>_smooth and <span class="html-italic">Sal</span>_rough. Pairwise alignment of the <span class="html-italic">Sal</span>_smooth and <span class="html-italic">Sal</span>_rough genomes was conducted using Mauve software. Colored blocks represent homologous (similar) regions, with connecting lines indicating shared sequences between the two genomes. Blocks below the center line indicate regions aligned in reverse complement (inverse) orientation. These homologous regions are called Locally Collinear Blocks (LCBs). A total of ten (<span class="html-italic">n</span> = 10) LCBs were identified, with a minimum weight of 3099, indicating strong homology and high similarity between the strains.</p>
Full article ">Figure 2
<p>Frameshift mutation in the <span class="html-italic">Wzz(fepE)</span> gene of <span class="html-italic">S.</span> Infantis rough strain. The grey highlighted chromatogram region shows an adenine (A) insertion after nucleotide position 32, extending the gene length in <span class="html-italic">Sal</span>_rough from 1137 to 1138 bp as observed in <span class="html-italic">Sal</span>_smooth. This insertion introduces an early stop codon at amino acid position 26, truncating the protein from 378 to 25 amino acids.</p>
Full article ">Figure 3
<p>Chicken embryo lethality rate of <span class="html-italic">S.</span> Infantis strains. Twelve (12) eggs were used in each experimental group except the negative control (10 eggs used) with an inoculum concentration of 10<sup>3</sup> CFU/mL and sterile Dulbecco’s phosphate-buffered saline (PBS) for the control group. Kaplan–Meier plot displaying the chicken embryo lethality when inoculated with sterile PBS (black line), smooth strain <span class="html-italic">Sal</span>_smooth (red line), and rough variant <span class="html-italic">Sal</span>_rough (blue line), for 5 days post-inoculation. Results show that the embryo lethality rate of the <span class="html-italic">Sal</span>_smooth group was higher than the negative control group and <span class="html-italic">Sal</span>_rough (<span class="html-italic">p &lt;</span> 0.05). No difference in embryo lethality rates was observed between the negative control and <span class="html-italic">Sal</span>_rough (<span class="html-italic">p</span> = 0.721). Statistical analysis was performed using the log-rank test.</p>
Full article ">
13 pages, 2987 KiB  
Article
Effect of Different Thawing Regimes on Cell Kinematics and Organelle Integrity of Nitrogen-Stored Wallachian Ram Spermatozoa
by Martin Ptáček, Filipp Georgijevič Savvulidi, Christopher LeBrun, Martina Janošíková, Temirkhan Kenzhebaev, Kairly Omashev, Beybit Kulataev and Nurlan Malmakov
Vet. Sci. 2024, 11(12), 602; https://doi.org/10.3390/vetsci11120602 - 27 Nov 2024
Viewed by 419
Abstract
Artificial insemination is an advanced reproductive technology used to increase the number of lambs born from elite sires to accelerate genetic gain in a flock [...] Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of semen thawing procedure design. T<sub>1–5</sub> = thawing temperature (where T<sub>1</sub>—45 °C; T<sub>2</sub>—50 °C; T<sub>3</sub>—55 °C; T<sub>4</sub>—60 °C; T<sub>5</sub>—65 °C).</p>
Full article ">Figure 2
<p>Total (TM) and progressive (PM) motility modified thawing protocols compared to the reference (39 °C/s) one. T0 = results detected immediately after thawing; T2 = results detected after 2 h of incubation in a water bath heated at 39 °C; black points with bounded lines indicate LSM values ± SE for modified thawing protocols; the red line with a grey marked area indicates the reference thawing protocol (LSM ± SE); * = indicates a significant difference in the modified thawing protocol to the reference thawing protocol at <span class="html-italic">p</span> &lt; 0.05 level of significance.</p>
Full article ">Figure 3
<p>Sperm viability (VIA) of modified thawing protocols compared to the reference (39 °C/s) one. T0 = results detected immediately after thawing; T2 = results detected after 2 h of incubation in a water bath heated at 39 °C; black points with bounded line indicate LSM values ± SE for modified thawing protocols; the red line with a grey marked area indicates the reference thawing protocol (LSM ± SE); * = indicates a significant difference in the modified thawing protocol to reference the thawing protocol at <span class="html-italic">p</span> &lt; 0.05 level of significance.</p>
Full article ">Figure 4
<p>(<b>A</b>)—Graphical illustration of the thawing curves for different thawing regimes. The temperature change inside the straw during thawing is shown. The red line represents the thawing curve of the control regime (39 degrees for 30 s). (<b>B</b>)—The approximation of thermal energy (calculated in Joules) absorbed by a single straw at different thawing regimes. The red dotted line represents the energy absorbed by a single straw at a reference thawing regime (39 degrees for 30 s).</p>
Full article ">
17 pages, 9821 KiB  
Article
Exploring Gene Expression and Alternative Splicing in Duck Embryonic Myoblasts via Full-Length Transcriptome Sequencing
by Jintao Wu, Shuibing Liu, Dongcheng Jiang, Ya’nan Zhou, Hongxia Jiang, Xiaoyun Xiao, Boqian Zha, Yukai Fang, Jie Huang, Xiaolong Hu, Huirong Mao, Sanfeng Liu and Biao Chen
Vet. Sci. 2024, 11(12), 601; https://doi.org/10.3390/vetsci11120601 - 27 Nov 2024
Viewed by 350
Abstract
The duck industry is vital for supplying high-quality protein, making research into the development of duck skeletal muscle critical for improving meat and egg production. In this study, we leveraged Oxford Nanopore Technologies (ONT) sequencing to perform full-length transcriptome sequencing of myoblasts harvested [...] Read more.
The duck industry is vital for supplying high-quality protein, making research into the development of duck skeletal muscle critical for improving meat and egg production. In this study, we leveraged Oxford Nanopore Technologies (ONT) sequencing to perform full-length transcriptome sequencing of myoblasts harvested from the leg muscles of duck embryos at embryonic day 13 (E13), specifically examining both the proliferative (GM) and differentiation (DM) phases. Our analysis identified a total of 5797 novel transcripts along with 2332 long non-coding RNAs (lncRNAs), revealing substantial changes in gene expression linked to muscle development. We detected 3653 differentially expressed genes and 2246 instances of alternative splicing, with key genes involved in essential pathways, such as ECM–receptor interaction and Notch signaling, prominently featured. Additionally, we constructed a protein–protein interaction network that highlighted critical regulators—MYOM3, MYL2, MYL1, TNNI2, and ACTN2—associated with the processes of proliferation and differentiation in myoblasts. This extensive transcriptomic investigation not only sheds light on the intricate molecular mechanisms driving skeletal muscle development in ducks but also provides significant insights for future breeding strategies aimed at enhancing the efficiency of duck production. The results emphasize the efficacy of ONT sequencing in uncovering complex regulatory networks within avian species, ultimately contributing to progress in animal husbandry. Full article
Show Figures

Figure 1

Figure 1
<p>Duck primary embryonic myogenesis. (<b>A</b>) Myoblasts after differential adhesion. (<b>B</b>) Differentiating myoblasts on 48 h (myotubes).</p>
Full article ">Figure 2
<p>Analysis of new transcripts and variable splicing. (<b>A</b>) Number of new transcripts and genes. (<b>B</b>) Type and number of novel transcripts. ‘o’ denotes regions on the same strand overlapping with reference exons, ‘j’ signifies at least one matching multi-exon, ‘x’ represents exon overlap on the opposite strand, ‘I’ indicates introns completely contained within the reference transcript, and ‘u’ denotes unknown new transcripts. (<b>C</b>) Nr annotation statistics. (<b>D</b>) GO enrichment results of novel transcripts. (<b>E</b>) KEGG enrichment results of novel transcripts.</p>
Full article ">Figure 3
<p>Alternative splicing types. SSR analysis and lncRNA identification. (<b>A</b>) Average percentage of each type of alternative splicing in GM samples. (<b>B</b>) Average percentage of each type of alternative splicing in DM samples. (<b>C</b>) Number and type of SSRs. (<b>D</b>) Venn diagram of predicted lncRNAs by the CNCI, CPC, and PLEK. (<b>E</b>) Statistics of lncRNA number.</p>
Full article ">Figure 4
<p>LncRNAs’ function prediction. (<b>A</b>) Number of cis-regulated lncRNA mRNA pairs. (<b>B</b>) Cis-acting lncRNAs on muscle-related genes. (<b>C</b>) Number of trans-regulated LncRNA mRNA pairs. (<b>D</b>) Trans-acting lncRNAs on muscle-related genes. (<b>E</b>) GO enrichment analysis of the cis-regulated target gene. (<b>F</b>) KEGG analysis of the cis-regulated target gene. (<b>G</b>) GO enrichment analysis of the trans-regulated target gene. (<b>H</b>) KEGG analysis of the trans-regulated target gene.</p>
Full article ">Figure 5
<p>Differentially expressed gene analysis. (<b>A</b>) Principal component analysis (PCA) of all samples using sequencing data. (<b>B</b>) Differentially expressed gene volcano map. (<b>C</b>) Differentially expressed gene heat map. (<b>D</b>) Differentially expressed gene GO enrichment analysis. (<b>E</b>) Differentially expressed gene KEGG analysis.</p>
Full article ">Figure 6
<p>Protein–protein interaction network diagram. Darker colors indicate greater numbers of neighbor nodes.</p>
Full article ">Figure 7
<p>RT-qPCR validation of the sequencing data. (<b>A</b>) Relative expression of mRNAs in GM and DM groups using RT-qPCR; (<b>B</b>) TPM of mRNAs in GM and DM groups in the ONT sequencing. TPM: Transcripts per million. Values are presented as the mean ± SEM. * indicates <span class="html-italic">p</span> &lt; 0.05, and ** indicates <span class="html-italic">p</span> &lt; 0.01, *** indicates <span class="html-italic">p</span> &lt; 0.001. GM indicates the proliferating myoblast group. DM indicates the differentiated myoblast group.</p>
Full article ">
11 pages, 1619 KiB  
Article
Molecular Investigations of Babesia caballi from Clinically Healthy Horses in Southwestern Romania
by Simona Giubega, Marius Stelian Ilie, Sorin Morariu, Mirela Imre, Cristian Dreghiciu, Tatiana Rugea, Simina Ivascu, Gheorghița Simion and Gheorghe Dărăbuș
Vet. Sci. 2024, 11(12), 600; https://doi.org/10.3390/vetsci11120600 - 27 Nov 2024
Viewed by 389
Abstract
Babesia caballi is a tick-borne hemoparasite that causes equine piroplasmosis. It has a significant economic impact, decreasing performance and affecting animal welfare. This study aimed to identify B. caballi DNA in the blood of horses from households in the southwestern and western regions [...] Read more.
Babesia caballi is a tick-borne hemoparasite that causes equine piroplasmosis. It has a significant economic impact, decreasing performance and affecting animal welfare. This study aimed to identify B. caballi DNA in the blood of horses from households in the southwestern and western regions of Romania. We included 310 animals, from which blood was collected via EDTA. To test the samples for the B. caballi parasite genome, we used real-time PCR and conventional PCR. The prevalence of B. caballi was 5.81% (18/310) in apparently healthy horses, suggesting that this parasite is enzootic in the regions studied, although veterinarians did not indicate any symptoms resembling clinical babesiosis. In Romania, there are insufficient epidemiologic data on equine babesiosis, and the results of the present study suggest the need for further investigations into the dynamics of transmission and to identify potential prevention and control strategies. Full article
(This article belongs to the Section Veterinary Microbiology, Parasitology and Immunology)
Show Figures

Figure 1

Figure 1
<p>Origin of the samples by county.</p>
Full article ">Figure 2
<p>Cycle threshold values in positive samples (blue) compared with their mean (orange) (29.74083).</p>
Full article ">Figure 3
<p>PCR gel electrophoresis of the 18S rRNA gene of <span class="html-italic">B. caballi</span> isolates. Line 1, 100 bp DNA ladder (BIOLINE<sup>®</sup> UK Ltd., London, UK); Line 2, positive control (<span class="html-italic">B. caballi</span>); Line 3, negative control; Lane 4, positive sample; Lines 5–6, negative samples; Lines 7–10, positive samples.</p>
Full article ">
17 pages, 983 KiB  
Systematic Review
A Systematic Review on the Impact of Vaccination for Respiratory Disease on Antibody Titer Responses, Health, and Performance in Beef and Dairy Cattle
by Hudson R. McAllister, Bradly I. Ramirez, Molly E. Crews, Laura M. Rey, Alexis C. Thompson, Sarah F. Capik and Matthew A. Scott
Vet. Sci. 2024, 11(12), 599; https://doi.org/10.3390/vetsci11120599 - 27 Nov 2024
Viewed by 638
Abstract
Bovine respiratory disease (BRD) is a multifactorial disease complex commonly affecting beef and dairy operations. Vaccination against major BRD-related pathogens is routinely performed for disease prevention; however, uniform reporting of health and performance outcomes is infrequent. Our objective was to evaluate the effect [...] Read more.
Bovine respiratory disease (BRD) is a multifactorial disease complex commonly affecting beef and dairy operations. Vaccination against major BRD-related pathogens is routinely performed for disease prevention; however, uniform reporting of health and performance outcomes is infrequent. Our objective was to evaluate the effect of commercially available BRD-pathogen vaccination on titer response with respect to health or performance in beef and dairy cattle. This study was conducted under Prisma 2020 guidelines for systematic reviews and PRESS guidelines utilizing five databases. Criteria for study inclusion were as follows: research conducted in the USA or Canada, between 1982 and 10 October 2022, on beef or dairy cattle, using a commercially available vaccine labeled for a respiratory pathogen of interest, which evaluated antibody titers alongside either performance or morbidity. A total of 3020 records underwent title and abstract evaluation. Full-text analysis was conducted on 466 reports; 101 studies were included in the final review. Approximately 74% of included studies were beef cattle-based versus 26% dairy cattle-based. This review aimed to assess how vaccination titer responses affect beef and dairy cattle health and performance, but varying study methods made comparisons difficult, highlighting the need for consistent reporting. Full article
Show Figures

Figure 1

Figure 1
<p>PRISMA flowchart for the systematic review on impact of vaccination for respiratory disease on antibody titer responses, health, and performance in beef and dairy cattle.</p>
Full article ">
9 pages, 1525 KiB  
Article
First Evidence of Cotinine in Canine Semen Reveals Tobacco Smoke Exposure
by Debora Groppetti, Giulia Pizzi, Elisa Giussani, Alessandro Pecile, Silvia Michela Mazzola, Valerio Bronzo and Eleonora Fusi
Vet. Sci. 2024, 11(12), 598; https://doi.org/10.3390/vetsci11120598 - 26 Nov 2024
Viewed by 398
Abstract
Tobacco smoke has numerous adverse effects on both human and animal health, including impaired reproductive function. Recent research has explored environmental exposure in dogs, investigating various biological matrices. However, no data are currently available on the presence of cotinine, a nicotine metabolite, in [...] Read more.
Tobacco smoke has numerous adverse effects on both human and animal health, including impaired reproductive function. Recent research has explored environmental exposure in dogs, investigating various biological matrices. However, no data are currently available on the presence of cotinine, a nicotine metabolite, in the canine ejaculate. This study aimed to evaluate the detectability of cotinine in the semen of dogs living with smoking owners. Additionally, seminal cotinine concentrations were correlated with those in serum and hair. To further examine the potential impact of smoking on canine fertility, the relationships between seminal cotinine, total sperm concentration, and antioxidant activity in plasma and semen were analyzed in exposed and non-exposed dogs. This study is the first to demonstrate the presence of cotinine in canine ejaculate and its correlation with blood and hair concentrations. While the potential toxic effect of cotinine on seminal parameters and male fertility in dogs requires further investigation, it is crucial to raise awareness among pet owners about the risks associated with domestic smoking for their animals. Full article
(This article belongs to the Section Veterinary Reproduction and Obstetrics)
Show Figures

Figure 1

Figure 1
<p>Comparison of seminal cotinine concentration between passive smokers and non-exposed dogs. S: passive-smoker dogs, N: non-exposed dogs, * indicates <span class="html-italic">p</span> = 0.0002.</p>
Full article ">Figure 2
<p>Receiver operating characteristic (ROC) curves illustrating the relationship between seminal cotinine in exposed and non-exposed dogs. ROC curves describe the tradeoff between sensitivity and specificity. The 45° diagonal of the ROC space is the random chance line. The respective area under the curve (AUC) values and level of significance are reported in the plot for each curve.</p>
Full article ">Figure 3
<p>Correlation between cotinine concentrations in serum, semen, and hair. * indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Total antioxidant capacity (TAC) in plasma and semen. S: passive-smoker dogs; N: non-exposed dogs.</p>
Full article ">
16 pages, 3263 KiB  
Article
Systemic Granulomatosis in the Meagre Argyrosomus regius: Fishing for a Plausible Etiology
by Claudio Murgia, Tiziana Cubeddu, Giovanni P. Burrai, Alberto Alberti, Luigi Bertolotti, Barbara Colitti, Marino Prearo, Paolo Pastorino, Giuseppe Esposito, Luciana Mandrioli, Gaspare Barbera, Marina Antonella Sanna, Marta Polinas, Esteban Soto and Elisabetta Antuofermo
Vet. Sci. 2024, 11(12), 597; https://doi.org/10.3390/vetsci11120597 - 26 Nov 2024
Viewed by 287
Abstract
Meagre (Argyrosomus regius) is one of the fast-growing species considered for sustainable aquaculture development along the Mediterranean and Eastern Atlantic coasts. The emergence of Systemic Granulomatosis (SG), a disease marked by multiple granulomas in various tissues, poses a significant challenge in [...] Read more.
Meagre (Argyrosomus regius) is one of the fast-growing species considered for sustainable aquaculture development along the Mediterranean and Eastern Atlantic coasts. The emergence of Systemic Granulomatosis (SG), a disease marked by multiple granulomas in various tissues, poses a significant challenge in meagre aquaculture. In the current study, we investigate the association of Mycobacterium spp. and SG in offshore aquaculture facilities in Sardinia, Italy. A total of 34 adult seemingly healthy meagre were arbitrarily collected and analyzed, combining histological, microbiological, molecular, metagenomics, and in situ techniques to investigate the presence of pathogens. Ziehl–Neelsen (ZN), periodic acid–schiff (PAS), and Giemsa stains were performed for the detection of acid-fast bacteria, common parasites, and fungi within granulomas, respectively. Granulomas were detected in 91% (31/34) of fish. The affected organs were kidney (88%), liver (47%), heart (41%), intestine (17.6%), and brain (5%). Acid-fast staining, along with Mycobacterium spp. specific quantitative PCR (qPCR), in situ hybridization (ISH) assay, and microbiological analyses showed negative results for the detection of Mycobacterium spp. and other bacteria implicated in granuloma formation. However, PCR amplification and sequencing of the 65-kDa heat shock protein gene revealed the presence of M. chelonae in 13% of both formalin-fixed and frozen liver tissues. Bacterial isolation failed to detect nontuberculous mycobacteria (NTM) and other bacteria typically associated with granulomas. Consistently, the use of an M. chelonae-specific probe in ISH failed to identify this bacterial species in granulomas. Collectively, results do not support the role of M. chelonae in the development of granulomas and suggest rejecting the hypothesis of a potential link between NTM and SG. Full article
(This article belongs to the Section Anatomy, Histology and Pathology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Affected meagre showing severe hemorrhages and bilateral exophthalmia. (<b>b</b>) Heart of meagre with visible white nodules (red arrows) on the epicardium. Bar: 0.2 cm.</p>
Full article ">Figure 2
<p>(<b>a</b>) Multifocal granulomas in the heart (Hematoxylin and eosin—H&amp;E. Bar: 100 µm). (<b>b</b>) Multifocal granulomas in the liver (H&amp;E. Bar: 100 µm). (<b>c</b>) High power field of a granuloma in the kidney characterized by a necrotic hypereosinophilic center (asterisk) surrounded by epithelioid and spindle cells arranged in concentric layers (arrows) (H&amp;E. Bar: 20 µm). (<b>d</b>) Negative Ziehl–Neelsen stain of a granuloma in the kidney (ZN. Bar: 20 µm). (<b>e</b>) Focal granuloma in the brain (thick arrows) (H&amp;E. Bar: 100 µm).</p>
Full article ">Figure 3
<p>Bar chart illustrating the relative abundance (in percentage) of the main microbial taxa at the phylum level in the brain, heart, spleen, kidney, and intestine of a meagre affected by systemic granulomatosis.</p>
Full article ">Figure 4
<p>Maximum Likelihood tree shows the hsp65 sequences obtained from systemic granulomatosis-affected meagre cluster with the <span class="html-italic">Mycobacterium chelonae</span>. <span class="html-italic">Nocardia farcinica</span> was selected as the outgroup. The other sequences are the most probable species causing mycobacteriosis in fish. Evolutionary analyses were conducted in MEGA7, performed using Maximum Likelihood method, Tamura–Nei model, and bootstrap 1000 replicates.</p>
Full article ">Figure 5
<p>(<b>a</b>) Granuloma in meagre’s liver without in situ hybridization signals (Hematoxylin counterstain. Bar: 20 µm). (<b>b</b>) Numerous, 1–2 microns in length, bacillary red rods (red chromogen) inside a granuloma of a <span class="html-italic">Carassius auratus</span> experimentally infected with <span class="html-italic">M. chelonae</span> (Hematoxylin counterstain. Bar: 20 µm).</p>
Full article ">
5 pages, 185 KiB  
Brief Report
European EHBP1L1 Genotyping Survey of Dyserythropoietic Anemia and Myopathy Syndrome in English Springer Spaniels
by Sarah Østergård Jensen, Alexandra Kehl and Urs Giger
Vet. Sci. 2024, 11(12), 596; https://doi.org/10.3390/vetsci11120596 - 26 Nov 2024
Viewed by 478
Abstract
Dyserythropoietic anemia and myopathy syndrome (DAMS) with neonatal losses was recently characterized as an autosomal recessive disorder caused by an EHBP1L1 frameshift variant in English Springer Spaniels (ESSPs). The frequency and dissemination of the mutation remained unknown. The EHBP1L1 protein is essential for [...] Read more.
Dyserythropoietic anemia and myopathy syndrome (DAMS) with neonatal losses was recently characterized as an autosomal recessive disorder caused by an EHBP1L1 frameshift variant in English Springer Spaniels (ESSPs). The frequency and dissemination of the mutation remained unknown. The EHBP1L1 protein is essential for muscle function, and the Rab8/10-EHBP1L1-Bin1-dynamin axis participates in nuclear polarization during the enucleation of erythroblasts. Lack of EHBP1L1 function decreases enucleation, leading to increased numbers of nucleated erythrocytes, which are characteristic of DAMS. A genotyping survey for the EHBP1L1 variant was conducted based upon submitted samples of ESSPs from Europe. DNA was extracted, and a real-time PCR assay, with allele-specific TaqMan probes for EHBP1L1 wild-type and frameshift deletion, was applied. Between September 2022 and August 2024, 803 samples were received from 18 European countries. The EHBP1L1 mutant allele frequency was 9.7%, including 4 homozygous dogs and 148 heterozygotes. The mutant EHBP1L1 allele was found in 13 countries. A total of 6 homozygous and 73 heterozygous ESSPs reported on an open database could be tracked to an original common ancestor. Although the survey is biased, it indicates that the mutant EHBP1L1 variant is disseminated in the breed and across Europe. The genotyping of ESSPs is recommended to diagnose DAMS and guide breeders. Full article
(This article belongs to the Section Veterinary Internal Medicine)
14 pages, 2373 KiB  
Article
Prevalence, Distribution and Risk Factors for Trematode Infections in Domesticated Ruminants in the Lake Victoria and Southern Highland Ecological Zones of Tanzania: A Cross-Sectional Study
by Godlisten Shedrack Materu, Jahashi Nzalawahe, Mita Eva Sengupta, Anna-Sofie Stensgaard, Abdul Katakweba, Birgitte J. Vennervald and Safari Kinung’hi
Vet. Sci. 2024, 11(12), 595; https://doi.org/10.3390/vetsci11120595 - 25 Nov 2024
Viewed by 441
Abstract
Trematode infections cause long-term suffering and debilitation, posing a significant threat to global animal health and production and leading to considerable economic losses. Studies on the epidemiology and control of these infections in Tanzania are limited. The few available studies have been conducted [...] Read more.
Trematode infections cause long-term suffering and debilitation, posing a significant threat to global animal health and production and leading to considerable economic losses. Studies on the epidemiology and control of these infections in Tanzania are limited. The few available studies have been conducted in abattoir settings. This study aimed to fill this knowledge gap by determining the prevalence, distribution, and risk factors for trematode infections in domesticated ruminants in two different ecological zones of Tanzania. A cross-sectional study was conducted in the Lake Victoria and the Southern highlands ecological zones of Tanzania. Rectal fecal samples were collected and examined for F. gigantica, Paramphistomidae, and S. bovis infections using the sedimentation technique. A total of 1367 domesticated ruminants were sampled and examined for trematode infections. The overall prevalence of trematode infections was found to be 65.7%. The individual prevalence of F. gigantica, Paramphistomidae, and S. bovis (based on egg morphology only) was 35.1%, 60.2%, and 3.1%, respectively. Adult cattle were more likely to be infected with Paramphistomidae (adjusted odds ratio, (AOR): 1.98; 95% confidence interval, (CI): 1.40–2.78) and S. bovis (AOR: 8.5; 95% CI: 1.12–64.19) than weaners. It was observed that trematode infections in domesticated ruminants are prevalent in the two ecological zones of Tanzania; therefore, effective and community-directed prevention and control strategies are highly needed to address trematode infections of domesticated ruminants in these areas. Full article
(This article belongs to the Section Veterinary Microbiology, Parasitology and Immunology)
Show Figures

Figure 1

Figure 1
<p>A map showing the study areas: (<b>A</b>) Tanzania administrative map, (<b>B</b>) Bariadi District Council, (<b>C</b>) Misungwi District Council and (<b>D</b>) Iringa Rural District Council.</p>
Full article ">Figure 2
<p><span class="html-italic">Fasciola</span> egg (<b>A</b>), <span class="html-italic">S. bovis</span> egg (<b>B</b>) and Paramphistomidae egg (<b>C</b>).</p>
Full article ">Figure 3
<p>Administrative map of Tanzania showing the distribution of trematode infections in the study villages in Misungwi, Bariadi and Iringa rural districts, Tanzania. (<b>A</b>) Tanzania administrative map, (<b>B</b>) Bariadi District Council, (<b>C</b>) Misungwi District Council and (<b>D</b>) Iringa Rural District Council.</p>
Full article ">
10 pages, 679 KiB  
Article
Echocardiographic Changes in Dogs with Stage B2 Myxomatous Mitral Valve Disease Treated with Pimobendan Monotherapy
by Andrew Crosland, Pablo Manuel Cortes-Sanchez, Siddharth Sudunagunta, Jonathan Bouvard, Elizabeth Bode, Geoff Culshaw and Joanna Dukes-McEwan
Vet. Sci. 2024, 11(12), 594; https://doi.org/10.3390/vetsci11120594 - 25 Nov 2024
Viewed by 506
Abstract
The present study aimed to evaluate the effects of chronic pimobendan monotherapy on cardiac size in dogs with stage B2 myxomatous mitral valve disease (MMVD). Data from 31 dogs diagnosed with MMVD and cardiomegaly (LA/Ao ≥ 1.6 and LVIDdn ≥ 1.7) were included. [...] Read more.
The present study aimed to evaluate the effects of chronic pimobendan monotherapy on cardiac size in dogs with stage B2 myxomatous mitral valve disease (MMVD). Data from 31 dogs diagnosed with MMVD and cardiomegaly (LA/Ao ≥ 1.6 and LVIDdn ≥ 1.7) were included. The intervention group were dogs treated with pimobendan (n = 24). Dogs not receiving any cardiac medication were controls (n = 7). Echocardiographic changes in left atrial and left ventricular dimensions were compared over time. There was significant group × time interaction for LVIDdN (p = 0.011) between diagnosis and initial follow-up (median 3–6 months). There was a significant reduction in LVIDdN over time in the pimobendan group (p = 0.038) but not in the control group (p = 0.216). There was no significant group × time interaction for LA/Ao, and there was no effect of group (p = 0.561), but LA/Ao in both groups decreased over time (p = 0.01). Restraint is advised when prescribing pimobendan based on the detection of a heart murmur where echocardiographic staging is an option. Some stage B2 dogs that received pimobendan no longer met the echocardiographic classification criteria for stage B2 MMVD and could have been misclassified as stage B1 and had their medication inappropriately withdrawn. We suggest these dogs are referred to as reverse remodelled stage B2. Full article
(This article belongs to the Section Veterinary Internal Medicine)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Box and whisker plot demonstrating the distribution of LVIDdN across the pimobendan group and control group at the time of enrollment (Time 0) and at first recheck (Time 1). Boxes represent the interquartile range, with the median indicated with the line dividing the box. Outliers are represented by solid dots. (<b>b</b>) Box and whisker plot demonstrating the distribution of LA/Ao across the pimobendan group and control group at the time of enrollment (Time 0) and at first recheck (Time 1). Boxes represent the interquartile range, with the median indicated with the line dividing the box. Outliers are represented by solid dots.</p>
Full article ">
17 pages, 1227 KiB  
Article
Mutational Landscape of KIT Proto-Oncogene Coding Sequence in 62 Canine Cutaneous and Subcutaneous Mast Cell Tumors
by Ludovica Montanucci, Elena Guidolin, Rosa Maria Lopparelli, Greta Mucignat, Marianna Pauletto, Mery Giantin and Mauro Dacasto
Vet. Sci. 2024, 11(12), 593; https://doi.org/10.3390/vetsci11120593 - 25 Nov 2024
Viewed by 529
Abstract
Canine mast cell tumors (MCTs) are common skin neoplasms with varying biological behaviors. The KIT proto-oncogene plays a key role in the development of these tumors, and internal tandem duplications on exon 11 are usually associated with more aggressive behavior, increased local recurrence, [...] Read more.
Canine mast cell tumors (MCTs) are common skin neoplasms with varying biological behaviors. The KIT proto-oncogene plays a key role in the development of these tumors, and internal tandem duplications on exon 11 are usually associated with more aggressive behavior, increased local recurrence, and decreased survival time. However, apart from exons 8–11 and 17, there is limited understanding of the overall KIT mutational landscape in canine MCTs. This work aims to analyze the entire KIT coding sequence (21 exons) in a cohort of 62 MCTs, which included 38 cutaneous and 24 subcutaneous tumors, and potentially identify new variants. In addition to confirming previously reported activating KIT mutations in exons 8, 9, and 11, we identified new variants in exons 2, 3, 5, 16, and the 3′ untranslated region (UTR). Notably, these last variants include an amino acid change (Asp/His) in exon 16. Additionally, we confirmed a differential prevalence of KIT variants in cutaneous and subcutaneous MCTs. These findings enhance our understanding of the KIT proto-oncogene coding sequence and provide valuable information for future confirmatory studies. Full article
(This article belongs to the Special Issue Genetic Diseases and Gene Mutation-Related Tumors in Small Animals)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the Type III tyrosine kinase receptor encoded by the <span class="html-italic">KIT</span> gene. Ig1–5 indicates the extracellular immunoglobulin-like domains; TMD indicates the transmembrane domain; JMD indicates the intracellular juxtamembrane domain; TK1 and TK2 indicate the tyrosine kinase domains. The PDB (Protein Data Bank) structures used for the visualizations are the structures of the human tyrosine kinase: 8DFP for the extracellular region (exons 1–9, top) and 7ZW8 for the intracellular region (exons 11–21, bottom).</p>
Full article ">Figure 2
<p>cDNA sequence of exons 11 and 12 and its protein translation. The black cDNA sequence indicates exon 11, while the blue cDNA sequence indicates exon 12. The individual codons are highlighted with alternating clear and pink shadows. The ITDs and the complex insertion–duplication are represented in each subsequent line. For each variant, the sequence of only the duplicated region is reported (bold, green). The sequence highlighted in yellow corresponds to the 7 bp insertion of INS-DUP579.</p>
Full article ">
19 pages, 3236 KiB  
Article
The Role of Dual Mutations G347E and E349D of the Pigeon Paramyxovirus Type 1 Hemagglutinin–Neuraminidase Protein In Vitro and In Vivo
by Yu Chen, Junhao Gong, Tiansong Zhan, Mingzhan Wang, Shunlin Hu and Xiufan Liu
Vet. Sci. 2024, 11(12), 592; https://doi.org/10.3390/vetsci11120592 - 25 Nov 2024
Viewed by 571
Abstract
Pigeon Newcastle disease (ND) is the most common viral infectious disease in the pigeon industry, caused by pigeon paramyxovirus type 1 (PPMV-1), a variant of chicken-origin Newcastle disease virus (NDV). Previous studies have identified significant amino acid differences between PPMV-1 and chicken-origin NDV [...] Read more.
Pigeon Newcastle disease (ND) is the most common viral infectious disease in the pigeon industry, caused by pigeon paramyxovirus type 1 (PPMV-1), a variant of chicken-origin Newcastle disease virus (NDV). Previous studies have identified significant amino acid differences between PPMV-1 and chicken-origin NDV at positions 347 and 349 in the hemagglutinin–neuraminidase (HN) protein, with PPMV-1 predominantly exhibiting glycine (G) at position 347 and glutamic acid (E) at position 349, while most chicken-origin NDVs show E at position 347 and aspartic acid (D) at position 349. However, the impact of these amino acid substitutions remains unclear. In this study, we generated a recombinant virus, NT-10-G347E/E349D, by introducing the G347E and E349D dual mutations into a PPMV-1 strain NT-10 using reverse genetics. The biological characteristics of NT-10 and NT-10-G347E/E349D were compared both in vitro and in vivo. In vitro, the G347E and E349D dual mutations reduce NT-10′s replication and neuraminidase activity in pigeon embryo fibroblast (PEF) cells while enhancing both in chicken embryo fibroblast (CEF) cells. Additionally, these mutations decrease NT-10′s binding affinity to the α-2,6 sialic acid receptor while significantly increasing its affinity for the α-2,3 receptor. In vivo, NT-10-G347E/E349D exhibited reduced pathogenicity in pigeons but increased pathogenicity in chickens compared to the parental NT-10 strain. The mutations also reduced the pigeon-to-pigeon transmission of NT-10 but enhanced its transmission from pigeons to chickens. Notably, significant antigenic differences were observed between NT-10 and NT-10-G347E/E349D, as an inactivated vaccine based on NT-10 provided full protection against NT-10 challenge in immunized pigeons but only 67% mortality protection against NT-10-G347E/E349D. Overall, these findings underscore the critical role of amino acids at positions 347 and 349 in PPMV-1 infection, pathogenicity, and transmission, providing a theoretical foundation for the scientific prevention and control of PPMV-1. Full article
(This article belongs to the Section Veterinary Microbiology, Parasitology and Immunology)
Show Figures

Figure 1

Figure 1
<p>Conservation analysis of amino acid residues at positions 345 to 353 in the HN protein of PPMV-1 and genotype XX NDV. A total of 136 NDV HN gene sequences, including 127 PPMV-1 strains and 9 genotype XX NDV strains, were downloaded from the NCBI database. Conservation analysis of amino acid residues at positions 345 to 353 in the HN protein was conducted for genotype XX NDV (<b>A</b>) and PPMV-1 (<b>B</b>) using WebLogo 3 software.</p>
Full article ">Figure 2
<p>Growth kinetics of NT-10 and NT-10-G347E/E349D in PEF and CEF cells. PEF (<b>A</b>) or CEF (<b>B</b>) cells were seeded in 6-well plates at a density of 2 × 10⁵ cells per well and infected with the indicated virus at a multiplicity of infection (MOI) of 0.01. The infected cells were cultured in Dulbecco’s modified Eagle’s medium supplemented with 1% FBS at 37 °C under 5% CO<sub>2</sub>. Cellular supernatants were collected at indicated time points post-infection (12 h, 24 h, 36 h, 48 h, and 72 h), and viral titers were quantified as TCID<sub>50</sub> using the Reed and Muench method. Error bars represent standard deviations (SDs) from triplicate analyses of three independent experiments. Statistical significance was assessed using a two-way ANOVA in GraphPad Prism 8 software. NS means no significant difference, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Sialic acid receptor-binding properties of NT-10 and NT-10-G347E/E349D. The direct binding of NT-10 and NT-10-G347E/E349D viruses to sialyl-glycopolymers containing varying concentrations of 3′SLN (α-2,3-linked sialic acids) (<b>A</b>) or 6′SLN (α-2,6-linked sialic acids) (<b>B</b>) was measured using a solid-phase receptor-binding assay. Viruses (2<sup>7</sup> HA units) were incubated with 96-well streptavidin-coated plates coated with biotinylated glycans at descending concentrations for 2 h at 4 °C. Post-incubation, plates were treated with antiserum specific to each virus, followed by horseradish peroxidase (HRP)-conjugated rabbit anti-chicken IgY antibody (1:2000). After washing, TMB substrate was added for color development, and the reaction was stopped by adding 1 mol/L H<sub>2</sub>SO<sub>4</sub>. The absorbance was measured at 450 nm using a microplate reader. All assays were performed in triplicate. Error bars represent SDs from triplicate analyses of three independent experiments. Statistical significance was evaluated using a two-way ANOVA in GraphPad Prism 8 software.</p>
Full article ">Figure 4
<p>NA activities of HN and G347E/E349D-mutated HN proteins in PEF and CEF cells. PEF and CEF cells were seeded in 6-well plates at a density of 2 × 10⁵ cells per well and transfected with 1 µg of pCA-HN or pCA-HN-G347E/E349D plasmids using the EL Transfection Reagent according to the manufacturer’s protocol. An empty vector (1 µg) was included as a control. At 72 hpt, cells were lysed using RIPA buffer, and HN protein levels in PEF (<b>A</b>) and CEF (<b>C</b>) cells were measured by Western blot. Neuraminidase activities in PEF (<b>B</b>) and CEF (<b>D</b>) cells were assessed using a neuraminidase assay kit. The neuraminidase activities were normalized to the values for pCA-HN transfected cells. Error bars represent SDs from triplicate analyses of three independent experiments. Statistical significance was analyzed using one-way ANOVA in GraphPad Prism 8 software. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (The original images for blots in <a href="#app1-vetsci-11-00592" class="html-app">Supplementary Figure S1</a>).</p>
Full article ">Figure 5
<p>Evaluation of the pathogenicity of NT-10 and NT-10-G347E/E349D in pigeons and chickens. One-month-old White King pigeons (<span class="html-italic">n</span> = 9) and four-week-old SPF chickens (<span class="html-italic">n</span> = 9) were intranasally infected with 100 μL containing 10⁶ EID50 of either NT-10 or NT-10-G347E/E349D. PBS was used as a control. Six birds in each group were monitored daily for survival over a 14-day period, and survival curves for pigeons (<b>A</b>) and chickens (<b>B</b>) were generated using GraphPad Prism 8 software. At 5 dpi, three pigeons (<b>C</b>) or chickens (<b>D</b>) from each group were euthanized, and lung, trachea, duodenum, and spleen tissues were collected for histopathological analysis. The tissues were fixed in 10% neutral-buffered formalin, dehydrated, embedded in paraffin, sectioned, and stained with hematoxylin and eosin (H&amp;E). Lesions, such as inflammatory cell infiltration and tissue damage, are marked with red arrows. Images were captured at 100× magnification with a scale bar representing 200 μm. Statistical significance was analyzed using the Gehan–Breslow–Wilcoxon test in GraphPad Prism software.</p>
Full article ">Figure 6
<p>Diagram of the experimental procedure. Briefly, groups of ten White King pigeons were inoculated via intranasal routes with 100 μL containing 10<sup>4</sup> EID<sub>50</sub> of either NT-10 or NT-10-G347E/E349D. At 12 hpi, infected pigeons were divided into two groups (5 pigeons per group), and 5 additional pigeons or SPF chickens were placed into the same isolation units for the direct-contact transmission studies. The cloacal swabs for virus shedding detection were collected from inoculated and contact birds at indicated time points. The virus in the collected swabs were detected in the SPF embryonated chicken eggs.</p>
Full article ">Figure 7
<p>Analysis of antigenic differences between NT-10 and NT-10-G347E/E349D. One-month-old pigeons (<span class="html-italic">n</span> = 6 per group) were intramuscularly inoculated with NT-10-based inactivated vaccines at a dose of 10⁴ EID50 per bird. Fourteen days post-vaccination, the pigeons were intranasally challenged with 100 μL containing 10⁷ EID50 of either NT-10 or NT-10-G347E/E349D. PBS was used as a control. Survival was monitored daily for 14 days, and survival curves were generated using GraphPad Prism 8 software (<b>A</b>). At 7 dpi, three pigeons from each group were euthanized, and lung, trachea, duodenum, and spleen tissues were collected for histopathological analysis. The tissues were fixed, embedded, sectioned, and stained with H&amp;E. Pathological lesions, such as bleeding, inflammation, and tissue degeneration, are marked with red arrows (<b>B</b>). Images were captured at 100× magnification with a scale bar representing 200 μm.</p>
Full article ">
19 pages, 2933 KiB  
Article
Expression of GnRH, Kisspeptin, and Their Specific Receptors in the Ovary and Uterus in Deslorelin-Treated Late-Prepubertal Bitches
by Muhammet Ali Karadağ, Aykut Gram, Sabine Schäfer-Somi, Selim Aslan and Duygu Kaya
Vet. Sci. 2024, 11(12), 591; https://doi.org/10.3390/vetsci11120591 - 25 Nov 2024
Viewed by 435
Abstract
In this study, the expression and localization of gonadotropin-releasing hormone (GnRH1) and kisspeptin (KISS1) and their specific receptors in canine ovarian and uterine tissues were investigated after the application of deslorelin acetate (Suprelorin®, 4.7 mg, Virbac, France) in the late prepubertal [...] Read more.
In this study, the expression and localization of gonadotropin-releasing hormone (GnRH1) and kisspeptin (KISS1) and their specific receptors in canine ovarian and uterine tissues were investigated after the application of deslorelin acetate (Suprelorin®, 4.7 mg, Virbac, France) in the late prepubertal period. We hypothesized that prolonged treatment of prepubertal dogs with deslorelin would alter the expression of GnRH and kisspeptin genes in the uterus and ovaries. Ovarian and uterine samples of 25 dogs with an average age of 7.8 ± 0.2 months and from mixed breeds were used. Following implant insertion, dogs entered estrus (EST; n = 6); dogs without estrus (N-EST; n = 10) comprised the experimental groups. Nine dogs with placebo implants served as a control (CONT). Ovarian and uterine tissues were investigated for expression of GnRH1, GnRHR, KISS1, and KISS1R/GPR54 mRNA and protein by using IHC and RT-qPCR. In the uterus, expression of GnRH1 significantly decreased in response to deslorelin treatment in the N-EST, compared with the control group. Compared with CONT, KISS1R expression in ovarian samples was significantly lower in the EST group. Uterine protein expression of GnRH1 appeared weaker in N-EST than in CONT. While GnRH1-system members and KISS1 protein were localized in the follicles at various stages and stroma, no or only weak signals were detected for KISS1R in the ovarian samples. Deslorelin-mediated induction of puberty by changing the expression of some of the GnRH and KISS1-system members seems to have an effect on ovarian and uterine functionality. Deslorelin implants can, therefore, not be considered a valuable alternative to induce fertile estrus in late-prepubertal bitches. However, further studies with a larger number of animals are needed to clarify the effect of deslorelin-mediated induction of puberty. Full article
Show Figures

Figure 1

Figure 1
<p>Study design.</p>
Full article ">Figure 2
<p>Mean serum E2 (<b>A</b>) and P4 (<b>B</b>) concentrations of the groups on the day of ovariohysterectomy. (EST: Estrus, N-EST: Non-Estrus, CONT: Control, the sensitivity of the measurements is 100%, and the specificity is 95.5%).</p>
Full article ">Figure 3
<p>Expression of GnRH1 (<b>A</b>), GnRHR (<b>B</b>), KISS1 (<b>C</b>), and KISS1R (<b>D</b>) in uterine tissue. (EST: Estrus, N-EST: Non-Estrus, CONT: Control). All numerical data are presented as mean ± SEM. <span class="html-italic">p</span> values &lt; 0.05 were considered significant and indicated. For comparison between groups, the parametric one-way ANOVA was followed by the Tukey multiple comparison post-hoc test.</p>
Full article ">Figure 4
<p>Expression of GnRH1 (<b>A</b>), GnRHR (<b>B</b>), KISS1 (<b>C</b>), and KISS1R (<b>D</b>) in ovarian tissue. (EST: Estrus, N-EST: Non-Estrus, CONT: Control). All numerical data are presented as mean ± SEM. <span class="html-italic">p</span> values &lt; 0.05 were considered significant and indicated. For comparison between groups, the parametric one-way ANOVA was followed by the Tukey multiple comparison post-hoc test.</p>
Full article ">Figure 5
<p>Immunolocalization of GnRH1, GnRHR, and KISS1 in uterine tissue. Immunoreactivity of GnRH1 in different areas of the uterus by groups; EST (<b>A</b>), N-EST (<b>E</b>), CONT (<b>I</b>), (bar: 50 μm-X20). Immunoreactivity of GnRHR in different areas of the uterus by groups: EST (<b>B</b>), N-EST (<b>F</b>), CONT (<b>J</b>), (bar: 50 μm-X20). Immunoreactivity of KISS1 in different areas of the uterus by groups: EST (<b>C</b>), N-EST (<b>G</b>), CONT (<b>K</b>), (bar: 50 μm-X20). No reaction was observed in the isotype controls (<b>D</b>,<b>H</b>,<b>L</b>, bar: 50 μm-20x) (ms: myometrial stroma, le: luminal epithelial cells, black thick arrow: endometrial stroma, arrowhead: myocytes, white arrow: endometrial glands, *: interstitial cells).</p>
Full article ">Figure 6
<p>Immunolocalization of GnRH1, GnRHR, and KISS1 in ovarian tissue. Immunoreactivity of GnRH1 in different areas of the ovary by groups; EST (<b>A</b>), N-EST (<b>E</b>), CONT (<b>I</b>), (bar: 20 μm-X40). Immunoreactivity of GnRHR in different areas of the ovary by groups: EST (<b>B</b>), N-EST (<b>F</b>), CONT (<b>J</b>), (bar: 20/50 μm-X20/40). Immunoreactivity of KISS1 in different areas of the ovary by groups: EST (<b>C</b>), N-EST (<b>G</b>), CONT (<b>K</b>), (bar: 20 μm-X40). No reaction was observed in the isotype controls (<b>D</b>,<b>H</b>,<b>L,</b> bar: bar: 20/50 μm-X20/40). (cl: corpus luteum, gf: Graafian follicle, sf: secondary follicle, pf: primary follicle, zp: zona pellucida, o: oocyt, black arrow: granulosa cell, arrowhead: theca cell, white arrow: endocrine interstitial cell).</p>
Full article ">
12 pages, 785 KiB  
Article
Epidemiology and Molecular Characterization of Entamoeba spp. in Non-Human Primates in Zoos in China
by Diya An, Shui Yu, Tingting Jiang, Jianhui Zhang, Qun Liu and Jing Liu
Vet. Sci. 2024, 11(12), 590; https://doi.org/10.3390/vetsci11120590 - 25 Nov 2024
Viewed by 350
Abstract
The genus Entamoeba infects both humans and NHPs. In zoos, visitors feeding significantly increases the frequency of human-to-NHP contact, thereby raising the risk of zoonotic transmission. In this study, six Entamoeba species were investigated and analyzed in the fecal samples of 14 NHP [...] Read more.
The genus Entamoeba infects both humans and NHPs. In zoos, visitors feeding significantly increases the frequency of human-to-NHP contact, thereby raising the risk of zoonotic transmission. In this study, six Entamoeba species were investigated and analyzed in the fecal samples of 14 NHP species from zoos in Beijing, Guiyang, Shijiazhuang, Tangshan, and Xingtai in China. A total of 19 out of 84 primate fecal samples tested positive for Entamoeba spp. by polymerase chain reaction (PCR). Among these, 14 samples contained mono-detections of E. coli (7/84), E. dispar (4/84), and E. polecki (3/84). Five samples were found to have mixed detections with two or three species, suggesting the potential for zoonotic transmission; however, no pathogenic E. histolytica, E. moshkovskii, or E. nuttalli were detected. This study provides new insights into parasitic detections in NHPs in Chinese zoos and offers valuable background information for the prevention and control of zoonotic parasitic diseases. Full article
(This article belongs to the Section Veterinary Food Safety and Zoonosis)
Show Figures

Figure 1

Figure 1
<p>Phenetic relationships of <span class="html-italic">Entamoeba</span> spp. Numbers on the branches are percent bootstrapping values from 1000 replicates, only bootstrap values &gt; 50 are indicated. The accession numbers utilized for the identification of <span class="html-italic">Entamoeba</span> spp. were AB444953 (<span class="html-italic">E. coli</span>), AB282661 (<span class="html-italic">E. dispar</span>), AF149907 (<span class="html-italic">E. hartmanni</span>), X65163 (<span class="html-italic">E. histolytica</span>), AF149906 (<span class="html-italic">E. moshkoskii</span>), AB282657 (<span class="html-italic">E. nuttalli</span>), AF149913 (<span class="html-italic">E. polecki-like</span> variant 1), and AF149912 (<span class="html-italic">E. polecki-like</span> variant 4).</p>
Full article ">
10 pages, 1122 KiB  
Article
Establishment of a Real-Time Reverse Transcription Recombinase-Aided Isothermal Amplification (qRT-RAA) Assay for the Rapid Detection of Bovine Respiratory Syncytial Virus
by Guanxin Hou, Siping Zhu, Hong Li, Chihuan Li, Xiaochen Liu, Chao Ren, Xintong Zhu, Qiumei Shi and Zhiqiang Zhang
Vet. Sci. 2024, 11(12), 589; https://doi.org/10.3390/vetsci11120589 - 24 Nov 2024
Viewed by 368
Abstract
Background: Bovine respiratory syncytial virus (BRSV) is a significant cause of bovine respiratory disease, resulting in significant losses to the cattle industry. For rapid detection of BRSV, a real-time recombinase-aided isothermal amplification assay (qRT-RAA) based on the F gene of BRSV was developed [...] Read more.
Background: Bovine respiratory syncytial virus (BRSV) is a significant cause of bovine respiratory disease, resulting in significant losses to the cattle industry. For rapid detection of BRSV, a real-time recombinase-aided isothermal amplification assay (qRT-RAA) based on the F gene of BRSV was developed in this study. Results: The developed qRT-RAA assay showed good exponential amplification of the target fragment in 20 min at a constant temperature of 39 °C. And this assay displayed a high specificity for BRSV, without cross-reactions with Infectious Bovine Rhinotracheitis Virus (IBRV), Bovine Parainfluenza Virus Type 3 (BPIV3), Bovine Viral Diarrhea Virus (BVDV), and Bovine Coronavirus (BCoV). With the standard RNA of BRSV serving as a template, the limit of detection for qRT-RAA was 102 copies/μL. We examined ninety-seven clinical samples from cattle with respiratory disease using this method and determined a positive rate of 7.2% (7/97), consistent with results using the classical PCR method reported previously. Conclusions: A qRT-RAA assay for BRSV detection was established in this study. The method is specific and sensitive and can be completed within 20 min at 39 °C. These works demonstrate that the generated qRT-RAA assay is an effective diagnostic tool for rapidly detecting BRSV in resource-limited settings, which may be applied for the clinical detection of BRSV. Full article
Show Figures

Figure 1

Figure 1
<p>Determination of optimal primers and reaction conditions for BRSV RT-RAA methods. (<b>A</b>) Lane M: DL2000 DNA Marker (Takara). Determination of optimal primer pairs for BRSV RT-RAA that target the <span class="html-italic">F</span> gene (numbers 1–3 correspond to BRSVF1/R1, BRSVF2/R2, and BRSVF3/R3). A 5 μL template containing 10<sup>4</sup> copies/μL of <span class="html-italic">F</span> gene RNA was added to each reaction system and carried out at 39 °C for 30 min. The amplification product was analyzed by electrophoretic in 1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) agarose gel. (<b>B</b>) Initial specificity assays. The primer pairs BRSV F1/R1 and 5 μL template were added (numbers 1–5 correspond to BRSV, BVDV, BPIV, IBRV, and BCoV). The reactions were performed at 39 °C for 30 min. (<b>C</b>) Results of optimal temperature screening for BRSV RT-RAA. Primer pair BRSV F1/R1 was selected for amplification using 5 μL template harboring 10<sup>4</sup> copies/μL of <span class="html-italic">F</span> gene RNA with an incubation time of 30 min and different incubation temperatures (numbers 1–4 correspond to 37 °C, 38 °C, 39 °C, and 40 °C). (<b>D</b>) Results of optimal time screening for BRSV real-time fluorescent RT-RAA. Primer pair BRSV F1/R1 was selected for basal-type amplification using <span class="html-italic">F</span> gene RNA as a template. The incubation temperature was 39 °C, and the incubation time was different (numbers 1–4 correspond to 10 min, 20 min, 30 min, and 40 min). All the amplification products were electrophoresed on 1% agarose gel.</p>
Full article ">Figure 2
<p>Evaluation of the BRSV RT-RAA detection method. RFU, Relative Fluorescence Unit. Each cycle denotes a 30 s reaction time. (<b>A</b>) The RT-RAA was performed with an RNA standard of 10<sup>10</sup> copies under 39 °C for 20 min. (<b>B</b>) Specificity analysis of RT-RAA assays. Nucleic acids of BRSV, BVDV, BPIV, IBRV, BCoV, and ddH<sub>2</sub>O were extracted and used as templates for RT-RAA assay under optimal conditions. (<b>C</b>) Sensitivity evaluation. RNA standards were serially 10-fold diluted to a range of 10<sup>8</sup> to 10<sup>0</sup> copies/μL and subjected to RT-RAA assays. (<b>D</b>) M: DL2000 DNA Marker (Takara). The former diluted RNA standards were taken as templates for typical PCR, as reported previously.</p>
Full article ">Figure 3
<p>Repeatability tests. RNA standards with concentrations of 10<sup>8</sup>, 10<sup>6</sup>, and 10<sup>4</sup> copies/μL were assayed using the RT-RAA system at 39 °C for 20 min. Each cycle denotes a 30 s reaction time. RFU is Relative Fluorescence Unit. (<b>A</b>) Inter-batch repeat test; (<b>B</b>) intra-batch repeat test.</p>
Full article ">
Previous Issue
Back to TopTop