[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

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

Search Results (43,126)

Search Parameters:
Keywords = genome

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 977 KiB  
Review
Tumour Heterogeneity and Disease Infiltration as Paradigms of Glioblastoma Treatment Resistance
by Pulkit Malhotra and Ruman Rahman
Onco 2024, 4(4), 349-358; https://doi.org/10.3390/onco4040024 (registering DOI) - 18 Oct 2024
Abstract
Isocitrate dehydrogenase wild-type glioblastoma, a Grade 4 malignant brain neoplasm, remains resistant to multimodal treatment, with a median survival of 16 months from diagnosis with no geographical bias. Despite increasing appreciation of intra-tumour genotypic variation and stem cell plasticity, such knowledge has yet [...] Read more.
Isocitrate dehydrogenase wild-type glioblastoma, a Grade 4 malignant brain neoplasm, remains resistant to multimodal treatment, with a median survival of 16 months from diagnosis with no geographical bias. Despite increasing appreciation of intra-tumour genotypic variation and stem cell plasticity, such knowledge has yet to translate to efficacious molecular targeted therapies in this post-genomic era. Critically, the manifestation of molecular heterogeneity and stem cell biological process within clinically relevant infiltrative disease is little understood. Here, we review the interactions between neural plasticity, intra-tumour heterogeneity and residual infiltrative disease, and we draw upon antibiotic resistance as an insightful analogy to further explain tumour heterogeneity. Full article
Show Figures

Figure 1

Figure 1
<p>Aspects of intra-tumour heterogeneity and its relationship with clonal evolution (Darwinian evolution: selection of cells that have mutated to exhibit more aggressive phenotypes, followed by natural selection of these cells conferred by external selection pressures) and the Cancer Stem Cell Model (in which heterogeneity encompasses the spectrum of differentiated cells and plasticity between CSC and non-CSC). Created using BioRender.</p>
Full article ">Figure 2
<p>Schematic outlining the analogous mechanism of tumour cells undergoing random mutations, as well as bacterial populations being selected to allow resistant populations to multiply. Created in BioRender.com.</p>
Full article ">
16 pages, 6383 KiB  
Article
From Diabetes to Dementia: Identifying Key Genes in the Progression of Cognitive Impairment
by Zhaoming Cao, Yage Du, Guangyi Xu, He Zhu, Yinchao Ma, Ziyuan Wang, Shaoying Wang and Yanhui Lu
Brain Sci. 2024, 14(10), 1035; https://doi.org/10.3390/brainsci14101035 (registering DOI) - 18 Oct 2024
Abstract
Objectives: To provide a basis for further research on the molecular mechanisms underlying type 2 diabetes-associated mild cognitive impairment (DCI) using two bioinformatics methods to screen key genes involved in the progression of mild cognitive impairment (MCI) and type 2 diabetes. Methods: RNA [...] Read more.
Objectives: To provide a basis for further research on the molecular mechanisms underlying type 2 diabetes-associated mild cognitive impairment (DCI) using two bioinformatics methods to screen key genes involved in the progression of mild cognitive impairment (MCI) and type 2 diabetes. Methods: RNA sequencing data of MCI and normal cognition groups, as well as expression profile and sample information data of clinical characteristic data of GSE63060, which contains 160 MCI samples and 104 normal samples, were downloaded from the GEO database. Hub genes were identified using weighted gene co-expression network analysis (WGCNA). Protein–protein interaction (PPI) analysis, combined with least absolute shrinkage and selection operator (LASSO) and receiver operating characteristic (ROC) curve analyses, was used to verify the genes. Moreover, RNA sequencing and clinical characteristic data for GSE166502 of 13 type 2 diabetes samples and 13 normal controls were downloaded from the GEO database, and the correlation between the screened genes and type 2 diabetes was verified by difference and ROC curve analyses. In addition, we collected clinical biopsies to validate the results. Results: Based on WGCNA, 10 modules were integrated, and six were correlated with MCI. Six hub genes associated with MCI (TOMM7, SNRPG, COX7C, UQCRQ, RPL31, and RPS24) were identified using the LASSO algorithm. The ROC curve was screened by integrating the GEO database, and revealed COX7C, SNRPG, TOMM7, and RPS24 as key genes in the progression of type 2 diabetes. Conclusions: COX7C, SNRPG, TOMM7, and RPS24 are involved in MCI and type 2 diabetes progression. Therefore, the molecular mechanisms of these four genes in the development of type 2 diabetes-associated MCI should be studied. Full article
Show Figures

Figure 1

Figure 1
<p>The workflow of the analysis process.</p>
Full article ">Figure 2
<p>Differentially expressed genes (DEGs) between MCI and normal control samples, and functional enrichment analysis. The heatmap and volcano plot of DEGs were obtained by cluster analysis. (<b>A</b>) In the horizontal axis, Con represents the normal control group, MCI represents the cognitive impairment group, and the vertical axis represents the DEGs. (<b>B</b>) Volcano plot of differently expressed genes. Red indicates up-regulated genes, and green indicates down-regulated genes. Blue dotted line: Threshold coordinate line (<b>C</b>) GO terms. (<b>D</b>) KEGG pathways. (<b>E</b>) GSEA analysis of DEGs in the MCI group.</p>
Full article ">Figure 3
<p>Identification of co-expressed modules and relationship of modules and disease status by WGCNA. (<b>A</b>) soft threshold filtering. (<b>B</b>) Hierarchical clustering tree of WGCNA co-expression network. (<b>C</b>) Correlation between modules and disease status. (<b>D</b>) Barplot of mean gene significance (GS) across modules. (<b>E</b>) Heatmap plot of coexpressed genes. (<b>F</b>) GO terms for biological processes (BP), GO terms for cellular components (CC), GO terms for molecular function (MF). (<b>G</b>) KEGG pathways.</p>
Full article ">Figure 3 Cont.
<p>Identification of co-expressed modules and relationship of modules and disease status by WGCNA. (<b>A</b>) soft threshold filtering. (<b>B</b>) Hierarchical clustering tree of WGCNA co-expression network. (<b>C</b>) Correlation between modules and disease status. (<b>D</b>) Barplot of mean gene significance (GS) across modules. (<b>E</b>) Heatmap plot of coexpressed genes. (<b>F</b>) GO terms for biological processes (BP), GO terms for cellular components (CC), GO terms for molecular function (MF). (<b>G</b>) KEGG pathways.</p>
Full article ">Figure 4
<p>PPI network construction and key genes screening. (<b>A</b>), Diagram of protein–protein interaction (PPI) network of the blue module. The degree of the node was reflected by its size and color. The larger the node, the deeper the color from yellow to purple, and the greater the degree. (<b>B</b>), the network diagram of the top 10 genes screened via the MCC algorithm. The darker color, from yellow to red, represented a greater score. (<b>C</b>), LASSO screening analysis of related genes in the blue module. (<b>D</b>), Difference analysis of COX7C, SNRPG, TOMM7, UQCRQ, RPL31, and RPS 24 expressions in MCI and normal control group. (<b>E</b>), ROC curve detection of COX7C, SNRPG, TOMM7, UQCRQ, RPL31, and RPS24 to predict the occurrence of MCI. Significance: ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>(<b>A</b>), Difference analysis of COX7C, SNRPG, TOMM7, UQCRQ, RPL31, and RPS24 expression in type 2 diabetes and normal groups. (<b>B</b>), ROC curve detection of COX7C, SNRPG, TOMM7, and RPS24 to predict the occurrence of type 2 diabetes. Significance identification: ns, <span class="html-italic">p</span> ≥ 0.05; *, <span class="html-italic">p</span>&lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5 Cont.
<p>(<b>A</b>), Difference analysis of COX7C, SNRPG, TOMM7, UQCRQ, RPL31, and RPS24 expression in type 2 diabetes and normal groups. (<b>B</b>), ROC curve detection of COX7C, SNRPG, TOMM7, and RPS24 to predict the occurrence of type 2 diabetes. Significance identification: ns, <span class="html-italic">p</span> ≥ 0.05; *, <span class="html-italic">p</span>&lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>(<b>A</b>) Validation of 4 diagnostic feature biomarkers in clinical tissues via Western blot analysis. The expression of (<b>B</b>), TOMM7, (<b>C</b>), SNRPG, (<b>D</b>), RPS24, and (<b>E</b>), COX7C. Statistical significance: *<span class="html-italic">p</span> &lt; 0.05; **<span class="html-italic">p</span> &lt; 0.01; ***<span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 9050 KiB  
Article
Genome-Wide Identification of MKK Gene Family and Response to Hormone and Abiotic Stress in Rice
by Fan Zhang, Jingjing Wang, Yiwei Chen, Junjun Huang and Weihong Liang
Plants 2024, 13(20), 2922; https://doi.org/10.3390/plants13202922 (registering DOI) - 18 Oct 2024
Abstract
Mitogen-activated protein kinase (MAPK/MPK) cascades are pivotal and highly conserved signaling modules widely distributed in eukaryotes; they play essential roles in plant growth and development, as well as biotic and abiotic stress responses. With the development of sequencing technology, the complete genome assembly [...] Read more.
Mitogen-activated protein kinase (MAPK/MPK) cascades are pivotal and highly conserved signaling modules widely distributed in eukaryotes; they play essential roles in plant growth and development, as well as biotic and abiotic stress responses. With the development of sequencing technology, the complete genome assembly of rice without gaps, T2T (Telomere-to-Telomere)—NIP (version AGIS-1.0), has recently been released. In this study, we used bioinformatic approaches to identify and analyze the rice MPK kinases (MKKs) based on the complete genome. A total of seven OsMKKs were identified, and their physical and chemical properties, chromosome localization, gene structure, subcellular localization, phylogeny, family evolution, and cis-acting elements were evaluated. OsMKKs can be divided into four subgroups based on phylogenetic relationships, and the family members located in the same evolutionary branch have relatively similar gene structures and conserved domains. Quantitative real-time PCR (qRT-PCR) revealed that all OsMKKs were highly expressed in rice seedling leaves. The expression levels of all OsMKKs were more or less altered under exogenous hormone and abiotic stress treatments, with OsMKK1, OsMKK6, and OsMKK3 being induced under almost all treatments, while the expression of OsMKK4 and OsMKK10-2 was repressed under salt and drought treatments and IAA treatment, respectively. In this study, we also summarized the recent progress in rice MPK cascades, highlighted their diverse functions, and outlined the potential MPK signaling network, facilitating further studies on OsMKK genes and rice MPK cascades. Full article
(This article belongs to the Section Plant Genetics, Genomics and Biotechnology)
Show Figures

Figure 1

Figure 1
<p>Chromosomal distribution of <span class="html-italic">OsMKK</span> genes in rice. The duplicated <span class="html-italic">OsMKK</span> genes are shown in red dashed line. Shadows of the same color belong to the same group.</p>
Full article ">Figure 2
<p>Phylogenetic tree of the 7 OsMKK proteins and 33 known functional MKK proteins from other plants. Green stars represent OsMKKs, cyan triangles represent AtMKKs, and pink checkmarks are known function MKKs.</p>
Full article ">Figure 3
<p>Gene structure and conserved protein motifs analysis of <span class="html-italic">OsMKK</span> genes. (<b>A</b>) ML phylogenetic tree analysis of OsMKKs. (<b>B</b>) Exon–intron structure of <span class="html-italic">OsMKKs</span>, where golden yellow boxes represent coding sequences (CDS), the blue boxes represent upstream/downstream sequences, and the black lines represent the introns. (<b>C</b>) The conserved motifs in OsMKK proteins. The ten conversed motifs are displayed in various unique colors. The gene and protein length are indicated by the scale at bottom. (<b>D</b>) Sequence logos of ten conserved domains. The conserved sequences of the different motifs are highlighted in different colored rectangles. (<b>E</b>) Sequence alignment and motif analysis of OsMKKs. Identical amino acids are shaded black, and similar amino acids are shaded purple. The P, C, and T loops, CCD, and the NTF2 domain are highlighted in colored rectangles (P loop: red; C loop: green; T loop: blue; CDD: pink; NTF2: yellow). The red stars show the active site, and the green stars indicate the phosphorylation site of OsMKK proteins. Species information can be found in <a href="#app1-plants-13-02922" class="html-app">Figure S1</a>.</p>
Full article ">Figure 4
<p>Predicted 3D models of OsMKK proteins. Models have been generated by Alpha 2 and visualized by rainbow color from N (blue) to C terminus (red) using PyMOL v2.5.8 software.</p>
Full article ">Figure 5
<p>Expression profiling of 7 <span class="html-italic">OsMKK</span> genes in different organs and tissues. The red color represents high-level expression, while the blue color represents low-level expression. DAP, days after pollination.</p>
Full article ">Figure 6
<p>Expression patterns of 7 <span class="html-italic">OsMKK</span> genes in the roots, stems, and leaves of rice seedlings. Data are represented as the mean ± SD of three independent replicates. Asterisks indicate statistically significant differences compared with root (** <span class="html-italic">p</span> &lt; 0.01; Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 7
<p>Cis-acting elements in the promoter of <span class="html-italic">OsMKK</span> genes. (<b>A</b>) Numbers of predicted cis-acting elements in <span class="html-italic">OsMKK</span> promoters are shown. (<b>B</b>) The distribution of predicted cis-acting elements on different gene promoters. Different colors represent different cis-acting elements.</p>
Full article ">Figure 8
<p>Expression levels of <span class="html-italic">OsMKK</span> genes under ABA, GA, IAA, salt, and drought stress treatments. Data are represented as the mean ± SD of three independent replicates. Different letters above bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05; Tukey’s test).</p>
Full article ">Figure 9
<p>OsMKKs are involved in plant growth and development and diverse biotic and abiotic stresses. (<b>A</b>–<b>D</b>) represent the processes in which each subgroup participates [<a href="#B3-plants-13-02922" class="html-bibr">3</a>,<a href="#B35-plants-13-02922" class="html-bibr">35</a>,<a href="#B36-plants-13-02922" class="html-bibr">36</a>,<a href="#B39-plants-13-02922" class="html-bibr">39</a>,<a href="#B40-plants-13-02922" class="html-bibr">40</a>,<a href="#B41-plants-13-02922" class="html-bibr">41</a>,<a href="#B42-plants-13-02922" class="html-bibr">42</a>,<a href="#B45-plants-13-02922" class="html-bibr">45</a>,<a href="#B46-plants-13-02922" class="html-bibr">46</a>,<a href="#B47-plants-13-02922" class="html-bibr">47</a>,<a href="#B48-plants-13-02922" class="html-bibr">48</a>,<a href="#B49-plants-13-02922" class="html-bibr">49</a>,<a href="#B50-plants-13-02922" class="html-bibr">50</a>,<a href="#B51-plants-13-02922" class="html-bibr">51</a>,<a href="#B52-plants-13-02922" class="html-bibr">52</a>,<a href="#B53-plants-13-02922" class="html-bibr">53</a>,<a href="#B54-plants-13-02922" class="html-bibr">54</a>,<a href="#B55-plants-13-02922" class="html-bibr">55</a>].</p>
Full article ">
19 pages, 1159 KiB  
Review
Pheochromocytoma–Paraganglioma Syndrome: A Multiform Disease with Different Genotype and Phenotype Features
by Mara Giacché, Maria Chiara Tacchetti, Claudia Agabiti-Rosei, Francesco Torlone, Francesco Bandera, Claudia Izzi and Enrico Agabiti-Rosei
Biomedicines 2024, 12(10), 2385; https://doi.org/10.3390/biomedicines12102385 (registering DOI) - 18 Oct 2024
Abstract
Pheochromocytoma and paraganglioma (PPGL) are rare tumors derived from the adrenal medulla and extra-adrenal chromaffin cells. Diagnosis is often challenging due to the great variability in clinical presentation; the complexity of management due to the dangerous effects of catecholamine excess and the potentially [...] Read more.
Pheochromocytoma and paraganglioma (PPGL) are rare tumors derived from the adrenal medulla and extra-adrenal chromaffin cells. Diagnosis is often challenging due to the great variability in clinical presentation; the complexity of management due to the dangerous effects of catecholamine excess and the potentially malignant behavior require in-depth knowledge of the pathology and multidisciplinary management. Nowadays, diagnostic ability has certainly improved and guidelines and consensus documents for treatment and follow-up are available. A major impulse to the development of this knowledge has come from the new findings on the genetic and molecular characteristics of PPGLs. Germline mutation in susceptibility genes is detected in 40% of subjects, with a mutation frequency of 10–12% also in patients with sporadic presentation and genetic testing should be incorporated within clinical care. PPGL susceptibility genes include “old genes” associated with Neurofibromatosis type 1 (NF1 gene), Von Hippel Lindau syndrome (VHL gene) and Multiple Endocrine Neoplasia type 2 syndrome (RET gene), the family of SDHx genes (SDHA, SDHB, SDHC, SDHD, SDHAF2), and genes less frequently involved such as TMEM, MAX, and FH. Each gene has a different risk of relapse, malignancy, and other organ involvement; for mutation carriers, affected or asymptomatic, it is possible to define a tailored long-life surveillance program according to the gene involved. In addition, molecular characterization of the tumor has allowed the identification of somatic mutations in other driver genes, bringing to 70% the PPGLs for which we know the mechanisms of tumorigenesis. This has expanded the catalog of tumor driver genes, which are identifiable in up to 70% of patients Integrated genomic and transcriptomic data over the last 10 years have revealed three distinct major molecular signatures, triggered by pathogenic variants in susceptibility genes and characterized by the activation of a specific oncogenic signaling: the pseudo hypoxic, the kinase, and the Wnt signaling pathways. These molecular clusters show a different biochemical phenotype and clinical behavior; they may also represent the prerequisite for implementing customized therapy and follow-up. Full article
(This article belongs to the Special Issue Adrenal Diseases: An Update)
Show Figures

Figure 1

Figure 1
<p>68GA-DOTA-SSA PET in SDHD-mutated patient showing bilateral HN PGLs.</p>
Full article ">Figure 2
<p>Metastatic PGL in SDHB carrier (68GA-DOTA-SSA PET).</p>
Full article ">Figure 3
<p>Sporadic giant right adrenal pheochromocytoma (CT scan and I123-MIBG Scintigraphy).</p>
Full article ">
9 pages, 1977 KiB  
Article
Genome-Wide Identification and Expression Analysis of the CLAVATA3/ESR-Related Gene Family in Tiger Nut
by Maria Gancheva, Nina Kon’kova, Alla Solovyeva, Lavrentii Danilov, Konstantin Gusev and Ludmila Lutova
Int. J. Plant Biol. 2024, 15(4), 1054-1062; https://doi.org/10.3390/ijpb15040074 (registering DOI) - 18 Oct 2024
Abstract
CLAVATA3 (CLV3)/EMBRYO SURROUNDING REGION (ESR)-related (CLE) genes encode a group of peptide hormones, which coordinate cell proliferation and differentiation in plants. Tiger nut (Cyperus esculentus L.) is a perennial monocot plant that produces oil-rich tubers. [...] Read more.
CLAVATA3 (CLV3)/EMBRYO SURROUNDING REGION (ESR)-related (CLE) genes encode a group of peptide hormones, which coordinate cell proliferation and differentiation in plants. Tiger nut (Cyperus esculentus L.) is a perennial monocot plant that produces oil-rich tubers. However, the mechanisms regulating tuber development in tiger nut are poorly understood, and nothing is known about CLE genes in tiger nut. In this study, we identified 34 CLE genes in the genomes, proteomes, and transcriptomes of C. esculentus (CeCLE). We analyzed their gene structures and expression profiles in different parts of the plant, at three stages of tuber development and in roots in response to dehydration stress. We found a relatively high expression level of CeCLE13 in growing tuber and suggested that the corresponding CLE peptide could be involved in the regulation of tuberization. We also analyzed CeCLE gene sequences in the genome of the most productive K-17 variety in the N. I. Vavilov All-Russian Institute of Plant Genetic Resources collection and found many single nucleotide polymorphisms, insertions, and deletions. Our data provides fundamental information for future research on tiger nut growth and tuberization. Full article
(This article belongs to the Section Plant Biochemistry and Genetics)
Show Figures

Figure 1

Figure 1
<p>Characteristics of <span class="html-italic">C. esculentus</span> L. K-17. (<b>a</b>) Morphology. Scale bar = 1 cm. (<b>b</b>) Valuable agronomic characteristics obtained previously [<a href="#B25-ijpb-15-00074" class="html-bibr">25</a>] and in this study.</p>
Full article ">Figure 2
<p>Phylogenetic tree and sequence analysis of CeCLE and <span class="html-italic">Arabidopsis thaliana</span> CLE (AtCLE) peptides. Multiple CLE domains are designated “d1” to “d8”. Default coloring in the Ugene [<a href="#B26-ijpb-15-00074" class="html-bibr">26</a>] is used.</p>
Full article ">Figure 3
<p>Distribution of signal peptides and CLE domains in the <span class="html-italic">CeCLE</span> genes. Multiple CLE domains are designated “d1” to “d8”.</p>
Full article ">Figure 4
<p>The TPM values of <span class="html-italic">CeCLE</span> genes in two cultivars (XJ and Jisha 1) with hierarchical clustering in different parts of tiger nut, at three stages of tuber development (40, 80, and 120 days after sowing (40 d, 80 d, 120 d Tuber)), and in roots in response to dehydration stress. The colors from green to pink represent the range of the TPM values from high to low, respectively.</p>
Full article ">
9 pages, 266 KiB  
Article
Machine Learning for the Genomic Prediction of Growth Traits in a Composite Beef Cattle Population
by El Hamidi Hay
Animals 2024, 14(20), 3014; https://doi.org/10.3390/ani14203014 (registering DOI) - 18 Oct 2024
Abstract
The adoption of genomic selection is prevalent across various plant and livestock species, yet existing models for predicting genomic breeding values often remain suboptimal. Machine learning models present a promising avenue to enhance prediction accuracy due to their ability to accommodate both linear [...] Read more.
The adoption of genomic selection is prevalent across various plant and livestock species, yet existing models for predicting genomic breeding values often remain suboptimal. Machine learning models present a promising avenue to enhance prediction accuracy due to their ability to accommodate both linear and non-linear relationships. In this study, we evaluated four machine learning models—Random Forest, Support Vector Machine, Convolutional Neural Networks, and Multi-Layer Perceptrons—for predicting genomic values related to birth weight (BW), weaning weight (WW), and yearling weight (YW), and compared them with other conventional models—GBLUP (Genomic Best Linear Unbiased Prediction), Bayes A, and Bayes B. The results demonstrated that the GBLUP model achieved the highest prediction accuracy for both BW and YW, whereas the Random Forest model exhibited a superior prediction accuracy for WW. Furthermore, GBLUP outperformed the other models in terms of model fit, as evidenced by the lower mean square error values and regression coefficients of the corrected phenotypes on predicted values. Overall, the GBLUP model delivered a superior prediction accuracy and model fit compared to the machine learning models tested. Full article
(This article belongs to the Section Animal Genetics and Genomics)
27 pages, 5002 KiB  
Article
The Diversity of Methylation Patterns in Serous Borderline Ovarian Tumors and Serous Ovarian Carcinomas
by Laura A. Szafron, Roksana Iwanicka-Nowicka, Piotr Sobiczewski, Marta Koblowska, Agnieszka Dansonka-Mieszkowska, Jolanta Kupryjanczyk and Lukasz M. Szafron
Cancers 2024, 16(20), 3524; https://doi.org/10.3390/cancers16203524 (registering DOI) - 18 Oct 2024
Abstract
Background: Changes in DNA methylation patterns are a pivotal mechanism of carcinogenesis. In some tumors, aberrant methylation precedes genetic changes, while gene expression may be more frequently modified due to methylation alterations than by mutations. Methods: Herein, 128 serous ovarian tumors [...] Read more.
Background: Changes in DNA methylation patterns are a pivotal mechanism of carcinogenesis. In some tumors, aberrant methylation precedes genetic changes, while gene expression may be more frequently modified due to methylation alterations than by mutations. Methods: Herein, 128 serous ovarian tumors were analyzed, including borderline ovarian tumors (BOTS) with (BOT.V600E) and without (BOT) the BRAF V600E mutation, low-grade (lg), and high-grade (hg) ovarian cancers (OvCa). The methylome of the samples was profiled with Infinium MethylationEPIC microarrays. Results: The biggest number of differentially methylated (DM) CpGs and regions (DMRs) was found between lgOvCa and hgOvCa. By contrast, the BOT.V600E tumors had the lowest number of DM CpGs and DMRs compared to all other groups and, in relation to BOT, their genome was strongly downmethylated. Remarkably, the ten most significant DMRs, discriminating BOT from lgOvCa, encompassed the MHC region on chromosome 6. We also identified hundreds of DMRs, being of potential use as predictive biomarkers in BOTS and hgOvCa. DMRs with the best discriminative capabilities overlapped the following genes: BAIAP3, IL34, WNT10A, NEU1, SLC44A4, and HMOX1, TCN2, PES1, RP1-56J10.8, ABR, NCAM1, RP11-629G13.1, AC006372.4, NPTXR in BOTS and hgOvCa, respectively. Conclusions: The global genome-wide hypomethylation positively correlates with the increasing aggressiveness of ovarian tumors. We also assume that the immune system may play a pivotal role in the transition from BOTS to lgOvCa. Given that the BOT.V600E tumors had the lowest number of DM CpGs and DMRs compared to all other groups, when methylome is considered, such tumors might be placed in-between BOT and OvCa. Full article
Show Figures

Figure 1

Figure 1
<p>Violin plots of methylation changes (average beta values) in the promoter and first-exon regions of the <span class="html-italic">TP53</span>, <span class="html-italic">MDM2</span>, and <span class="html-italic">CDKN1A</span> genes (the remaining significant results are presented in <a href="#app1-cancers-16-03524" class="html-app">Supplementary Figure S3</a>). The values range from 0 to 1 (where 0 means no methylation and 1 denotes 100% methylation of CpGs detected in the region). Each analysis is supplemented with the results of two non-parametric statistical tests: the Kruskal–Wallis test (to determine overall methylation differences between the groups) and the Wilcoxon rank sum test to identify differences between particular groups; NS—non-significant result. Low p-values are displayed in exponential notation (e–n), in which e (exponent) multiplies the preceding number by 10 to the minus nth power.</p>
Full article ">Figure 2
<p>Differentially methylated CpGs. (<b>A</b>): the upset plot demonstrating the number of differentially methylated CpGs in each inter-tumor-group comparison (blue bars) and the number of such CpGs (red bars) for the specific intersection of tumor groups (all sets included in the given intersection are indicated with black dots, that are connected with a line if the intersection contains more than one set). (<b>B</b>–<b>G</b>): the distribution of M-values for the most differentiating CpGs for each inter-tumor-group comparison, followed by genomic locations and gene names with strand identificators shown in brackets. M-value is the log2 of the ratio between signal intensities for probes specific to methylated (numerator) and unmethylated (denominator) cytosines in the given CpG site. The higher the M-value, the higher the methylation level.</p>
Full article ">Figure 3
<p>Context plots depicting the most significant DMR for each inter-tumor-group comparison. Each plot title contains encompassed gene name(s) with the DNA strand identifier (+/−), on which the coding sequence of each gene is located. Below, a chromosome ideogram, graphical representation of the genomic range, and DMR location within the genome are shown. These are followed by a line + dot plot demonstrating the distribution of beta values for each CpG and sample (dot) along with mean values for each CpG (line). The visualization of Dnase I hypersensitive sites (DHSS) and transcription factor binding sites (TFBS) is also provided for the assessment of transcriptional activity in each DMR. (<b>A</b>): BOT vs. BOT.V600E (chr11:g.both 47269539–47270908); (<b>B</b>): BOT vs. lgOvCa (chr6:g.both 32935236–32943025); (<b>C</b>): BOT vs. hgOvCa (chr2:g.both 63275602–63285097); (<b>D</b>): BOT.V600E vs. lgOvCa (chr6:g.both 30651511–30654559); (<b>E</b>): BOT.V600E vs. hgOvCa (chr1:g. 2221807–2222674); (<b>F</b>): lgOvCa vs. hgOvCa (chr10:g.both 134977981–134981930).</p>
Full article ">Figure 4
<p>Nominated regression analyses for selected DMRs in hgOvCa. (<b>A</b>–<b>F</b>): Cox regression analysis (OS) in the subgroup of tumors with TP53 accumulation for the <span class="html-italic">HMOX1(+)/NA(−)</span> genes. (<b>A</b>,<b>B</b>): AUC plot for uni- and multivariable models obtained before (<b>A</b>) and after (<b>B</b>) a bootstrap-based cross-validation of the original data set. A red dashed line in B indicates the same time point which was used to draw the time-dependent ROC curve (<b>C</b>). An optimal cutoff point for this ROC curve, was calculated based on the multivariable model using the Youden index. Discrimination sensitivity and specificity values for this cutoff point are also provided. (<b>D</b>): Kaplan-Meier survival curves obtained for the patients divided into two categories (risk higher (high) or lower (low) than for the ROC curve (<b>C</b>)-estimated cutoff point) based on the risk of death, calculated using the multivariable model. The Kaplan-Meier curves are supplemented with the result of the log-rank test, as well. Box (<b>E</b>) and bar (<b>F</b>) plots depicting mean methylation beta values within the DMR in patients with the high or low risk of death. (<b>G</b>–<b>I</b>): logistic regression analysis (CR) for a DMR in unknown gene(s), in the subgroup of patients treated with the TP regimen. (<b>G</b>): ROC curves for uni- and multivariable logistic regression models. Box (<b>H</b>) and bar (<b>I</b>) plots depicting mean methylation beta values within the DMR in patients with (1) and without (0) CR. RT: residual tumor; TP: taxane/platinum chemotherapy; CR: complete remission. Low p-values are displayed in exponential notation (e−n), in which e (exponent) multiplies the preceding number by 10 to the minus nth power.</p>
Full article ">Figure 5
<p>A nominated logistic regression analysis for a DMR in the <span class="html-italic">BAIAP3(+)/NA(−)</span> gene in the whole group of BOTS patients (Full table). (<b>A</b>): ROC curves for uni- and multivariable logistic regression models; Box (<b>B</b>) and bar (<b>C</b>) plots depicting mean methylation beta values within the DMR in tumors with (Yes) and without (No) microinvasion/non-invasive implants.</p>
Full article ">
7 pages, 598 KiB  
Review
Genetic Variation, Polyploidy, Hybridization Influencing the Aroma Profiles of Rosaceae Family
by Xi Chen, Yu Zhang, Weihua Tang, Geng Zhang, Yuanhua Wang and Zhiming Yan
Genes 2024, 15(10), 1339; https://doi.org/10.3390/genes15101339 (registering DOI) - 18 Oct 2024
Abstract
Background: The fragrance and aroma of Rosaceae plants are complex traits influenced by a multitude of factors, with genetic variation standing out as a key determinant which is largely impacted by polyploidy. Polyploidy serves as a crucial evolutionary mechanism in plants, significantly boosting [...] Read more.
Background: The fragrance and aroma of Rosaceae plants are complex traits influenced by a multitude of factors, with genetic variation standing out as a key determinant which is largely impacted by polyploidy. Polyploidy serves as a crucial evolutionary mechanism in plants, significantly boosting genetic diversity and fostering speciation. Objective: This review focuses on the Rosaceae family, emphasizing how polyploidy influences the production of volatile organic compounds (VOCs), which are essential for the aromatic characteristics of economically important fruits like strawberries, apples, and cherries. The review delves into the biochemical pathways responsible for VOC biosynthesis, particularly highlighting the roles of terpenoids, esters alcohols, aldehydes, ketones, phenolics, hydrocarbons, alongside the genetic mechanisms that regulate these pathways. Key enzymes, such as terpene synthases and alcohol acyltransferases, are central to this process. This review further explores how polyploidy and hybridization can lead to the development of novel metabolic pathways, contributing to greater phenotypic diversity and complexity in fruit aromas. It underscores the importance of gene dosage effects, isoenzyme diversity, and regulatory elements in determining VOC profiles. Conclusions: These findings provide valuable insights for breeding strategies aimed at improving fruit quality and aligning with consumer preferences. Present review not only elucidates the complex interplay between genomic evolution and fruit aroma but also offers a framework for future investigations in plant biology and agricultural innovation. Full article
(This article belongs to the Special Issue Genetics Studies on Crop Agronomy Traits Improvement)
Show Figures

Figure 1

Figure 1
<p>The variations in some VOCs production because of enzyme isoforms: (<b>a</b>) <span class="html-italic">Rosa chinensis</span> phenylpropanoid pathway, different caffeic acid O-methyltransferase (RcOMT) isoforms; (<b>b</b>) AAT isoforms from diploid strawberry and octoploid strawberry; (<b>c</b>) two members of BAHD family of acyltransferases acetyl-CoA:benzylalcohol O-acetyltransferase (BEAT) and benzoyl-CoA:benzyl alcohol benzoyl transferase (BEBT); (<b>d</b>), terpene synthase (TPS) isoforms (+)-(3S)-linalool nerolidol synthase 1&amp;2 (RcLIN-NERS1&amp;2).</p>
Full article ">
12 pages, 1797 KiB  
Article
Investigating the Quantification Capabilities of a Nanopore-Based Sequencing Platform for Food Safety Application via External Standards of Lambda DNA and Lambda Spiked Beef
by Sky Harper, Katrina L. Counihan, Siddhartha Kanrar, George C. Paoli, Shannon Tilman and Andrew G. Gehring
Foods 2024, 13(20), 3304; https://doi.org/10.3390/foods13203304 (registering DOI) - 18 Oct 2024
Viewed by 107
Abstract
Six hundred million cases of disease and roughly 420,000 deaths occur globally each year due to foodborne pathogens. Current methods to screen and identify pathogens in swine, poultry, and cattle products include immuno-based techniques (e.g., immunoassay integrated biosensors), molecular methods (e.g., DNA hybridization [...] Read more.
Six hundred million cases of disease and roughly 420,000 deaths occur globally each year due to foodborne pathogens. Current methods to screen and identify pathogens in swine, poultry, and cattle products include immuno-based techniques (e.g., immunoassay integrated biosensors), molecular methods (e.g., DNA hybridization and PCR assays), and traditional culturing. These methods are often used in tandem to screen, quantify, and characterize samples, prolonging real-time comprehensive analysis. Next-generation sequencing (NGS) is a relatively new technology that combines DNA-sequencing chemistry and bioinformatics to generate and analyze large amounts of short- or long-read DNA sequences and whole genomes. The goal of this project was to evaluate the quantitative capabilities of the real-time NGS Oxford Nanopore Technologies’ MinION sequencer through a shotgun-based sequencing approach. This investigation explored the correlation between known amounts of the analyte (lambda DNA as a pathogenic bacterial surrogate) with data output, in both the presence and absence of a background matrix (Bos taurus DNA). A positive linear correlation was observed between the concentration of analyte and the amount of data produced, number of bases sequenced, and number of reads generated in both the presence and absence of a background matrix. In the presence of bovine DNA, the sequenced data were successfully mapped to the NCBI lambda reference genome. Furthermore, the workflow from pre-extracted DNA to target identification took less than 3 h, demonstrating the potential of long-read sequencing in food safety as a rapid method for screening, identification, and quantification. Full article
(This article belongs to the Special Issue Advances in Foodborne Pathogen Analysis and Detection)
Show Figures

Figure 1

Figure 1
<p>Amount of lambda DNA vs. the number of reads produced with relative error represented. Two lines of best fit are shown, a logarithmic and linear line.</p>
Full article ">Figure 2
<p>Amount of lambda DNA vs. the total amount of bases sequenced (Mb). A (linear) line of best fit was generated with this Equation: y = 0.24x + 26. The R<sup>2</sup> value was 0.99, which suggested a strong correlation.</p>
Full article ">Figure 3
<p>Amount of lambda DNA vs. the number of sequences mapped to lambda phage genome (%). A linear line of best fit was generated with this Equation: y = 0.013x + 1.4. The R<sup>2</sup> value was 0.95.</p>
Full article ">Figure 4
<p>Amount of lambda DNA vs. the number of bases mapped to lambda phage genome (%). A linear line of best fit was generated with this Equation: y = 0.0381x + 5.06. The R<sup>2</sup> value was 0.991.</p>
Full article ">Figure 5
<p>Amount of lambda DNA vs. coverage of the number (#) of sequences (<b>top</b>) and the number (#) of bases (<b>bottom</b>) to Lambdap22.</p>
Full article ">
19 pages, 759 KiB  
Review
Hepatocellular-Carcinoma-Derived Organoids: Innovation in Cancer Research
by Carlo Airola, Maria Pallozzi, Eleonora Cesari, Lucia Cerrito, Leonardo Stella, Claudio Sette, Felice Giuliante, Antonio Gasbarrini and Francesca Romana Ponziani
Cells 2024, 13(20), 1726; https://doi.org/10.3390/cells13201726 (registering DOI) - 18 Oct 2024
Viewed by 109
Abstract
Hepatocellular carcinomas (HCCs) are highly heterogeneous malignancies. They are characterized by a peculiar tumor microenvironment and dense vascularization. The importance of signaling between immune cells, endothelial cells, and tumor cells leads to the difficult recapitulation of a reliable in vitro HCC model using [...] Read more.
Hepatocellular carcinomas (HCCs) are highly heterogeneous malignancies. They are characterized by a peculiar tumor microenvironment and dense vascularization. The importance of signaling between immune cells, endothelial cells, and tumor cells leads to the difficult recapitulation of a reliable in vitro HCC model using the conventional two-dimensional cell cultures. The advent of three-dimensional organoid tumor technology has revolutionized our understanding of the pathogenesis and progression of several malignancies by faithfully replicating the original cancer genomic, epigenomic, and microenvironmental landscape. Organoids more closely mimic the in vivo environment and cell interactions, replicating factors such as the spatial organization of cell surface receptors and gene expression, and will probably become an important tool in the choice of therapies and the evaluation of tumor response to treatments. This review aimed to describe the ongoing and potential applications of organoids as an in vitro model for the study of HCC development, its interaction with the host’s immunity, the analysis of drug sensitivity tests, and the current limits in this field. Full article
(This article belongs to the Section Tissues and Organs)
Show Figures

Figure 1

Figure 1
<p>Clinical applications of HCC-derived organoids. Hepatocellular carcinoma liver organoids’ current applications: Considering the complexity of these three-dimensional models, organoids are increasingly used in cancer research in order to mimic real-life cell-to-cell interactions in tumors. Concerning hepatocellular carcinoma, organoids are useful for the in vitro and in vivo study of cancer cell and cancer stem cell behavior, to evaluate cell signaling, and to unravel the specific alterations in inflammatory, metabolic, and proliferative pathways that lead to tumorigenesis, growth maintenance, immune suppression, angiogenesis, and the mechanisms of resistance and tumor escape through the replication and analysis of the tumor microenvironment. Organoids can also be used for drug screening, for investigating the key drivers of HCC development, and to identify markers of aggressiveness.</p>
Full article ">
14 pages, 1809 KiB  
Article
Causal Relationship Between Sjögren’s Syndrome and Gut Microbiota: A Two-Sample Mendelian Randomization Study
by Xinrun Wang, Minghui Liu and Weiping Xia
Biomedicines 2024, 12(10), 2378; https://doi.org/10.3390/biomedicines12102378 (registering DOI) - 18 Oct 2024
Viewed by 138
Abstract
Background: Gut microbiota have been previously reported to be related to a variety of immune diseases. However, the causal connection between Sjögren’s syndrome (SS) and gut microbiota has yet to be clarified. Methods: We employed a two-sample Mendelian randomization (MR) analysis to evaluate [...] Read more.
Background: Gut microbiota have been previously reported to be related to a variety of immune diseases. However, the causal connection between Sjögren’s syndrome (SS) and gut microbiota has yet to be clarified. Methods: We employed a two-sample Mendelian randomization (MR) analysis to evaluate the causal connection between gut microbiota and SS, utilizing summary statistics from genome-wide association studies (GWASs) obtained from the MiBioGen and FinnGen consortia. The inverse variance weighted (IVW) approach represents the primary method of Mendelian randomization (MR) analysis. Sensitivity analysis was used to eliminate instrumental variables heterogeneity and horizontal pleiotropy. In addition, we performed an analysis using independent GWAS summary statistics for SS from the European Bioinformatics Institute (EBI) dataset for further verify our results. Results: IVW results demonstrated that the phylum Lentisphaerae (OR = 0.79, 95% CI: 0.63–0.99, p = 0.037), class Deltaproteobacteria (OR = 0.67, 95% CI: 0.47–0.96, p = 0.030), family Porphyromonadaceae (OR = 0.60, 95% CI: 0.38–0.94, p = 0.026), genus Eubacterium coprostanoligenes group (OR = 0.61, 95% CI: 0.4–0.93, p = 0.021), genus Blautia (OR = 0.62, 95% CI: 0.43–0.90, p = 0.012), genus Butyricicoccus (OR = 0.61, 95% CI: 0.42–0.90, p = 0.012), genus Escherichia.Shigella (OR = 0.7, 95% CI: 0.49–0.99, p = 0.045) and genus Subdoligranulum (OR = 0.61, 95% CI: 0.44–0.86, p = 0.005) exhibited protective effects on SS. Relevant heterogeneity of horizontal pleiotropy or instrumental variables was not detected. Furthermore, repeating our results with an independent cohort provided by the EBI dataset, only the genus Eubacterium coprostanoligenes group remained significantly associated with the protective effect on SS (OR = 0.41, 95% CI: 0.18–0.91, p = 0.029). Two-step MR analysis further revealed that genus Eubacterium coprostanoligenes group exerts its protective effect by reducing CXCL6 levels in SS (OR, 0.87; 95% CI = 0.76–0.99, p = 0.033). Conclusions: Our study using two-sample MR analysis identified a causal association between multiple genera and SS. A two-step MR result calculated that genus Eubacterium coprostanoligenes group mediated its protective effect by reducing CXCL6 levels in SS. However, the datasets available from the MiBioGen and FinnGen consortia do not provide sufficient information or comprehensive demographic data for subgroup analyses. Additional validation using various omics technologies is necessary to comprehend the development of SS in the intricate interplay between genes and the environment over a period of time. Full article
Show Figures

Figure 1

Figure 1
<p>Scatter plots comparing the genetic correlations between gut microbiota and SS using various MR techniques. (<b>A</b>) The impact of the phylum <span class="html-italic">Lentisphaerae</span> on SS. (<b>B</b>) The impact of the class <span class="html-italic">Deltaproteobacteria</span> on SS. (<b>C</b>) The impact of the family <span class="html-italic">Porphyromonadaceae</span> on SS. (<b>D</b>) The impact of the genus <span class="html-italic">Eubacterium coprostanoligenes</span> on SS. (<b>E</b>) The impact of the genus <span class="html-italic">Blautia</span> on SS. (<b>F</b>) The impact of the genus <span class="html-italic">Butyricicoccus</span> on SS. (<b>G</b>) The impact of the genus <span class="html-italic">Escherichia.Shigella</span> on SS. (<b>H</b>) The impact of the genus <span class="html-italic">Subdoligranulum</span> on SS. The inclines of the line indicate the individual causal impact of each method. MR: Mendelian randomization; SS: Sjögren’s syndrome.</p>
Full article ">Figure 2
<p>Funnel plots used to evaluate the pleiotropy of observed causal connections between gut microbiota and SS for (<b>A</b>) phylum <span class="html-italic">Lentisphaerae</span>; (<b>B</b>) class <span class="html-italic">Deltaproteobacteria</span>; (<b>C</b>) family <span class="html-italic">Porphyromonadaceae</span>; (<b>D</b>) genus <span class="html-italic">Eubacterium coprostanoligenes</span>; (<b>E</b>) genus <span class="html-italic">Blautia</span>; (<b>F</b>) genus <span class="html-italic">Butyricicoccus</span>; (<b>G</b>) genus <span class="html-italic">Escherichia.Shigella</span>; and (<b>H</b>) genus <span class="html-italic">Subdoligranulum</span>.</p>
Full article ">Figure 3
<p>Leave-one-out analysis allowed us to assess the impact of individual SNPs on the relationship between gut microbiota and SS risk. (<b>A</b>) phylum <span class="html-italic">Lentisphaerae</span>; (<b>B</b>) class <span class="html-italic">Deltaproteobacteria</span>; (<b>C</b>) family <span class="html-italic">Porphyromonadaceae</span>; (<b>D</b>) genus <span class="html-italic">Eubacterium coprostanoligenes</span>; (<b>E</b>) genus <span class="html-italic">Blautia</span>; (<b>F</b>) genus <span class="html-italic">Butyricicoccus</span>; (<b>G</b>) genus <span class="html-italic">Escherichia.Shigella</span>; and (<b>H</b>) genus <span class="html-italic">Subdoligranulum</span>. This approach demonstrates how each specific SNP affects the overall result. SNPs, Single nucleotide polymorphism; SS: Sjögren’s syndrome.</p>
Full article ">
13 pages, 4004 KiB  
Essay
Genome-Wide Identification and Expression Analysis of the PsTPS Gene Family in Pisum sativum
by Hao Yuan, Baoxia Liu, Guwen Zhang, Zhijuan Feng, Bin Wang, Yuanpeng Bu, Yu Xu, Zhihong Sun, Na Liu and Yaming Gong
Horticulturae 2024, 10(10), 1104; https://doi.org/10.3390/horticulturae10101104 (registering DOI) - 18 Oct 2024
Viewed by 166
Abstract
This study aimed to explore the role of the trehalose-6-phosphate synthase (TPS) gene family in the adaptation of peas to environmental stress. A comprehensive analysis of the PsTPS gene family identified 20 genes with conserved domains and specific chromosomal locations. Phylogenetic [...] Read more.
This study aimed to explore the role of the trehalose-6-phosphate synthase (TPS) gene family in the adaptation of peas to environmental stress. A comprehensive analysis of the PsTPS gene family identified 20 genes with conserved domains and specific chromosomal locations. Phylogenetic analysis delineated evolutionary relationships, while gene structure analysis revealed compositional insights, and motif analysis provided functional insights. Cis-regulatory element identification predicted gene regulation patterns. Tissue-specific and stress-induced expression profiling highlighted eight genes with ubiquitous expression, with PsTPS15 and PsTPS18 displaying elevated expression levels in roots, nodules, and young stems, and PsTPS13 and PsTPS19 expression downregulated in seeds. Transcriptome analysis identified a differential expression of 20 PsTPS genes, highlighting the significance of 14 genes in response to drought and salinity stress. Notably, under drought conditions, the expression of PsTPS4 and PsTPS6 was initially upregulated and then downregulated, whereas that of PsTPS15 and PsTPS19 was downregulated. Salinity stress notably altered the expression of PsTPS4, PsTPS6, and PsTPS19. Taken together, these findings elucidate the regulatory mechanisms of the PsTPS gene family and their potential as genetic targets for enhancing crop stress tolerance. Full article
(This article belongs to the Section Genetics, Genomics, Breeding, and Biotechnology (G2B2))
Show Figures

Figure 1

Figure 1
<p>Chromosomal locations of the <span class="html-italic">PsTPS</span> genes on the seven pea chromosomes. The distribution of <span class="html-italic">PsTPS</span> genes is relatively sparse, and they are not distributed on every chromosome. The highest distribution of <span class="html-italic">PsTPS</span> genes is observed on Chr5, which contains seven genes.</p>
Full article ">Figure 2
<p>Phylogenetic tree incorporating TPS proteins from <span class="html-italic">Pisum sativum</span> L, <span class="html-italic">Arabidopsis</span>, and <span class="html-italic">Glycine max</span>. The tree of the <span class="html-italic">TPS</span> gene family was constructed by the IQ-TREE 2 software (Version 2.2.0) using the maximum likelihood (ML) method with 1000 bootstrap replicates. The color of the outer ring and branches denote <span class="html-italic">TPS</span> subfamilies.</p>
Full article ">Figure 3
<p>The phylogenetic relationship, conserved motifs, and gene structure of <span class="html-italic">PsTPSs</span>. (<b>a</b>) The maximum likelihood (ML) phylogenetic tree of PsTPS proteins was constructed using a full-length sequence with 1000 bootstrap replicates; (<b>b</b>) Distribution of conserved motifs in PsTPS proteins. A total of 10 motifs were predicted, and the scale bar represents 100 aa; (<b>c</b>) Distribution of the TPS domain in PsTPSs; (<b>d</b>) The gene structures of <span class="html-italic">PsTPSs</span>, including introns (black lines) and exons (green rectangles). The scale bar indicates 1000 bp.</p>
Full article ">Figure 4
<p>CREs on the putative promoters of <span class="html-italic">PsTPSs</span>. (<b>a</b>) Distribution of CREs identified in the 2000 bp upstream promoter region of <span class="html-italic">PsTPSs</span>; (<b>b</b>) The number of CREs on the putative promoters of <span class="html-italic">PsTPSs</span>. Numbers in the heatmap represent the number of elements.</p>
Full article ">Figure 5
<p>Syntenic analyses of <span class="html-italic">TPS</span> genes in <span class="html-italic">Pisum sativum</span>, <span class="html-italic">Arabidopsis</span>, <span class="html-italic">G. max</span>. (<b>a</b>) Seven chromosomes from <span class="html-italic">Pisum sativum</span> (Ps1–Ps7) are mapped, with chromosome length expressed as Mb. Lines denote syntenic <span class="html-italic">TPS</span> gene pairs on the chromosomes. (<b>b</b>) The seven chromosomes of <span class="html-italic">Pisum sativum</span> (Ps1–7), five chromosomes of <span class="html-italic">A. thaliana</span> (At1–5), and twenty chromosomes of <span class="html-italic">G. max</span> (Gm1–20) are mapped. Lines represent syntenic <span class="html-italic">TPS</span> gene pairs.</p>
Full article ">Figure 6
<p>Predicted protein–protein interaction networks of PsTPS proteins with other proteins using the STRING tool. Interactions between proteins are represented by gray lines.</p>
Full article ">Figure 7
<p>Expression profiles of the eight <span class="html-italic">PsTPS</span> genes. The color scale from blue to red indicates increasing log2-transformed FPKM values.</p>
Full article ">Figure 8
<p>Transcriptome analysis depicting the expression levels of 14 <span class="html-italic">PsTPS</span> genes in <span class="html-italic">Pisum sativum</span> under drought stress conditions induced by 10%, 20%, and 30% PEG6000 and salt stress induced by 100 mM, 200 mM, and 300 mM NaCl. Each experiment was conducted independently with a minimum of three replicates. “CK_0h” denotes the control group.</p>
Full article ">
12 pages, 2909 KiB  
Review
Exploring Fish Parvalbumins through Allergen Names and Gene Identities
by Johannes M. Dijkstra, Annette Kuehn, Eiji Sugihara and Yasuto Kondo
Genes 2024, 15(10), 1337; https://doi.org/10.3390/genes15101337 (registering DOI) - 18 Oct 2024
Viewed by 143
Abstract
Parvalbumins are the main source of food allergies in fish meat, with each fish possessing multiple different parvalbumins. The naming convention of these allergens in terms of allergen codes (numbers) is species-specific. Allergen codes for parvalbumin isoallergens and allergen variants are based on [...] Read more.
Parvalbumins are the main source of food allergies in fish meat, with each fish possessing multiple different parvalbumins. The naming convention of these allergens in terms of allergen codes (numbers) is species-specific. Allergen codes for parvalbumin isoallergens and allergen variants are based on sequence identities relative to the first parvalbumin allergen discovered in that particular species. This means that parvalbumins with similar allergen codes, such as catfish Pan h 1.0201 and redfish Seb m 1.0201, are not necessarily the most similar proteins, or encoded by the same gene. Here, we aim to elucidate the molecular basis of parvalbumins. We explain the complicated genetics of fish parvalbumins in an accessible manner for fish allergen researchers. Teleost or modern bony fish, which include most commercial fish species, have varying numbers of up to 22 parvalbumin genes. All have derived from ten parvalbumin genes in their common ancestor. We have named these ten genes “parvalbumin 1-to-10” (PVALB1-to-PVALB10), building on earlier nomenclature established for zebrafish. For duplicated genes, we use variant names such as, for example, “PVALB2A and PVALB2B”. As illustrative examples of our gene identification system, we systematically analyze all parvalbumin genes in two common allergy-inducing species in Japan: red seabream (Pagrus major) and chum salmon (Oncorhynchus keta). We also provide gene identifications for known parvalbumin allergens in various fish species. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Parvalbumins have six α-helices A-to-F and two “EF-hand” domains for binding Ca<sup>2+</sup> ions (indicated as spheres). (<b>A</b>) The structure, in cartoon format, of common carp pvalb4_(Chr.A3) (a β2-parvalbumin; PDB accession 4CPV) [<a href="#B7-genes-15-01337" class="html-bibr">7</a>], which was the first parvalbumin of which the structure was elucidated [<a href="#B8-genes-15-01337" class="html-bibr">8</a>]. Different α-helices are in different colors. (<b>B</b>) Superimposition of various parvalbumin structures, in ribbon format, reveals a common structure. Light pink, human α-parvalbumin (1RK9); pink, pike pvalb7 α-parvalbumin (2PAS); magenta, spotless smooth-hound shark SPV-I α-parvalbumin (5ZGM); green, human oncomodulin (1TTX); splitpea green, chicken CPV3-oncomodulin (2KYF); soft purple, chicken ATH β2-parvalbumin (3FS7); cyan, Atlantic cod pvalb2 β2-parvalbumin (2MBX); green cyan, pike pvalb3 β2-parvalbumin (1PVB); aquamarine, common carp pvalb4_(Chr.A3) β2-parvalbumin (4CPV); light teal, spotless smooth-hound shark SPV-II β2-parvalbumin (5ZH6). (<b>C</b>) The structure, in ribbon format, of common carp pvalb4_(Chr.A3) (PDB accession 4CPV), shows in black those residues that are well conserved throughout EF-hand domain family molecules and in gray other residues that are well conserved throughout parvalbumins; the sidechains of these residues are shown in sticks format. This figure is used, with permission, from our open access article [<a href="#B9-genes-15-01337" class="html-bibr">9</a>], and the figures were created with the help of Pymol 2.5.2 software (<a href="https://pymol.org/2/" target="_blank">https://pymol.org/2/</a> (accessed on 27 October 2022)).</p>
Full article ">Figure 2
<p>Parvalbumins in extant teleost fish derive from ten parvalbumin genes in their common ancestor and belong to three different ancient parvalbumin lineages. (<b>A</b>) The immediate ancestor of extant teleost fish possessed at least the genes <span class="html-italic">PVALB1</span>-to-<span class="html-italic">PVALB10</span>, spread over four different chromosomal regions deriving from two whole-genome duplication (WGD) events. The parvalbumin genes are indicated with thick-lined boxes that are pointed in the gene direction and are colored magenta for α-parvalbumins, green for oncomodulins, and different kinds of blue for <span class="html-italic">PVALB1</span>-to-4, e<span class="html-italic">PVALB5</span>, and <span class="html-italic">PVALB10</span>. Neighboring non-parvalbumin genes are indicated by lower boxes with their name abbreviations inside. (<b>B</b>) Parvalbumin gene organization in red seabream and chum salmon, with the direction of the depicted scaffolds adjusted for homogenization. For relevant genomic region information, or Genbank accession numbers providing access to such information, see <a href="#app1-genes-15-01337" class="html-app">Supplementary Files S1 and S2</a>. Most symbols are as in (<b>A</b>), and the boxes with dashed lines and Ψ symbols indicate probable pseudogenes. (<b>C</b>) A condensed part of a phylogenetic tree created by the Maximum Likelihood method using 209 parvalbumin amino acid sequences of fishes and other species. Only the teleost fish sequences are indicated here, with between brackets the number of teleost sequences condensed in the respective part of the tree. For the complete tree and sequence information, see [<a href="#B9-genes-15-01337" class="html-bibr">9</a>]. The percentage of trees in which the associated taxa clustered together is shown next to the branches if &gt;50. Percentages of aa identity, calculated with the help of Clustal Omega (<a href="https://www.ebi.ac.uk/jdispatcher/msa/clustalo" target="_blank">https://www.ebi.ac.uk/jdispatcher/msa/clustalo</a> (accessed on 25 March 2024)), are indicated per cluster.</p>
Full article ">Figure 3
<p>Parvalbumin amino acid consensus sequences. Consensus sequences were created using WebLogo 2.8.2 (<a href="https://weblogo.berkeley.edu/logo.cgi" target="_blank">https://weblogo.berkeley.edu/logo.cgi</a> (accessed on 25 March 2024)) software for analysis of the parvalbumin sequences listed in [<a href="#B9-genes-15-01337" class="html-bibr">9</a>], which tried to provide a broad overview of parvalbumin sequences while focusing on teleost parvalbumins. (<b>A</b>) Sequence logo for all analyzed 209 parvalbumin sequences, with helices indicated above the alignment based on the structure of common carp pvalb4_(Chr.A3) protein (PDB database 4CPV). (<b>B</b>) Frequency plots for residues at positions that help to distinguish between the α-parvalbumins (<span class="html-italic">n</span> = 45; 30 from teleosts), oncomodulins (<span class="html-italic">n</span> = 43; 30 from teleosts), and β2-parvalbumins (<span class="html-italic">n</span> = 121; 87 from teleosts). (<b>C</b>) Frequency plots for residues at positions that help to distinguish between the combined pvalb1-to-4 sequences (<span class="html-italic">n</span> = 64), pvalb5 (<span class="html-italic">n</span> = 13), and pvalb10 (<span class="html-italic">n</span> = 10) of teleosts. (<b>D</b>) Frequency plots for residues at positions that help to distinguish between teleost pvalb1 (<span class="html-italic">n</span> = 15), pvalb2 (<span class="html-italic">n</span> = 10), pvalb3 (<span class="html-italic">n</span> = 17), and pvalb4 (<span class="html-italic">n</span> = 22). The letters represent amino acids and their sizes correspond with their level of conservation. For a discussion of the structural importance of these characteristic residues, see [<a href="#B9-genes-15-01337" class="html-bibr">9</a>]. *, many parvalbumins are a bit shorter and do not have a residue at position 109.</p>
Full article ">Figure 4
<p>Percentages of amino acid identity between parvalbumins of red seabream, Atlantic cod, chum salmon, chicken, and human. Colors highlight comparisons between parvalbumins belonging to the same family: β2-parvalbumins (teleost pvalb1-to-4), blue (cyan); α-parvalbumins, pink; oncomodulins, green.</p>
Full article ">
15 pages, 3932 KiB  
Article
Whole Genome Sequencing Analysis of Model Organisms Elucidates the Association Between Environmental Factors and Human Cancer Development
by Shinya Hasegawa, Yutaka Shoji, Mamoru Kato, Asmaa Elzawahry, Momoko Nagai, Min Gi, Shugo Suzuki, Hideki Wanibuchi, Sachiyo Mimaki, Katsuya Tsuchihara and Yukari Totsuka
Int. J. Mol. Sci. 2024, 25(20), 11191; https://doi.org/10.3390/ijms252011191 (registering DOI) - 17 Oct 2024
Viewed by 206
Abstract
Determining a novel etiology and mechanism of human cancer requires extraction of characteristic mutational signatures derived from chemical substances. This study explored the mutational signatures of N-nitroso bile acid conjugates using Salmonella strains. Exposing S. typhimurium TA1535 to N-nitroso-glycine/taurine bile acid [...] Read more.
Determining a novel etiology and mechanism of human cancer requires extraction of characteristic mutational signatures derived from chemical substances. This study explored the mutational signatures of N-nitroso bile acid conjugates using Salmonella strains. Exposing S. typhimurium TA1535 to N-nitroso-glycine/taurine bile acid conjugates induced a predominance of C:G to T:A transitions. Two mutational signatures, B1 and B2, were extracted. Signature B1 is associated with N-nitroso-glycine bile acid conjugates, while Signature B2 is linked to N-nitroso-taurine bile acid conjugates. Signature B1 revealed a strong transcribed strand bias with GCC and GCT contexts, and the mutation pattern of N-nitroso-glycine bile acid conjugates in YG7108, which lacks O6-methylguanine DNA methyltransferases, matched that of the wild-type strain TA1535, suggesting that O6-methyl-deoxyguanosine contributes to mutations in the relevant regions. COSMIC database-based similarity analysis revealed that Signature B1 closely resembled SBS42, which is associated with occupational cholangiocarcinoma caused by overexposure to 1,2-dichlolopropane (1,2-DCP) and/or dichloromethane (DCM). Moreover, the inflammatory response pathway was induced by 1,2-DCP exposure in a human cholangiocyte cell line, and iNOS expression was positive in occupational cholangiocarcinomas. These results suggest that 1,2-DCP triggers an inflammatory response in biliary epithelial cells by upregulating iNOS and N-nitroso-glycine bile acid conjugate production, resulting in cholangiocarcinoma via DNA adduct formation. Full article
(This article belongs to the Special Issue Molecular Progression of Genome-Related Diseases)
Show Figures

Figure 1

Figure 1
<p>Mutagenicity of four types of <span class="html-italic">N</span>-nitroso bile acid conjugates, NO-TCA (<b>a</b>), NO-TDCA (<b>b</b>), NO-GCA (<b>c</b>), and NO-GDCA (<b>d</b>) in <span class="html-italic">Salmonella typhimurium</span> TA1535 and YG7108 without metabolic activation systems. Mutagenic activities observed in the wild-type TA1535 (blue) and YG7108 (orange), which lack <span class="html-italic">O</span><sup>6</sup>-methylguanine DNA methyltransferases (MGMT), were estimated using the number of revertant colonies.</p>
Full article ">Figure 2
<p>Levels of alkylating DNA adducts in TA1535 and YG7108 after incubation with NO-GDCA.</p>
Full article ">Figure 3
<p>Trinucleotide mutational patterns for four types of <span class="html-italic">N</span>-nitroso bile acid conjugates, NO-TCA (<b>a</b>), NO-TDCA (<b>b</b>), NO-GCA (<b>c</b>), and NO-GDCA (<b>d</b>) in TA1535. The 96-pattern of mutational types are indicated by different colors. The vertical axis indicates the number of mutations. For <span class="html-italic">N</span>-nitroso-taurine conjugates, ACC, CCT, and TCG contexts were predominant, whereas for <span class="html-italic">N</span>-nitroso-glycine conjugates, GCC and GCT contexts were predominant.</p>
Full article ">Figure 4
<p>Extraction of mutational signatures from the 48 samples of <span class="html-italic">S. typhimurium</span> TA1535 exposed to <span class="html-italic">N</span>-nitroso bile acid conjugates by NMF analysis. Two mutational signatures, Signature B1 (<b>a</b>) and Signature B2 (<b>b</b>), were identified. The mutational signatures were normalized using the trinucleotide frequency in the genome. The horizontal axis represents the 96-pattern of mutational types in the same order as in <a href="#ijms-25-11191-f003" class="html-fig">Figure 3</a>, and the vertical axis indicates the percentage of mutations attributed to a specific signature. Signature B1 features a unique trinucleotide mutational context of GCC and GCT, whereas Signature B2 has a prominent trinucleotide context of ACC, CCT, and TCG.</p>
Full article ">Figure 5
<p>Strand bias pattern of the C to T transition in Signature B1/B2 (<b>a</b>). Green triangles indicate a strong strand bias on a transcribed strand, and purple triangles indicate that on an untranscribed strand. Formation of candidate DNA adducts that contribute to inducing a strand bias from bile acid conjugates (<b>b</b>).</p>
Full article ">Figure 6
<p>Comparison of Signature B1 and SBS42, a mutational signature of occupational cholangiocarcinoma caused by overexposure to 1,2-DCP (<b>a</b>). Significant strand bias for the C:G to T:A mutations were observed in the transcribed (T, blue) and untranscribed (UT, orange) strands in both Signature B1 and SBS42 (<b>b</b>). Red arrows indicated that similar transcriptional strand biases patterns.</p>
Full article ">Figure 6 Cont.
<p>Comparison of Signature B1 and SBS42, a mutational signature of occupational cholangiocarcinoma caused by overexposure to 1,2-DCP (<b>a</b>). Significant strand bias for the C:G to T:A mutations were observed in the transcribed (T, blue) and untranscribed (UT, orange) strands in both Signature B1 and SBS42 (<b>b</b>). Red arrows indicated that similar transcriptional strand biases patterns.</p>
Full article ">Figure 7
<p>Possible mechanisms of 1,2-DCP associated with occupational cholangiocarcinoma. Excessive exposure to 1,2-DCP could induce inflammation and mediate the <span class="html-italic">N</span>-nitrosation of glycine bile acid conjugates. These <span class="html-italic">N</span>-nitroso-glycine bile acid conjugates likely affect biliary epithelial cells, ultimately inducing occupational cholangiocarcinoma.</p>
Full article ">Figure 8
<p>Expression of iNOS in occupational cholangiocarcinoma. Representative data of immunohistochemical staining for iNOS from occupational cholangiocarcinoma is shown. iNOS staining was localized to the cytoplasm of inflammatory and cancer cells and tended to be overexpressed for cancer cells in areas with significant inflammatory cell infiltration. Bars are 100 μm.</p>
Full article ">
17 pages, 1815 KiB  
Article
Decoding the Genetic Basis of Mast Cell Hypersensitivity and Infection Risk in Hypermobile Ehlers-Danlos Syndrome
by Purusha Shirvani, Arash Shirvani and Michael F. Holick
Curr. Issues Mol. Biol. 2024, 46(10), 11613-11629; https://doi.org/10.3390/cimb46100689 (registering DOI) - 17 Oct 2024
Viewed by 207
Abstract
Hypermobile Ehlers-Danlos syndrome (hEDS) is a connective tissue disorder marked by joint hypermobility, skin hyperextensibility, and tissue fragility. Recent studies have linked hEDS with mast cell activation syndrome (MCAS), suggesting a genetic interplay affecting immune regulation and infection susceptibility. This study aims to [...] Read more.
Hypermobile Ehlers-Danlos syndrome (hEDS) is a connective tissue disorder marked by joint hypermobility, skin hyperextensibility, and tissue fragility. Recent studies have linked hEDS with mast cell activation syndrome (MCAS), suggesting a genetic interplay affecting immune regulation and infection susceptibility. This study aims to decode the genetic basis of mast cell hypersensitivity and increased infection risk in hEDS by identifying specific genetic variants associated with these conditions. We conducted whole-genome sequencing (WGS) on 18 hEDS participants and 7 first-degree relatives as controls, focusing on identifying genetic variants associated with mast cell dysregulation. Participants underwent clinical assessments to document hEDS symptoms and mast cell hypersensitivity, with particular attention to past infections and antihistamine response. Our analysis identified specific genetic variants in MT-CYB, HTT, MUC3A, HLA-B and HLA-DRB1, which are implicated in hEDS and MCAS. Protein–protein interaction (PPI) network analysis revealed significant interactions among identified variants, highlighting their involvement in pathways related to antigen processing, mucosal protection, and collagen synthesis. Notably, 61.1% of the hEDS cohort reported recurrent infections compared to 28.5% in controls, and 72.2% had documented mast cell hypersensitivity versus 14.2% in controls. These findings provide a plausible explanation for the complex interplay between connective tissue abnormalities and immune dysregulation in hEDS. The identified genetic variants offer insights into potential therapeutic targets for modulating mast cell activity and improving patient outcomes. Future research should validate these findings in larger cohorts and explore the functional implications of these variants to develop effective treatment strategies for hEDS and related mast cell disorders. Full article
Show Figures

Figure 1

Figure 1
<p>Protein–protein interaction network highlighting mast cell hypersensitivity pathways. This figure illustrates the protein–protein interaction network, which demonstrated significantly more interactions than expected for a random set of proteins, with a PPI enrichment <span class="html-italic">p</span>-value of 1.49 × 10<sup>−10</sup>. The Markov cluster algorithm (MCL) identified at least six distinct clusters within the network. MCL is a method used to cluster proteins based on their interaction patterns within a protein-protein interaction network. This approach helps identify groups of proteins that interact more frequently with each other than with those outside the group, suggesting functional relatedness. The first cluster is involved in antigen processing and the presentation of endogenous peptide antigens and MHC protein complexes (red border). The second cluster relates to the defective GALNT3 causing hyperphosphatemic familial tumoral calcinosis (HFTC), including genes such as MUC3A, MUC16, MUC19, and ZNF717 (green circle). These MUC genes are major glycoprotein components of mucus gels, providing a protective barrier against particles and infectious agents at mucosal surfaces and potentially involved in ligand binding and intracellular signaling. The third cluster is associated with collagen chain trimerization and extracellular matrix structural constituents conferring tensile strength, including genes such as COL4A2, COL6A2 and MMP16 (yellow circle). The fourth cluster relates to retinoid and cholesterol metabolism, including genes such as LPL and LRP2 (blue circle). The fifth cluster is associated with mitochondrial complex I assembly model OXPHOS system, including genes such as MT-ND1 and ACAD9 (green rectangle). The last cluster relates to triplet repeat expansion, including genes such as SPTA1 (black circle). This refers to proteins encoded by a gene which has a triplet repeat expansion, i.e., the increase of triplet (trinucleotide) repeats within the gene sequence. The length of such repeats is frequently polymorphic, and there is often a correlation between repeat length and disease severity.</p>
Full article ">Figure 2
<p>Protein–protein interaction network capturing hEDS-specific genes with potential relationships to established EDS genes. This figure illustrates the protein–protein interaction network, which showed significantly more interactions than expected for a random set of proteins, with a PPI enrichment <span class="html-italic">p</span>-value of 1.49 × 10<sup>−10</sup>. The red nodes represent the known genes associated with different types of EDS, while the green nodes represent genes with variations specific to hEDS subjects that have relationships with these known genes. All of these genes, except PHACTR1, are involved in collagen chain trimerization and extracellular matrix structural constituents conferring tensile strength. PHACTR1 (phosphatase and actin regulator 1) binds actin monomers (G actin) and plays a role in various processes, including the regulation of actin cytoskeleton dynamics, actin stress fibers formation, cell motility and survival, tubule formation by endothelial cells, and regulation of PPP1CA activity. It is also involved in the regulation of cortical neuron migration and dendrite arborization. To simplify the figure, pathways related to HLA and information repeated from <a href="#cimb-46-00689-f001" class="html-fig">Figure 1</a> have been removed from this pathway.</p>
Full article ">Figure 3
<p>Protein–protein interaction network capturing MCAS-specific genetic variants in hEDS with potential relationships to established MCAS genes. This figure illustrates the protein–protein interaction network, which demonstrated significantly more interactions than would be expected for a random set of proteins, with a PPI enrichment <span class="html-italic">p</span>-value of 1.0 × 10<sup>−16</sup>. The red and green nodes represent known genes associated with mast cell activation syndrome (MCAS), while the yellow nodes represent genes with variations specific to hEDS subjects that have relationships with these known MCAS genes. The red nodes are involved in pathways related to hematopoietic or lymphoid organ development, whereas the yellow nodes participate in inflammatory responses and the positive regulation of interleukin-10 production. The results demonstrate that all known genes related to mast cell activation syndrome or mast cell hypersensitivity are interconnected, as anticipated. We identified additional genes within this network, including TLR1, RET, HP, ZNF521, and CCR5. Notably, ZNF521, a transcription factor, was also identified in the pathway depicted in <a href="#cimb-46-00689-f002" class="html-fig">Figure 2</a>. It plays a role alongside RUNX2 in regulating osteoblast differentiation. To simplify the figure, pathways related to HLA and information repeated from <a href="#cimb-46-00689-f001" class="html-fig">Figure 1</a> and <a href="#cimb-46-00689-f002" class="html-fig">Figure 2</a> have been removed from this pathway.</p>
Full article ">
Back to TopTop