[go: up one dir, main page]

Next Issue
Volume 17, February
Previous Issue
Volume 16, December
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 17, Issue 1 (January 2016) – 134 articles

Cover Story (view full-size image): The Helix pomatia Cd-metallothionein (HpCdMT) is a paradigmatic Cd-specific MT. This work studies how the amino acid sequence of the linker connecting the two nine-Cys HpCdMT moieties influences its ability for metal coordination. Hence, we characterized the metal complexes yielded by two HpCdMT constructs containing longer linkers than the -KT- dipeptide of the wild type form. One mutant reproduced the linker sequence of another gastropod MT, the limpet M. crenulata, with a clear polar character; and the other derived from a plant MT (wheat Ec-1), had a clear apolar composition. Image provided by Òscar Palacios.View this article.
  • 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:
4916 KiB  
Article
Unraveling Molecular Differences of Gastric Cancer by Label-Free Quantitative Proteomics Analysis
by Peng Dai, Qin Wang, Weihua Wang, Ruirui Jing, Wei Wang, Fengqin Wang, Kazem M. Azadzoi, Jing-Hua Yang and Zhen Yan
Int. J. Mol. Sci. 2016, 17(1), 69; https://doi.org/10.3390/ijms17010069 - 21 Jan 2016
Cited by 32 | Viewed by 6452
Abstract
Gastric cancer (GC) has significant morbidity and mortality worldwide and especially in China. Its molecular pathogenesis has not been thoroughly elaborated. The acknowledged biomarkers for diagnosis, prognosis, recurrence monitoring and treatment are lacking. Proteins from matched pairs of human GC and adjacent tissues [...] Read more.
Gastric cancer (GC) has significant morbidity and mortality worldwide and especially in China. Its molecular pathogenesis has not been thoroughly elaborated. The acknowledged biomarkers for diagnosis, prognosis, recurrence monitoring and treatment are lacking. Proteins from matched pairs of human GC and adjacent tissues were analyzed by a coupled label-free Mass Spectrometry (MS) approach, followed by functional annotation with software analysis. Nano-LC-MS/MS, quantitative real-time polymerase chain reaction (qRT-PCR), western blot and immunohistochemistry were used to validate dysregulated proteins. One hundred forty-six dysregulated proteins with more than twofold expressions were quantified, 22 of which were first reported to be relevant with GC. Most of them were involved in cancers and gastrointestinal disease. The expression of a panel of four upregulated nucleic acid binding proteins, heterogeneous nuclear ribonucleoprotein hnRNPA2B1, hnRNPD, hnRNPL and Y-box binding protein 1 (YBX-1) were validated by Nano-LC-MS/MS, qRT-PCR, western blot and immunohistochemistry assays in ten GC patients’ tissues. They were located in the keynotes of a predicted interaction network and might play important roles in abnormal cell growth. The label-free quantitative proteomic approach provides a deeper understanding and novel insight into GC-related molecular changes and possible mechanisms. It also provides some potential biomarkers for clinical diagnosis. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Venn diagrams of total identified proteins and dysregulated proteins. (<b>A</b>) Total proteins identified from three cases in tumor or adjacent tissues respectively. One thousand seven hundred forty-four proteins identified appear in both tumor and adjacent tissues; (<b>B</b>) The number of proteins with more than twofold differential expression in three cases, respectively, and the number of proteins shared in two or three cases.</p>
Full article ">Figure 2
<p>The hierarchical heatmap of 146 dysregulated proteins analyzed by Ingenuity Pathway Analysis (IPA)<b>.</b> The major boxes represent specific family or category of related functions. The smaller squares within the major boxes represent the number of proteins. Each individual square represent a specific protein. Colored squares indicate protein predicted state: increasing (orange), or decreasing (blue). Darker colors indicate higher absolute Z-scores.</p>
Full article ">Figure 3
<p>The functional annotation of dysregulated proteins was analyzed by Protein Analysis Through Evolutionary Relationships (PANTHER), Database for Annotation, Visualization and Integrated Discovery (DAVID), STRING and Reactome. (<b>A</b>) Protein Classes; (<b>B</b>) Biological Process; and (<b>C</b>) Molecular Function of 146 dysregulated proteins were summarized in a pie chart by PANTHER; (<b>D</b>) Molecular function; and (<b>E</b>) Biological process based on the 65 upregulated proteins were depicted in a bar graph by DAVID; (<b>F</b>) Pathway analysis of 146 dysregulated proteins was indicated by PANTHER, DAVID, STRING and Reactome. For each category, the percentage or <span class="html-italic">p</span>-value of dysregulated proteins is indicated.</p>
Full article ">Figure 4
<p>Protein-protein interactions (Evidence Mode) of dysregulated proteins were predicted by STRING. (<b>A</b>) Protein-protein interaction network formed with 146 dysregulated proteins. The three possible systematic dynamic clusters were indicated in red circles; (<b>B</b>) The network predicted 65 upregulated proteins. Some important proteins dispersed and located in the keynotes were marked with a red box. Different line colors represent the types of evidence for the association.</p>
Full article ">Figure 4 Cont.
<p>Protein-protein interactions (Evidence Mode) of dysregulated proteins were predicted by STRING. (<b>A</b>) Protein-protein interaction network formed with 146 dysregulated proteins. The three possible systematic dynamic clusters were indicated in red circles; (<b>B</b>) The network predicted 65 upregulated proteins. Some important proteins dispersed and located in the keynotes were marked with a red box. Different line colors represent the types of evidence for the association.</p>
Full article ">Figure 5
<p>Expression levels of hnRNPs and YBX-1 in GC and adjacent tissues. (<b>A</b>) qRT-PCR (<span class="html-italic">n</span> = 10) results showing the mRNA expression of hnRNPs and YBX-1. The ratio below the dotted line represented down-expression in GC tissues; otherwise represented up-expression in GC tissues; (<b>B</b>) Western blots (<span class="html-italic">n</span> = 10) of hnRNPs and YBX-1. N represent adjacent tissue and T represent tumor tissue; (<b>C</b>) Grayscale scanning of western blots bands. The ratio was compared to β-actin and statistically analyzed. Significance of differences between GC and adjacent tissues are displayed by ** <span class="html-italic">p</span>-value &lt; 0.01 or * <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 6
<p>Representative immunohistochemical staining for sectioned formalin fixed GC and adjacent tissues. Specific antibodies of Anti-hnRNPA2B1 (Santa Cruz, TX, USA), Anti-hnRNPD (Proteintech, Chicago, IL, USA), Anti-hnRNPL (Santa Cruz) and Anti-YBX-1 (Santa Cruz) were hybridized respectively. IHC results showed that the morphology of tubular glands disappeared in cancer sections when compared to adjacent tissues. Cancer sections have stronger and higher density nuclei staining, high ratios of nucleus/cytoplasmic area, different shaped nuclei including megakaryocytes and polykaryocytes (arrows). Weak cytoplasmic staining were only seen in hnRNPA2B1, hnRNPD and YBX-1 hybridized normal sections. The magnification is 400×; scale bar: 20 μm.</p>
Full article ">
813 KiB  
Editorial
Acknowledgement to Reviewers of International Journal of Molecular Sciences in 2015
by International Journal of Molecular Sciences Editorial Office
Int. J. Mol. Sci. 2016, 17(1), 142; https://doi.org/10.3390/ijms17010142 - 21 Jan 2016
Viewed by 10630
Abstract
The editors of International Journal of Molecular Sciences would like to express their sincere gratitude to the following reviewers for assessing manuscripts in 2015. [...] Full article
1126 KiB  
Review
An Overview of Direct Somatic Reprogramming: The Ins and Outs of iPSCs
by Siddharth Menon, Siny Shailendra, Andrea Renda, Michael Longaker and Natalina Quarto
Int. J. Mol. Sci. 2016, 17(1), 141; https://doi.org/10.3390/ijms17010141 - 21 Jan 2016
Cited by 35 | Viewed by 12335
Abstract
Stem cells are classified into embryonic stem cells and adult stem cells. An evolving alternative to conventional stem cell therapies is induced pluripotent stem cells (iPSCs), which have a multi-lineage potential comparable to conventionally acquired embryonic stem cells with the additional benefits of [...] Read more.
Stem cells are classified into embryonic stem cells and adult stem cells. An evolving alternative to conventional stem cell therapies is induced pluripotent stem cells (iPSCs), which have a multi-lineage potential comparable to conventionally acquired embryonic stem cells with the additional benefits of being less immunoreactive and avoiding many of the ethical concerns raised with the use of embryonic material. The ability to generate iPSCs from somatic cells provides tremendous promise for regenerative medicine. The breakthrough of iPSCs has raised the possibility that patient-specific iPSCs can provide autologous cells for cell therapy without the concern for immune rejection. iPSCs are also relevant tools for modeling human diseases and drugs screening. However, there are still several hurdles to overcome before iPSCs can be used for translational purposes. Here, we review the recent advances in somatic reprogramming and the challenges that must be overcome to move this strategy closer to clinical application. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Totipotency: After fertilization, Embryonic Stem Cells (ESCs) maintain the ability to form all three germ layers as well as extra-embryonic tissues or placental cells and are termed as totipotent. Pluripotency: These more specialized cells of the blastocyst stage maintain the ability to self-renew and differentiate into the three germ layers and down many lineages but do not form extra-embryonic tissues or placental cells. Reprogrammed somatic cells, iPSCs, also demonstrate the ability to self-renew and differentiate into all three germ layers <span class="html-italic">in vivo</span> and <span class="html-italic">in vitro</span>. Thus, iPSCs are also considered to be pluripotent stem cells. Multipotency: Adult or somatic stem cells are undifferentiated cells found in postnatal tissues. These specialized cells are considered to be multipotent; with very limited ability to self-renew and are committed to lineage specific differentiation.</p>
Full article ">Figure 2
<p>The direct reprogramming of somatic cells to induced pluripotent stem cells (iPSCs) can be achieved through the ectopic expression of specific transcription factors: <span class="html-italic">Oct-4</span>, <span class="html-italic">C-Myc</span>, <span class="html-italic">Sox-2</span> and <span class="html-italic">Klf-4</span>, the “Yamanaka or OSKM Factors”. Delivery of these factors can be accomplished through a variety of methods. Initial methods developed integrating retroviral or lentiviral vectors. More recent strategies utilized non-integrating methods, further categorized by the use of DNA such as plasmids, mini circles/episomal, adenovirus, Sendai virus, and non-DNA based procedures such as, mRNAs, microRNAs (miRs), small molecules and bioactive proteins. Once reprogrammed to a pluripotent state, iPSCs can be expanded <span class="html-italic">in vitro</span> and subsequently differentiated to ectoderm, mesoderm and endoderm linages for use in cell therapy, disease modeling and drug discovery.</p>
Full article ">
2291 KiB  
Article
Elucidating the Diversity of Aquatic Microdochium and Trichoderma Species and Their Activity against the Fish Pathogen Saprolegnia diclina
by Yiying Liu, Christin Zachow, Jos M. Raaijmakers and Irene De Bruijn
Int. J. Mol. Sci. 2016, 17(1), 140; https://doi.org/10.3390/ijms17010140 - 21 Jan 2016
Cited by 15 | Viewed by 9201
Abstract
Animals and plants are increasingly threatened by emerging fungal and oomycete diseases. Amongst oomycetes, Saprolegnia species cause population declines in aquatic animals, especially fish and amphibians, resulting in significant perturbation in biodiversity, ecological balance and food security. Due to the prohibition of several [...] Read more.
Animals and plants are increasingly threatened by emerging fungal and oomycete diseases. Amongst oomycetes, Saprolegnia species cause population declines in aquatic animals, especially fish and amphibians, resulting in significant perturbation in biodiversity, ecological balance and food security. Due to the prohibition of several chemical control agents, novel sustainable measures are required to control Saprolegnia infections in aquaculture. Previously, fungal community analysis by terminal restriction fragment length polymorphism (T-RFLP) revealed that the Ascomycota, specifically the genus Microdochium, was an abundant fungal phylum associated with salmon eggs from a commercial fish farm. Here, phylogenetic analyses showed that most fungal isolates obtained from salmon eggs were closely related to Microdochium lycopodinum/Microdochium phragmitis and Trichoderma viride species. Phylogenetic and quantitative PCR analyses showed both a quantitative and qualitative difference in Trichoderma population between diseased and healthy salmon eggs, which was not the case for the Microdochium population. In vitro antagonistic activity of the fungi against Saprolegnia diclina was isolate-dependent; for most Trichoderma isolates, the typical mycoparasitic coiling around and/or formation of papilla-like structures on S. diclina hyphae were observed. These results suggest that among the fungal community associated with salmon eggs, Trichoderma species may play a role in Saprolegnia suppression in aquaculture. Full article
(This article belongs to the Special Issue Fish Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Isolation of salmon egg-associated fungi and oomycetes on potato dextrose agar plates. One or two salmon eggs from a <span class="html-italic">Saprolegnia</span>-infected batch (diseased, replicate No. 1–6) or a healthy batch (replicate No. 7–12) were placed onto the agar plates to allow fungal outgrowth.</p>
Full article ">Figure 2
<p>Phylogenetic tree of ITS rRNA sequences of nine <span class="html-italic">Microdochium</span> isolates from salmon eggs and reference strains. The phylogenetic analyses were conducted in Mega 5 using the Kimura 2-parameter method [<a href="#B48-ijms-17-00140" class="html-bibr">48</a>] to compute evolutionary distances. The bootstrap values indicated at the nodes are based on 1000 bootstrap replicates. Branch values lower than 50% are hidden. Closed and open circles indicate <span class="html-italic">Microdochium</span> isolates from <span class="html-italic">Saprolegnia</span>-infected (diseased) or healthy salmon egg samples, respectively. Red, blue and black colors indicate strains from terrestrial/plant, aquatic and unknown sources of isolation, respectively. The scale bar indicates an evolutionary distance of 0.01 nucleotide substitution per sequence position. Twenty-five ITS sequences of good quality and at least 550 bp of reference strains of <span class="html-italic">Microdochium</span> were downloaded from GenBank; their strain names are preceded by the accession numbers.</p>
Full article ">Figure 3
<p>Detection of total fungi, <span class="html-italic">Microdochium</span> and <span class="html-italic">Trichoderma</span> in salmon egg samples by quantitative PCR. Total fungal community was detected using ITS4-ITS9 primers, <span class="html-italic">Microdochium lycopodinum/Microdochium phragmitis</span> species were detected by MPF-MPR primers, and <span class="html-italic">Trichoderma</span> species were detected by ITS1TrF-ITS4TrR primers. The concentration of DNA template of each sample was normalized at 5 ± 1 ng·μL<sup>−1</sup>. Closed and open circles indicate DNA samples extracted from <span class="html-italic">Saprolegnia</span>-infected (diseased) or healthy samples, respectively. The lane on the left is the size marker and the band size (bp) is indicated next to each band.</p>
Full article ">Figure 4
<p><span class="html-italic">In vitro</span> activities of <span class="html-italic">Microdochium</span> and <span class="html-italic">Trichoderma</span> isolates. (<b>a</b>) Dual culture of <span class="html-italic">Microdochium</span> isolate 749F1 or <span class="html-italic">Trichoderma</span> isolate 762F1b with <span class="html-italic">Saprolegnia diclina</span> 1152F4 on 1/5th strength potato dextrose agar (1/5PDA). <span class="html-italic">S. diclina</span> and the fungal isolates were pre-grown on 1/5PDA. A hyphal plug of <span class="html-italic">S. diclina</span> was inoculated on the left side of the fresh 1/5PDA and a hyphal plug of <span class="html-italic">Microdochium</span> isolate 749F1 or <span class="html-italic">Trichoderma</span> isolate 762F1b was inoculated on the right side. The dual cultures were incubated for six days at 20–25 °C. The black arrows indicate <span class="html-italic">S. diclina</span> plugs and the white arrows indicate <span class="html-italic">Microdochium</span> or <span class="html-italic">Trichoderma</span> plugs; (<b>b</b>) Microscopic pictures of the hyphal interaction between <span class="html-italic">Trichoderma</span> and <span class="html-italic">S. diclina</span>. The black arrows indicate hyphae of <span class="html-italic">S. diclina</span>. The white arrows indicate the coiling of <span class="html-italic">Trichoderma</span> hyphae around <span class="html-italic">S. diclina</span> hyphae (<b>top</b> pictures) or the formation of papilla-like structure of <span class="html-italic">Trichoderma</span> hyphae around <span class="html-italic">S. diclina</span> hyphae (<b>bottom</b> pictures). Scale bars represent 10 μm.</p>
Full article ">Figure 5
<p>Phylogenetic tree of ITS rRNA sequences of nine <span class="html-italic">Trichoderma</span> isolates and reference strains. The phylogenetic analyses were conducted in Mega 5 using the Tamura 3-parameter method [<a href="#B70-ijms-17-00140" class="html-bibr">70</a>] to compute evolutionary distances. The bootstrap values indicated at the nodes are based on 1000 bootstrap replicates. Branch values lower than 50% are hidden. Closed and open circles indicate isolates from <span class="html-italic">Saprolegnia</span>-infected (diseased) or healthy samples, respectively. Red, blue and black colors indicate strains from terrestrial/plant, aquatic and unknown sources of isolation, respectively. The scale bar indicates an evolutionary distance of 0.01 nucleotide substitution per sequence position. Outer labels describe section names based on the list of species in ISTH website [<a href="#B64-ijms-17-00140" class="html-bibr">64</a>] and only sections contained at least five strains are indicated. 104 ITS sequences of good quality and at least 550 bp of reference strains of <span class="html-italic">Trichoderma</span> were downloaded from GenBank; their corresponding strain names are preceded by the accession numbers. Strain names followed by “(T)” indicate type strains.</p>
Full article ">Figure 6
<p>Phylogenetic tree of <span class="html-italic">tef1</span> sequences of nine <span class="html-italic">Trichoderma viride</span> isolates and reference strains. The phylogenetic analyses were conducted in Mega 5 using the Tamura-Nei method [<a href="#B71-ijms-17-00140" class="html-bibr">71</a>] to compute evolutionary distances. The bootstrap values indicated at the nodes are based on 1000 bootstrap replicates. Branch values lower than 50% are hidden. Closed and open circles indicate isolates from <span class="html-italic">Saprolegnia</span>-infected (diseased) or healthy samples, respectively. Red, blue and black colors indicate strains from terrestrial/plant, aquatic and unknown sources of isolation, respectively. The scale bar indicates an evolutionary distance of 0.01 nucleotide substitution per sequence position. 11 sequences of good quality and at least 1200 bp of reference strains of <span class="html-italic">Trichoderma</span> were downloaded from GenBank; their strain names are preceded by the accession numbers. Strain name followed by “(T)” indicates type strain.</p>
Full article ">
1567 KiB  
Review
Cadmium Protection Strategies—A Hidden Trade-Off?
by Adolf Michael Sandbichler and Martina Höckner
Int. J. Mol. Sci. 2016, 17(1), 139; https://doi.org/10.3390/ijms17010139 - 21 Jan 2016
Cited by 81 | Viewed by 8341
Abstract
Cadmium (Cd) is a non-essential transition metal which is introduced into the biosphere by various anthropogenic activities. Environmental pollution with Cd poses a major health risk and Cd toxicity has been extensively researched over the past decades. This review aims at changing the [...] Read more.
Cadmium (Cd) is a non-essential transition metal which is introduced into the biosphere by various anthropogenic activities. Environmental pollution with Cd poses a major health risk and Cd toxicity has been extensively researched over the past decades. This review aims at changing the perspective by discussing protection mechanisms available to counteract a Cd insult. Antioxidants, induction of antioxidant enzymes, and complexation of Cd to glutathione (GSH) and metallothionein (MT) are the most potent protective measures to cope with Cd-induced oxidative stress. Furthermore, protection mechanisms include prevention of endoplasmic reticulum (ER) stress, mitophagy and metabolic stress, as well as expression of chaperones. Pre-exposure to Cd itself, or co-exposure to other metals or trace elements can improve viability under Cd exposure and cells have means to reduce Cd uptake and improve Cd removal. Finally, environmental factors have negative or positive effects on Cd toxicity. Most protection mechanisms aim at preventing cellular damage. However, this might not be possible without trade-offs like an increased risk of carcinogenesis. Full article
(This article belongs to the Special Issue Metal Metabolism in Animals)
Show Figures

Figure 1

Figure 1
<p>Cell density assay (Hoechst 33342) with Z3 zebrafish cells in control experiments using different culture media. Cell numbers were measured once after the 18 h treatment and once after the 6 h recovery period. L-15−: L-15 complete media without FBS L-15+: L-15 complete media. Cell numbers were normalized to 10,000 cells of the Hank’s buffered salt solution (HBSS) control. Statistical analysis was performed prior to data normalization using a <span class="html-italic">t</span>-test. Groups were compared to HBSS treatment (* <span class="html-italic">p</span> ≤ 0.05). Values are mean ± standard error from 3 biological replicates.</p>
Full article ">Figure 2
<p>Cell density assay (Hoechst 33342) with Z3 zebrafish cells. (<b>A</b>) Effect of vitamin C (VC) and <span class="html-italic">N</span>-acetylcysteine (NAC) on HBSS incubated cells; (<b>B</b>) Recovery from HBSS and Cd treatment using NAC; (<b>C</b>) Recovery from HBSS and Cd treatment using VC. Cell numbers were normalized to 10,000 cells of the HBSS control. Statistical analysis was performed prior to data normalization using a <span class="html-italic">t</span>-test. Exposures were compared to HBSS treatment (* <span class="html-italic">p</span> ≤ 0.05). Square bracket indicates statistical significance from comparison of normalized data. Values are mean ± standard error from 3 biological replicates.</p>
Full article ">
695 KiB  
Review
The TEAD Family and Its Oncogenic Role in Promoting Tumorigenesis
by Yuhang Zhou, Tingting Huang, Alfred S. L. Cheng, Jun Yu, Wei Kang and Ka Fai To
Int. J. Mol. Sci. 2016, 17(1), 138; https://doi.org/10.3390/ijms17010138 - 21 Jan 2016
Cited by 144 | Viewed by 15904
Abstract
The TEAD family of transcription factors is necessary for developmental processes. The family members contain a TEA domain for the binding with DNA elements and a transactivation domain for the interaction with transcription coactivators. TEAD proteins are required for the participation of coactivators [...] Read more.
The TEAD family of transcription factors is necessary for developmental processes. The family members contain a TEA domain for the binding with DNA elements and a transactivation domain for the interaction with transcription coactivators. TEAD proteins are required for the participation of coactivators to transmit the signal of pathways for the downstream signaling processes. TEADs also play an important role in tumor initiation and facilitate cancer progression via activating a series of progression-inducing genes, such as CTGF, Cyr61, Myc and Gli2. Recent studies have highlighted that TEADs, together with their coactivators, promote or even act as the crucial parts in the development of various malignancies, such as liver, ovarian, breast and prostate cancers. Furthermore, TEADs are proposed to be useful prognostic biomarkers due to the ideal correlation between high expression and clinicopathological parameters in gastric, breast, ovarian and prostate cancers. In this review, we summarize the functional role of TEAD proteins in tumorigenesis and discuss the key role of TEAD transcription factors in the linking of signal cascade transductions. Improved knowledge of the TEAD proteins will be helpful for deep understanding of the molecular mechanisms of tumorigenesis and identifying ideal predictive or prognostic biomarkers, even providing clinical translation for anticancer therapy in human cancers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The regulatory cascade of TEAD family in cancer cells as transcription factors. YAP/TAZ, vgll, and p160 family proteins are the main binding partners for TEADs to activate the downstream transcription as transcription co-activators. <span class="html-italic">CTGF</span>, <span class="html-italic">Cyr61</span>, <span class="html-italic">Myc</span>, <span class="html-italic">AREG</span> and <span class="html-italic">AXL</span> are the downstream targets of YAP and TAZ by interaction with TEADs. vgll directly regulates the expression of <span class="html-italic">IGFBP-5</span>, <span class="html-italic">VEGFA</span> and <span class="html-italic">IRF2BP2</span> through TEADs. In colon cancer, TEAD4 also regulates Vimentin expression in a YAP-independent manner.</p>
Full article ">
468 KiB  
Review
The Clinical Significance of Phosphorylated Heat Shock Protein 27 (HSPB1) in Pancreatic Cancer
by Mitsuru Okuno, Seiji Adachi, Osamu Kozawa, Masahito Shimizu and Ichiro Yasuda
Int. J. Mol. Sci. 2016, 17(1), 137; https://doi.org/10.3390/ijms17010137 - 21 Jan 2016
Cited by 17 | Viewed by 7750
Abstract
Pancreatic cancer is one of most aggressive forms of cancer. After clinical detection it exhibits fast metastatic growth. Heat shock protein 27 (HSP27; HSPB1) has been characterized as a molecular chaperone which modifies the structures and functions of other proteins in cells when [...] Read more.
Pancreatic cancer is one of most aggressive forms of cancer. After clinical detection it exhibits fast metastatic growth. Heat shock protein 27 (HSP27; HSPB1) has been characterized as a molecular chaperone which modifies the structures and functions of other proteins in cells when they are exposed to various stresses, such as chemotherapy. While the administration of gemcitabine, an anti-tumor drug, has been the standard treatment for patients with advanced pancreatic cancer, accumulating evidence shows that HSP27 plays a key role in the chemosensitivity to gemcitabine. In addition, phosphorylated HSP27 induced by gemcitabine has been associated with the inhibition of pancreatic cancer cell growth. In this review, we summarize the role of phosphorylated HSP27, as well as HSP27, in the regulation of chemosensitivity in pancreatic cancer. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic representation of the role of HSP27 in the sensitivity to gemcitabine in human pancreatic cancer. Gemcitabine induces the phosphorylation of HSP27 via the MAPK-MAPKAPK-2 pathway and phosphorylated HSP27 leads cells to growth suppression in pancreatic cancer.</p>
Full article ">
886 KiB  
Article
Column Selection for Biomedical Analysis Supported by Column Classification Based on Four Test Parameters
by Alina Plenis, Natalia Rekowska and Tomasz Bączek
Int. J. Mol. Sci. 2016, 17(1), 136; https://doi.org/10.3390/ijms17010136 - 21 Jan 2016
Cited by 2 | Viewed by 4378
Abstract
This article focuses on correlating the column classification obtained from the method created at the Katholieke Universiteit Leuven (KUL), with the chromatographic resolution attained in biomedical separation. In the KUL system, each column is described with four parameters, which enables estimation of the [...] Read more.
This article focuses on correlating the column classification obtained from the method created at the Katholieke Universiteit Leuven (KUL), with the chromatographic resolution attained in biomedical separation. In the KUL system, each column is described with four parameters, which enables estimation of the FKUL value characterising similarity of those parameters to the selected reference stationary phase. Thus, a ranking list based on the FKUL value can be calculated for the chosen reference column, then correlated with the results of the column performance test. In this study, the column performance test was based on analysis of moclobemide and its two metabolites in human plasma by liquid chromatography (LC), using 18 columns. The comparative study was performed using traditional correlation of the FKUL values with the retention parameters of the analytes describing the column performance test. In order to deepen the comparative assessment of both data sets, factor analysis (FA) was also used. The obtained results indicated that the stationary phase classes, closely related according to the KUL method, yielded comparable separation for the target substances. Therefore, the column ranking system based on the FKUL-values could be considered supportive in the choice of the appropriate column for biomedical analysis. Full article
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The FA plot of the variables (<b>A</b>) and objects (<b>B</b>) established based on the auto-scaled KUL test parameters calculated for the 18 columns studied.</p>
Full article ">Figure 2
<p>The FA plot for the variables (<b>A</b>) and objects (<b>B</b>) established for the auto-scaled retention parameters (<span class="html-italic">t<sub>R</sub></span> and <span class="html-italic">R<sub>s</sub></span>) of the compounds of interest during the column performance test based on the LC analysis of M<sub>0</sub> and its two metabolites.</p>
Full article ">
1220 KiB  
Review
Photodynamic Therapy in Non-Gastrointestinal Thoracic Malignancies
by Biniam Kidane, Dhruvin Hirpara and Kazuhiro Yasufuku
Int. J. Mol. Sci. 2016, 17(1), 135; https://doi.org/10.3390/ijms17010135 - 21 Jan 2016
Cited by 9 | Viewed by 6608
Abstract
Photodynamic therapy has a role in the management of early and late thoracic malignancies. It can be used to facilitate minimally-invasive treatment of early endobronchial tumours and also to palliate obstructive and bleeding effects of advanced endobronchial tumours. Photodynamic therapy has been used [...] Read more.
Photodynamic therapy has a role in the management of early and late thoracic malignancies. It can be used to facilitate minimally-invasive treatment of early endobronchial tumours and also to palliate obstructive and bleeding effects of advanced endobronchial tumours. Photodynamic therapy has been used as a means of downsizing tumours to allow for resection, as well as reducing the extent of resection necessary. It has also been used successfully for minimally-invasive management of local recurrences, which is especially valuable for patients who are not eligible for radiation therapy. Photodynamic therapy has also shown promising results in mesothelioma and pleural-based metastatic disease. As new generation photosensitizers are being developed and tested and methodological issues continue to be addressed, the role of photodynamic therapy in thoracic malignancies continues to evolve. Full article
(This article belongs to the Special Issue Advances in Photodynamic Therapy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photodynamic therapy (PDT) for radiographically occult early stage lung cancer. (<b>A</b>) A 65-year-old male presented with abnormal sputum cytology and normal chest X-ray; (<b>B</b>) Bronchoscopy revealed an endobronchial abnormality at the orifice of the left upper division anterior segmental bronchus (arrow). Biopsy was positive for squamous cell carcinoma; and (<b>C</b>) Bronchoscopy six months post PDT (120 J/cm<sup>2</sup>) shows complete response with no residual tumour.</p>
Full article ">Figure 2
<p>PDT for central type lung cancer. (<b>A</b>) Endobronchial tumour (squamous cell carcinoma) completely obstructing the left upper lobe bronchus and extending into the left main stem bronchus; (<b>B</b>) complete response to PDT with only scarring on follow up bronchoscopy one year post PDT.</p>
Full article ">Figure 3
<p>PDT for recurrent lung cancer. (<b>A</b>) A 70-year-old male with history of left upper lobectomy for stage IA T1aN0M0 squamous cell carcinoma. Bronchoscopy performed 10 years post resection for the investigation of abnormal sputum cytology revealed abnormal bronchial mucosa in the left lower lobe basal lateral segmental bronchus; (<b>B</b>) Narrow band imaging shows dotted vessels within the abnormal mucosa; (<b>C</b>) Autofluorescence bronchoscopy shows abnormal fluorescence in the same area; and (<b>D</b>) Follow-up bronchoscopy six months post PDT (180 J/cm<sup>2</sup>) shows only scarring from the treatment with complete response.</p>
Full article ">
818 KiB  
Review
Structure Prediction: New Insights into Decrypting Long Noncoding RNAs
by Kun Yan, Yasir Arfat, Dijie Li, Fan Zhao, Zhihao Chen, Chong Yin, Yulong Sun, Lifang Hu, Tuanmin Yang and Airong Qian
Int. J. Mol. Sci. 2016, 17(1), 132; https://doi.org/10.3390/ijms17010132 - 21 Jan 2016
Cited by 51 | Viewed by 8100
Abstract
Long noncoding RNAs (lncRNAs), which form a diverse class of RNAs, remain the least understood type of noncoding RNAs in terms of their nature and identification. Emerging evidence has revealed that a small number of newly discovered lncRNAs perform important and complex biological [...] Read more.
Long noncoding RNAs (lncRNAs), which form a diverse class of RNAs, remain the least understood type of noncoding RNAs in terms of their nature and identification. Emerging evidence has revealed that a small number of newly discovered lncRNAs perform important and complex biological functions such as dosage compensation, chromatin regulation, genomic imprinting, and nuclear organization. However, understanding the wide range of functions of lncRNAs related to various processes of cellular networks remains a great experimental challenge. Structural versatility is critical for RNAs to perform various functions and provides new insights into probing the functions of lncRNAs. In recent years, the computational method of RNA structure prediction has been developed to analyze the structure of lncRNAs. This novel methodology has provided basic but indispensable information for the rapid, large-scale and in-depth research of lncRNAs. This review focuses on mainstream RNA structure prediction methods at the secondary and tertiary levels to offer an additional approach to investigating the functions of lncRNAs. Full article
(This article belongs to the Collection Regulation by Non-coding RNAs)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The graphical abstract of this review.</p>
Full article ">Figure 2
<p>Mechanisms of lncRNA action in transcriptional regulation. (<b>a</b>) Transcription of the lncRNA SRG1 inhibits the expression of the SER3 gene by interfering with the binding of RNA polymerase II to DNA; (<b>b</b>) The expression of the p15 antisense RNA, the lncRNA of a tumor suppressor gene, results in the silencing of the p15 gene through the induction of heterochromatin formation, which persisted after the p15 antisense RNA was turned off; (<b>c</b>) lncRNA binds to the major DHFR promoter and IIB, a general transcriptional factor, to form a stable and specific complex to dissociate the preinitiation complex from the major DHFR promoter; (<b>d</b>) As a response to stress, the RNA-binding protein TLS, under allosteric modulation via lncRNA upstream of CCND1, binds to chromatin-binding protein (CBP) and inhibits CBP/P300 HAT activities on CCND1; (<b>e</b>) The lncRNA Evf2, a crucial co-enhancer of regulatory proteins involved in transcription, cooperates with the Dlx2 protein to activate the Dlx5/6 enhancer in a target gene; (<b>f</b>) In response to heat shock, the lncRNA HSR1 (heat shock RNA-1) promotes the trimerization of HSF1 (heat-shock transcription factor 1), and consequently the translation factor EIF interacts with HSR1 and HSF1 to forms a complex to facilitate the expression of heat-shock protein (HSP); (<b>g</b>) NFAT is nuclear factor of activated T cells. The lncRNA NRON (noncoding repressor of NFAT) may form a complex with importin proteins to regulate the subcellular localization of NFAT. The knockdown of NRON increases the expression and activity of NFAT; (<b>h</b>) The lncRNA metastasis-associated lung adenocarcinoma transcript 1(MALAT1) has been shown to be abnormally expressed in many human cancers. The nascent MALAT1 transcript is cleaved by RNase P to produce the 3′ end of the mature MALAT1 transcript and the 5′ end of the small RNA; (<b>i</b>) Several studies have elucidated that some lncRNAs can act as microRNA sponges to competitively bind to microRNAs and decrease microRNA-induced tumorsphere differentiation.</p>
Full article ">
2691 KiB  
Article
VLDL from Metabolic Syndrome Individuals Enhanced Lipid Accumulation in Atria with Association of Susceptibility to Atrial Fibrillation
by Hsiang-Chun Lee, Hsin-Ting Lin, Liang-Yin Ke, Chi Wei, Yi-Lin Hsiao, Chih-Sheng Chu, Wen-Ter Lai, Shyi-Jang Shin, Chu-Huang Chen, Sheng-Hsiung Sheu and Bin-Nan Wu
Int. J. Mol. Sci. 2016, 17(1), 134; https://doi.org/10.3390/ijms17010134 - 20 Jan 2016
Cited by 14 | Viewed by 6312
Abstract
Metabolic syndrome (MetS) represents a cluster of metabolic derangements. Dyslipidemia is an important factor in MetS and is related to atrial fibrillation (AF). We hypothesized that very low density lipoproteins (VLDL) in MetS (MetS-VLDL) may induce atrial dilatation and vulnerability to AF. VLDL [...] Read more.
Metabolic syndrome (MetS) represents a cluster of metabolic derangements. Dyslipidemia is an important factor in MetS and is related to atrial fibrillation (AF). We hypothesized that very low density lipoproteins (VLDL) in MetS (MetS-VLDL) may induce atrial dilatation and vulnerability to AF. VLDL was therefore separated from normal (normal-VLDL) and MetS individuals. Wild type C57BL/6 male mice were divided into control, normal-VLDL (nVLDL), and MetS-VLDL (msVLDL) groups. VLDL (15 µg/g) and equivalent volumes of saline were injected via tail vein three times a week for six consecutive weeks. Cardiac chamber size and function were measured by echocardiography. MetS-VLDL significantly caused left atrial dilation (control, n = 10, 1.64 ± 0.23 mm; nVLDL, n = 7, 1.84 ± 0.13 mm; msVLDL, n = 10, 2.18 ± 0.24 mm; p < 0.0001) at week 6, associated with decreased ejection fraction (control, n = 10, 62.5% ± 7.7%, vs. msVLDL, n = 10, 52.9% ± 9.6%; p < 0.05). Isoproterenol-challenge experiment resulted in AF in young msVLDL mice. Unprovoked AF occurred only in elderly msVLDL mice. Immunohistochemistry showed excess lipid accumulation and apoptosis in msVLDL mice atria. These findings suggest a pivotal role of VLDL in AF pathogenesis for MetS individuals. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Metabolic syndrome-very low density lipoproteins (MetS-VLDL) is cytotoxic and provokes oxidative stress, with greater internalization in HL-1 atrial myocytes. (<b>A</b>) HL-1 cells were treated with normal-VLDL or MetS-VLDL at different test concentrations (3.125, 6.25, 12.5 and 25 mg/dL) for 24 h. OD values at 450 nm indicating viability were significantly lower in the msVLDL group (* <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 4 for each group); (<b>B</b>) HL-1 cells treated with MetS-VLDL showed significantly reduced cell viability (% of control) with a concentration of 25 mg/mL (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 4); (<b>C</b>) DCF fluorescence (excitation at 480 nm and emission at 520 nm) indicated total cytosolic oxidant activity (values of % control; <span class="html-italic">n</span> = 3 for each group). MetS-VLDL significantly increased oxidative stress at 25 mg/dL (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01); (<b>D</b>) Representative images for control, nVLDL, msVLDL, and msVLDL with VLDL receptor (VLDLR) antibody (msVLDL + Ab) groups (<span class="html-italic">n</span> = 3 for each group); (<b>E</b>) Internalization of DiI-labeled VLDL particles (red) increased in size and number in MetS-VLDL treated HL-1 cells (msVLDL) compared to normal-VLDL treated cells (nVLDL) (* <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3). Pre-treatment with VLDLR Ab for 24 h reduced internalization (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Both VLDLs caused LV dilation but only MetS-VLDL caused left atrial dilation. (<b>A</b>) Echocardiography of murine heart. Left atrium (LA) and left ventricle (LV) were identified in B-mode; (<b>B</b>) M-mode images for measurements of diameters of aortic root (AO), LA and LV. LA was significantly enlarged in the MetS-VLDL injection group (msVLDL) (<span class="html-italic">n</span> = 6) but not in the normal-VLDL injection group (nVLDL) (<span class="html-italic">n</span> = 7) or the control group (<span class="html-italic">n</span> = 5); (<b>C</b>) Significant LA enlargement developed as early as 4–6 weeks after injection in the msVLDL group. LV dilatation developed significantly until 6 weeks. (msVLDL <span class="html-italic">vs.</span> control, $ <span class="html-italic">p</span> &lt; 0.05; msVLDL <span class="html-italic">vs.</span> nVLDL, # <span class="html-italic">p</span> &lt; 0.05; nVLDL <span class="html-italic">vs.</span> control, * <span class="html-italic">p</span> &lt; 0.05); (<b>D</b>) No significant difference in body weight of the groups.</p>
Full article ">Figure 3
<p>Isoproterenol-induced and unprovoked atrial fibrillation (AF) were observed only in msVLDL mice. (<b>A</b>–<b>D</b>) Representative tracings of young mice after ISO injection show abnormalities including normal regular sinus rhythm in the control group (<span class="html-italic">n</span> = 5), premature atrial complex (PAC) in the nVLDL group, premature ventricular complex (PVC *) and AF (absence of clear P waves and irregular RR intervals) in the msVLDL group (<span class="html-italic">n</span> = 5); (<b>E</b>) Heart rate responses after isoproterenol injection were not different among groups; (<b>F</b>) For elderly mice, spontaneous, unprovoked AF was noted in the msVLDL group (<span class="html-italic">n</span> = 6) with an incidence of 50%. PAC was observed in one mouse in the nVLDL group (<span class="html-italic">n</span> = 5). All control mice (<span class="html-italic">n</span> = 5) had sinus rhythm. $ <span class="html-italic">p</span> &lt; 0.001 for msVLDL <span class="html-italic">vs.</span> control and # <span class="html-italic">p</span> &lt; 0.001 for msVLDL <span class="html-italic">vs.</span> nVLDL.</p>
Full article ">Figure 4
<p>Apoptosis in atrial tissue of msVLDL mice. Representative <span class="html-italic">in situ</span> terminal deoxynucleotidyl transferase (TUNEL) staining of atrial tissues from control (<b>left</b>), nVLDL (<b>middle</b>), and msVLDL (<b>right</b>) (<span class="html-italic">n</span> = 3 for each groups). Normal nuclei with DAPI staining appear blue. Condensed or fragmented nuclei appeared bright green and indicate cells undergoing apoptosis. Arrows indicate apoptotic atrial myocytes in the msVLDL group. The scale bars indicate 100 µm.</p>
Full article ">Figure 5
<p>Greater lipid accumulation in atrial tissue of msVLDL mice. (<b>A</b>) Representative Oil-Red-O-stained sections of atrial tissues from control (<b>left</b>), nVLDL (<b>middle</b>), and msVLDL (<b>right</b>). Each red rectangle indicates the area to be magnified (20×). Tiny red lipid droplets in controls were few but the number increased in nVLDL and msVLDL atria. Some lipid droplets increased in size in the msVLDL; (<b>B</b>) Lipid droplets were significantly increased in the VLDL groups, especially in the msVLDL group (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 3 for each group).</p>
Full article ">Figure 6
<p>Potential mechanism by which VLDL promotes AF in MetS. In MetS, the biochemical properties of VLDL are changed. MetS-VLDL can induce cellular reactive oxygen species, atrial myocyte cytotoxicity, and excess lipid accumulation resulting in subsequent gene dysregulation corresponding to metabolic derangement. Structural and potentially electrical remodeling initiated by MetS-VLDL in concert contribute to AF vulnerability and development.</p>
Full article ">
712 KiB  
Review
Iron Homeostasis in Health and Disease
by Raffaella Gozzelino and Paolo Arosio
Int. J. Mol. Sci. 2016, 17(1), 130; https://doi.org/10.3390/ijms17010130 - 20 Jan 2016
Cited by 249 | Viewed by 19829
Abstract
Iron is required for the survival of most organisms, including bacteria, plants, and humans. Its homeostasis in mammals must be fine-tuned to avoid iron deficiency with a reduced oxygen transport and diminished activity of Fe-dependent enzymes, and also iron excess that may catalyze [...] Read more.
Iron is required for the survival of most organisms, including bacteria, plants, and humans. Its homeostasis in mammals must be fine-tuned to avoid iron deficiency with a reduced oxygen transport and diminished activity of Fe-dependent enzymes, and also iron excess that may catalyze the formation of highly reactive hydroxyl radicals, oxidative stress, and programmed cell death. The advance in understanding the main players and mechanisms involved in iron regulation significantly improved since the discovery of genes responsible for hemochromatosis, the IRE/IRPs machinery, and the hepcidin-ferroportin axis. This review provides an update on the molecular mechanisms regulating cellular and systemic Fe homeostasis and their roles in pathophysiologic conditions that involve alterations of iron metabolism, and provides novel therapeutic strategies to prevent the deleterious effect of its deficiency/overload. Full article
(This article belongs to the Special Issue Metal Metabolism in Animals)
Show Figures

Figure 1

Figure 1
<p>The importance of Iron in pathophysiologic conditions. Essential to ensure survival, disruption of iron homeostasis has been shown to be involved in a variety of pathophysiological conditions, which include anemia and iron-overload related disorders. In particular, the importance of tissue iron accumulation in inflammation and infection, cancer, genetic, cardiovascular and neurodegenerative diseases continuously increases.</p>
Full article ">
6385 KiB  
Article
Computational Analysis of Structure-Based Interactions for Novel H1-Antihistamines
by Yinfeng Yang, Yan Li, Yanqiu Pan, Jinghui Wang, Feng Lin, Chao Wang, Shuwei Zhang and Ling Yang
Int. J. Mol. Sci. 2016, 17(1), 129; https://doi.org/10.3390/ijms17010129 - 19 Jan 2016
Cited by 18 | Viewed by 8557
Abstract
As a chronic disorder, insomnia affects approximately 10% of the population at some time during their lives, and its treatment is often challenging. Since the antagonists of the H1 receptor, a protein prevalent in human central nervous system, have been proven as [...] Read more.
As a chronic disorder, insomnia affects approximately 10% of the population at some time during their lives, and its treatment is often challenging. Since the antagonists of the H1 receptor, a protein prevalent in human central nervous system, have been proven as effective therapeutic agents for treating insomnia, the H1 receptor is quite possibly a promising target for developing potent anti-insomnia drugs. For the purpose of understanding the structural actors affecting the antagonism potency, presently a theoretical research of molecular interactions between 129 molecules and the H1 receptor is performed through three-dimensional quantitative structure-activity relationship (3D-QSAR) techniques. The ligand-based comparative molecular similarity indices analysis (CoMSIA) model (Q2 = 0.525, R2ncv = 0.891, R2pred = 0.807) has good quality for predicting the bioactivities of new chemicals. The cross-validated result suggests that the developed models have excellent internal and external predictability and consistency. The obtained contour maps were appraised for affinity trends for the investigated compounds, which provides significantly useful information in the rational drug design of novel anti-insomnia agents. Molecular docking was also performed to investigate the mode of interaction between the ligand and the active site of the receptor. Furthermore, as a supplementary tool to study the docking conformation of the antagonists in the H1 receptor binding pocket, molecular dynamics simulation was also applied, providing insights into the changes in the structure. All of the models and the derived information would, we hope, be of help for developing novel potent histamine H1 receptor antagonists, as well as exploring the H1-antihistamines interaction mechanism. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of activities (p<span class="html-italic">K<sub>i</sub></span>) for the training and the test sets <span class="html-italic">versus</span> the numbers of compounds. The training and the test sets are colored blue and orange, respectively.</p>
Full article ">Figure 2
<p>The ligand-based correlation plots of the predicted <span class="html-italic">versus</span> the actual p<span class="html-italic">K<sub>i</sub></span> values using the training (filled red triangles) and the test (filled black dots) set compounds based on the optimal CoMSIA model.</p>
Full article ">Figure 3
<p>Contour maps of CoMSIA combined with compound 49. (<b>A</b>) Contour maps in steric (green/yellow) fields. Green and yellow contours represent regions where bulky groups will increase and decrease the activity, respectively; (<b>B</b>) Contour maps in electrostatic (red/blue) fields. Red and blue contours represent regions where negative- and positive-charged substituents will decrease and increase the activity, respectively; (<b>C</b>) Contour maps in hydrophobic (yellow/gray) fields. Yellow and gray contours represent regions where the hydrophobic and hydrophilic groups will increase their activity; (<b>D</b>) Contour maps in H-bond (HB) donor (cyan/purple) fields. Cyan and purple contours represent regions where HB donor substituents will enhance and decrease the activity, respectively.</p>
Full article ">Figure 4
<p>(<b>A</b>) Binding poses of co-crystallized (magenta) and re-docked (green) compound doxepin; (<b>B</b>) overlap of the compound 49 (orange) and experimental doxepin (green; PDB code: 3RZE) conformation.</p>
Full article ">Figure 5
<p>Docked conformation of compound 49 into histamine H<sub>1</sub> receptor. The projection highlights the structure of the active site with compound 49, which is displayed in sticks.</p>
Full article ">Figure 6
<p>(<b>A</b>) Plot of the root-mean-square deviation (RMSD) of docked complex/ligand <span class="html-italic">versus</span> the MD simulation time in the MD-simulated structures; (<b>B</b>) view of the superimposed backbone atoms of the average structure for the MD simulations and the initial structure of the docking for the complex. Compound 49 is represented as a carbon-chain in green for the initial complex and a carbon-chain in orange for the average structure, respectively.</p>
Full article ">Figure 7
<p>Plot of the MD-simulated structures of the binding site with compound 49. H-bonds are shown as dotted black lines; amino acid residues in the active site are represented as sticks.</p>
Full article ">Figure 8
<p>(<b>A</b>) Compound 49 used as the template molecule for alignments, with the common framework marked in blue bold. The substituent containing a protonated –NMe<sub>2</sub> group at the position-18 is depicted in a red oval, which would be more desirable for potent antagonism activity; (<b>B</b>–<b>D</b>) show the results of Alignment-I, -II and -III of all molecules, respectively. All compounds in these panels are colored white for common carbon, blue for nitrogen, red for oxygen, yellow for sulfur and cyan for hydrogen atoms, respectively.</p>
Full article ">Figure 9
<p>Proposed hypothetical histamine H<sub>1</sub>-receptor active site models. The structure-activity relationship is taken from the results of 3D-QSAR, docking and MD simulation studies for compound 49.</p>
Full article ">
1296 KiB  
Review
Lung Regeneration: Endogenous and Exogenous Stem Cell Mediated Therapeutic Approaches
by Khondoker M. Akram, Neil Patel, Monica A. Spiteri and Nicholas R. Forsyth
Int. J. Mol. Sci. 2016, 17(1), 128; https://doi.org/10.3390/ijms17010128 - 19 Jan 2016
Cited by 55 | Viewed by 17871
Abstract
The tissue turnover of unperturbed adult lung is remarkably slow. However, after injury or insult, a specialised group of facultative lung progenitors become activated to replenish damaged tissue through a reparative process called regeneration. Disruption in this process results in healing by fibrosis [...] Read more.
The tissue turnover of unperturbed adult lung is remarkably slow. However, after injury or insult, a specialised group of facultative lung progenitors become activated to replenish damaged tissue through a reparative process called regeneration. Disruption in this process results in healing by fibrosis causing aberrant lung remodelling and organ dysfunction. Post-insult failure of regeneration leads to various incurable lung diseases including chronic obstructive pulmonary disease (COPD) and idiopathic pulmonary fibrosis. Therefore, identification of true endogenous lung progenitors/stem cells, and their regenerative pathway are crucial for next-generation therapeutic development. Recent studies provide exciting and novel insights into postnatal lung development and post-injury lung regeneration by native lung progenitors. Furthermore, exogenous application of bone marrow stem cells, embryonic stem cells and inducible pluripotent stem cells (iPSC) show evidences of their regenerative capacity in the repair of injured and diseased lungs. With the advent of modern tissue engineering techniques, whole lung regeneration in the lab using de-cellularised tissue scaffold and stem cells is now becoming reality. In this review, we will highlight the advancement of our understanding in lung regeneration and development of stem cell mediated therapeutic strategies in combating incurable lung diseases. Full article
(This article belongs to the Special Issue Stem Cell Activation in Adult Organism)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Epithelial-mesenchymal cross-talk and governing signalling during early development and branching morphogenesis of lung. Factors are represented only at sites where the expression is most abundant. Fibroblast growth factor-10 (FGF-10) is highly expressed in the distal mesenchyme and acts as a chemotactic focus for the epithelium during lung budding. FGF-10 also regulates Sox9 expression in the distal epithelial progenitors and induces bone morphogenetic protein (BMP)-4 expression. Sox2 expression in the proximal epithelium is under regulation of histone deacetylases 1/2 (HDAC1/2) signalling. FGF-10 expression in the mesenchyme is regulated by Wnt/β-catenin signalling (red arrow). A high concentration of BMP-4 signal also serves to locally inhibit endoderm proliferation, thereby inducing the lateral outgrowth of new airway branches. Sonic hedgehog (Shh) at the distal tips functions to downregulate FGF-10 expression in the mesenchyme, which limits local budding. Transforming growth factor-β (TGF-β) signalling also prevents local budding, by decreasing endodermal proliferation and by stimulating synthesis of matrix components at branch points. Solid arrows indicate sources from and influences on cells/molecules; dotted arrow indicates direction of patterning.</p>
Full article ">Figure 2
<p>Endogenous stem cells in the airways and alveoli with their differentiation in the postnatal lung. (<b>A</b>) The trachea and bronchi of the rodent and human lung are lined with multiple epithelial lineages. Basal cells are located in this region and can generate secretory club and ciliated cell lineages. Notch signalling is crucial for differentiation of basal cells and also suppresses ciliated cell differentiation. HDAC1 and HDAC2 are essential for secretory epithelial regeneration; (<b>B</b>) The bronchiolar lining epithelium of rodents lacks basal cells but contains variant club cells, secretory cells and ciliated cells; (<b>C</b>) Progenitor cell populations and their differentiated progeny in the lung alveolus. The alveolar epithelium consists of alveolar epithelial cell AECI and AECII cells. AECII cells can generate AECI cells during homeostasis and after injury. Generation of alveolar epithelium by other cells, such as BASCs has yet to be supported by lineage tracing. The AECII to AECI differentiation signalling is not clear (marked by “?”), but AECII self-renewal and proliferation occurs via EGFR/KRAS signalling. Curved arrows indicate differentiation and line arrows indicate molecular or physical stimulus.</p>
Full article ">
1329 KiB  
Review
The Complex Relationship between Metals and Carbonic Anhydrase: New Insights and Perspectives
by Maria Giulia Lionetto, Roberto Caricato, Maria Elena Giordano and Trifone Schettino
Int. J. Mol. Sci. 2016, 17(1), 127; https://doi.org/10.3390/ijms17010127 - 19 Jan 2016
Cited by 64 | Viewed by 8098
Abstract
Carbonic anhydrase is a ubiquitous metalloenzyme, which catalyzes the reversible hydration of CO2 to HCO3 and H+. Metals play a key role in the bioactivity of this metalloenzyme, although their relationships with CA have not been completely clarified [...] Read more.
Carbonic anhydrase is a ubiquitous metalloenzyme, which catalyzes the reversible hydration of CO2 to HCO3 and H+. Metals play a key role in the bioactivity of this metalloenzyme, although their relationships with CA have not been completely clarified to date. The aim of this review is to explore the complexity and multi-aspect nature of these relationships, since metals can be cofactors of CA, but also inhibitors of CA activity and modulators of CA expression. Moreover, this work analyzes new insights and perspectives that allow translating new advances in basic science on the interaction between CA and metals to applications in several fields of research, ranging from biotechnology to environmental sciences. Full article
(This article belongs to the Special Issue Metal Metabolism in Animals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The reversible hydration of carbon dioxide to bicarbonate catalyzed by CAs by means of a metal (M)-hydroxide mechanism. Modified from Berg [<a href="#B17-ijms-17-00127" class="html-bibr">17</a>]. (<b>1</b>) The release of a proton from the zinc-bound water generates the zinc-bound OH<sup>−</sup>; (<b>2</b>) A CO<sub>2</sub> molecule binds to the active site and is positioned for optimal interaction with the zinc-bound OH<sup>−</sup>; (<b>3</b>) The hydroxide ion attacks the carbonyl of CO<sub>2</sub>, producing HCO<sub>3</sub><sup>−</sup>; (<b>4</b>) The release of HCO<sub>3</sub><sup>−</sup> regenerates the enzyme.</p>
Full article ">Figure 2
<p>Human CAII: in detail, the metal binding site with the zinc ion as a sphere, the direct ligand histidines, H94, H96, H119, and the water molecule. Modified from Mahon <span class="html-italic">et al</span> [<a href="#B31-ijms-17-00127" class="html-bibr">31</a>] and from Dutta and Goodsell [<a href="#B32-ijms-17-00127" class="html-bibr">32</a>].</p>
Full article ">Figure 3
<p>From <span class="html-italic">in vivo</span> and <span class="html-italic">in vivo</span> assessment of the sensitivity of CA to trace metals to environmental monitoring.</p>
Full article ">
3116 KiB  
Article
Statin Therapy and the Development of Cerebral Amyloid Angiopathy—A Rodent in Vivo Approach
by Björn Reuter, Alexander Venus, Saskia Grudzenski, Patrick Heiler, Lothar Schad, Matthias Staufenbiel, Michael G. Hennerici and Marc Fatar
Int. J. Mol. Sci. 2016, 17(1), 126; https://doi.org/10.3390/ijms17010126 - 19 Jan 2016
Cited by 5 | Viewed by 7069
Abstract
Background: Cerebral amyloid angiopathy (CAA) is characterized by vascular deposition of amyloid β (Aβ) with a higher incidence of cerebral microbleeds (cMBs) and spontaneous hemorrhage. Since statins are known for their benefit in vascular disease we tested for the effect on CAA. Methods: [...] Read more.
Background: Cerebral amyloid angiopathy (CAA) is characterized by vascular deposition of amyloid β (Aβ) with a higher incidence of cerebral microbleeds (cMBs) and spontaneous hemorrhage. Since statins are known for their benefit in vascular disease we tested for the effect on CAA. Methods: APP23-transgenic mice received atorvastatin-supplemented food starting at the age of eight months (n = 13), 12 months (n = 7), and 16 months (n = 6), respectively. Controls (n = 16) received standard food only. At 24 months of age cMBs were determined with T2*-weighted 9.4T magnetic resonance imaging and graded by size. Results: Control mice displayed an average of 35 ± 18.5 cMBs (mean ± standard deviation), compared to 29.3 ± 9.8 in mice with eight months (p = 0.49), 24.9 ± 21.3 with 12 months (p = 0.26), and 27.8 ± 15.4 with 16 months of atorvastatin treatment (p = 0.27). In combined analysis treated mice showed lower absolute numbers (27.4 ± 15.6, p = 0.16) compared to controls and also after adjustment for cMB size (p = 0.13). Conclusion: Despite to a non-significant trend towards fewer cMBs our results failed to provide evidence for beneficial effects of long-term atorvastatin treatment in the APP23-transgenic mouse model of CAA. A higher risk for bleeding complications was not observed. Full article
(This article belongs to the Special Issue Amyloid-beta and Neurological Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Number of cerebral microbleeds (cMBs) in MRI and APP23-tg mice treated with atorvastatin for eight months (<span class="html-italic">n</span> = 6), 12 months (<span class="html-italic">n</span> = 7), and 16 months (<span class="html-italic">n</span> = 13), respectively. Compared to controls (<span class="html-italic">n</span> = 17) and in pooled analysis no significant differences between the groups were observed; (<b>B</b>) Histologically-assessed numbers of cMBs per slide in untreated mice (<span class="html-italic">n</span> = 6) did not differ significantly from those in mice treated for 16 months with atorvastatin (<span class="html-italic">n</span> = 7); (<b>C</b>) In these mice total numbers of thioflavin S positive vessels, graded by severity of amyloid β (Aβ) deposits, and after calculation of a cerebral amyloid angiopathy (CAA) severity score showed no significant difference between the groups.</p>
Full article ">Figure 2
<p>Time schedule of the study. APP23-tg mice received either standard food over the whole study period (controls, <span class="html-italic">n</span> = 17 with brain imaging and histology) or atorvastatin-supplemented food starting at the age of eight months (<span class="html-italic">n</span> = 13), 12 months (<span class="html-italic">n</span> = 7), and 16 months (<span class="html-italic">n</span> = 6), respectively.</p>
Full article ">Figure 3
<p>T2* weighted and time of flight magnetic resonance angiography (TOF-MRA) image of brain vessels and vessel associated cMBs in an APP23-tg mouse obtained by 9.4 T animal scanner. (<b>A</b>) Representative T2* weighted image with two hypointense areas (arrows); and (<b>B</b>) brepresentative TOF-MRA image of corresponding brain section from the same animal with a hypointense signal representing a cMB and a hyperintense signal representing a brain vessel.</p>
Full article ">Figure 4
<p>T2* weighted image of cMBs in an APP23-tg mouse obtained with 9.4 T animal scanner. Representative T2* weighted image of an APP23-tg mouse shows multiple cMBs which were graded by size (<b>left</b> image). Representative graded cMBs are shown belonging to group A ≤ 100 µm (red arrow), B = 150 ≤ x ≤ 200 µm and C &gt; 200 µm (<b>right</b> images).</p>
Full article ">Figure 5
<p>Fluorescence analysis of affected brain vessels demonstrating Aβ burden. Representative images of Aβ affected thioflavin S stained vessels in APP-tg mice. Vessels were graded into score 1, 2 or 3 depending on degree of Aβ burden. Blue = DAPI (4′,6-Diamidin-2-Phenylindol); red = STL (biotinylated Solanum Tuberosum (Potato) Lectin); green = thioflavin S.</p>
Full article ">
3440 KiB  
Article
Influence of Bxpel1 Gene Silencing by dsRNA Interference on the Development and Pathogenicity of the Pine Wood Nematode, Bursaphelenchus xylophilus
by Xiu-Wen Qiu, Xiao-Qin Wu, Lin Huang and Jian-Ren Ye
Int. J. Mol. Sci. 2016, 17(1), 125; https://doi.org/10.3390/ijms17010125 - 19 Jan 2016
Cited by 20 | Viewed by 5559
Abstract
As the causal agent of pine wilt disease (PWD), the pine wood nematode (PWN), Bursaphelenchus xylophilus, causes huge economic losses by devastating pine forests worldwide. The pectate lyase gene is essential for successful invasion of their host plants by plant-parasitic nematodes. To [...] Read more.
As the causal agent of pine wilt disease (PWD), the pine wood nematode (PWN), Bursaphelenchus xylophilus, causes huge economic losses by devastating pine forests worldwide. The pectate lyase gene is essential for successful invasion of their host plants by plant-parasitic nematodes. To demonstrate the role of pectate lyase gene in the PWD process, RNA interference (RNAi) is used to analyze the function of the pectate lyase 1 gene in B. xylophilus (Bxpel1). The efficiency of RNAi was detected by real-time PCR. The result demonstrated that the quantity of B. xylophilus propagated with control solution treatment was 62 times greater than that soaking in double-stranded RNA (dsRNA) after B. xylophilus inoculation in Botrytis cinerea for the first generation (F1). The number of B. xylophilus soaking in control solution was doubled compared to that soaking in Bxpel1 dsRNA four days after inoculation in Pinus thunbergii. The quantity of B. xylophilus was reduced significantly (p < 0.001) after treatment with dsRNAi compared with that using a control solution treatment. Bxpel1 dsRNAi reduced the migration speed and reproduction of B. xylophilus in pine trees. The pathogenicity to P. thunbergii seedling of B. xylophilus was weaker after soaking in dsRNA solution compared with that after soaking in the control solution. Our results suggest that Bxpel1 gene is a significant pathogenic factor in the PWD process and this basic information may facilitate a better understanding of the molecular mechanism of PWD. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The phenotype of <span class="html-italic">B. xylophilus</span> after soaking in doble distilled water (ddH<sub>2</sub>O) (<b>A</b>) and control solution (<b>B</b>) for 48 h. Scale bars = 50 μm.</p>
Full article ">Figure 1 Cont.
<p>The phenotype of <span class="html-italic">B. xylophilus</span> after soaking in doble distilled water (ddH<sub>2</sub>O) (<b>A</b>) and control solution (<b>B</b>) for 48 h. Scale bars = 50 μm.</p>
Full article ">Figure 2
<p>The phenotype of <span class="html-italic">B. xylophilus</span> after soaking in <span class="html-italic">Bxpel1</span> double-stranded RNA (dsRNA) for 48 h (<b>A</b>,<b>B</b>). Scale bars = 50 μm.</p>
Full article ">Figure 3
<p>Effects of double-stranded RNA interference (dsRNAi) on <span class="html-italic">B. xylophilus</span> propagation. (<b>A</b>) The growth of the first generation of <span class="html-italic">B. xylophilus</span> on <span class="html-italic">B. cinerea</span>; (<b>B</b>) The growth of the second generation of <span class="html-italic">B. xylophilus</span> on <span class="html-italic">B. cinerea.</span></p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">GFP</span> gene dsRNA on propagation of <span class="html-italic">B. xylophilus</span>. (<b>A</b>) <span class="html-italic">B. xylophilus</span> inoculating on <span class="html-italic">B. cinerea</span> on the first day; (<b>B</b>) <span class="html-italic">B. xylophilus</span> inoculating on <span class="html-italic">B. cinerea</span> on the seventh day<span class="html-italic">.</span></p>
Full article ">Figure 5
<p>Effects of dsRNAi on <span class="html-italic">B. xylophilus</span> reproduction. Vertical bars indicate standard deviation of the means (<span class="html-italic">n</span> = 3), and different letters indicate significant differences at a level of <span class="html-italic">p</span> &lt; 0.05 based on Tukey’s test. F1: first generation; F2: second generation.</p>
Full article ">Figure 6
<p>Expression of the <span class="html-italic">Bxpel1</span> gene in <span class="html-italic">B. xylophilus.</span> Vertical bars indicate standard deviation of the means (<span class="html-italic">n</span> = 3), and different letters indicate significant differences at a level of <span class="html-italic">p</span> &lt; 0.05 based on Tukey’s test. F3: third generation; F4: fourth generation; F5: fifth generation; F6: sixth generation; F7: seventh generation.</p>
Full article ">Figure 7
<p>Wilting symptoms development of <span class="html-italic">P. thunbergii</span> after inoculation with <span class="html-italic">B. xylophilus</span> at 1 day (<b>A</b>); 10 days (<b>B</b>); 20 days (<b>C</b>); and 30 days (<b>D</b>). The pine trees were inoculated with PWNs that had been soaked in control solution (CK1), ddH<sub>2</sub>O (CK2), or <span class="html-italic">Bxpel1</span> dsRNA (<span class="html-italic">Bxpel1</span>).</p>
Full article ">
499 KiB  
Review
Molecular Regulation of Adipogenesis and Potential Anti-Adipogenic Bioactive Molecules
by Dorothy Moseti, Alemu Regassa and Woo-Kyun Kim
Int. J. Mol. Sci. 2016, 17(1), 124; https://doi.org/10.3390/ijms17010124 - 19 Jan 2016
Cited by 506 | Viewed by 17830
Abstract
Adipogenesis is the process by which precursor stem cells differentiate into lipid laden adipocytes. Adipogenesis is regulated by a complex and highly orchestrated gene expression program. In mammalian cells, the peroxisome proliferator-activated receptor γ (PPARγ), and the CCAAT/enhancer binding proteins (C/EBPs) such as [...] Read more.
Adipogenesis is the process by which precursor stem cells differentiate into lipid laden adipocytes. Adipogenesis is regulated by a complex and highly orchestrated gene expression program. In mammalian cells, the peroxisome proliferator-activated receptor γ (PPARγ), and the CCAAT/enhancer binding proteins (C/EBPs) such as C/EBPα, β and δ are considered the key early regulators of adipogenesis, while fatty acid binding protein 4 (FABP4), adiponectin, and fatty acid synthase (FAS) are responsible for the formation of mature adipocytes. Excess accumulation of lipids in the adipose tissue leads to obesity, which is associated with cardiovascular diseases, type II diabetes and other pathologies. Thus, investigating adipose tissue development and the underlying molecular mechanisms is vital to develop therapeutic agents capable of curbing the increasing incidence of obesity and related pathologies. In this review, we address the process of adipogenic differentiation, key transcription factors and proteins involved, adipogenic regulators and potential anti-adipogenic bioactive molecules. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular regulation of adipogenesis.</p>
Full article ">
3037 KiB  
Article
Genipin Derivatives Protect RGC-5 from Sodium Nitroprusside-Induced Nitrosative Stress
by Rikang Wang, Jiaqiang Zhao, Lei Zhang, Lizhi Peng, Xinyi Zhang, Wenhua Zheng and Heru Chen
Int. J. Mol. Sci. 2016, 17(1), 117; https://doi.org/10.3390/ijms17010117 - 19 Jan 2016
Cited by 5 | Viewed by 5376
Abstract
CHR20 and CHR21 are a pair of stable diastereoisomers derived from genipin. These stereoisomers are activators of neuronal nitric oxide synthase (nNOS) and endothelial nitric oxide synthase (eNOS). In the rat retinal ganglion (RGC-5) cell model these compounds are non-toxic. Treatment of RGC-5 [...] Read more.
CHR20 and CHR21 are a pair of stable diastereoisomers derived from genipin. These stereoisomers are activators of neuronal nitric oxide synthase (nNOS) and endothelial nitric oxide synthase (eNOS). In the rat retinal ganglion (RGC-5) cell model these compounds are non-toxic. Treatment of RGC-5 with 750 μM of sodium nitroprusside (SNP) produces nitrosative stress. Both genipin derivatives, however, protect these cells against SNP-induced apoptic cell death, although CHR21 is significantly more potent than CHR20 in this regard. With Western blotting we showed that the observed neuroprotection is primarily due to the activation of protein kinase B (Akt)/eNOS and extracellular signal-regulated kinase (ERK1/2) signaling pathways. Therefore, LY294002 (a phosphatidylinositol 3-kinase (PI3K) inhibitor) or PD98059 (a MAPK-activating enzyme inhibitor) abrogated the protective effects of CHR20 and CHR21. Altogether, our results show that in our experimental setup neuroprotection by the diasteromeric pair is mediated through the PI3K/Akt/eNOS and ERK1/2 signaling pathways. Further studies are needed to establish the potential of these compounds to prevent ntric oxide (NO)-induced toxicity commonly seen in many neurodegenerative diseases. Full article
(This article belongs to the Special Issue Neuroprotective Strategies 2015)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of genipin and <b>CHR20/21</b>. “a” is a symbol in the IUPAC’ nomemclature rule to represent the first atom linked to the outside ring.</p>
Full article ">Figure 2
<p>The neuro-effects of SNP on RGC-5 cells. Cells were exposed to different concentrations of SNP as indicated for 24 h. Cell viability was evaluated by MTT assay. The values were expressed as percentage of control, which is set to 100%. The percentage of MTT activity was presented as mean ± SD for six replicates. <b>*</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group.</p>
Full article ">Figure 3
<p>Cytotoxicity and MTT activity of <b>CHR20</b> and <b>CHR21</b> on RGC-5 cells. Cells that were cultured in RPMI-1640 medium with low serum (0.2% of FBS) for 24 h were set as the control group. RGC-5 cells with 0.2% FBS were treated with <b>CHR20</b> (<b>A</b>) and <b>CHR21</b> (<b>B</b>) at concentrations of 3, 10 and 30 µM, respectively for 24 h. Cell viability was determined by MTT assay and presented as mean ± SD for six replicates. * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group.</p>
Full article ">Figure 4
<p>Effects of <b>CHR20/21</b> against SNP–induced insults in RGC-5 cells. RGC-5 cells were pre-treated with SNP at a concentration of 750 µmol/L for 0.5, 1, 2, and 4 h, respectively, then exposed to various concentrations of <b>CHR20</b> (<b>A</b>) and <b>CHR21</b> (<b>B</b>), respectively, for another 24 h. MTT activity was determined by MTT assay. The percentage of MTT activity was presented as mean ± SD for six replicates. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> control group, <b>**</b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP-treated group.</p>
Full article ">Figure 5
<p><b>CHR20</b> and <b>CHR21</b> protected RGC-5 cells from apoptosis induced by SNP. RGC-5 cells were pretreated with/without <b>CHR20/CHR21</b> for 2 h, respectively, after treatment of 750 μM SNP for 24 h and the cells were stained with Hoechst 33258 as described in Materials and Methods. (<b>A</b>) Morphological changes shown by fluorescence microscope (200×) image analysis. (a) control group; (b) 30 μM <b>CHR20</b> group; (c) 30 μM <b>CHR21</b> group; (d) 750 μM SNP group; (e) 750 μM SNP + 30 μM <b>CHR20</b> group; (f) 750 μM SNP + 30 μM <b>CHR21</b> group. The arrow indicates nuclear fragmentation or chromatin condensation; (<b>B</b>) Histogram showing the apoptosis rate in RGC-5 cells. The number of apoptic cells was about 1.5 × 10<sup>4</sup> cells/well in 25 visual fields in 48-well culture plate, counted by high content screening system (ArrayScanVTI, Thermo Fisher Scientific, Waltham, MA, USA). The percentage of apoptotic cells was calculated as ratio of apoptotic cells and the total number of cells counted. Data are given as mean ± SD for three individual experiments. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group. <b>**</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> SNP group; (<b>C</b>) Effects of <b>CHR20/CHR21</b> on cleaved caspase-3. The expression of cleaved caspase-3 was determined by western blotting as described in Materials and Methods.</p>
Full article ">Figure 6
<p><b>CHR20/CHR21</b> increase phosphorylation levels in PI3K/Akt and MEK/ERK1/2. RGC-5 cells were treated with 10 μM <b>CHR20</b> and <b>CHR21</b>, repectively for 5–80 min or 0.3–30 μM <b>CHR20/CHR21</b> for 40 min. By applying Western blotting with antibodies including anti-phospho-Akt (Ser473), anti-phospho-ERK1/2, anti-phospho-and eNOS (Ser1177), respectively, the phosphorylation of the relevant proteins were determined. (<b>A</b>) <b>CHR20</b> and <b>CHR21</b> time-dependently increased the phosphorylation levels of Akt and ERK1/2 in RGC-5 cells; (<b>B</b>) <b>CHR20</b> and <b>CHR21</b> dose-dependently induced the phosphorylation of Akt and ERK1/2 in RGC-5 cells; (<b>C</b>) Effects of pathway inhibitors on the phosphorylation of Akt, eNOS and ERK1/2 affected by <b>CHR21</b>. The density of the blot in each lane was presented as mean ± standard deviation. Each data was calculated based on at least three individual experiments. <b>*</b> <span class="html-italic">p</span> &lt; 0.05, <b>**</b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> control group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR21</b> pre-treated group. Blots were quantified using ImageJ software. pd: PD98059; ly: LY294002; p-eNOS: phosphorylated protein level of eNOS; pAkt: phosphorylated protein level of Akt; pErk: phosphorylated protein level of ERK; CHR: abbreviation of Chen HeRu; T-Erk: total protein level of ERK; p-NOS: phosphorylated protein level of NOS; T-NOS: total protein level of NOS.</p>
Full article ">Figure 6 Cont.
<p><b>CHR20/CHR21</b> increase phosphorylation levels in PI3K/Akt and MEK/ERK1/2. RGC-5 cells were treated with 10 μM <b>CHR20</b> and <b>CHR21</b>, repectively for 5–80 min or 0.3–30 μM <b>CHR20/CHR21</b> for 40 min. By applying Western blotting with antibodies including anti-phospho-Akt (Ser473), anti-phospho-ERK1/2, anti-phospho-and eNOS (Ser1177), respectively, the phosphorylation of the relevant proteins were determined. (<b>A</b>) <b>CHR20</b> and <b>CHR21</b> time-dependently increased the phosphorylation levels of Akt and ERK1/2 in RGC-5 cells; (<b>B</b>) <b>CHR20</b> and <b>CHR21</b> dose-dependently induced the phosphorylation of Akt and ERK1/2 in RGC-5 cells; (<b>C</b>) Effects of pathway inhibitors on the phosphorylation of Akt, eNOS and ERK1/2 affected by <b>CHR21</b>. The density of the blot in each lane was presented as mean ± standard deviation. Each data was calculated based on at least three individual experiments. <b>*</b> <span class="html-italic">p</span> &lt; 0.05, <b>**</b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> control group. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR21</b> pre-treated group. Blots were quantified using ImageJ software. pd: PD98059; ly: LY294002; p-eNOS: phosphorylated protein level of eNOS; pAkt: phosphorylated protein level of Akt; pErk: phosphorylated protein level of ERK; CHR: abbreviation of Chen HeRu; T-Erk: total protein level of ERK; p-NOS: phosphorylated protein level of NOS; T-NOS: total protein level of NOS.</p>
Full article ">Figure 7
<p>Effects of NOS on the action of <b>CHR21</b>. (<b>A</b>) Endothelial nitrioxide synthase (eNOS) mediates the action of <b>CHR21</b> on MTT activity in RGC-5 cells. Cells pre-exposed to 7-nitroindazole (7-NI, 50 μM, 30 min), an specific inhibitor for neuronal nitric oxide synthase (nNOS), <span class="html-italic">N</span>(5)-(-iminoethyl)-<span class="html-small-caps">l</span>-ornithine (<span class="html-small-caps">l</span>-NIO, 100 μM, 30 min), an inhibitor of endothelial NOS (eNOS), and compound 1400-W (20 μM, 30 min), a potent inhibitor of inducible NOS (iNOS), respectively, were treated with <b>CHR20/CHR21</b>. The MTT activity of cells determined by MTT assay. Only <span class="html-small-caps">l</span>-NIO attenuated the protective effects of <b>CHR20/21</b> on RGC-5 cells with statistically significant; while 7-NIO and 1400 W had no statistically significant effect. <b><sup>&amp;&amp;</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> control; <b><sup>#</sup><sup>#</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP group; <b>*</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> <b>CHR20</b> + SNP or <b>CHR21</b> + SNP (<span class="html-italic">n</span> = 3); (<b>B</b>) Effects of 7-NI, <span class="html-small-caps">l</span>-NIO and 1400 W on the NOS activities in RGC-5 cells. RGC-5 cells were treated with 7-NI, <span class="html-small-caps">l</span>-NIO and 1400 W, respectively. By applying Typed NOS Detection Kit, determination of the activities of nNOS, eNOS (endothelial NOS) and iNOS was carried out. One nmol of NO formed per mL cell lysate in one minute was defined as U/mL. With <b>*</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group, the difference is considered statistically significant. All values are expressed as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>Both the PI3K/Akt/eNOS pathway and the ERK1/2 pathways were involved in the neuroprotective effects of <b>CHR20/21</b>. Pre-incubation of LY294002, Akt VIII inhibitor, and PD98059 at a dose of 10 µM, respectively blocked the neuroprotective effects of <b>CHR20</b> (<b>A</b>) and <b>CHR21</b> (<b>B</b>) in RGC-5 cells. MTT activity was determined by MTT assay. Each experiment was performed in triplicate. <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP group; <b>**</b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR20/CHR21</b> + SNP group; <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>&amp;&amp;</sup> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR20/CHR21</b> + PD group, or <b>CHR20/CHR21</b> + LY group, respectively; (<b>C</b>) SNP inhibited the phosphorylations of Akt, eNOS and ERK1/2, respectively. And LY294002 and PD98059 reversed the effects of <b>CHR21</b>. The density of each lane was presented as mean ± standard deviation for three individual experiments. <b>*</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group, <b><sup>&amp;&amp;</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP group, <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR21</b> + SNP group. Blots were quantified using Image J software. Similar results were obtained with <b>CHR20</b> (data not shown).</p>
Full article ">Figure 8 Cont.
<p>Both the PI3K/Akt/eNOS pathway and the ERK1/2 pathways were involved in the neuroprotective effects of <b>CHR20/21</b>. Pre-incubation of LY294002, Akt VIII inhibitor, and PD98059 at a dose of 10 µM, respectively blocked the neuroprotective effects of <b>CHR20</b> (<b>A</b>) and <b>CHR21</b> (<b>B</b>) in RGC-5 cells. MTT activity was determined by MTT assay. Each experiment was performed in triplicate. <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP group; <b>**</b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR20/CHR21</b> + SNP group; <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>&amp;&amp;</sup> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR20/CHR21</b> + PD group, or <b>CHR20/CHR21</b> + LY group, respectively; (<b>C</b>) SNP inhibited the phosphorylations of Akt, eNOS and ERK1/2, respectively. And LY294002 and PD98059 reversed the effects of <b>CHR21</b>. The density of each lane was presented as mean ± standard deviation for three individual experiments. <b>*</b> <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> control group, <b><sup>&amp;&amp;</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> SNP group, <b><sup>##</sup></b> <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">vs.</span> <b>CHR21</b> + SNP group. Blots were quantified using Image J software. Similar results were obtained with <b>CHR20</b> (data not shown).</p>
Full article ">
882 KiB  
Article
Technological Implications of Modifying the Extent of Cell Wall-Proanthocyanidin Interactions Using Enzymes
by Ana Belén Bautista-Ortín, Rim Ben Abdallah, Liliana Del Rocío Castro-López, María Dolores Jiménez-Martínez and Encarna Gómez-Plaza
Int. J. Mol. Sci. 2016, 17(1), 123; https://doi.org/10.3390/ijms17010123 - 18 Jan 2016
Cited by 27 | Viewed by 5926
Abstract
The transference and reactivity of proanthocyanidins is an important issue that affects the technological processing of some fruits, such as grapes and apples. These processes are affected by proanthocyanidins bound to cell wall polysaccharides, which are present in high concentrations during the processing [...] Read more.
The transference and reactivity of proanthocyanidins is an important issue that affects the technological processing of some fruits, such as grapes and apples. These processes are affected by proanthocyanidins bound to cell wall polysaccharides, which are present in high concentrations during the processing of the fruits. Therefore, the effective extraction of proanthocyanidins from fruits to their juices or derived products will depend on the ability to manage these associations, and, in this respect, enzymes that degrade these polysaccharides could play an important role. The main objective of this work was to test the role of pure hydrolytic enzymes (polygalacturonase and cellulose) and a commercial enzyme containing these two activities on the extent of proanthocyanidin-cell wall interactions. The results showed that the modification promoted by enzymes reduced the amount of proanthocyanidins adsorbed to cell walls since they contributed to the degradation and release of the cell wall polysaccharides, which diffused into the model solution. Some of these released polysaccharides also presented some reactivity towards the proanthocyanidins present in a model solution. Full article
(This article belongs to the Special Issue Phenolics and Polyphenolics 2015)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of the size exclusion chromatogram of the original proanthocyanidin in solution and that for those remaining after (<b>A</b>) the first (CW + PA) or second experiment (2CW + PA) as well as those remaining after reacting with the polysaccharides found in the model solution (S + PA); (<b>B</b>) the first (W + PG + PA) or second experiment ((2CW + PG) + PA) when polygalacturonase was present, as well as those remaining after reacting with the polysaccharides found in the model solution (S(PG) + PA); (<b>C</b>) the first (CW + CEL + PA) or second experiment ((2CW + CEL) + PA) when cellulase was present, as well as those remaining after reacting with the polysaccharides found in the model solution (S(CEL) + PA); and (<b>D</b>) the first (CW + CE + PA) or second experiment ((2CW + CE) + PA) when a commercial enzyme was present, as well as those remaining after reacting with the polysaccharides found in the model solution (S(CE) + PA).</p>
Full article ">
2979 KiB  
Article
Metformin Restores Parkin-Mediated Mitophagy, Suppressed by Cytosolic p53
by Young Mi Song, Woo Kyung Lee, Yong-ho Lee, Eun Seok Kang, Bong-Soo Cha and Byung-Wan Lee
Int. J. Mol. Sci. 2016, 17(1), 122; https://doi.org/10.3390/ijms17010122 - 16 Jan 2016
Cited by 71 | Viewed by 10227
Abstract
Metformin is known to alleviate hepatosteatosis by inducing 5’ adenosine monophosphate (AMP)-kinase-independent, sirtuin 1 (SIRT1)-mediated autophagy. Dysfunctional mitophagy in response to glucolipotoxicities might play an important role in hepatosteatosis. Here, we investigated the mechanism by which metformin induces mitophagy through restoration of the [...] Read more.
Metformin is known to alleviate hepatosteatosis by inducing 5’ adenosine monophosphate (AMP)-kinase-independent, sirtuin 1 (SIRT1)-mediated autophagy. Dysfunctional mitophagy in response to glucolipotoxicities might play an important role in hepatosteatosis. Here, we investigated the mechanism by which metformin induces mitophagy through restoration of the suppressed Parkin-mediated mitophagy. To this end, our ob/ob mice were divided into three groups: (1) ad libitum feeding of a standard chow diet; (2) intraperitoneal injections of metformin 300 mg/kg; and (3) 3 g/day caloric restriction (CR). HepG2 cells were treated with palmitate (PA) plus high glucose in the absence or presence of metformin. We detected enhanced mitophagy in ob/ob mice treated with metformin or CR, whereas mitochondrial spheroids were observed in mice fed ad libitum. Metabolically stressed ob/ob mice and PA-treated HepG2 cells showed an increase in expression of endoplasmic reticulum (ER) stress markers and cytosolic p53. Cytosolic p53 inhibited mitophagy by disturbing the mitochondrial translocation of Parkin, as demonstrated by immunoprecipitation. However, metformin decreased ER stress and p53 expression, resulting in induction of Parkin-mediated mitophagy. Furthermore, pifithrin-α, a specific inhibitor of p53, increased mitochondrial incorporation into autophagosomes. Taken together, these results indicate that metformin treatment facilitates Parkin-mediated mitophagy rather than mitochondrial spheroid formation by decreasing the inhibitory interaction with cytosolic p53 and increasing degradation of mitofusins. Full article
(This article belongs to the Special Issue Modulators of Endoplasmic Reticulum Stress)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Metformin induced mitophagy. Electron microscopy was performed on hepatocytes from <span class="html-italic">ob</span>/<span class="html-italic">ob</span> mice under <span class="html-italic">ad libitum</span> feeding of chow diet (<b>A</b>-<b>a</b>), caloric restriction (<b>A</b>-<b>b</b>), and <span class="html-italic">ad libitum</span> feeding of chow diet with 300 mg/kg metformin treatment (<b>A</b>-<b>c</b>). In the <span class="html-italic">ad libitum</span> group, spheroid mitochondria were readily detected. However, in the metformin-treated group, autophagic double membranes were detected, indicating mitophagy; (<b>B</b>) co-localization between a marker of autophagosomes (LC3) and a mitochondrial marker (MitoTracker Red) was assessed in HepG2 cells stably transfected with green fluorescent protein (GFP)-LC3. Although there was little co-localization between mitochondria and autophagosomes in the 0.25 mM palmitate-treated HepG2 cells, LC3-labeled structures were seen surrounding the fragmented mitochondria in cells treated with 0.5 mM metformin and 0.25 mM palmitate. Scale bar = 20 μm.</p>
Full article ">Figure 2
<p>Metformin restored the Parkin-mediated mitophagy inhibited by glucolipotoxicity-induced cytosolic p53. Western blotting analysis was performed to assess cytosolic p53 levels with the cytosolic fractions isolated from palmitate- (<b>A</b>-<b>a</b>) or metformin- (<b>A</b>-<b>b</b>) and 35 mM glucose-treated HepG2 cells; the expression of cytosolic p53 protein was assessed following 0.5 mM metformin and/or 0.25 mM palmitate with 35 mM glucose treatment of HepG2 cells (<b>B</b>-<b>a</b>) or in hepatocytes from <span class="html-italic">ob</span>/<span class="html-italic">ob</span> mice with <span class="html-italic">ad libitum</span> feeding of chow diet, caloric restriction, and <span class="html-italic">ad libitum</span> feeding of chow diet with 300 mg/kg metformin treatment (<b>B</b>-<b>b</b>); (<b>C</b>) Expression of Parkin (green) on mitochondria (MitoTracker Red) was observed with a confocal microscopy. The bottom panels show enlarged views of the boxed areas. The yellow dots indicate the mitochondria that co-localize with Parkin. The cytosolic and mitochondrial fractionation experiments demonstrated Parkin translocation from the cytosol to mitochondria depending on the conditions used in the HepG2 cells; a co-immunoprecipitation assay was conducted to investigate the interaction between p53 and Parkin, depending on the conditions used in the HepG2 cells (<b>D</b>-<b>a</b>) or in hepatocytes from mice (<b>D</b>-<b>b</b>). GAPDH or COX IV was used for normalization. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively). Scale bar: white = 20 μm, yellow = 10 μm. Asterisks (* <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) indicate significant differences. Abbreviation: Mfn1, mitofusin1.</p>
Full article ">Figure 2 Cont.
<p>Metformin restored the Parkin-mediated mitophagy inhibited by glucolipotoxicity-induced cytosolic p53. Western blotting analysis was performed to assess cytosolic p53 levels with the cytosolic fractions isolated from palmitate- (<b>A</b>-<b>a</b>) or metformin- (<b>A</b>-<b>b</b>) and 35 mM glucose-treated HepG2 cells; the expression of cytosolic p53 protein was assessed following 0.5 mM metformin and/or 0.25 mM palmitate with 35 mM glucose treatment of HepG2 cells (<b>B</b>-<b>a</b>) or in hepatocytes from <span class="html-italic">ob</span>/<span class="html-italic">ob</span> mice with <span class="html-italic">ad libitum</span> feeding of chow diet, caloric restriction, and <span class="html-italic">ad libitum</span> feeding of chow diet with 300 mg/kg metformin treatment (<b>B</b>-<b>b</b>); (<b>C</b>) Expression of Parkin (green) on mitochondria (MitoTracker Red) was observed with a confocal microscopy. The bottom panels show enlarged views of the boxed areas. The yellow dots indicate the mitochondria that co-localize with Parkin. The cytosolic and mitochondrial fractionation experiments demonstrated Parkin translocation from the cytosol to mitochondria depending on the conditions used in the HepG2 cells; a co-immunoprecipitation assay was conducted to investigate the interaction between p53 and Parkin, depending on the conditions used in the HepG2 cells (<b>D</b>-<b>a</b>) or in hepatocytes from mice (<b>D</b>-<b>b</b>). GAPDH or COX IV was used for normalization. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively). Scale bar: white = 20 μm, yellow = 10 μm. Asterisks (* <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) indicate significant differences. Abbreviation: Mfn1, mitofusin1.</p>
Full article ">Figure 3
<p>Metformin restored Parkin-mediated mitophagy inhibited by glucolipotoxicity-induced ER stress. The expression of ER stress markers was assessed following 0.5 mM metformin and/or 0.25 mM palmitate with 35 mM glucose treatment of HepG2 cells (<b>A</b>-<b>a</b>) or in hepatocytes from <span class="html-italic">ob</span>/<span class="html-italic">ob</span> mice with <span class="html-italic">ad libitum</span> feeding of chow diet, caloric restriction, or <span class="html-italic">ad libitum</span> feeding of chow diet with 300 mg/kg metformin treatment (<b>A</b>-<b>b</b>) using Western blotting analysis. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively); (<b>B</b>) the expression of ER stress markers and p53 protein was assessed in response to thapsigargin in HepG2 cells; (<b>C</b>-<b>a</b>) cell viability was assessed in HepG2 cells exposed to 0.1 μM thapsigargin and/or 0.5 mM metformin using CCK-8. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 5 independent experiments); (<b>C</b>-<b>b</b>) the expression of an ER stress marker was also examined under these conditions; (<b>C</b>-<b>c</b>) the cytosolic and mitochondrial fractionation experiments demonstrated Parkin translocation from the cytosol to mitochondria, depending on the conditions used. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively). GAPDH or COX IV was used for normalization. Asterisks (* <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) indicate significant differences. Abbreviation: Mfn1, mitofusin1; CCK-8, cell counting kit-8.</p>
Full article ">Figure 3 Cont.
<p>Metformin restored Parkin-mediated mitophagy inhibited by glucolipotoxicity-induced ER stress. The expression of ER stress markers was assessed following 0.5 mM metformin and/or 0.25 mM palmitate with 35 mM glucose treatment of HepG2 cells (<b>A</b>-<b>a</b>) or in hepatocytes from <span class="html-italic">ob</span>/<span class="html-italic">ob</span> mice with <span class="html-italic">ad libitum</span> feeding of chow diet, caloric restriction, or <span class="html-italic">ad libitum</span> feeding of chow diet with 300 mg/kg metformin treatment (<b>A</b>-<b>b</b>) using Western blotting analysis. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively); (<b>B</b>) the expression of ER stress markers and p53 protein was assessed in response to thapsigargin in HepG2 cells; (<b>C</b>-<b>a</b>) cell viability was assessed in HepG2 cells exposed to 0.1 μM thapsigargin and/or 0.5 mM metformin using CCK-8. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 5 independent experiments); (<b>C</b>-<b>b</b>) the expression of an ER stress marker was also examined under these conditions; (<b>C</b>-<b>c</b>) the cytosolic and mitochondrial fractionation experiments demonstrated Parkin translocation from the cytosol to mitochondria, depending on the conditions used. Values displayed are mean ± SEM (<span class="html-italic">n</span> = 3 independent experiments, respectively). GAPDH or COX IV was used for normalization. Asterisks (* <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) indicate significant differences. Abbreviation: Mfn1, mitofusin1; CCK-8, cell counting kit-8.</p>
Full article ">Figure 4
<p>A p53 inhibitor restored the Parkin-mediated mitophagy inhibited by glucolipotoxicity. (<b>A</b>-<b>a</b>) the expression of cytosolic p53 was assessed in HepG2 cells treated with 50 μM PFT-α, a p53 inhibitor, and/or 0.25 mM palmitate with 35 mM glucose using Western blotting analysis; (<b>A</b>-<b>b</b>) the cytosolic and mitochondrial fractionation experiments demonstrated Parkin translocation from the cytosol to mitochondria, depending on the conditions used. Exposure time of Western blotting was long, compared to <a href="#ijms-17-00122-f002" class="html-fig">Figure 2</a>C; (<b>B</b>) co-localization was assessed between a marker of autophagosomes (LC3) and a mitochondrial marker (MitoTracker Red) in HepG2 cells stably transfected with GFP-LC3, depending on the conditions used. LC3-labeled structures were observed surrounding the fragmented mitochondria in cells treated with PFT-α and palmitate, although there was little co-localization between mitochondria and autophagosomes in cells treated with 0.25 mM palmitate alone. Scale bar = 20 μm. <span class="html-italic">Abbreviation</span>: Mfn1, mitofusin1; PFT-α, pifithrin-α.</p>
Full article ">Figure 5
<p>Summary of the working thesis of this study. See the text for details. Black and red arrows indicate the effects of glucolipotoxicity and the drugs (metformin or PFTα), respectively.</p>
Full article ">
6977 KiB  
Article
Alternative Splicing in Adhesion- and Motility-Related Genes in Breast Cancer
by Rosanna Aversa, Anna Sorrentino, Roberta Esposito, Maria Rosaria Ambrosio, Angela Amato, Alberto Zambelli, Alfredo Ciccodicola, Luciana D’Apice and Valerio Costa
Int. J. Mol. Sci. 2016, 17(1), 121; https://doi.org/10.3390/ijms17010121 - 16 Jan 2016
Cited by 16 | Viewed by 6470
Abstract
Breast cancer is the most common tumor and the second leading cause of cancer death among woman, mainly caused by the metastatic spread. Tumor invasiveness is due to an altered expression of adhesion molecules. Among them, semaphorins are of peculiar interest. Cancer cells [...] Read more.
Breast cancer is the most common tumor and the second leading cause of cancer death among woman, mainly caused by the metastatic spread. Tumor invasiveness is due to an altered expression of adhesion molecules. Among them, semaphorins are of peculiar interest. Cancer cells can manipulate alternative splicing patterns to modulate the expression of adhesion- and motility-related molecules, also at the isoform level. In this study, combining RNA-Sequencing on MCF-7 to targeted experimental validations—in human breast cell lines and breast tumor biopsies—we identified 12 new alternative splicing transcripts in genes encoding adhesion- and motility-related molecules, including semaphorins, their receptors and co-receptors. Among them, a new SEMA3F transcript is expressed in all breast cell lines and breast cancer biopsies, and is translated into a new semaphorin 3F isoform. In silico analysis predicted that most of the new putative proteins lack functional domains, potentially missing some functions and acquiring new ones. Our findings better describe the extent of alternative splicing in breast cancer and highlight the need to further investigate adhesion- and motility-related molecules to gain insights into breast cancer progression. Full article
(This article belongs to the Special Issue Pre-mRNA Splicing 2015)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Overview of RNA-Seq data processing and of the computational and experimental approach used for the identification and validation of the new spliced transcripts; (<b>B</b>) Barplot with log<sub>2</sub> normalized expression values (FPKM) of semaphorin/Plexin/Neuropilin—and of the adhesion markers <span class="html-italic">CDH1</span> and <span class="html-italic">CD44</span>—according to RNA-Seq data.</p>
Full article ">Figure 2
<p>Schematic representation of <span class="html-italic">de novo</span> identified transcripts for adhesion-related genes and their differential expression by RT-PCR in breast cell lines (a representative image is shown for each gene). (<b>A</b>) <span class="html-italic">THBS1</span> new transcript; (<b>B</b>) <span class="html-italic">RAP1GAP</span> new transcript; (<b>C</b>) <span class="html-italic">LGALS1</span> new transcript; (<b>D</b>) <span class="html-italic">CD47</span> new transcript. For each new transcript, in the upper part is shown the genomic region encompassing the gene; red arrows indicate the direction of transcription. Below, the exon/intron structure of the gene is depicted. Red cross indicates the exons that are spliced out in the new transcript. The electropherogram and the nucleotide sequence above refer to the new splice junction identified. Images of PCR amplicons on agarose gel from three independent replicates experiments were acquired. For all panels, M is 100 bp DNA marker. The black arrows indicate the PCR band corresponding to the newly identified transcripts. Pixel density of each band (only for the newly identified mRNA isoforms) was quantified by ImageJ software and compared to the pixel density of the 100 bp DNA marker (at known concentration) to normalize data across replicates and cell lines. GAPDH gel bands are reported below to show that the starting amount of template was the same for all analyzed samples. Values are reported as AU = Arbitrary Units in the boxplots. Dark horizontal lines represent the mean, the box represents the 25th and 75th percentiles and the whiskers the 5th and 95th percentiles. Asterisks indicate significant <span class="html-italic">p</span> values (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Schematic representation of <span class="html-italic">de novo</span> identified transcripts of motility-related genes and their expression by RT-PCR in breast cell lines (a representative image is shown for each gene). (<b>A</b>) <span class="html-italic">RHOA</span>; (<b>B</b>) <span class="html-italic">RHOD</span>; (<b>C</b>) <span class="html-italic">CASK</span>; (<b>D</b>) <span class="html-italic">JAG2</span>; (<b>E</b>) <span class="html-italic">CTTN</span> new transcripts. Details about the gel bands and their quantification are provided in <a href="#ijms-17-00121-f002" class="html-fig">Figure 2</a> legend. Asterisks indicate significant <span class="html-italic">p</span> values (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Schematic representation of <span class="html-italic">de novo</span> identified transcripts for (<b>A</b>) <span class="html-italic">SEMA3C</span> (<b>B</b>) <span class="html-italic">PLXNB1</span> genes. Nucleotide sequenceand their expression by RT-PCR in breast cell lines are also depicted. Details about the gel bands and their quantification are provided in <a href="#ijms-17-00121-f002" class="html-fig">Figure 2</a>’s legend. Asterisks indicate significant <span class="html-italic">p</span> values (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>(<b>A</b>) Scheme of the new <span class="html-italic">SEMA3F</span> transcript and expression by RT-PCR on MCF-7 cell line, showing the canonical (307 bp) and the new alternative transcript (149 bp) of <span class="html-italic">SEMA3F</span>; M is 100 bp DNA marker; (<b>B</b>) Boxplots with the relative expression of both annotated and new <span class="html-italic">SEMA3F</span> transcript in breast cell lines measured with qRT-PCR (<span class="html-italic">n</span> = 3); (<b>C</b>) On the <b>left</b>, immunoreactive bands—corresponding to the canonical (about 90 kDa) and the new isoform (about 70 kDa) of semaphorin 3F—detected by Western Blot on whole breast cell lines lysates. On the <b>right</b>, the densitometry results (AU = Arbitrary Units) by ImageJ of the Western Blot bands, after normalization with actin, are shown as bar graphs (<span class="html-italic">n</span> = 2); (<b>D</b>) A detail of the protein alignment of the canonical and the new semaphorin 3F protein isoform is shown. The dashed boxes represent the functional domains that are deleted in the novel protein sequence. Asterisks indicate significant <span class="html-italic">p</span> values (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Semaphorin 3F expression in BC tumor biopsies (<span class="html-italic">n</span> = 18). (<b>A</b>,<b>B</b>) qRT-PCR analysis of the new (LN626688) and the annotated (NM_004186.3) <span class="html-italic">SEMA3F</span> transcripts, respectively. Samples have been ordered according to tumor grade (G2, G3); # indicates Triple Negative biopsies; asterisks (*) indicate the sample with the lowest absolute expression, used as reference; (<b>C</b>) Relative ratio of Δ<span class="html-italic">C</span><sub>t</sub> values between skipping and canonical mRNA isoforms of <span class="html-italic">SEMA3F</span> across breast cancer subtypes. TNM: pT1 (<span class="html-italic">n</span> = 7); pT2 (<span class="html-italic">n</span> = 9); pT3 (<span class="html-italic">n</span> = 2). Tumor grade: G2 (<span class="html-italic">n</span> = 11); G3 (<span class="html-italic">n</span> = 7). ER<sup>+</sup>/PR<sup>+</sup> (<span class="html-italic">n</span> = 15); TN (triple negative samples) (<span class="html-italic">n</span> = 3); (<b>D</b>) Six representative images by immunohistochemistry assay on FFPE slices from breast cancer specimens are shown; N-terminal anti-Sema3F (that recognizes both the canonical and the new splicing variant) primary antibody was used. Scale Bar (in red) value = 33.76 µm.</p>
Full article ">
5840 KiB  
Article
Platelet-Rich Plasma Increases the Levels of Catabolic Molecules and Cellular Dedifferentiation in the Meniscus of a Rabbit Model
by Hye-Rim Lee, Oog-Jin Shon, Se-Il Park, Han-Jun Kim, Sukyoung Kim, Myun-Whan Ahn and Sun Hee Do
Int. J. Mol. Sci. 2016, 17(1), 120; https://doi.org/10.3390/ijms17010120 - 16 Jan 2016
Cited by 33 | Viewed by 7236
Abstract
Despite the susceptibility to frequent intrinsic and extrinsic injuries, especially in the inner zone, the meniscus does not heal spontaneously owing to its poor vascularity. In this study, the effect of platelet-rich plasma (PRP), containing various growth factors, on meniscal mechanisms was examined [...] Read more.
Despite the susceptibility to frequent intrinsic and extrinsic injuries, especially in the inner zone, the meniscus does not heal spontaneously owing to its poor vascularity. In this study, the effect of platelet-rich plasma (PRP), containing various growth factors, on meniscal mechanisms was examined under normal and post-traumatic inflammatory conditions. Isolated primary meniscal cells of New Zealand white (NZW) rabbits were incubated for 3, 10, 14 and 21 days with PRP(−), 10% PRP (PRP(+)), IL(+) or IL(+)PRP(+). The meniscal cells were collected and examined using reverse-transcription polymerase chain reaction (RT-PCR). Culture media were examined by immunoblot analyses for matrix metalloproteinases (MMP) catabolic molecules. PRP containing growth factors improved the cellular viability of meniscal cells in a concentration-dependent manner at Days 1, 4 and 7. However, based on RT-PCR, meniscal cells demonstrated dedifferentiation, along with an increase in type I collagen in the PRP(+) and in IL(+)PRP(+). In PRP(+), the aggrecan expression levels were lower than in the PRP(−) until Day 21. The protein levels of MMP-1 and MMP-3 were higher in each PRP group, i.e., PRP(+) and IL(+)PRP(+), at each culture time. A reproducible 2-mm circular defect on the meniscus of NZW rabbit was used to implant fibrin glue (control) or PRP in vivo. After eight weeks, the lesions in the control and PRP groups were occupied with fibrous tissue, but not with meniscal cells. This study shows that PRP treatment of the meniscus results in an increase of catabolic molecules, especially those related to IL-1α-induced inflammation, and that PRP treatment for an in vivo meniscus injury accelerates fibrosis, instead of meniscal cartilage. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Assessment of individual growth factors in platelet-rich plasma (PRP) by semi-quantitative RT-PCR. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> individual Y rabbits. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> individual O rabbits; (<b>B</b>) Evaluation of individual growth factors containing PRP from young rabbits (Y) and old rabbits (O) by ELISA. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> individual Y rabbits. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> individual O rabbits; (<b>C</b>) Cellular viability after PRP treatments. Each value is expressed as the mean ± standard error. The <span class="html-italic">p</span>-value is approximate (from a chi-square distribution). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> the untreated cells. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> the 0.1% PRP-treated cells.</p>
Full article ">Figure 2
<p>(<b>A</b>) mRNA expression in meniscal cells with/without PRP. Aggrecan mRNA expression decreased in a time-dependent manner in PRP(+). Type II collagen gene expression decreased in PRP(+) compared to PRP(−) after Day 10. In initial examinations, MMP-3, MMP-9, and MMP-13 displayed increased expression in the PRP(+); (<b>B</b>) Protein expression analysis in culture media of meniscal cells with/without PRP. PRP(+) showed a relative upregulation in the protein levels of MMP-1, MMP-3, and TIMP-1 compared to PRP(−). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> PRP(−).</p>
Full article ">Figure 2 Cont.
<p>(<b>A</b>) mRNA expression in meniscal cells with/without PRP. Aggrecan mRNA expression decreased in a time-dependent manner in PRP(+). Type II collagen gene expression decreased in PRP(+) compared to PRP(−) after Day 10. In initial examinations, MMP-3, MMP-9, and MMP-13 displayed increased expression in the PRP(+); (<b>B</b>) Protein expression analysis in culture media of meniscal cells with/without PRP. PRP(+) showed a relative upregulation in the protein levels of MMP-1, MMP-3, and TIMP-1 compared to PRP(−). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> PRP(−).</p>
Full article ">Figure 3
<p>(<b>A</b>) mRNA expression in IL-1α-treated meniscal cells with/without PRP. mRNA expression of aggrecan and type II collagen was generally low compared to that observed in IL(+). The aggrecan and type II collagen expression levels were lower in IL(+)PRP(+) than IL(+); (<b>B</b>) Immunoblot analysis in IL-1α-treated culture media of meniscal cells with/without PRP. At every time point, protein expression levels of MMP-1, MMP-3, and TIMP-1 were higher in IL(+)PRP(+) than in IL(+), and there was no significant difference in their time courses. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> IL(+).</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>) mRNA expression in IL-1α-treated meniscal cells with/without PRP. mRNA expression of aggrecan and type II collagen was generally low compared to that observed in IL(+). The aggrecan and type II collagen expression levels were lower in IL(+)PRP(+) than IL(+); (<b>B</b>) Immunoblot analysis in IL-1α-treated culture media of meniscal cells with/without PRP. At every time point, protein expression levels of MMP-1, MMP-3, and TIMP-1 were higher in IL(+)PRP(+) than in IL(+), and there was no significant difference in their time courses. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 <span class="html-italic">versus</span> IL(+).</p>
Full article ">Figure 4
<p>(<b>A</b>,<b>B</b>) Time sequential mRNA expression of each growth factor from meniscal cells with and without either PRP or IL-1α. The changes in the mRNA levels of TGF-β, VEGF, PDGF-D, and FGF-2 showed similar patterns between PRP(−) and PRP(+), as well as between IL-treated groups; (<b>C</b>,<b>D</b>) Time sequential protein expression of each growth factor from culture media of meniscal cells with and without either PRP or IL-1α. Protein levels of growth factors were higher in PRP(+) and IL(+)PRP(+)than in the other groups. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3.</p>
Full article ">Figure 4 Cont.
<p>(<b>A</b>,<b>B</b>) Time sequential mRNA expression of each growth factor from meniscal cells with and without either PRP or IL-1α. The changes in the mRNA levels of TGF-β, VEGF, PDGF-D, and FGF-2 showed similar patterns between PRP(−) and PRP(+), as well as between IL-treated groups; (<b>C</b>,<b>D</b>) Time sequential protein expression of each growth factor from culture media of meniscal cells with and without either PRP or IL-1α. Protein levels of growth factors were higher in PRP(+) and IL(+)PRP(+)than in the other groups. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3.</p>
Full article ">Figure 5
<p>Effects of PRP on apoptosis in meniscal cells. The PRP(+) had no change in the ratio of Bax to Bcl2 mRNA in all examined periods, but the value of Bax/Bcl2 increased significantly in a time-dependent manner in IL(+)PRP(+). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 <span class="html-italic">versus</span> Day 3.</p>
Full article ">Figure 6
<p>Histopathological analysis of full-thickness defects of the meniscus. The defective lesions of control and PRP-treated groups were completely replaced by fibrous tissue, instead of meniscal cartilage, at four and eight weeks. PRP-treated lesions were relatively thickened with hypercellularity of fibroblasts when compared to the control (black full line arrow: repair site, black dotted line arrow: non-defective site) H &amp; E, magnifications 40× and 100×. <b>A</b> 4wk (40×)-Control; <b>B</b> 4wk (100×)-Control; <b>C</b> 4wk (40×)-PRP; <b>D</b> 4wk (100×)-PRP; <b>E</b> 8wk (40×)-Control; <b>F</b> 8wk (100×)-Control; <b>G</b> 8wk (40×)-PRP; <b>H</b> 8wk (100×)-PRP.</p>
Full article ">
3003 KiB  
Article
Cloning and Transcriptional Activity of the Mouse Omi/HtrA2 Gene Promoter
by Dan Liu, Xin Liu, Ye Wu, Wen Wang, Xinliang Ma and Huirong Liu
Int. J. Mol. Sci. 2016, 17(1), 119; https://doi.org/10.3390/ijms17010119 - 16 Jan 2016
Cited by 10 | Viewed by 6074
Abstract
HtrA serine peptidase 2 (HtrA2), also named Omi, is a pro-apoptotic protein that exhibits dramatic changes in expression levels in a variety of disorders, including ischemia/reperfusion injury, cancer, and neurodegeneration. In our study, Omi/HtrA2 protein levels were high in the heart, brain, kidney [...] Read more.
HtrA serine peptidase 2 (HtrA2), also named Omi, is a pro-apoptotic protein that exhibits dramatic changes in expression levels in a variety of disorders, including ischemia/reperfusion injury, cancer, and neurodegeneration. In our study, Omi/HtrA2 protein levels were high in the heart, brain, kidney and liver, with elevated heart/brain expression in aging mice. A similar expression pattern was observed at the mRNA level, which suggests that the regulation of Omi/HtrA2 is predominately transcriptional. Promoter binding by transcription factors is the main influencing factor of transcription, and to identify specific promoter elements that contribute to the differential expression of mouse Omi/HtrA2, we constructed truncated Omi/HtrA2 promoter/luciferase reporter vectors and analyzed their relative luciferase activity; it was greatest in the promoter regions at −1205~−838 bp and −146~+93 bp, with the −838~−649 bp region exhibiting negative regulatory activity. Bioinformatics analysis suggested that the Omi/HtrA2 gene promoter contains a CpG island at −709~+37 bp, and eight heat shock transcription factor 1 (HSF1) sites, two Sp1 transcription factor (SP1)sites, one activator protein (AP) site, seven p53 sites, and four YY1 transcription factor(YY1) sites were predicted in the core areas. Furthermore, we found that p53 and HSF1 specifically binds to the Omi/HtrA2 promoter using chromatin immunoprecipitation analysis. These results provide a foundation for understanding Omi/HtrA2 regulatory mechanisms, which could further understanding of HtrA-associated diseases. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Tissue specific expression of HtrA serine peptidase 2 (Omi/HtrA2). Omi/HtrA2 protein expression in young and aging mice was detected by Western blotting (<b>A</b>), and the relative levels were quantified using ImageJ software (<b>B</b>). Omi/HtrA2 mRNA expression was detected by quantitative RT-PCR (<b>C</b>). <span class="html-italic">n</span> = 5 per group. * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> young.</p>
Full article ">Figure 2
<p>Verification of mouse Omi/HtrA2 promoter luciferase expression plasmids. Plasmids were verified by double enzyme digestion with <span class="html-italic">Kpn</span>I and <span class="html-italic">Hind</span>III, which releases a vector fragment and a variable-sized promoter insert. M: DL2000 Marker; <b>1</b>: pGL3-239; <b>2</b>: pGL3-472; <b>3</b>: pGL3-742; <b>4</b>: pGL3-931; <b>5</b>: pGL3-1298. Luc: luciferase expression plasmids.</p>
Full article ">Figure 3
<p>Relative luciferase activity of full length and truncated mouse Omi/HtrA2 promoter plasmids. Luciferase assays were performed in NIH3T3 cells (<b>A</b>); H9c2 cells (<b>B</b>); and HEK-293 cells (<b>C</b>). The different letters indicate the significant differences between the plasmids (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Chromatin Immunoprecipitation(ChIP) was performed with chromatin prepared from mouse heart. The promoter region containing the Heat shock transcription factor 1 (HSF1) and p53 site was analysed by PCR following immunoprecipitation with the HSF1 and p53 antibodies. Results of amplification for soluble chromatin before immunoprecipitation are shown as positive control (input) and negative control (IgG).</p>
Full article ">
4177 KiB  
Article
Ultra-Fast Glyco-Coating of Non-Biological Surfaces
by Eleanor Williams, Katie Barr, Elena Korchagina, Alexander Tuzikov, Stephen Henry and Nicolai Bovin
Int. J. Mol. Sci. 2016, 17(1), 118; https://doi.org/10.3390/ijms17010118 - 16 Jan 2016
Cited by 13 | Viewed by 6480
Abstract
The ability to glycosylate surfaces has medical and diagnostic applications, but there is no technology currently recognized as being able to coat any surface without the need for prior chemical modification of the surface. Recently, a family of constructs called function-spacer-lipids (FSL) has [...] Read more.
The ability to glycosylate surfaces has medical and diagnostic applications, but there is no technology currently recognized as being able to coat any surface without the need for prior chemical modification of the surface. Recently, a family of constructs called function-spacer-lipids (FSL) has been used to glycosylate cells. Because it is known that lipid-based material can adsorb onto surfaces, we explored the potential and performance of cell-labelling FSL constructs to “glycosylate” non-biological surfaces. Using blood group A antigen as an indicator, the performance of a several variations of FSL constructs to modify a large variety of non-biological surfaces was evaluated. It was found the FSL constructs when optimised could in a few seconds glycosylate almost any non-biological surface including metals, glass, plastics, rubbers and other polymers. Although the FSL glycan coating was non-covalent, and therefore temporary, it was sufficiently robust with appropriate selection of spacer and surface that it could capture anti-glycan antibodies, immobilize cells (via antibody), and withstand incubation in serum and extensive buffer washing, making it suitable for diagnostic and research applications. Full article
(This article belongs to the Special Issue Glycosylation and Glycoproteins)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagrams of the two primary blood group A function-spacer-lipid (FSL) constructs used in this paper. The upper schematic Atri-Ad-DOPE shows an FSL with a trisaccharide generic blood group A antigen and a short 2 nm adipate spacer while the lower schematic Atetra-CMG-DOPE shows a type 2 chain specific A tetrasaccharide FSL with a longer 7 nm carboxymethylglycine spacer. Both constructs have the same dioleoylphosphatidylethanolamine (DOPE) lipid tail.</p>
Full article ">Figure 2
<p>Photographic images of 20 representative coupons (see also <a href="#ijms-17-00118-t001" class="html-table">Table 1</a>) each spotted with 1 µL of Atri-Ad-DOPE (12 o’clock spot) and Atetra-CMG-DOPE (6 o’clock spot). Schematic diagram shows experimental layout. FSL spotted coupons were allowed to air dry then visualised by EIA using monoclonal anti-A and precipitating chromogenic substrate (purple).</p>
Full article ">Figure 3
<p>An example of microplate prepared with blood group <b>A</b> and <b>B</b> FSL constructs printed on paper (MG paper—cellulose esters) and prepared as a microplate to determine specificity of monoclonal reagents. Reprint from [<a href="#B8-ijms-17-00118" class="html-bibr">8</a>]. Copyright 2014 with permission from John Wiley and Sons. Alphanumeric characters appear when developed in the EIA reaction wherever the monoclonal reagent has bound to the FSL. Area outside of the printed area is the “internal” negative control and/or blank. In this example antibodies in co-ordinates <b>A1</b>, <b>A4</b>, <b>B3</b> have anti-A specificity, those in <b>A2</b>, <b>B1</b> and <b>B4</b> have anti-B specificity while those in <b>A3</b> and <b>B2</b> anti-A + B specificity. This EIA technique either with printed alphanumeric characters or spots of FSL constructs on a variety of different surfaces and optionally prepared as microplates was the basis for most experiments.</p>
Full article ">Figure 4
<p>Examples of 21 different surfaces reacted with Atri-Ad-DOPE. FSL constructs on most surfaces were inkjet printed as words identifying each surface material. The microspheres and glass fibre were coated with FSL construct by immersion, while a paintbrush was used to apply the FSL construct to gold foil. The blood group A FSL constructs bound to the various surfaces were visualised by EIA using monoclonal anti-A and a precipitating chromogenic substrate.</p>
Full article ">Figure 5
<p>Capture of red blood cells onto printed glyco-FSLs. Six different ABO related FSL constructs based on adipate and CMG spacers were printed onto five different surfaces. The upper row of images show the presence of these FSL constructs on surfaces using EIA and monoclonal antibodies directed against both the A and B antigens. The middle row of images show blood type specific binding of these same surfaces when reacted with monoclonal IgM anti-A and used to capture blood group A red cells. Of particular note is the poor EIA and strong red cell binding reaction of the silver membrane, highlighting the inability of some surfaces to show strong reactivity by the EIA method. The lower row of images are SEM 2000× magnifications of the edge of the printed areas, showing delineation between the printed blood group A FSL + IgM anti-A capture of blood group A red cells area and unprinted areas. It can be seen by its ability to immobilise RBCs that the monoclonal anti-A used had higher cross reactive affinity for the Forssman disaccharide (2FSad) than for the pentasaccharide (5FSad). This is expected as the Forssman antigen is a cross-reactive target of the anti-A reagent used, and the pentasccahride with a more complete Forssman structure will have less off-target binding [<a href="#B9-ijms-17-00118" class="html-bibr">9</a>].</p>
Full article ">Figure 6
<p>Attachment of blood group A red cells onto 20 µm polycarbonate microspheres coated with Atri-Ad-DOPE and IgM anti-A. Inset (<b>a</b>) shows light microscopy image at 1000× magnification while insets (<b>b</b>) &amp; (<b>c</b>) show higher magnifications under SEM (post glutaraldehyde fixation). Although the main image and inset (<b>a</b>) appears to show cells only on the perimeter of the microsphere, this is an artefact due to the microscopic plane of focus. Varying the plane of focus reveals red cells evenly distributed over the microspheres. No binding occurred on controls.</p>
Full article ">Figure 7
<p>Incorporation of Atri-Ad-DOPE into nanofibres during the electrospinning process. Cellulose acetate (CA) and polyvinyl butyral (PVB) nanofibres were electrospun with Atri-Ad-DOPE and Btri-Ad-DOPE constructs included in the pre-spinning liquid polymers. Spun nanofibres were twisted into cords and then immunostained. Staining with monoclonal anti-A by the EIA method was able to demonstrate incorporation of the blood group A FSL construct into both nanofibres during the electrospinning process.</p>
Full article ">Figure 8
<p>Effect of molar concentration on the surface binding characteristics of Atri-Ad-DOPE and Atetra-CMG-DOPE when applied as a 1 µL spots on MG paper.</p>
Full article ">Figure 9
<p>Effect of application buffer pH on surface binding characteristics of Atri-Ad-DOPE and Atetra-CMG-DOPE when applied as a 1 µL spots (50 µmol/L) on paper. Columns show the time of contact in seconds of the FSL with the surface before being washed with PBS. The last clearly positive reaction (defined as even reactivity over the spot area) in each column is indicated by a yellow circle.</p>
Full article ">Figure 10
<p>Effect of application buffer ionic concentration on the surface binding characteristics of Atri-Ad-DOPE and Atetra-CMG-DOPE when applied as a 1 µL spots (50 µmol/L) on paper. Columns show the time of contact in seconds of the FSL with the surface before being washed away with PBS. The last clearly positive reaction in each column is indicated by a yellow circle.</p>
Full article ">Figure 11
<p>Effect of FSL application buffer (PBS, surfactant (Tw20), 70% ethanol) on the binding characteristics of Atri-Ad-DOPE and Atetra-CMG-DOPE when applied as a 1 µL spots (50 µmol/L) on paper, stainless steel and polyester film. Columns show the time of contact in seconds of the FSL with the surface before being washed away with PBS. The last clearly positive reaction in each column is indicated by a yellow circle.</p>
Full article ">Figure 12
<p>Effect of contact with human serum on the binding characteristics of Atri-Ad-DOPE and Atetra-CMG-DOPE when applied as a 1 µL spots (50 µmol/L) on paper, and exposed to serum for up to 24 h. After exposure to serum the presence of blood group A FSL constructs was determined by EIA with monoclonal anti-A. Serum concentration values relate to 100% being undiluted serum and 20% is serum diluted 1:5.</p>
Full article ">Figure 13
<p>Schematic diagrams of two additional blood group A function-spacer-lipid (FSL) constructs used in this paper. The upper schematic Atetra-Ad-DOPE shows an FSL with a tetrasaccharide blood group type 2 A antigen with an adipate spacer while the lower schematic is of an A trisaccharide with a cholesterol lipid tail (Atri-Chol).</p>
Full article ">Figure 14
<p>Elution profiles of four blood group A FSL variants printed on paper and exposed to 50%, 70% and 96% ethanol and methanol for 1 h. The FSL variants Atri-Ad-DOPE (3A0ad), Atri-Ad-sterol (3A0as), Atetra-Ad-DOPE (4A2ad) and Atetra-CMG-DOPE (4A2cd) when exposed to water (H<sub>2</sub>O) for 1 h all stained strongly by EIA. Exposure to 50% alcohol was effective at removing most constructs, but increasing concentrations of alcohol were less effective. Differences observed can be partially attributable to the type of lipid tail, spacer and size of the saccharide moiety. These results are probably only valid for this type of surface (paper).</p>
Full article ">Figure 15
<p>Detergent/surfactant stability of blood group A FSL on paper and nylon nanofibres. FSL variants Atri-Ad-DOPE (printed as FSL-Atri in image) and Atetra-CMG-DOPE (printed as FSL-Atetra in image) were printed on both paper and nylon nanofibres, exposed to deionised water (H<sub>2</sub>O), 70% methanol (MeOH), 5% Tween 20 or 5% Triton X-100 for 1 h, washed then immunostained by EIA.</p>
Full article ">
1416 KiB  
Article
Direct Analysis in Real Time (DART) of an Organothiophosphate at Ultrahigh Resolution by Fourier Transform Ion Cyclotron Resonance Mass Spectrometry and Tandem Mass Spectrometry
by Laszlo Prokai and Stanley M. Stevens
Int. J. Mol. Sci. 2016, 17(1), 116; https://doi.org/10.3390/ijms17010116 - 16 Jan 2016
Cited by 7 | Viewed by 6461
Abstract
Direct analysis in real time (DART) is a recently developed ambient ionization technique for mass spectrometry to enable rapid and sensitive analyses with little or no sample preparation. After swab-based field sampling, the organothiophosphate malathion was analyzed using DART-Fourier transform ion cyclotron resonance [...] Read more.
Direct analysis in real time (DART) is a recently developed ambient ionization technique for mass spectrometry to enable rapid and sensitive analyses with little or no sample preparation. After swab-based field sampling, the organothiophosphate malathion was analyzed using DART-Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometry (MS) and tandem mass spectrometry (MS/MS). Mass resolution was documented to be over 800,000 in full-scan MS mode and over 1,000,000 for an MS/MS product ion produced by collision-induced dissociation of the protonated analyte. Mass measurement accuracy below 1 ppm was obtained for all DART-generated ions that belonged to the test compound in the mass spectra acquired using only external mass calibration. This high mass measurement accuracy, achievable at present only through FTMS, was required for unequivocal identification of the corresponding molecular formulae. Full article
(This article belongs to the Special Issue Fourier Transform Mass Spectrometry in Molecular Sciences)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of the sampling method: (<b>a</b>) About 1.3 ± 0.1 mL of aqueous 0.4% <span class="html-italic">w</span>/<span class="html-italic">v</span> malathion solution was sprayed from approximately 30 cm distance from a household spray bottle (foreground) by a single application of the lever onto a concrete wall in the background; (<b>b</b>) After swabbing the surface across the sprayed area using cotton-tipped applicators with their wooden handle glued in scintillation-vial caps (on the <b>left</b>), the swabs were screwed into the vials (on the <b>right</b>) to protect from sample loss and cross-contamination.</p>
Full article ">Figure 2
<p>(<b>a</b>) Direct analysis in real time (DART) mass spectrum of a sample collected by swabbing from the field (<a href="#ijms-17-00116-f001" class="html-fig">Figure 1</a>) using FT-ICR with desired <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> set to 100,000 FWHM at <span class="html-italic">m/z</span> 400 (Mass spectrum averaged for the entire acquisition period); actual <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> (FWHM) achieved for (<b>b</b>) the protonated malathion (nominal <span class="html-italic">m/z</span> 331) and (<b>c</b>) DART fragment ion of the compound (nominal <span class="html-italic">m/z</span> 285) at set mass resolution of 100,000 at <span class="html-italic">m/z</span> 400 (The insets show the actual <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> with desired mass resolution set to 500,000 at <span class="html-italic">m/z</span> 400).</p>
Full article ">Figure 3
<p>Product ion MS/MS of protonated malathion (<span class="html-italic">m/z</span> 331) from a sample collected by swabbing from the field (<a href="#ijms-17-00116-f001" class="html-fig">Figure 1</a>) followed by DART–MS using a hybrid mass spectrometer with CID in the linear ion trap and fragment ions detected in the FT-ICR (set <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> of 100,000 FWHM at <span class="html-italic">m/z</span> 400). Inset: The fragment ion <span class="html-italic">m/z</span> 285 (set <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> of 500,000 FWHM at <span class="html-italic">m/z</span> 400). The fragmentation leading to the observed ion is indicated by the thick dashed line on the displayed structure of the compound.</p>
Full article ">Figure 4
<p>Relative abundances of the measured A+1 and A+2 isotope peaks (bottom trace), as well as isotopic fine structure of the A+2 isotope peak (inset over the bottom trace labeled “Measured”) for the protonated analyte also affords the best match with those predicted for C<sub>10</sub>H<sub>20</sub>O<sub>6</sub>PS<sub>2</sub>. The isotope peak abundances and isotopic fine structures for the indicated elemental compositions listed in <a href="#ijms-17-00116-t001" class="html-table">Table 1</a><b>a</b> (colored traces and insets, respectively) were predicted presuming <span class="html-italic">M</span>/Δ<span class="html-italic">M</span> of 120,000, FWHM.</p>
Full article ">
6530 KiB  
Article
Cerebellar Expression of the Neurotrophin Receptor p75 in Naked-Ataxia Mutant Mouse
by Maryam Rahimi Balaei, Xiaodan Jiao, Niloufar Ashtari, Pegah Afsharinezhad, Saeid Ghavami and Hassan Marzban
Int. J. Mol. Sci. 2016, 17(1), 115; https://doi.org/10.3390/ijms17010115 - 15 Jan 2016
Cited by 13 | Viewed by 6740
Abstract
Spontaneous mutation in the lysosomal acid phosphatase 2 (Acp2) mouse (nax—naked-ataxia mutant mouse) correlates with severe cerebellar defects including ataxia, reduced size and abnormal lobulation as well as Purkinje cell (Pc) degeneration. Loss of Pcs in the nax cerebellum is compartmentalized [...] Read more.
Spontaneous mutation in the lysosomal acid phosphatase 2 (Acp2) mouse (nax—naked-ataxia mutant mouse) correlates with severe cerebellar defects including ataxia, reduced size and abnormal lobulation as well as Purkinje cell (Pc) degeneration. Loss of Pcs in the nax cerebellum is compartmentalized and harmonized to the classic pattern of gene expression of the cerebellum in the wild type mouse. Usually, degeneration starts in the anterior and posterior zones and continues to the central and nodular zones of cerebellum. Studies have suggested that the p75 neurotrophin receptor (NTR) plays a role in Pc degeneration; thus, in this study, we investigated the p75NTR pattern and protein expression in the cerebellum of the nax mutant mouse. Despite massive Pc degeneration that was observed in the nax mouse cerebellum, p75NTR pattern expression was similar to the HSP25 pattern in nax mice and comparable with wild type sibling cerebellum. In addition, immunoblot analysis of p75NTR protein expression did not show any significant difference between nax and wild type sibling (p > 0.5). In comparison with wild type counterparts, p75NTR pattern expression is aligned with the fundamental cytoarchitecture organization of the cerebellum and is unchanged in the nax mouse cerebellum despite the severe neurodevelopmental disorder accompanied with Pc degeneration. Full article
(This article belongs to the Special Issue Mechanisms of Neurodegeneration)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Frontal sections of an adult mouse cerebellum immunostained with p75NTR, CaBP and HSP25 at P28. The lobules are shown by Roman numerals. (<b>A</b>,<b>B</b>) The cerebellar cortex immunoperoxidase stained by p75NTR shows that the Pc somata in the Purkinje cell layer and dendrites in the molecular layer are immunoreactive (arrow; black in A, and white in B). Some Pc somata expressing p75NTR are weak (A, arrowhead) or absent (B, arrowhead); (<b>C</b>–<b>E</b>) Double immunostaining of a frontal section at lobule IX and X using p75NTR (red) and CaBP (green). Stripes of immunoreactive Pcs are clear in the vermis; (<b>F</b>) Immunostaining of a frontal section at the cerebellum using p75NTR. Immunoreactivity is not present in the AZ and PZ. Stripes of immunoreactive Pcs are clear in the vermis of the CZ (VII) and NZ (X); (<b>G</b>) Immunostaining of a frontal section through the lobule VII using HSP25. Stripes of immunoreactive Pcs are clear in the vermis of the CZ. Abbreviations: ml, molecular layer; Pcl, Purkinje cell layer; gl, granular layer.Scale bars: <b>A</b> = 100 µm; <b>B</b> = 20 µm; <b>E</b> = 250 μm (applies to <b>C</b>–<b>E</b>); <b>F</b> = 500 µm; <b>G</b> = 250 µm.</p>
Full article ">Figure 2
<p>Sagittal sections of P4 wt<sup>+/+</sup> (<b>A</b>–<b>C</b>) and <span class="html-italic">nax</span><sup>−/−</sup> (<b>D</b>–<b>I</b>) and sagittal sections of P6 wt (<b>J</b>) and <span class="html-italic">nax</span> (<b>K</b>) cerebella immunostained with p75NTR, CaBP and Pax6. The lobules are shown by Roman numerals. (<b>A</b>–<b>C</b>) The wt cerebellum shows normal lobules (<b>A</b>) with expression of p75NTR (red) in the entire egz (external germinal zone)and dispersing through the Pc layers. A higher magnification of “A” shown in “B” and Pax6 immunostaining is indicated egz in “C”; (<b>D</b>–<b>F</b>) The <span class="html-italic">nax</span> cerebellum shows a small cerebellum with underdeveloped lobulation (<b>B</b>). P75NTR expression (red) is present in the underdeveloped egz and in the developing Pcl (<b>B</b>). Immunostaining with Pax6 shows the thin layer of the egz in the <span class="html-italic">nax</span> cerebellum (<b>F</b>); (<b>G</b>–<b>I</b>) Magnified views of white box in E showing CaBP (<b>G</b>), p75NTR (<b>H</b>), and merged image (<b>I</b>); (<b>J</b>–<b>K</b>) The sagittal sections immunoperoxidase stained by p75NTR shows immunoreactivity of the Pc in wt (<b>J</b>) and <span class="html-italic">nax</span> (<b>K</b>) cerebella. Abbreviations: egz, external germinal zone; Pcl, Purkinje cell layer. Scale bars: <b>D</b> = 500 µm (applies to <b>A</b> and <b>D</b>); <b>E</b> = 50 µm (applies to <b>B</b> and <b>E</b>); <b>F</b> = 40 μm (applies to <b>C</b> and <b>F</b>); <b>J</b> = 50 µm; <b>K</b> = 50 µm.</p>
Full article ">Figure 3
<p>Frontal sections at the cerebellum of wt (<b>A</b>,<b>B</b>) and <span class="html-italic">nax</span> (<b>C</b>,<b>D</b>) at P6 immunoperoxidase stained with p75NTR. (<b>A</b>,<b>B</b>) The wt cerebellum shows normal vermis (v) and hemisphere (h) lobulation (<b>A</b>) with uniform expression of p75NTR in the entire cerebellar cortex. A higher magnification of black box (b) in “<b>A</b>” indicates localization of p75NTR expression in Pcl and EGZ; (<b>C</b>,<b>D</b>) Underdeveloped <span class="html-italic">nax</span> cerebellum exhibits uniform p75NTR expression in putative vermis and in the hemisphere. Magnified views of black box (d) in “<b>C</b>” showing p75NTR in multilayer Pc and thin or lack of EGZ. The p75NTR expression shows the outline of cerebellar nuclei (cn) both in wt (<b>A</b>) and <span class="html-italic">nax</span> (<b>C</b>) sections. Abbreviations: egz, external germinal zone; Pcl, Purkinje cell layer; h, hemisphere; cn, cerebellar nuclei; gl, granular layer. Scale bars: <b>A</b> and <b>C</b> = 500 µm; <b>D</b> = 100 µm (applies to <b>B</b> and <b>D</b>).</p>
Full article ">Figure 4
<p>Transverse sections of <span class="html-italic">nax</span><sup>−/−</sup> (<b>A</b>–<b>D</b>) and wt sibling cerebella at P22 immunostained with p75NTR, CaBP and HSP25. (<b>A</b>,<b>B</b>) P75NTR immunostaining shows an array of parasagittal stripes in the CZ and NZ of the <span class="html-italic">nax</span> cerebellum, with high magnification of the CZ shown in <b>B</b>; (<b>C</b>,<b>D</b>) HSP25 immunostaining of the <span class="html-italic">nax</span> cerebellum shows parasagittal stripes in the CZ and NZ, <b>D</b> is a higher magnification of <b>C</b>; (<b>E</b>) At a high magnification, p75NTR expression is shown in the CZ of the wt<sup>+/+</sup> sibling. Scale bars: <b>C</b> = 1 mm (applies to <b>A</b> and <b>C</b>); <b>D</b> = 500 µm (applies to <b>B</b> and <b>D</b>); <b>E</b> = 250 μm.</p>
Full article ">Figure 5
<p>Western blot analysis of p75NTR expression during cerebellar development at P5, P9 and P18. This experiment was repeated over three different litters for each postnatal days in wt and <span class="html-italic">nax</span> siblings (P5, P9 and P18, wt; <span class="html-italic">n</span> = 9 and <span class="html-italic">nax</span>; <span class="html-italic">n</span> = 9). (<b>A</b>) Immunoblots of total cell lysate from wt sibling and <span class="html-italic">nax</span> mouse cerebellum indicate significant down-regulation in the wt sibling and also in the <span class="html-italic">nax</span> cerebellum from early postnatal development to P18. However, there is no considerable difference in p75NTR expression between the wt and <span class="html-italic">nax</span> cerebellum, despite the higher p75NTR expression in <span class="html-italic">nax</span> compared with wt. (<span class="html-italic">p</span> &gt; 0.05). Protein loading was confirmed using GAPDH (<span class="html-italic">n</span> = 18); (<b>B</b>) The data in the bar graph are presented as the mean ± SEM, and statistical analysis was performed using Kruskal-Wallis test followed by Dunnett’s multiple comparison test using Mann-Whitney test. Abbreviation: ns, not significant.</p>
Full article ">Figure 5 Cont.
<p>Western blot analysis of p75NTR expression during cerebellar development at P5, P9 and P18. This experiment was repeated over three different litters for each postnatal days in wt and <span class="html-italic">nax</span> siblings (P5, P9 and P18, wt; <span class="html-italic">n</span> = 9 and <span class="html-italic">nax</span>; <span class="html-italic">n</span> = 9). (<b>A</b>) Immunoblots of total cell lysate from wt sibling and <span class="html-italic">nax</span> mouse cerebellum indicate significant down-regulation in the wt sibling and also in the <span class="html-italic">nax</span> cerebellum from early postnatal development to P18. However, there is no considerable difference in p75NTR expression between the wt and <span class="html-italic">nax</span> cerebellum, despite the higher p75NTR expression in <span class="html-italic">nax</span> compared with wt. (<span class="html-italic">p</span> &gt; 0.05). Protein loading was confirmed using GAPDH (<span class="html-italic">n</span> = 18); (<b>B</b>) The data in the bar graph are presented as the mean ± SEM, and statistical analysis was performed using Kruskal-Wallis test followed by Dunnett’s multiple comparison test using Mann-Whitney test. Abbreviation: ns, not significant.</p>
Full article ">
464 KiB  
Article
Herb-Induced Liver Injury in the Berlin Case-Control Surveillance Study
by Antonios Douros, Elisabeth Bronder, Frank Andersohn, Andreas Klimpel, Reinhold Kreutz, Edeltraut Garbe and Juliane Bolbrinker
Int. J. Mol. Sci. 2016, 17(1), 114; https://doi.org/10.3390/ijms17010114 - 15 Jan 2016
Cited by 31 | Viewed by 10765
Abstract
Herb-induced liver injury (HILI) has recently attracted attention due to increasing reports of hepatotoxicity associated with use of phytotherapeutics. Here, we present data on HILI from the Berlin Case-Control Surveillance Study. The study was initiated in 2000 to investigate the serious toxicity of [...] Read more.
Herb-induced liver injury (HILI) has recently attracted attention due to increasing reports of hepatotoxicity associated with use of phytotherapeutics. Here, we present data on HILI from the Berlin Case-Control Surveillance Study. The study was initiated in 2000 to investigate the serious toxicity of drugs including herbal medicines. Potential cases of liver injury were ascertained in more than 180 Departments of all 51 Berlin hospitals from October 2002 to December 2011. Drug or herb intake was assessed through a standardized face-to-face interview. Drug or herbal aetiology was assessed based on the updated Council for International Organizations of Medical Sciences scale. In ten of all 198 cases of hepatotoxicity included in the study, herbal aetiology was assessed as probable (once ayurvedic herb) or possible (Valeriana five times, Mentha piperita once, Pelargonium sidoides once, Hypericum perforatum once, Eucalyptus globulus once). Mean age was 56.4 ± 9.7 years, and the predominant pattern of liver injury was hepatocellular. No cases of acute liver failure or death were observed. This case series corroborates known risks for ayurvedic herbs, supports the suspected association between Valeriana use and liver injury, and indicates a hepatotoxic potential for herbs such as Pelargonium sidoides, Hypericum perforatum or Mentha piperita that were rarely associated with liver injury before. However, given that possible causality does not prove clinical significance, further studies in this field are needed. Full article
(This article belongs to the Special Issue Drug, Herb, and Dietary Supplement Hepatotoxicity)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of lead substances of assessed phytotherapeutics.</p>
Full article ">
1286 KiB  
Article
In Silico Insight into Potential Anti-Alzheimer’s Disease Mechanisms of Icariin
by Zhijie Cui, Zhen Sheng, Xinmiao Yan, Zhiwei Cao and Kailin Tang
Int. J. Mol. Sci. 2016, 17(1), 113; https://doi.org/10.3390/ijms17010113 - 15 Jan 2016
Cited by 14 | Viewed by 6169
Abstract
Herbal compounds that have notable therapeutic effect upon Alzheimer's disease (AD) have frequently been found, despite the recent failure of late-stage clinical drugs. Icariin, which is isolated from Epimedium brevicornum, is widely reported to exhibit significant anti-AD effects in in vitro and in [...] Read more.
Herbal compounds that have notable therapeutic effect upon Alzheimer's disease (AD) have frequently been found, despite the recent failure of late-stage clinical drugs. Icariin, which is isolated from Epimedium brevicornum, is widely reported to exhibit significant anti-AD effects in in vitro and in vivo studies. However, the molecular mechanism remains thus far unclear. In this work, the anti-AD mechanisms of icariin were investigated at a target network level assisted by an in silico target identification program (INVDOCK). The results suggested that the anti-AD effects of icariin may be contributed by: attenuation of hyperphosphorylation of tau protein, anti-inflammation and regulation of Ca2+ homeostasis. Our results may provide assistance in understanding the molecular mechanism and further developing icariin into promising anti-AD agents. Full article
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of icariin docked into acetylcholinesterase (AChE) A chain (<b>a</b>,<b>c</b>) and PDE5 A chain (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 2
<p>Icariin’s overall anti-Alzheimer's disease (AD) mechanistic network. Light Red ovals represent predicted icariin’s targets. Blue ovals represent indirectly regulated genes by icariin with experimental results. Yellow arrows represent indirect effect from icariin on these genes. Red arrows represent direct effect from icariin on these targets. The direction of arrows refers to icariin’s effects on targets (activate/upregulate or inhibit/downregulate). A green oval represents approved therapeutic target for AD.</p>
Full article ">Figure 3
<p>Average shortest path calculation of inter-subnetworks.</p>
Full article ">
1926 KiB  
Article
Transcription Factor Sp1 Promotes the Expression of Porcine ROCK1 Gene
by Ruirui Zhang, Xiaoting Feng, Mengsi Zhan, Cong Huang, Kun Chen, Xiaoyin Tang, Tingting Kang, Yuanzhu Xiong and Minggang Lei
Int. J. Mol. Sci. 2016, 17(1), 112; https://doi.org/10.3390/ijms17010112 - 15 Jan 2016
Cited by 10 | Viewed by 6048
Abstract
Rho-associated, coiled-coil containing protein kinase 1 (ROCK1) gene plays a crucial role in maintaining genomic stability, tumorigenesis and myogenesis. However, little is known about the regulatory elements governing the transcription of porcine ROCK1 gene. In the current study, the transcription start [...] Read more.
Rho-associated, coiled-coil containing protein kinase 1 (ROCK1) gene plays a crucial role in maintaining genomic stability, tumorigenesis and myogenesis. However, little is known about the regulatory elements governing the transcription of porcine ROCK1 gene. In the current study, the transcription start site (TSS) was identified by 5’-RACE, and was found to differ from the predicted one. The region in ROCK1 promoter which is critical for promoter activity was investigated via progressive deletions. Site-directed mutagenesis indicated that the region from −604 to −554 bp contains responsive elements for Sp1. Subsequent experiments showed that ROCK1 promoter activity is enhanced by Sp1 in a dose-dependent manner, whereas treatment with specific siRNA repressed ROCK1 promoter activity. Electrophoretic mobility shift assay (EMSA), DNA pull down and chromatin immunoprecipitation (ChIP) assays revealed Sp1 can bind to this region. qRT-PCR and Western blotting research followed by overexpression or inhibition of Sp1 indicate that Sp1 can affect endogenous ROCK1 expression at both mRNA and protein levels. Overexpression of Sp1 can promote the expression of myogenic differentiation 1(MyoD), myogenin (MyoG), myosin heavy chain (MyHC). Taken together, we conclude that Sp1 positively regulates ROCK1 transcription by directly binding to the ROCK1 promoter region (from −604 to −532 bp) and may affect the process of myogenesis. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>5’-Deletion analysis of the porcine <span class="html-italic">ROCK1</span> promoter activity. Schematic representation of the progressive deletions of porcine <span class="html-italic">ROCK1</span> 5’-flanking region in pGL3-Basic vector and the relative activities of <span class="html-italic">ROCK1</span> promoter corresponding to the progressive deletions. The predicted transcription start site (TSS, the red arrow in the figure) was set +1, differs from the TSS in NCBI database. The pGL3-control/basic vectors were used as positive/negative control, while pRL-TK was used as internal control. Data were expressed as means ± SD of three replicates.</p>
Full article ">Figure 2
<p>Site-directed mutation of Sp1 binding sites in <span class="html-italic">ROCK1</span>-P5 fragment. (<b>A</b>) Schematic structure of site-directed mutagenesis in the putative Sp1 binding sites (the black slash) of porcine <span class="html-italic">ROCK1</span> gene. LUC represents the <span class="html-italic">Luciferase</span> gene in the vectors; (<b>B</b>,<b>C</b>) Luciferase activity of site-directed mutagenesis in PK and C2C12 cells. Statistical differences of relative activities were analyzed in the same cells; ** <span class="html-italic">p</span> &lt; 0.01, data were expressed as means ± SD of three replicates.</p>
Full article ">Figure 3
<p>Binding of Sp1 with the <span class="html-italic">ROCK1</span>-P5 fragment was analyzed <span class="html-italic">in vitro</span> and <span class="html-italic">in vivo</span>. The first (<b>A</b>), the second (<b>B</b>) and the third (<b>C</b>) biotin-labeled probes were incubated with the NE of PK cells. Lane 1 was the negative control without NE; the reagents were incubated in the absence competitor probes in Lane 2 or in presence of 50× excess competitor (Lane 3)/competitor-mutant (Lane 4) probes, respectively; (<b>D</b>) The three probes were incubated with PK NE, respectively; (<b>E</b>) Proteins of PK extracted from DNA-pull down materials were detected by Western blot. The total non-denaturing proteins/Streptavidin MagneSphere<sup>®</sup> Paramagnetic Particles were taken as positive/negative control (PC/NC). The three potential Sp1 binding sites were named as Sp1.1, Sp1.2, and Sp1.3 in (<b>D</b>,<b>E</b>). The competitor/competitor-mutant probes were 50-fold excess and arrows indicated the specific DNA-protein complex bands; (<b>F</b>) Schematic diagram of the Sp1 binding sites in the porcine <span class="html-italic">ROCK1</span>-P5 fragment; (<b>G</b>) ChIP assay of Sp1 binding to porcine <span class="html-italic">ROCK1</span>-P5 fragment in PK cells. The <span class="html-italic">in vivo</span> interaction of Sp1 and Sp3 with porcine <span class="html-italic">ROCK1</span> promoter was determined by ChIP assay, in which Normal mouse IgG was used as negative control. DNA isolated from immunoprecipitated materials was used for PCR amplification, whereas total chromatin was used as input (positive control). The antibodies used in ChIP assay were listed in the right of the figure and the corresponding amplification product obtained here was 107 bp.</p>
Full article ">Figure 4
<p>Sp1 stimulates the expression of porcine <span class="html-italic">ROCK1</span> gene. (<b>A</b>) Over-expression of <span class="html-italic">Sp1</span> up-regulated <span class="html-italic">ROCK1</span> luciferase activity; (<b>B</b>) Over-expression efficacy of Sp1; (<b>C</b>,<b>D</b>) Over-expression of Sp1 stimulated <span class="html-italic">ROCK1</span> expression at mRNA and protein level; (<b>E</b>) The interferences efficacy of siRNA; (<b>F</b>) Suppressing <span class="html-italic">Sp1</span> reduced the <span class="html-italic">ROCK1</span> promoter activity; (<b>G</b>,<b>H</b>) Inhibition of <span class="html-italic">Sp1</span> suppressed <span class="html-italic">ROCK1</span> expression at mRNA and protein level. The amount of plasmid was kept constant by the addition of pcDNA3.1 (+) vector. The data were obtained both in PK and C2C12 cells and expressed as means ± SD of three replicates. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p><span class="html-italic">Sp1</span> promotes the process of myogenesis. Over-expression of <span class="html-italic">Sp1</span> significantly stimulated <span class="html-italic">MyoD</span>, <span class="html-italic">MyoG</span>, <span class="html-italic">MyHC</span> mRNA expression in C2C12 cells, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop