[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (29)

Search Parameters:
Keywords = glycosylated serum albumin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 11024 KiB  
Article
ACE Phenotyping in Human Blood and Tissues: Revelation of ACE Outliers and Sex Differences in ACE Sialylation
by Enikő E. Enyedi, Pavel A. Petukhov, Alexander J. Kozuch, Steven M. Dudek, Attila Toth, Miklós Fagyas and Sergei M. Danilov
Biomedicines 2024, 12(5), 940; https://doi.org/10.3390/biomedicines12050940 - 23 Apr 2024
Viewed by 1044
Abstract
Angiotensin-converting enzyme (ACE) metabolizes a number of important peptides participating in blood pressure regulation and vascular remodeling. Elevated ACE expression in tissues (which is generally reflected by blood ACE levels) is associated with an increased risk of cardiovascular diseases. Elevated blood ACE is [...] Read more.
Angiotensin-converting enzyme (ACE) metabolizes a number of important peptides participating in blood pressure regulation and vascular remodeling. Elevated ACE expression in tissues (which is generally reflected by blood ACE levels) is associated with an increased risk of cardiovascular diseases. Elevated blood ACE is also a marker for granulomatous diseases. Decreased blood ACE activity is becoming a new risk factor for Alzheimer’s disease. We applied our novel approach—ACE phenotyping—to characterize pairs of tissues (lung, heart, lymph nodes) and serum ACE in 50 patients. ACE phenotyping includes (1) measurement of ACE activity with two substrates (ZPHL and HHL); (2) calculation of the ratio of hydrolysis of these substrates (ZPHL/HHL ratio); (3) determination of ACE immunoreactive protein levels using mAbs to ACE; and (4) ACE conformation with a set of mAbs to ACE. The ACE phenotyping approach in screening format with special attention to outliers, combined with analysis of sequencing data, allowed us to identify patient with a unique ACE phenotype related to decreased ability of inhibition of ACE activity by albumin, likely due to competition with high CCL18 in this patient for binding to ACE. We also confirmed recently discovered gender differences in sialylation of some glycosylation sites of ACE. ACE phenotyping is a promising new approach for the identification of ACE phenotype outliers with potential clinical significance, making it useful for screening in a personalized medicine approach. Full article
(This article belongs to the Section Molecular and Translational Medicine)
Show Figures

Figure 1

Figure 1
<p>ACE phenotyping in human lung homogenates. ACE activity in 18 samples of human homogenates (1:9 weight/volume ratio, diluted 1/20 in PBS, 9 males and 9 females) was quantified using a spectrofluorometric assay with Z-Phe-His-Leu (<b>A</b>) and Hip-His-Leu (not shown) as substrates. ACE activity from human lung homogenates was precipitated by mAb 9B9 (<b>B</b>) and by mAbs 1G12, 2H9 and 2D1 (not shown). (<b>C</b>) Ratio of the rate of hydrolysis of the two substrates (ZPHL/HHL ratio) in the tested samples. (<b>D</b>) 1G12/9B9 binding ratio. (<b>E</b>) 2H9/2D1 binding ratio. Data expressed as % of individual ACE activity in solution (<b>A</b>) or precipitated by mAbs (<b>B</b>–<b>E</b>) from mean values for all samples. Bars with significant changes in % of control ACE activity are colored as follows: increase more than 20%—with orange, &gt;50%—with brown, more than 2-fold—with red, decrease more than 20%—yellow, more than 50%—light blue. Grey bars represent data within normal values between 80 and 120% of control. Mean values (+SD) from 2–5 experiments (each made in triplicates).</p>
Full article ">Figure 2
<p>ACE phenotyping in human serum. In total, 18 samples, diluted 1/15 in PBS, which are the pairs from the same patients as tested lung tissues, were performed exactly as in <a href="#biomedicines-12-00940-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Conformational fingerprinting of ACE from patient S13. ACE activity was precipitated from lung homogenates of patient S13 (diluted 1/20) versus pooled lung homogenates of patients 2 to 9 (<b>A</b>) by 20 mAbs to different epitopes of human somatic ACE. Analogous conformational fingerprinting of blood ACE from patient S13 (<b>B</b>) and patient TX-18 (<b>C</b>) was performed with 7 mAbs to ACE. Data presented as ratio (%) of ACE activity precipitation from lung (<b>A</b>) to that from control lungs or serum (<b>B</b>,<b>C</b>) or that from control serum. Mean values (+SD) from 2 experiments (each made in triplicates). Bars are colored as in <a href="#biomedicines-12-00940-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 4
<p>Effect of human serum albumin on lung ACE activity from patient S13. (<b>A</b>,<b>B</b>) Lung homogenate of patient S13 (red) and control pooled lung homogenate (black) were titrated with different concentrations of human serum albumin (HSA)—free fatty acid-bound albumin (HAS + FFA) and FFA-depleted HSA-preparation (HSA-FFA). After equilibration, ACE activity was determined with fluorometric assay with fluorogenic substrate (Abz-FRK(Dnp)P-OH), as described earlier [<a href="#B19-biomedicines-12-00940" class="html-bibr">19</a>]. Data expressed as a % of residual activity. (<b>C</b>) Superposition of ligand-free (gray, PDB: 1E78) and fatty acid-bound (wheat, PDB:1E7H) human albumin. Fatty acids are rendered as van der Waals spheres in red. The sequence 138-144, Acein-1 [<a href="#B36-biomedicines-12-00940" class="html-bibr">36</a>], is rendered in magenta and light pink colors in ligand-free and fatty acid-bound albumin, respectively. The sequence 210–218, Albutensin A [<a href="#B37-biomedicines-12-00940" class="html-bibr">37</a>], is rendered in blue in ligand-free and fatty acid-bound albumin.</p>
Full article ">Figure 5
<p>Albumin docking to ACE dimer. Human somatic ACE model presented as a dimer [<a href="#B25-biomedicines-12-00940" class="html-bibr">25</a>] was used for modeling human albumin docking to ACE. Docked structure of fatty acid-bound (cyan, PDB:1E7H) albumin to a dimer model of ACE [<a href="#B16-biomedicines-12-00940" class="html-bibr">16</a>]. Two molecules of albumin can bind to the ACE dimer, but only one albumin molecule is shown to show which ACE epitopes interact with albumin. Fatty acids are rendered as van der Waals spheres in red. The subdomains I and II of N domain of ACE are rendered with grey and wheat colors, respectively. The subdomains I and II of C domain of ACE are rendered with pink and yellow colors, respectively. Epitope for mAb 2D1 showed schematically as a circle with 600 A<sup>2</sup>.</p>
Full article ">Figure 6
<p>Inhibitory effect of human serum albumin (HSA) on ACE activity. Lung and spleen homogenates (1/20 dilution in PBS) were incubated with HSA at final concentration 10 mg/mL and residual ACE activity was determined fluorometrically with ZPHL (<b>A</b>) and HHL (<b>B</b>) as substrates. (<b>C</b>) Ratio of the rates of hydrolysis by two substrates. Data are mean ± SD from 3 independent experiments performed in triplicates. Three spleen tissues were from patients with Gaucher disease. ACE activity was precipitated from spleen and lung homogenates of unrelated patients (<b>D</b>) and spleen homogenates from patients with Gaucher disease (<b>E</b>). Data are presented as binding ratios for each mAb (mean ± SD from 3 independent experiments performed in triplicates). (<b>F</b>) Model of ACE dimer (adapted from [<a href="#B25-biomedicines-12-00940" class="html-bibr">25</a>]) where docking of HSA (from <a href="#biomedicines-12-00940-f005" class="html-fig">Figure 5</a>) was shown by black arrow on one monomer and putative ACE binding protein preventing albumin binding was shown on another monomer in this ACE dimer.</p>
Full article ">Figure 7
<p>Model of CCL18 docking to ACE. A model of human somatic ACE (PDB 7Q3Y [<a href="#B42-biomedicines-12-00940" class="html-bibr">42</a>]) was used for modeling CCL18 docking to ACE. The surface ACE is colored in beige. CCL18 structure (PDB 4MHE) is shown in gray. The epitopes for mAbs on the N and C domains of ACE are shown in different colors. Asparagine (N) residues of the putative glycosylation sites are highlighted in lime green. The N and C terminal residues of CCL18 are highlighted in black.</p>
Full article ">
19 pages, 6613 KiB  
Article
Influence of Desialylation on the Drug Binding Affinity of Human Alpha-1-Acid Glycoprotein Assessed by Microscale Thermophoresis
by Tino Šeba, Robert Kerep, Tin Weitner, Dinko Šoić, Toma Keser, Gordan Lauc and Mario Gabričević
Pharmaceutics 2024, 16(2), 230; https://doi.org/10.3390/pharmaceutics16020230 - 5 Feb 2024
Viewed by 1178
Abstract
Human serum alpha-1-acid glycoprotein (AAG) is an acute-phase plasma protein involved in the binding and transport of many drugs, especially basic and lipophilic substances. The sialic acid groups that terminate the N-glycan chains of AAG have been reported to change in response to [...] Read more.
Human serum alpha-1-acid glycoprotein (AAG) is an acute-phase plasma protein involved in the binding and transport of many drugs, especially basic and lipophilic substances. The sialic acid groups that terminate the N-glycan chains of AAG have been reported to change in response to numerous health conditions and may have an impact on the binding of drugs to AAG. In this study, we quantified the binding between native and desialylated AAG and seven drugs from different pharmacotherapeutic groups (carvedilol, diltiazem, dipyridamole, imipramine, lidocaine, propranolol, vinblastine) using microscale thermophoresis (MST). This method was chosen due to its robustness and high sensitivity, allowing precise quantification of molecular interactions based on the thermophoretic movement of fluorescent molecules. Detailed glycan analysis of native and desialylated AAG showed over 98% reduction in sialic acid content for the enzymatically desialylated AAG. The MST results indicate that desialylation generally alters the binding affinity between AAG and drugs, leading to either an increase or decrease in Kd values, probably due to conformational changes of AAG caused by the different sialic acid content. This effect is also reflected in an increased denaturation temperature of desialylated AAG. Our findings indicate that desialylation impacts free drug concentrations differently, depending on the binding affinity of the drug with AAG relative to human serum albumin (HSA). For drugs such as dipyridamole, lidocaine, and carvedilol, which have a higher affinity for AAG, desialylation significantly changes free drug concentrations. In contrast, drugs such as propranolol, imipramine, and vinblastine, which have a strong albumin binding, show only minimal changes. It is noteworthy that the free drug concentration of dipyridamole is particularly sensitive to changes in AAG concentration and glycosylation, with a decrease of up to 15% being observed, underscoring the need for dosage adjustments in personalized medicine. Full article
Show Figures

Figure 1

Figure 1
<p>Normalized UPLC chromatogram of fluorescently labeled and purified N-glycans. Fluorescence was recorded using an excitation wavelength of 250 nm and an emission wavelength of 482 nm. The numbers in the figure correspond to the Peak No. in <a href="#pharmaceutics-16-00230-t001" class="html-table">Table 1</a>. The top panel represents native human AAG sample; bottom panel represents desialylated human AAG sample.</p>
Full article ">Figure 2
<p>The interaction of drugs with native (green dots) and desialylated AAG (orange dots) was assessed using MST. A titration series of drugs was performed while the labeled AAG was kept constant (20 nM). The solid line represents the theoretical fit of the data. Error bars indicate the standard deviation for each data point, which was calculated based on two independent measurements.</p>
Full article ">Figure 3
<p>The recorded fluorescence signal of native (green line) and desialylated (orange line) AAG (1 mg/mL) as a function of temperature. The measurements were taken in 25 mM sodium phosphate buffer at pH 7.4. Inflection temperatures of native (green dot) and desialylated (orange dot) AAG are 67.0 and 72.2 °C, respectively.</p>
Full article ">Figure 4
<p>Percentage of free drug at chosen <span class="html-italic">C</span><sub>max</sub> (<b>top panel</b>) and <span class="html-italic">C</span><sub>min</sub> (<b>bottom panel</b>) therapeutic values depending on the plasma concentrations of native AAG + 45 mg/mL HSA (AAG + s) or desialylated AAG + 45 mg/mL HSA (AAG − s).</p>
Full article ">Figure 5
<p>Percentage difference of free drug at <span class="html-italic">C</span><sub>max</sub> and <span class="html-italic">C</span><sub>min</sub>. Values are calculated versus the native AAG as reference.</p>
Full article ">
19 pages, 5302 KiB  
Article
Anti-Aggregative and Protective Effects of Vicenin-2 on Heat and Oxidative Stress-Induced Damage on Protein Structures
by Giuseppe Tancredi Patanè, Lisa Lombardo, Stefano Putaggio, Ester Tellone, Silvana Ficarra, Davide Barreca, Giuseppina Laganà, Laura De Luca and Antonella Calderaro
Int. J. Mol. Sci. 2023, 24(24), 17222; https://doi.org/10.3390/ijms242417222 - 7 Dec 2023
Cited by 1 | Viewed by 926
Abstract
Vicenin-2, a flavonoid categorized as a flavones subclass, exhibits a distinctive and uncommon C-glycosidic linkage. Emerging evidence challenges the notion that deglycosylation is not a prerequisite for the absorption of C-glycosyl flavonoid in the small intestine. Capitalizing on this experimental insight and considering [...] Read more.
Vicenin-2, a flavonoid categorized as a flavones subclass, exhibits a distinctive and uncommon C-glycosidic linkage. Emerging evidence challenges the notion that deglycosylation is not a prerequisite for the absorption of C-glycosyl flavonoid in the small intestine. Capitalizing on this experimental insight and considering its biological attributes, we conducted different assays to test the anti-aggregative and antioxidant capabilities of vicenin-2 on human serum albumin under stressful conditions. Within the concentration range of 0.1–25.0 μM, vicenin-2 effectively thwarted the heat-induced HSA fibrillation and aggregation of HSA. Furthermore, in this study, we have observed that vicenin-2 demonstrated protective effects against superoxide anion and hydroxyl radicals, but it did not provide defense against active chlorine. To elucidate the underlying mechanisms, behind this biological activity, various spectroscopy techniques were employed. UV-visible spectroscopy revealed an interaction between HSA and vicenin-2. This interaction involves the cinnamoyl system found in vicenin-2, with a peak of absorbance observed at around 338 nm. Further evidence of the interaction comes from circular dichroism spectrum, which shows that the formation of bimolecular complex causes a reduction in α-helix structures. Fluorescence and displacement investigations indicated modifications near Trp214, identifying Sudlow’s site I, similarly to the primary binding site. Molecular modeling revealed that vicenin-2, in nonplanar conformation, generated hydrophobic interactions, Pi-pi stacking, and hydrogen bonds inside Sudlow’s site I. These findings expand our understanding of how flavonoids bind to HSA, demonstrating the potential of the complex to counteract fibrillation and oxidative stress. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Representative fluorescence microscopy images of HSA fibrils in the absence or presence (0.0–25.0 µM) of vicenin-2. (<b>B</b>) UV-visible spectra of Congo red alone or with HSA treated at high temperature. (<b>C</b>) Variation in the maximum absorbance of Congo red–HSA in the absence or presence of 0.0–25.0 µM of vicenin-2 after treatment to induce fibrillation. (a) HSA alone incubated at 338 K; (b) HSA alone incubated at 310 K. Asterisks (**) indicate a significant difference with respect to control (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The impact of vicenin-2 on protein degradation induced by superoxide anion, hydroxyl radical, and active chlorine. HSA was subjected to electrophoretic separation on polyacrylamide gel electrophoresis (PAGE), with incubation in the absence or presence of different radical alone, or with the addition of 12.5–100.0 μM of the flavonoid. The samples were subjected a 40 min incubation at 310 K and were subsequently analyzed by 7.5% polyacrylamide-gel electrophoresisThe integrated density of each band is presented as a percentage of the untreated HSA sample in the experiments involving superoxide anion, hydroxyl radical, and active chlorine. The histograms depict the data as means ± S.D. (n = 3). Asterisks (**) indicate a significant difference with respect to control (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>UV-visible absorption spectra of vicenin-2 (<b>—</b> 76.5 µM) in the absence or presence of increasing HSA concentrations (<b><span style="color:#2E74B5">—</span></b> 9.4 µM, <b><span style="color:#FF33CC">—</span></b> 18.8 µM, <b><span style="color:lime">—</span></b> 37.7 µM, <b><span style="color:red">—</span></b> 75.5 µM). The inset shows the variation of absorbance obtained with three different experiments at the maximum absorbance of Band I. Asterisks (**) indicate a significant difference with respect to control (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Fluorescence emission spectra of HSA (1.5 × 10<sup>−5</sup> mol/L) in the absence or presence of the same molar concentration (1.5 × 10<sup>−5</sup> mol/L) of vicenin-2 at 298 K and the maximum tested concentration of the flavonoid (<b>A</b>). Stern–Volmer (<b>B</b>) and modified Stern–Volmer (<b>C</b>) plots for the vicenin-2–HSA complex at three different temperatures. Analysis of binding equilibrium, thermodynamics, and acting forces. Plots of log(<span class="html-italic">F</span>0 − <span class="html-italic">F</span>)/<span class="html-italic">F</span> as a function of <span class="html-italic">log</span>[<span class="html-italic">Q</span>] for the binding of vicenin-2 to HSA at the temperature of 310 K (<b>D</b>), as well as van’t Hoff plot (<b>E</b>) and effect of site-specific markers on the fluorescence of HSA–vicenin-2 complex (<b>F</b>). Data are the results of three different experiments, expressed as mean ± SD.</p>
Full article ">Figure 5
<p>Superimposition of the crystalized warfarin (yellow sticks) to the rigid docking (<b>A</b>) and the induced fit docking (<b>B</b>) poses (purple sticks).</p>
Full article ">Figure 6
<p>An overview of the HSA 3D structure and identification of Sudlow’s site I (<b>A</b>). Magnification focused on the binding site and the interactions of the vicenin-2 with the residues forming the pockets (<b>B</b>).</p>
Full article ">Figure 7
<p>The plots of RMSD of the complex HSA–vicenin-2 (<b>A</b>) and the free HSA (<b>B</b>). On the x-axis, the simulation time is depicted. The left y-axis illustrates the HAS RMSD progression, while the y-axis on the right plots the RMSD evolution of vicenin-2 within the binding site during the simulation, indicating the degree of stability exhibited by vicenin-2.</p>
Full article ">Figure 7 Cont.
<p>The plots of RMSD of the complex HSA–vicenin-2 (<b>A</b>) and the free HSA (<b>B</b>). On the x-axis, the simulation time is depicted. The left y-axis illustrates the HAS RMSD progression, while the y-axis on the right plots the RMSD evolution of vicenin-2 within the binding site during the simulation, indicating the degree of stability exhibited by vicenin-2.</p>
Full article ">Figure 8
<p>Molecular interaction in the complex. (<b>A</b>) Protein–ligand interactions are classified into four categories distinguishable by the color of the bars: hydrophobic interactions (purple), hydrogen bonds (green), ionic contacts (fuchsia), and water bridges (blue). (<b>B</b>) An in-depth examination of the contacts between ligand atoms and residues, considering interactions that occur more than 40% of the time.</p>
Full article ">Figure 9
<p>CD spectra of free HSA and the vicenin-2/HSA complex at T = 298 K in phosphate buffer of pH 7.4, 20 mM.</p>
Full article ">
12 pages, 4190 KiB  
Article
Recombinant Human CD19 in CHO-K1 Cells: Glycosylation Patterns as a Quality Attribute of High Yield Processes
by Magdalena Billerhart, Monika Hunjadi, Vanessa Hawlin, Clemens Grünwald-Gruber, Daniel Maresch, Patrick Mayrhofer and Renate Kunert
Int. J. Mol. Sci. 2023, 24(13), 10891; https://doi.org/10.3390/ijms241310891 - 30 Jun 2023
Cited by 2 | Viewed by 2330
Abstract
CD19 is an essential protein in personalized CD19-targeting chimeric antigen receptor (CAR)-T cell-based cancer immunotherapies and CAR-T cell functionality evaluation. However, the recombinant expression of this “difficult to-express” (DTE) protein is challenging, and therefore, commercial access to the protein is limited. We have [...] Read more.
CD19 is an essential protein in personalized CD19-targeting chimeric antigen receptor (CAR)-T cell-based cancer immunotherapies and CAR-T cell functionality evaluation. However, the recombinant expression of this “difficult to-express” (DTE) protein is challenging, and therefore, commercial access to the protein is limited. We have previously described the successful stable expression of our soluble CD19-AD2 fusion protein of the CD19 extracellular part fused with human serum albumin domain 2 (AD2) in CHO-K1 cells. The function, stability, and secretion rate of DTE proteins can be improved by culture conditions, such as reduced temperature and a shorter residence time. Moreover, glycosylation, as one of the most important post-translational modifications, represents a critical quality attribute potentially affecting CAR-T cell effector function and thus impacting therapy’s success. In this study, we increased the production rate of CD19-AD2 by 3.5-fold through applying hypothermic culture conditions. We efficiently improved the purification of our his-tagged CD19-AD2 fusion protein via a Ni-NTA-based affinity column using a stepwise increase in the imidazole concentration. The binding affinity to commercially available anti-CD19 antibodies was evaluated via Bio-Layer Interferometry (BLI). Furthermore, we revealed glycosylation patterns via Electrospray Ionization Mass Spectrometry (ESI–MS), and five highly sialylated and multi-antennary N-glycosylation sites were identified. In summary, we optimized the CD19-AD2 production and purification process and were the first to characterize five highly complex N-glycosylation sites. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>The semi-continuous perfusion cultivation process at 37 °C (right panel) vs. 32 °C (left panel). (<b>A</b>) VCD and viability of all process days. (<b>B</b>) The volumetric titer of CD19-AD2 determined from the daily harvested supernatant is displayed. The asterisk indicates the temperature shift from 37 °C to 32 °C, and a vertical dashed line indicates the start of feed supplementation on day four.</p>
Full article ">Figure 2
<p>CD19-AD2 protein purification. (<b>A</b>) The HisTrap profile example for the stepwise elution of CD19-AD2 fusion protein with 25% and 45% of 500 mM imidazole, containing elution buffer (pH 7.4). The grey highlighter indicates the peak fractions of host cell proteins (fraction 1) and CD19-AD2 (fractions 2 and 3). (<b>B</b>) The quality of the purified protein was verified by non-reducing SDS-PAGE/Coomassie staining. Equivolumetric amounts of the pooled culture supernatant, 1:5 diluted (32 °C) or 1:10 diluted (37°) concentrated and rebuffered supernatant, as well as elution fractions 1 to 3, were loaded. The area between the horizontal lines represents the highly glycosylated CD19-AD2.</p>
Full article ">Figure 3
<p>Real-time binding sensorgrams of the immobilized biotinylated anti-CD19 antibodies (ligand concentration 10 µg/mL) and CD19-AD2 construct (analyte). (<b>A</b>) FMC63-btn; (<b>B</b>) HIB19-btn; (<b>C</b>) 3B10-btn; (<b>D</b>) 4G7-btn were captured on SA biosensors and dipped in wells containing the analyte CD19-AD2 at different concentrations between 10 and 600 nM. The spectral shift, corresponding to the thickness of the biolayer, differs for individual analyses. The binding signals (blue sensorgrams) were obtained by subtracting the signals from ligand-coated SA biosensors dipped in wells with buffer (non-specific binding control). Fitted curves are depicted as red lines and were obtained by global fitting using a 1:1 ligand model.</p>
Full article ">Figure 4
<p>Glycosylation pattern of CD19-AD2 expressed in CHO-K1 cells cultivated at 37 °C vs. 32 °C. The five N-glycosylation sites at position N67, N106, N119, N162, and N246 are characterized. (<b>A</b>) Color-coded bar graph of the site-specific glycosylation as relative proportions of found glycoforms with a table legend. (<b>B</b>) Representative structures of sialylated N-glycans found and their percentual representation are displayed extra in table format (threshold of prevalence &gt; 4.5%).</p>
Full article ">Figure 4 Cont.
<p>Glycosylation pattern of CD19-AD2 expressed in CHO-K1 cells cultivated at 37 °C vs. 32 °C. The five N-glycosylation sites at position N67, N106, N119, N162, and N246 are characterized. (<b>A</b>) Color-coded bar graph of the site-specific glycosylation as relative proportions of found glycoforms with a table legend. (<b>B</b>) Representative structures of sialylated N-glycans found and their percentual representation are displayed extra in table format (threshold of prevalence &gt; 4.5%).</p>
Full article ">
16 pages, 2173 KiB  
Article
Age-Related Changes in Post-Translational Modifications of Proteins from Whole Male and Female Skeletal Elements
by Elizabeth Johnston and Michael Buckley
Molecules 2023, 28(13), 4899; https://doi.org/10.3390/molecules28134899 - 21 Jun 2023
Cited by 3 | Viewed by 2040
Abstract
One of the key questions in forensic cases relates to some form of age inference, whether this is how old a crime scene is, when in time a particular crime was committed, or how old the victim was at the time of the [...] Read more.
One of the key questions in forensic cases relates to some form of age inference, whether this is how old a crime scene is, when in time a particular crime was committed, or how old the victim was at the time of the crime. These age-related estimations are currently achieved through morphological methods with varying degrees of accuracy. As a result, biomolecular approaches are considered of great interest, with the relative abundances of several protein markers already recognized for their potential forensic significance; however, one of the greatest advantages of proteomic investigations over genomics ones is the wide range of post-translational modifications (PTMs) that make for a complex but highly dynamic resource of information. Here, we explore the abundance of several PTMs including the glycosylation, deamidation, and oxidation of several key proteins (collagen, fetuin A, biglycan, serum albumin, fibronectin and osteopontin) as being of potential value to the development of an age estimation tool worthy of further evaluation in forensic contexts. We find that glycosylations lowered into adulthood but deamidation and oxidation increased in the same age range. Full article
(This article belongs to the Special Issue Mass Spectrometry-Driven Advancements in Forensic Science)
Show Figures

Figure 1

Figure 1
<p>Number of detected glycosylations of lysine in whole rat proteomes for both male and female rats. Males and females showed similar numbers throughout life, with a peak abundance seen at 6–8 weeks.</p>
Full article ">Figure 2
<p>Age-related changes in galactosyl (K) and glucosylgalactosyl (K) modifications for both male and female rats. (<b>A</b>) Collagen-I-α-1 with males and females showing similar increase in abundance throughout life. (<b>B</b>) Collagen-I-α-2, showed a much shallower increase in abundance, except for galactosyl (K) in female, which rapidly peaked at 8–10 weeks. (<b>C</b>) Collagen-II-α-1, which steadily decrease in abundance for both male and females.</p>
Full article ">Figure 3
<p>Abundance of N-linked glycosylations in the overall proteome in both female (<b>A</b>) and male (<b>B</b>) rats. (<b>A</b>) Female rats showed a relatively even but rapid decrease in all N-glycosylations from 3 to 4 weeks old. (<b>B</b>) Males also showed a similar decrease from 3 to 4 weeks, but less rapid than females.</p>
Full article ">Figure 4
<p>Abundance of O-linked glycosylations in the overall proteome for male and female rats. O-linked glycosylations are present on serine and threonine amino acids. (<b>A</b>,<b>B</b>) serine-related O-linked glycosylations in females and males, respectively, and both showed a decrease throughout life, but females began with an initially higher abundance in Hex(S) than males; and (<b>C</b>,<b>D</b>) threonine-related O-linked glycosylations in females and males, respectively, which showed a steady decline for both sexes.</p>
Full article ">Figure 5
<p>Graphs showing the proteins that displayed changes in abundance of PTMs across age. (<b>A</b>) Deamidation CO1A2, COBA1, FETUA, PGS1 and ALBU showed such a change. (<b>B</b>) Oxidation of lysine changes showed changes in CO1A1, CO1A2, CO2A1, COBA1, CO5A1, FETUA, PGS1 and OSTP, thus showing the highest number of proteins affected. (<b>C</b>) Oxidation of proline showed changes in CO1A1, CO1A2, COBA1, CO5A1, FETUA, PGS1, FETUA and PGS1. (<b>D</b>) Oxidation of methionine showed changes in CO1A1, COBA1, PGS1 AND ALBU. (<b>E</b>) Arg-&gt;GluSA had changes in CO1A1, CO2A1 AND COBA1. (<b>F</b>) Lys-&gt;Allysine had changes in CO1A1, CO1A2, CO2A1, COBA1 AND CO5A1. Pro-&gt;pyrrolidinone (<b>G</b>) and Pro-&gt;pyrrolidone (<b>H</b>) had the same proteins showing age-related changes: CO1A1, CO1A2, CO2A1 and COBA1. Graphs showing PTMs with proteins that showed no age-related change can be found in <a href="#app1-molecules-28-04899" class="html-app">Supplementary Material Figure S1</a>.</p>
Full article ">
12 pages, 1379 KiB  
Article
Effectiveness of One-Year Pemafibrate Therapy on Non-Alcoholic Fatty Liver Disease Refractory to Long-Term Sodium Glucose Cotransporter-2 Inhibitor Therapy: A Pilot Study
by Satoshi Shinozaki, Toshiyuki Tahara, Kouichi Miura, Alan Kawarai Lefor and Hironori Yamamoto
Life 2023, 13(6), 1327; https://doi.org/10.3390/life13061327 - 5 Jun 2023
Cited by 1 | Viewed by 2141
Abstract
Background: Both pemafibrate and sodium glucose cotransporter-2 (SGLT2) inhibitor can decrease serum transaminase levels in patients with non-alcoholic fatty liver disease (NAFLD) complicated with dyslipidemia and type 2 diabetes mellitus (T2DM), respectively. However, the effectiveness of combined therapy has been rarely reported. Methods: [...] Read more.
Background: Both pemafibrate and sodium glucose cotransporter-2 (SGLT2) inhibitor can decrease serum transaminase levels in patients with non-alcoholic fatty liver disease (NAFLD) complicated with dyslipidemia and type 2 diabetes mellitus (T2DM), respectively. However, the effectiveness of combined therapy has been rarely reported. Methods: This is a two-center retrospective observational study. NAFLD patients complicated with T2DM treated with pemafibrate for >1 year were included, in whom prior treatment with SGLT2 inhibitor > 1 year failed to normalize serum alanine aminotransferase (ALT) levels. Hepatic inflammation, function, and fibrosis were assessed by ALT, albumin-bilirubin (ALBI) score, and Mac-2 binding protein glycosylation isomer (M2BPGi) levels, respectively. Results: Seven patients were included. The median duration of prior treatment with SGLT2 inhibitors was 2.3 years. During the one year before starting pemafibrate therapy, the therapy did not significantly change hepatic enzymes. All patients received pemafibrate 0.1 mg twice daily without dose escalations. During one year of pemafibrate therapy, triglyceride, aspartate aminotransferase, ALT, γ-glutamyl transpeptidase, ALBI score, and M2BPGi levels significantly improved (p < 0.05), although weight or hemoglobin A1c did not significantly change. Conclusions: One year of pemafibrate therapy improves markers of hepatic inflammation, function, and fibrosis in NAFLD patients in whom long-term SGLT2 inhibitor therapy failed to normalize serum ALT. Full article
(This article belongs to the Section Physiology and Pathology)
Show Figures

Figure 1

Figure 1
<p>Study flowchart.</p>
Full article ">Figure 2
<p>Changes in parameters during a two-year treatment course including one year of sodium glucose cotransporter-2 (SGLT2) inhibitor therapy and one year of pemafibrate add-on therapy. The changes in each parameter shown include mean value of (<b>a</b>) weight, (<b>b</b>) aspartate aminotransferase, (<b>c</b>) alanine aminotransferase, (<b>d</b>) γ-glutamyl transpeptidase, and (<b>e</b>) serum triglycerides. All statistical analyses were performed using Friedman’s test. Dark black line: mean; grey line: each patient.</p>
Full article ">Figure 2 Cont.
<p>Changes in parameters during a two-year treatment course including one year of sodium glucose cotransporter-2 (SGLT2) inhibitor therapy and one year of pemafibrate add-on therapy. The changes in each parameter shown include mean value of (<b>a</b>) weight, (<b>b</b>) aspartate aminotransferase, (<b>c</b>) alanine aminotransferase, (<b>d</b>) γ-glutamyl transpeptidase, and (<b>e</b>) serum triglycerides. All statistical analyses were performed using Friedman’s test. Dark black line: mean; grey line: each patient.</p>
Full article ">
15 pages, 33494 KiB  
Article
Alterations in the Glycan Composition of Serum Glycoproteins in Attention-Deficit Hyperactivity Disorder
by Kristína Kianičková, Lucia Pažitná, Paras H. Kundalia, Zuzana Pakanová, Marek Nemčovič, Peter Baráth, Eva Katrlíková, Ján Šuba, Jana Trebatická and Jaroslav Katrlík
Int. J. Mol. Sci. 2023, 24(10), 8745; https://doi.org/10.3390/ijms24108745 - 14 May 2023
Cited by 4 | Viewed by 1879
Abstract
Changes in protein glycosylation are associated with most biological processes, and the importance of glycomic analysis in the research of disorders is constantly increasing, including in the neurodevelopmental field. We glycoprofiled sera in 10 children with attention-deficit hyperactivity disorder (ADHD) and 10 matching [...] Read more.
Changes in protein glycosylation are associated with most biological processes, and the importance of glycomic analysis in the research of disorders is constantly increasing, including in the neurodevelopmental field. We glycoprofiled sera in 10 children with attention-deficit hyperactivity disorder (ADHD) and 10 matching healthy controls for 3 types of samples: whole serum, sera after depletion of abundant proteins (albumin and IgG), and isolated IgG. The analytical methods used were a lectin-based glycoprotein microarray enabling high-throughput glycan analysis and matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry (MS) as a standard method for the identification of glycan structures. For microarray analysis, the samples printed on microarray slides were incubated with biotinylated lectins and detected using the fluorescent conjugate of streptavidin by a microarray scanner. In the ADHD patient samples, we found increased antennary fucosylation, decreased di-/triantennary N-glycans with bisecting N-acetylglucosamine (GlcNAc), and decreased α2-3 sialylation. The results obtained by both independent methods were consistent. The study’s sample size and design do not allow far-reaching conclusions to be drawn. In any case, there is a strong demand for a better and more comprehensive diagnosis of ADHD, and the obtained results emphasize that the presented approach brings new horizons to studying functional associations of glycan alterations in ADHD. Full article
(This article belongs to the Section Macromolecules)
Show Figures

Figure 1

Figure 1
<p>Graphic representation of relative signal intensities and normalized relative signal intensities (inserted figures) of interactions for the (<b>A</b>) sera, (<b>B</b>) DS (depleted sera without albumin and IgG), and (<b>C</b>) IgG fraction with lectins measured by the lectin-based glycoprotein microarray. Statistically significant differences (<span class="html-italic">t</span>-test, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05) are shown between the ADHD and control groups.</p>
Full article ">Figure 1 Cont.
<p>Graphic representation of relative signal intensities and normalized relative signal intensities (inserted figures) of interactions for the (<b>A</b>) sera, (<b>B</b>) DS (depleted sera without albumin and IgG), and (<b>C</b>) IgG fraction with lectins measured by the lectin-based glycoprotein microarray. Statistically significant differences (<span class="html-italic">t</span>-test, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05) are shown between the ADHD and control groups.</p>
Full article ">Figure 2
<p>Representative MS spectra of samples labeled as A6 for (<b>A</b>) serum, (<b>B</b>) DS, and (<b>C</b>) IgG, with marked signals of N-glycans present in all spectra of individual groups of samples. The <span class="html-italic">m/z</span> range is 1500–4000.</p>
Full article ">
10 pages, 1696 KiB  
Article
Direct Detection of Glycated Human Serum Albumin and Hyperglycosylated IgG3 in Serum, by MALDI-ToF Mass Spectrometry, as a Predictor of COVID-19 Severity
by Ray K. Iles, Jason K. Iles, Jonathan Lacey, Anna Gardiner and Raminta Zmuidinaite
Diagnostics 2022, 12(10), 2521; https://doi.org/10.3390/diagnostics12102521 - 17 Oct 2022
Cited by 8 | Viewed by 1989
Abstract
The prefusion spike protein of SARS-CoV-2 binds advanced glycation end product (AGE)-glycated human serum albumin (HSA) and a higher mass (hyperglycosylated/glycated) immunoglobulin (Ig) G3, as determined by matrix assisted laser desorption mass spectrometry (MALDI-ToF). We set out to investigate if the total blood [...] Read more.
The prefusion spike protein of SARS-CoV-2 binds advanced glycation end product (AGE)-glycated human serum albumin (HSA) and a higher mass (hyperglycosylated/glycated) immunoglobulin (Ig) G3, as determined by matrix assisted laser desorption mass spectrometry (MALDI-ToF). We set out to investigate if the total blood plasma of patients who had recovered from acute respiratory distress syndrome (ARDS) as a result of COVID-19, contained more glycated HSA and higher mass (glycosylated/glycated) IgG3 than those with only clinically mild or asymptomatic infections. A direct serum dilution, and disulphide bond reduction, method was developed and applied to plasma samples from SARS-CoV-2 seronegative (n = 30) and seropositive (n = 31) healthcare workers (HCWs) and 38 convalescent plasma samples from patients who had been admitted with acute respiratory distress (ARDS) associated with COVID-19. Patients recovering from COVID-19 ARDS had significantly higher mass AGE-glycated HSA and higher mass IgG3 levels. This would indicate that increased levels and/or ratios of hyper-glycosylation (probably terminal sialic acid) IgG3 and AGE glycated HSA may be predisposition markers for the development of COVID-19 ARDS as a result of SARS-CoV2 infection. Furthermore, rapid direct analysis of serum/plasma samples by MALDI-ToF for such humoral immune correlates of COVID-19 presents a feasible screening technology for the most at risk; regardless of age or known health conditions. Full article
(This article belongs to the Section Pathology and Molecular Diagnostics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mass spectra profile, 40,000 <span class="html-italic">m</span>/<span class="html-italic">z</span> to 85,000 <span class="html-italic">m</span>/<span class="html-italic">z</span>, of human reduced plasma samples (treated with Tris(2-carboxyethyl)phosphine (TCEP)) in order to reveal immunoglobulin heavy chains (IgM Hc ‘75,000 <span class="html-italic">m</span>/<span class="html-italic">z</span>, IgA ~56,000 <span class="html-italic">m</span>/<span class="html-italic">z</span> not found, IgG3 54,000 <span class="html-italic">m</span>/<span class="html-italic">z</span> and unconfirmed IgX, thought to be IgG4 Hc, at 48,000 <span class="html-italic">m</span>/<span class="html-italic">z</span>. HSA resolved at 66,4000+ <span class="html-italic">m</span>/<span class="html-italic">z</span> and transferrin at 79,000 <span class="html-italic">m</span>/<span class="html-italic">z</span>.</p>
Full article ">Figure 2
<p>Box and whisker plots of relative intensitities and variance in peak apex molecular mass of IgG1 heavy chains (IgG1 Hc), IgG3 heavy chains (IgG3 Hc) and human serum albumin (HSA) for the different sample groups: Blue represents data from SARS-CoV-2 seronegative HCWs, Orange from SARS-CoV-2 seropositive HCWs having recovered from COVID-19 with mild symptoms and Red sample data from convalescent patients recovering from COVID-19 ARDS. The table to the right is numeric data from the plots detailing the mean and median values of peak intensity (arbitrary units (AU)) and molecular mass (<span class="html-italic">m</span>/<span class="html-italic">z</span>) of IgG1 Hc, IgG3 Hc and HSA for the respective groups.</p>
Full article ">Figure 3
<p>Diagrammatic representation of IgG1 (<b>left</b>) and IgG3 (<b>right</b>) illustrating the differences in heavy chain structure with special reference to the larger neck domain and the O-linked glycosylation sites found there. Affinity binding: –, none; +,weak; ++, strong; and +++, very strong.</p>
Full article ">Figure 4
<p>Spike protein sialic acid glycan binding (S1 subunit NTD-domain) role in (<b>A</b>) mucosal epithelial attachment and infection of respiratory cells and (<b>B</b>) potential role in binding IgG3 at the neck region and AGE glycated HSA, during viremia, thereby aiding immune evasion and deposition of reactive complex that may give rise to vascular pathologies.</p>
Full article ">
15 pages, 5185 KiB  
Article
Engineering of a Long-Acting Bone Morphogenetic Protein-7 by Fusion with Albumin for the Treatment of Renal Injury
by Mei Takano, Shota Toda, Hiroshi Watanabe, Rui Fujimura, Kento Nishida, Jing Bi, Yuki Minayoshi, Masako Miyahisa, Hitoshi Maeda and Toru Maruyama
Pharmaceutics 2022, 14(7), 1334; https://doi.org/10.3390/pharmaceutics14071334 - 24 Jun 2022
Cited by 1 | Viewed by 1935
Abstract
The bone morphogenetic protein-7 (BMP7) is capable of inhibiting TGF-β/Smad3 signaling, which subsequently results in protecting the kidney from renal fibrosis, but its lower blood retention and osteogenic activity are bottlenecks for its clinical application. We report herein on the fusion of carbohydrate-deficient [...] Read more.
The bone morphogenetic protein-7 (BMP7) is capable of inhibiting TGF-β/Smad3 signaling, which subsequently results in protecting the kidney from renal fibrosis, but its lower blood retention and osteogenic activity are bottlenecks for its clinical application. We report herein on the fusion of carbohydrate-deficient human BMP7 and human serum albumin (HSA-BMP7) using albumin fusion technology and site-directed mutagenesis. When using mouse myoblast cells, no osteogenesis was observed in the glycosylated BMP7 derived from Chinese hamster ovary cells in the case of unglycosylated BMP7 derived from Escherichia coli and HSA-BMP7. On the contrary, the specific activity for the Smad1/5/8 phosphorylation of HSA-BMP7 was about 25~50-times lower than that for the glycosylated BMP7, but the phosphorylation activity of the HSA-BMP7 was retained. A pharmacokinetic profile showed that the plasma half-life of HSA-BMP7 was similar to that for HSA and was nearly 10 times longer than that of BMP7. In unilateral ureteral obstruction mice, weekly dosing of HSA-BMP7 significantly attenuated renal fibrosis, but the individual components, i.e., HSA or BMP7, did not. HSA-BMP7 also attenuated a cisplatin-induced acute kidney dysfunction model. The findings reported herein indicate that HSA-BMP7 has the potential for use in clinical applications for the treatment of renal injuries. Full article
(This article belongs to the Section Drug Delivery and Controlled Release)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of the HSA-BMP7 fusion protein. Structural diagram (image view) of the HSA-BMP7 fusion protein containing 3 engineered point mutations at positions 10, 29 and 80 (N10Q, N29Q and N80Q) at which wild-type human BMP7 has N-glycosylated sites. (GGGGSGGGGS) indicates the polyglycine and serine linker (<b>A</b>). Reduced SDS-PAGE of HSA-BMP7 and Western blot analysis of HSA-BMP7 (<b>B</b>). The transferred PVDF membranes were incubated with primary antibodies against HSA or human BMP7. Equal amount of proteins (1 μg/lane) were electrophoresed. Far-UV circular dichroism spectra of HSA and HSA-BMP7 (<b>C</b>). The protein concentration was 2.5 μM in PBS (pH 7.4).</p>
Full article ">Figure 2
<p>In vitro activity of HSA-BMP7 relative to native glycosylated BMP7 derived from CHO cells. Effect of HSA-BMP7 on the phosphorylation of Smad1/5/8 in HK-2 cells (<b>A</b>). HK-2 cells were incubated in 6-well plates in K-SFM medium containing 5 ng/mL of human recombinant EGF and 0.05 μg/mL of bovine pituitary extract at 37 °C for 24 h, and then incubated with BMP7 (10 nM) or HSA-BMP7 (250 nM or 500 nM) for 60 min in K-SFM. Phosphorylation of Smad1/5/8 protein HK-2 cells was determined by Western blot. Representative Western blot bands and their semi-quantitative data are shown. Results are the means ± SD (<span class="html-italic">n</span> = 3). Effect of HSA-BMP7 on the TGF-β stimulated mRNA expression of α-SMA in HK-2 cells (<b>B</b>). α-SMA mRNA in HK-2 cells was determined by quantitative RT-PCR in the presence of 3 ng/mL TGF-β with or without BMP7 or HSA-BMP7 for 48 h. Results are the means ± SD (<span class="html-italic">n</span> = 3–10). * <span class="html-italic">p</span> &lt; 0.05, compared with control (untreated). # <span class="html-italic">p</span> &lt; 0.05, compared with TGF-β treatment. Effect of HSA-BMP7 on alkaline phosphatase (ALP) activity in C2C12 cells (<b>C</b>). C2C12 cells were incubated in 24-well plates in DMEM with 10% heat inactivated FBS at 37 °C for 24 h. Glycosylated BMP7 derived from CHO cells, unglycosylated BMP7 derived from <span class="html-italic">E. coli</span>. or HSA-BMP7 was added to the DMEM with 5% FBS. The cells were incubated for 9 days. The results are the means ± SD (<span class="html-italic">n</span> = 3–6). * <span class="html-italic">p</span> &lt; 0.05, compared with control (untreated). # <span class="html-italic">p</span> &lt; 0.05, compared with 30 nM BMP7 (CHO).</p>
Full article ">Figure 3
<p>Plasma level of <sup>125</sup>I-labeled HSA-BMP7, HSA or BMP7 after intravenous administration to mice. <sup>125</sup>I-labeled proteins were injected through the tail vein of the mice. Results are the means ± SD (<span class="html-italic">n</span> = 3–4).</p>
Full article ">Figure 4
<p>Effect of HSA-BMP7 on renal fibrosis in obstructive nephropathy. Experimental protocol for studying the effect of HSA-BMP7 against unilateral ureteral obstruction (UUO) induced renal fibrosis in mice (<b>A</b>). ICR mice underwent unilateral ligation. Two ligatures, 5 mm apart, were placed in the upper two-thirds of the ureter. Saline, HSA-BMP7, HSA or BMP7 (100 nmol/kg) was administered intravenously just after and 7 days after UUO. Masson trichrome staining of the kidneys of UUO mice with saline, HSA-BMP7, HSA or BMP7 treatment (100 nmol/kg) (<b>B</b>). Representative photomicrographs of Masson trichrome stained kidney section of healthy (non-obstructed kidney) and obstructed kidney. Results are the mean ± SD (<span class="html-italic">n</span> = 4–6). ** <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. ## <span class="html-italic">p</span> &lt; 0.05, compared with UUO with saline treatment. Immunostaining of α-SMA of the kidney of UUO mice with saline or HSA-BMP7 treatment (100 nmol/kg) (<b>C</b>). Representative photomicrographs of α-SMA-stained kidney sections of healthy (non-obstructed kidney) and obstructed kidney. Results are the mean ± SD (<span class="html-italic">n</span> = 4–6). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. # <span class="html-italic">p</span> &lt; 0.05, compared with UUO with saline treatment. Effect of HSA-BMP7 on hydroxyproline levels in kidney of UUO mice (<b>D</b>). Results are the means ± SD (<span class="html-italic">n</span> = 5–7). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. # <span class="html-italic">p</span> &lt; 0.05, compared with UUO with a saline treatment. mRNA expression of Col1a2, α-SMA or TGF-β in kidney of UUO mice with a saline or HSA-BMP7 treatment (<b>E</b>). Col1A2, α-SMA or TGF-β mRNA was determined by quantitative RT-PCR. Results are the mean ± SD (<span class="html-italic">n</span> = 4–7). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. NS, not significant difference.</p>
Full article ">Figure 4 Cont.
<p>Effect of HSA-BMP7 on renal fibrosis in obstructive nephropathy. Experimental protocol for studying the effect of HSA-BMP7 against unilateral ureteral obstruction (UUO) induced renal fibrosis in mice (<b>A</b>). ICR mice underwent unilateral ligation. Two ligatures, 5 mm apart, were placed in the upper two-thirds of the ureter. Saline, HSA-BMP7, HSA or BMP7 (100 nmol/kg) was administered intravenously just after and 7 days after UUO. Masson trichrome staining of the kidneys of UUO mice with saline, HSA-BMP7, HSA or BMP7 treatment (100 nmol/kg) (<b>B</b>). Representative photomicrographs of Masson trichrome stained kidney section of healthy (non-obstructed kidney) and obstructed kidney. Results are the mean ± SD (<span class="html-italic">n</span> = 4–6). ** <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. ## <span class="html-italic">p</span> &lt; 0.05, compared with UUO with saline treatment. Immunostaining of α-SMA of the kidney of UUO mice with saline or HSA-BMP7 treatment (100 nmol/kg) (<b>C</b>). Representative photomicrographs of α-SMA-stained kidney sections of healthy (non-obstructed kidney) and obstructed kidney. Results are the mean ± SD (<span class="html-italic">n</span> = 4–6). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. # <span class="html-italic">p</span> &lt; 0.05, compared with UUO with saline treatment. Effect of HSA-BMP7 on hydroxyproline levels in kidney of UUO mice (<b>D</b>). Results are the means ± SD (<span class="html-italic">n</span> = 5–7). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. # <span class="html-italic">p</span> &lt; 0.05, compared with UUO with a saline treatment. mRNA expression of Col1a2, α-SMA or TGF-β in kidney of UUO mice with a saline or HSA-BMP7 treatment (<b>E</b>). Col1A2, α-SMA or TGF-β mRNA was determined by quantitative RT-PCR. Results are the mean ± SD (<span class="html-italic">n</span> = 4–7). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. NS, not significant difference.</p>
Full article ">Figure 5
<p>Effect of HSA-BMP7 on cisplatin-induced nephropathy. Experimental protocol for studying the effect of HSA-BMP7 against cisplatin-induced nephropathy mice (<b>A</b>). Saline or HSA-BMP7 (200 nmol/kg) was administered intravenously at 30 min before administering an intraperitoneal injection of cisplatin (15 mg/kg). Changes in the levels of blood urea nitrogen (BUN), serum creatinine (SCr) and creatinine clearance (CCr) after an intraperitoneal injection of cisplatin (15 mg/kg) (<b>B</b>). Results are the means ± SD (<span class="html-italic">n</span> = 4–7). * <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice. Histological assessment and TUNEL staining of the kidney of cisplatin-treated mice (<b>C</b>,<b>D</b>). Representative photomicrograghs of PAS and TUNEL stained kidney sections. Original magnifications: ×800 (<b>C</b>); ×400 (<b>D</b>). Scale bars represent 100 μm. The injured areas are indicated by arrows. The semi-quantitative analyses were also performed. Results are the mean ± SD (<span class="html-italic">n</span> = 4–7). ** <span class="html-italic">p</span> &lt; 0.05, compared with healthy mice.</p>
Full article ">
11 pages, 875 KiB  
Article
Glycated Albumin and Glycated Albumin/HbA1c Predict the Progression of Coronavirus Disease 2019 from Mild to Severe Disease in Korean Patients with Type 2 Diabetes
by Jeongseon Yoo, Youngah Choi, Shin Ae Park, Ji Yeon Seo, Chul Woo Ahn and Jaehyun Han
J. Clin. Med. 2022, 11(9), 2327; https://doi.org/10.3390/jcm11092327 - 21 Apr 2022
Cited by 3 | Viewed by 2000
Abstract
Hyperglycemia is among the main risk factors for severe COVID-19. We evaluated the association of glycated albumin (GA) and GA/HbA1c ratio with progression of COVID-19 from mild to severe disease in patients with type 2 diabetes mellitus (T2DM). Our retrospective study included 129 [...] Read more.
Hyperglycemia is among the main risk factors for severe COVID-19. We evaluated the association of glycated albumin (GA) and GA/HbA1c ratio with progression of COVID-19 from mild to severe disease in patients with type 2 diabetes mellitus (T2DM). Our retrospective study included 129 patients aged over 18 years with COVID-19 and T2DM who did not have any need of oxygen supplement. Of these, 59 patients whose COVID-19 was aggravated and required oxygen supplementation eventually were classified as having severe disease. Clinical and laboratory data were compared between mild and severe cases. The median of GA (18.4% vs. 20.95%, p = 0.0013) and GA/HbA1c (2.55 vs. 2.68, p = 0.0145) were higher in severe disease than in mild disease and positively correlated with C-reactive protein (Kendal Tau coefficient 0.200 and 0.126, respectively; all p < 0.05). Multiple logistic regression analysis showed that GA (odds ratio (OR), 1.151; 95% confidence interval (CI), 1.024–1.294) and GA/HbA1c (OR, 8.330; 95% CI, 1.786–38.842) increased the risk of severe disease. Patients with GA 20% or higher were 4.03 times more likely to progress from mild to severe disease. GA and GA/HbA1c ratio predicted progression of COVID-19 from mild to severe disease in patients with T2DM. Full article
(This article belongs to the Topic Infectious Diseases)
Show Figures

Figure 1

Figure 1
<p>Kendall Tau correlations of glucose control status and inflammatory markers with d-dimer. FBS, fasting blood sugar; GA, glycated albumin; CRP, C-reactive protein.</p>
Full article ">
12 pages, 1694 KiB  
Article
Human Serum Extracellular Vesicle Proteomic Profile Depends on the Enrichment Method Employed
by Mikel Azkargorta, Ibon Iloro, Iraide Escobes, Diana Cabrera, Juan M. Falcon-Perez, Felix Elortza and Felix Royo
Int. J. Mol. Sci. 2021, 22(20), 11144; https://doi.org/10.3390/ijms222011144 - 15 Oct 2021
Cited by 5 | Viewed by 1991
Abstract
The proteomic profiling of serum samples supposes a challenge due to the large abundance of a few blood proteins in comparison with other circulating proteins coming from different tissues and cells. Although the sensitivity of protein detection has increased enormously in the last [...] Read more.
The proteomic profiling of serum samples supposes a challenge due to the large abundance of a few blood proteins in comparison with other circulating proteins coming from different tissues and cells. Although the sensitivity of protein detection has increased enormously in the last years, specific strategies are still required to enrich less abundant proteins and get rid of abundant proteins such as albumin, lipoproteins, and immunoglobulins. One of the alternatives that has become more promising is to characterize circulating extracellular vesicles from serum samples that have great interest in biomedicine. In the present work, we enriched the extracellular vesicles fraction from human serum by applying different techniques, including ultracentrifugation, size-exclusion chromatography, and two commercial precipitation methods based on different mechanisms of action. To improve the performance and efficacy of the techniques to promote purity of the preparations, we have employed a small volume of serum samples (<100 mL). The comparative proteomic profiling of the enriched preparations shows that ultracentrifugation procedure yielded a larger and completely different set of proteins than other techniques, including mitochondrial and ribosome related proteins. The results showed that size exclusion chromatography carries over lipoprotein associated proteins, while a polymer-based precipitation kit has more affinity for proteins associated with granules of platelets. The precipitation kit that targets glycosylation molecules enriches differentially protein harboring glycosylation sites, including immunoglobulins and proteins of the membrane attack complex. Full article
(This article belongs to the Special Issue Extracellular Vesicles as a New Source of Liquid Biopsy 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the EV separation design. Each serum was diluted (the 100 µL loaded for each test contains 80 µL of undiluted serum), cleared and subsequently divided in four aliquots processed by four different techniques, ultracentrifugation (UC), INV (Total Exosome Purification kit, Invitrogen, Thermo Fisher Scientific, Waltham, MA, USA), GAG (Exo-GAG precipitation solution, Nasabiotech, A Coruña, Spain) and SEC (size exclusion chromatography). The results presented in this study were obtained from four independent biological replicates.</p>
Full article ">Figure 2
<p>Western blot analysis of human EV-enriched serum preparations by using different procedures (UC, ultracentrifugation, SEC3, fractions 2–4 from SEC, SEC8 fractions 7–9 from SEC, INV, Total isolation solution from Invitrogen, Invitrogen, Thermo Fisher Scientific, Waltham, MA, USA, GAG, Exo-GAG precipitation solution from Nasabiotech, A Coruña, Spain). * In the case of the sample INV, it was not possible to assay the sample against IgG, due to overload of protein, as shown in the small picture of the gel at the right corner.</p>
Full article ">Figure 3
<p>Differences in the proteomic profile enrichment achieved by each technique. (<b>A</b>) Unsupervised heatmap of normalized intensity values for each technique and identified protein. The color intensity correlates with the value of normalized abundance. (<b>B</b>) Proteins significatively enriched for each technique (sense of the comparison y axe vs. x axe) according to ANOVA and Tukey’s post hoc analysis (significance is considered for <span class="html-italic">p</span> value &lt; 0.05, <span class="html-italic">n</span> = 4). The color intensity correlates with the number of molecules enriched.</p>
Full article ">Figure 4
<p>(<b>A</b>) The upper graph shows the number of proteins annotated in UNIPROT as harboring glycosylation sites (in black) within the total number of proteins detected for each technique (at least two peptides should be detected in 3 biological replicates). (<b>B</b>) The graph shows the number of proteins with glycosylated sites within the number of enriched or underrepresented proteins for each technique compared with the rest or between SEC3 and INV.</p>
Full article ">Figure 5
<p>Summary of the cell component analysis associated with the proteins overrepresented in the preparations obtained with different isolation techniques.</p>
Full article ">
12 pages, 2623 KiB  
Article
How Does Glycation Affect Binding Parameters of the Albumin-Gliclazide System in the Presence of Drugs Commonly Used in Diabetes? In Vitro Spectroscopic Study
by Katarzyna Wiglusz, Ewa Żurawska-Płaksej, Anna Rorbach-Dolata and Agnieszka Piwowar
Molecules 2021, 26(13), 3869; https://doi.org/10.3390/molecules26133869 - 24 Jun 2021
Cited by 5 | Viewed by 2371
Abstract
In this research, the selected drugs commonly used in diabetes and its comorbidities (gliclazide, cilazapril, atorvastatin, and acetylsalicylic acid) were studied for their interactions with bovine serum albumin—native and glycated. Two different spectroscopic methods, fluorescence quenching and circular dichroism, were utilized to elucidate [...] Read more.
In this research, the selected drugs commonly used in diabetes and its comorbidities (gliclazide, cilazapril, atorvastatin, and acetylsalicylic acid) were studied for their interactions with bovine serum albumin—native and glycated. Two different spectroscopic methods, fluorescence quenching and circular dichroism, were utilized to elucidate the binding interactions of the investigational drugs. The glycation process was induced in BSA by glucose and was confirmed by the presence of advanced glycosylation end products (AGEs). The interaction between albumin and gliclazide, with the presence of another drug, was confirmed by calculation of association constants (0.11–1.07 × 104 M−1). The nature of changes in the secondary structure of a protein depends on the drug used and the degree of glycation. Therefore, these interactions may have an influence on pharmacokinetic parameters. Full article
(This article belongs to the Section Bioorganic Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of gliclazide (<b>a</b>), cilazapril (<b>b</b>), atorvastatin (<b>c</b>) and acetylsalicylic acid (<b>d</b>).</p>
Full article ">Figure 2
<p>Emission spectra of nonglycated and differently glycated BSA at excitation wavelength 335 nm (<b>a</b>) and 370 nm (<b>b</b>). Relative increase in fluorescence intensity is shown in the inserts. Native BSA (black lines and bars), g10_BSA (blue lines and bars) and g30_BSA (red lines and bars) signify the use of 0, 10 and 30 mM glucose for glycation.</p>
Full article ">Figure 3
<p>Fluorescence emission spectra of native BSA, g10_BSA and g30_BSA in the presence of cilazapril (<b>A</b>), atorvastatin (<b>B</b>) and acetylsalicylic acid (<b>C</b>) recorded at λ<sub>ex</sub> 295 nm.</p>
Full article ">Figure 4
<p>Changes in fluorescence intensity of BSA solutions (nonglycated and glycated at different degrees) after the addition of 14 μM gliclazide (molar ratio 1:7) in the presence of another drug observed at 295 nm excitation wavelength. Fluorescence is normalized to the value observed for the BSA-DRUG (cilazapril, atorvastatin or acetylsalicylic acid) complex and expressed as percentage. The colors of the bars refer to the following: native BSA—black, g10_BSA—blue, g30_BSA—red. Statistical significance was calculated using ANOVA, and <span class="html-italic">P</span> &lt; 0.05 was considered significant.</p>
Full article ">Figure 5
<p>The Stern-Volmer plots for the interaction of gliclazide with BSA in the presence of ci-lazapril (<b>A</b>), atorvastatin (<b>B</b>) and acetylsalicylic acid (<b>C</b>): native BSA (■), g10_BSA (●), g30_BSA (▲) at λ<sub>ex</sub> = 295 nm.</p>
Full article ">Figure 6
<p>Circular dichroism spectra of protein (2 μM): native BSA, g10_BSA, g30_BSA with CIL (8 μM) and GLICL (14 μM), pH 7.4.</p>
Full article ">Scheme 1
<p>Schematic work for the study of binding interactions of the albumin-gliclazide-drug system [<a href="#B46-molecules-26-03869" class="html-bibr">46</a>].</p>
Full article ">
9 pages, 654 KiB  
Article
Serum Levels of Fibroblast Growth Factor 21 Are Positively Associated with Aortic Stiffness in Patients with Type 2 Diabetes Mellitus
by Sin-Yi Huang, Du-An Wu, Jen-Pi Tsai and Bang-Gee Hsu
Int. J. Environ. Res. Public Health 2021, 18(7), 3434; https://doi.org/10.3390/ijerph18073434 - 26 Mar 2021
Cited by 3 | Viewed by 2023
Abstract
Aortic stiffness (AS), assessed using carotid–femoral pulse wave velocity (cfPWV), is associated with cardiovascular disease in type 2 diabetes mellitus (T2DM). The relationship between serum fibroblast growth factor 21 (FGF-21) and AS in T2DM patients was evaluated. Fasting serum FGF-21 levels of 130 [...] Read more.
Aortic stiffness (AS), assessed using carotid–femoral pulse wave velocity (cfPWV), is associated with cardiovascular disease in type 2 diabetes mellitus (T2DM). The relationship between serum fibroblast growth factor 21 (FGF-21) and AS in T2DM patients was evaluated. Fasting serum FGF-21 levels of 130 T2DM patients were measured using an enzyme immunoassay kit. A validated tonometry system was used to measure cfPWV (>10 m/s indicated AS). Of these T2DM patients, 34.6% were defined as the AS group. T2DM patients with AS were older; exhibited higher systolic blood pressure, diastolic blood pressure, and body fat mass; higher triglyceride, fasting glucose, glycosylated hemoglobin, and creatinine levels; higher urine albumin-to-creatinine ratios and serum FGF-21 levels; and lower estimated glomerular filtration rates. The FGF-21 level (odds ratio = 1.005, 95% confidence interval: 1.002–1.009, p = 0.002) as well as systolic blood pressure was an independent predictor of AS and positively correlated to cfPWV values (β = 0.369, p < 0.001) in T2DM patients. For T2DM patients, serum FGF-21 level could be a predictor for AS. Full article
(This article belongs to the Section Global Health)
Show Figures

Figure 1

Figure 1
<p>Receiver operating characteristic curve analysis of FGF-21 to predict aortic stiffness of T2DM patients. The area under the curve indicated the predictive power of FGF-21 for aortic stiffness of T2DM patients.</p>
Full article ">
19 pages, 5189 KiB  
Article
In Vitro Infection with Hepatitis B Virus Using Differentiated Human Serum Culture of Huh7.5-NTCP Cells without Requiring Dimethyl Sulfoxide
by Connie Le, Reshma Sirajee, Rineke Steenbergen, Michael A. Joyce, William R. Addison and D. Lorne Tyrrell
Viruses 2021, 13(1), 97; https://doi.org/10.3390/v13010097 - 12 Jan 2021
Cited by 9 | Viewed by 4159
Abstract
An estimated two billion people worldwide have been infected with hepatitis B virus (HBV). Despite the high infectivity of HBV in vivo, a lack of easily infectable in vitro culture systems hinders studies of HBV. Overexpression of the sodium taurocholate co-transporting polypeptide (NTCP) [...] Read more.
An estimated two billion people worldwide have been infected with hepatitis B virus (HBV). Despite the high infectivity of HBV in vivo, a lack of easily infectable in vitro culture systems hinders studies of HBV. Overexpression of the sodium taurocholate co-transporting polypeptide (NTCP) bile acid transporter in hepatoma cells improved infection efficiency. We report here a hepatoma cell culture system that does not require dimethyl sulfoxide (DMSO) for HBV infection. We overexpressed NTCP in Huh7.5 cells and allowed these cells to differentiate in a medium supplemented with human serum (HS) instead of fetal bovine serum (FBS). We show that human serum culture enhanced HBV infection in Huh7.5-NTCP cells, e.g., in HS cultures, HBV pgRNA levels were increased by as much as 200-fold in comparison with FBS cultures and 19-fold in comparison with FBS+DMSO cultures. Human serum culture increased levels of hepatocyte differentiation markers, such as albumin secretion, in Huh7.5-NTCP cells to similar levels found in primary human hepatocytes. N-glycosylation of NTCP induced by culture in human serum may contribute to viral entry. Our study demonstrates an in vitro HBV infection of Huh7.5-NTCP cells without the use of potentially toxic DMSO. Full article
(This article belongs to the Section Animal Viruses)
Show Figures

Figure 1

Figure 1
<p>Overexpression of NTCP in Huh7.5 cells. (<b>A</b>) Lentiviral-transduced puromycin-selected Huh7.5-NTCP cell line expressed more NTCP mRNA than the parental Huh7.5 cell line. RT-qPCR was used to measure NTCP and hypoxanthine-guanine phosphoribosyltransferase (HPRT) mRNA levels. Huh7.5-NTCP cells expressed more cell surface NTCP than parental Huh7.5 cells as illustrated with (<b>B</b>) flow cytometry and (<b>C</b>) immunofluorescence (IF) microscopy. Immunofluorescent staining of NTCP is shown in red, and the DAPI (4′,6-diamidino-2-phenylindole) stain of nuclei is shown in blue. Images show a single plane/z-stack. The scale bars are 10 μm. (<b>A</b>,<b>B</b>) Average values with error bars (±SD) derived from three experiments are plotted. Unpaired Student’s <span class="html-italic">t</span>-test was used for statistical analysis. * <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>Enhancement of HBV replication by human serum culture. Human serum culture increased HBV (<b>A</b>) pgRNA, (<b>B</b>) cccDNA, and (<b>C</b>) HBV surface antigen (HBsAg) levels from Huh7.5-NTCP cells. Huh7.5-NTCP cells were cultured in the media supplemented with FBS or HS and with or without the addition of DMSO during HBV infection. Samples were collected on day 14 (<b>A</b>,<b>B</b>) or day 7 (<b>C</b>) post-infection. Pregenomic RNA was measured using RT-qPCR from 10 ng of total RNA. Covalently closed circular DNA was quantified using q-PCR from 10 ng of gDNA. HBsAg was measured in a culture supernatant using enzyme-linked immunosorbent assay (ELISA). Average values with error bars (±SD) derived from three experiments are plotted. One-way analysis of variance (ANOVA) was used with the Bonferroni correction for multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Sustained infection by HBV in Huh7.5-NTCP cells cultured in human serum. Huh7.5-NTCP cells cultured in the medium supplemented with FBS or HS were infected with 500 genome equivalents per cell with or without 2% DMSO during infection. HBV pgRNA in the cells was repeatedly measured using RT-qPCR every 4–5 days for 50 days after HBV infection. Average values with error bars (±SD) derived from three experiments are plotted.</p>
Full article ">Figure 4
<p>Enhancement of HBV replication and expression of hepatocyte markers in Huh7.5-NTCP cells cultured in human serum. (<b>A</b>–<b>C</b>) Huh7.5-NTCP cells were cultured for various lengths of time in a medium supplemented with FBS or HS. Cells maintained in HS-supplemented media were infected after the indicated number of days in HS-containing media. During HBV infection, DMSO was either absent (−) or present (+). Samples were collected on day 14 post-infection for (<b>A</b>) RT-qPCR analysis of pgRNA or (<b>B</b>) nanoluciferase reporter luminescence analysis. (<b>A</b>,<b>B</b>) One-way analysis of variance (ANOVA) was used with the Bonferroni correction for multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Secreted human albumin concentration after 6 h and 24 h was determined using ELISA. Average values (±SD) derived from three experiments are plotted. Two-way analysis of variance (ANOVA) was used with the Bonferroni correction for multiple comparison test. Blue *, <span class="html-italic">p</span> &lt; 0.01 compared to FBS albumin secretion in 6 h. Black *, <span class="html-italic">p</span> &lt; 0.01 compared to FBS albumin secretion in 24 h.</p>
Full article ">Figure 5
<p>Reduction of HBV infection by MyrB, an entry inhibitor. MyrB was added to culture at 300 nM 30 min prior to infection and remained during HBV infection and one day post-infection. Cell monolayers and the culture supernatant were collected on day 7 post-infection for (<b>A</b>) RT-qPCR analysis of pgRNA and (<b>B</b>) ELISA of the surface antigen (HBsAg). Average values with error bars (±SD) derived from three experiments are plotted.</p>
Full article ">Figure 6
<p>Changes in NTCP mRNA levels and surface NTCP protein expression under various culture conditions. Huh7.5 or Huh7.5-NTCP cells were (<b>A</b>) not infected with the virus (mock), or (<b>B</b>) infected with HBV. Samples were collected on day 7 post-infection for RT-qPCR analyses of NTCP mRNA and HPRT mRNA levels. ΔΔCT values were calculated to determine fold changes in NTCP mRNA expression normalized to that of the Huh7.5 cells cultured in the medium containing FBS. Huh7.5-NTCP cells were analyzed with flow cytometry to assess (<b>C</b>) cell surface expression of NTCP based on median fluorescence intensity and (<b>D</b>) the percentage of cells expressing NTCP. Average values with error bars (±SD) derived from three independent experiments are plotted. One-way analysis of variance (ANOVA) was used with the Bonferroni correction for multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 7
<p>NTCP glycosylation and inhibition by tunicamycin. (<b>A</b>) Western blot analyses of NTCP glycosylation in Huh7.5-NTCP cells that were uninfected (mock) or infected with HBV. (<b>B</b>) Inhibition of N-glycosylation with tunicamycin suppressed HBV infection. Huh7.5-NTCP cells were incubated with 1 μg/mL tunicamycin for 2.5 h, followed by washing four times with PBS prior to infection. The cells were infected with nanoluciferase-expressing HBV (HBVNL) (MOI 500). Luminescence in relative light units (RLU) per well was measured to indicate nanoluciferase (NL) activity. Average values with error bars (±SD) derived from three experiments are plotted.</p>
Full article ">
10 pages, 656 KiB  
Article
Changed Profile of Serum Transferrin Isoforms in Primary Biliary Cholangitis
by Agnieszka Grytczuk, Alicja Bauer, Ewa Gruszewska, Bogdan Cylwik and Lech Chrostek
J. Clin. Med. 2020, 9(9), 2894; https://doi.org/10.3390/jcm9092894 - 8 Sep 2020
Cited by 5 | Viewed by 2020
Abstract
Liver damage affects the synthesis of proteins and glycoproteins, and alters their posttranslational modification, such as glycosylation changing the serum profile of glycoprotein isoforms. The retention of hydrophobic bile acids in the course of cholestatic liver diseases is a major cause of liver [...] Read more.
Liver damage affects the synthesis of proteins and glycoproteins, and alters their posttranslational modification, such as glycosylation changing the serum profile of glycoprotein isoforms. The retention of hydrophobic bile acids in the course of cholestatic liver diseases is a major cause of liver damage in primary biliary cholangitis (PBC). The study objective was to determine the serum profile of transferrin isoforms in primary biliary cholangitis and compare it to transferrin isoforms profile in extrahepatic cholestasis. The study was carried out in 76 patients with PBC and 40 healthy blood donors. Transferrin isoforms were analyzed by the capillary electrophoresis method. The mean relative concentrations of disialotransferrin and trisialotransferrin in PBC were significantly lower than those in the healthy subjects (p < 0.001, p = 0.011; respectively). None of the transferrin isoforms changed according to the disease severity evaluated by the Ludwig scoring system. However, the disease stage affected the activity of alkaline phosphatase (ALP) and γ-glutamyl transferase (GGT), and albumin level (p = 0.002; p = 0.013 and p = 0.005, respectively). Our results indicate that serum profile of transferrin isoforms alters primary biliary cholangitis and differs in comparison to transferrin isoforms profile in extrahepatic cholestasis. The decreased concentrations of lower sialylated isoforms of transferrin (low percentage share in total transferrin level) are not associated with the histological stage of disease. Full article
(This article belongs to the Special Issue Gallbladder and Biliary Tract Diseases)
Show Figures

Figure 1

Figure 1
<p>Total transferrin concentration (<b>top</b>) and bile acids (<b>bottom</b>) in PBC patients according to Ludwig’s scale. C—control group, 1–4—stages of liver injury according to Ludwig’s scale. Comparison between stages of Ludwig’s scoring system: ANOVA rank Kruskal–Wallis test (H, <span class="html-italic">p</span>).</p>
Full article ">Figure 2
<p>The relative concentration of 2-sialotransferrin (<b>top</b>) and 3-sialotransferrin (<b>bottom</b>) in PBC patients according to Ludwig’s scale. C—control group, 1–4—stages of liver injury according to Ludwig’s scale, SE—standard error. Comparison between stages of Ludwig’s scale: ANOVA rank Kruskal–Wallis test (H, <span class="html-italic">p</span>).</p>
Full article ">
Back to TopTop