[go: up one dir, main page]

 
 

Advances in Artificial and Biological Membranes, Volume II

A special issue of Membranes (ISSN 2077-0375). This special issue belongs to the section "Membrane Applications".

Deadline for manuscript submissions: closed (30 November 2023) | Viewed by 13988

Special Issue Editors


E-Mail Website
Guest Editor
1. Faculty of Materials Science and Ceramics, AGH University of Science and Technology, al. Mickiewicza 30, 30-059 Kraków, Poland
2. Johan Gadolin Process Chemistry Centre, c/o Centre for Process Analytical Chemistry and Sensor Technology (ProSens), Åbo Akademi University, Biskopsgatan 8, 20500 Åbo-Turku, Finland
Interests: sensor technology; electroanalysis; analytical/clinical chemistry; ion-selective electrodes; ion-sensor architectures; membranes and electroactive materials for ion-sensors
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
Department of Chemistry, Durham University, Lower Mountjoy, Stockton Road, Durham DH1 3LE, UK
Interests: electrochemistry; electrochemical sensors in real media; liquid-liquid interface; biosensors; biofilms

E-Mail Website
Guest Editor
Department of Biophysics, Warsaw University of Life Sciences – SGGW, 159 Nowoursynowska St., 02-776 Warsaw, Poland
Interests: biological membranes; ion channels and transporters; epithelial transport of ions and water; mitochondria bioelectrochemistry; sensors; microfluidic systems; mathematical modeling
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
1. School of Chemistry and Chemical Engineering, Guangzhou University, Guangzhou 510006, China
2. School of Chemical Engineering and Technology, Sun Yat-sen University, Guangzhou 510276, China
Interests: electrochemical sensors and materials; wearable devices
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues

Ion sensors, conventionally known as ion-selective membrane electrodes, were founded a hundred years ago through the invention of a pH electrode with a glass membrane (in 1906, Cremer, and in 1909, Haber and Klemensiewicz). This electrode type having a symmetric design with an internal contact created by a solution, paved the way to the present architectures of a large family of chemical sensors and biosensors.

Today, there are numerous sensors with acting membranes composed of glass, solid-state, plastics, and composites, as well many internal contacts composed of conducting polymers, carbon nanotubes, graphene, conducting clays, and composites. For this reason, the sensors can be miniaturized, created service-free, and produced by a mass fabrication technique such as 3D printing. Responses are now treated by models able to access time-and-space domains of the sensors. Supported by advanced modeling, ion sensors and biosensors are now deliberately calibration-free, may undergo automatic quality checks, and can act in ad hoc and routine applications such as theaters, hospitals, sports, water control, etc.

The heart of the sensor is always the same: an artificial or biological membrane or film able to develop a response to analyte due to selective, fast, and reversible processes at the sample–membrane interface or membrane/film bulk.

The aim of this Special Issue is to present retrospective outlooks as well as novel waves and the progress that has been recently realized in sensor technology, all aspects contributing to the successful advancement of the design, understanding, and application of ion sensors and biosensors being of interest, the submission site ready to receive your sensor science contributions.

Prof. Dr. Andrzej Lewenstam
Prof. Dr. Ritu Kataky
Prof. Dr. Krzysztof Dołowy
Prof. Dr. Li Niu
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Membranes is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2200 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • ion-sensor architectures
  • membranes and electroactive materials for ion-sensors
  • new fabrication schemes
  • response interpretation and modeling
  • new applications: wearable, disposable, remotely controlled ion-sensors
  • routine ion-sensors application

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Related Special Issue

Published Papers (15 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research, Review

5 pages, 201 KiB  
Editorial
Special Issue “Advances in Artificial and Biological Membranes: Mechanisms of Ionic Sensitivity, Ion-Sensor Designs, and Applications for Ion Measurement”
by Andrzej Lewenstam and Krzysztof Dołowy
Membranes 2020, 10(12), 427; https://doi.org/10.3390/membranes10120427 - 15 Dec 2020
Cited by 3 | Viewed by 1860
Abstract
Ion sensors, conventionally known as ion-selective membrane electrodes, were devised 100 years ago with the invention of a pH electrode with a glass membrane (in 1906 Cremer, in 1909 Haber and Klemensiewicz) [...] Full article

Research

Jump to: Editorial, Review

14 pages, 5523 KiB  
Article
Understanding Bidirectional Water Transport across Bronchial Epithelial Cell Monolayers: A Microfluidic Approach
by Miroslaw Zajac, Slawomir Jakiela and Krzysztof Dolowy
Membranes 2023, 13(12), 901; https://doi.org/10.3390/membranes13120901 - 6 Dec 2023
Viewed by 1834
Abstract
Deciphering the dynamics of water transport across bronchial epithelial cell monolayers is pivotal for unraveling respiratory physiology and pathology. In this study, we employ an advanced microfluidic system to explore bidirectional water transport across 16HBE14σ bronchial epithelial cells. Previous experiments unveiled electroneutral multiple [...] Read more.
Deciphering the dynamics of water transport across bronchial epithelial cell monolayers is pivotal for unraveling respiratory physiology and pathology. In this study, we employ an advanced microfluidic system to explore bidirectional water transport across 16HBE14σ bronchial epithelial cells. Previous experiments unveiled electroneutral multiple ion transport, with chloride ions utilizing transcellular pathways and sodium ions navigating both paracellular and transcellular routes. Unexpectedly, under isoosmotic conditions, rapid bidirectional movement of Na+ and Cl was observed, leading to the hypothesis of a substantial transport of isoosmotic solution (145 mM NaCl) across cell monolayers. To validate this conjecture, we introduce an innovative microfluidic device, offering a 500-fold sensitivity improvement in quantifying fluid flow. This system enables the direct measurement of minuscule fluid volumes traversing cell monolayers with unprecedented precision. Our results challenge conventional models, indicating a self-regulating mechanism governing water transport that involves the CFTR channel and anion exchangers. In healthy subjects, equilibrium is achieved at an apical potential of Δφap = −30 mV, while subjects with cystic fibrosis exhibit modulation by an anion exchanger, reaching equilibrium at [Cl] = 67 mM in the airway surface liquid. This nuanced electrochemical basis for bidirectional water transport in bronchial epithelia sheds light on physiological intricacies and introduces a novel perspective for understanding respiratory conditions. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) A comprehensive illustration of the measurement system is presented, with detailed descriptions provided in the legend. (<b>B</b>) The physical manifestation of the polycarbonate chamber, designed to house the epithelial cell insert, is depicted. (<b>C</b>) An actual representation of the microfluidic chip, featuring two T-arms designed for measuring the mobility of an oil droplet within a KHS fluid environment. (<b>D</b>) A graphical representation illustrating the relationship between oil-droplet mobility and length for three distinct capillary numbers. In this specific experiment, the continuous phase was represented by KHS1. (<b>E</b>) A graphical depiction of the dependence of oil-droplet mobility on capillary numbers <math display="inline"><semantics> <mrow> <mtext> </mtext> <mi>C</mi> <mi>a</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math> across different continuous phases: KHS1, KHS2, KHS3. (Mobility (β) is defined as the velocity of the droplet phase relative to the average velocity of the continuous phase.)</p>
Full article ">Figure 2
<p>The figure delineates water transport across the 16HBE14σ cell monolayer under the influence of chloride gradients. Panel (<b>A</b>) demonstrates water transport across cell monolayers with the apical chamber filled with a low-chloride solution (KHS3) and the basolateral chamber filled with a high-chloride solution (KHS1). In Panel (<b>B</b>), water transport occurs in the reversed chloride gradient, with the apical chamber filled with a high-chloride solution (KHS1) and the basolateral chamber filled with a low-chloride solution (KHS3). The data are presented as mean ± SD (<span class="html-italic">n</span> = 5). Positive values indicate basolateral to apical water flux, while negative values indicate apical to basolateral water flux.</p>
Full article ">Figure 3
<p>Water transport across 16HBE14σ cell monolayers was assessed under sodium gradients. (<b>A</b>) Water transport was examined in cell monolayers with the apical chamber filled with a low-sodium solution (KHS2), while the basolateral chamber contained a high-sodium solution (KHS1). (<b>B</b>) Conversely, water transport was studied in cell monolayers with a reversed sodium gradient. In this case, the apical chamber was filled with a high-sodium solution (KHS1), and the basolateral chamber contained a low-sodium solution (KHS2). The data are presented as mean ± SD (<span class="html-italic">n</span> = 4). Positive values indicate basolateral to apical water flux, while negative values denote apical to basolateral water flux.</p>
Full article ">Figure 4
<p>The study investigated the initial and final concentrations of sodium, chloride, choline, and gluconate ions on either side of the cell layer, as well as the direction of solution transport across epithelial cell monolayers. The experiments were conducted under the influence of different gradients: (<b>A</b>) basolateral-to-apical sodium gradient; (<b>B</b>) apical-to-basolateral sodium gradient; (<b>C</b>) basolateral-to-apical chloride gradient; (<b>D</b>) apical-to-basolateral chloride gradient. Final concentrations were determined in the 5th minute after the cessation of solution flow.</p>
Full article ">Figure 5
<p>Ionic transport direction in the CFTR channel relative to apical potential difference.</p>
Full article ">Figure 6
<p>Directional preference of the Cl/HCO<sub>3</sub> exchanger in response to varied chloride concentrations in the airway surface liquid (ASL) or medium.</p>
Full article ">
11 pages, 2904 KiB  
Article
Changes in Ion Transport across Biological Membranes Exposed to Particulate Matter
by Jakub Hoser, Adrianna Dabrowska, Miroslaw Zajac and Piotr Bednarczyk
Membranes 2023, 13(9), 763; https://doi.org/10.3390/membranes13090763 - 29 Aug 2023
Viewed by 1097
Abstract
The cells of living organisms are surrounded by the biological membranes that form a barrier between the internal and external environment of the cells. Cell membranes serve as barriers and gatekeepers. They protect cells against the entry of undesirable substances and are the [...] Read more.
The cells of living organisms are surrounded by the biological membranes that form a barrier between the internal and external environment of the cells. Cell membranes serve as barriers and gatekeepers. They protect cells against the entry of undesirable substances and are the first line of interaction with foreign particles. Therefore, it is very important to understand how substances such as particulate matter (PM) interact with cell membranes. To investigate the effect of PM on the electrical properties of biological membranes, a series of experiments using a black lipid membrane (BLM) technique were performed. L-α-Phosphatidylcholine from soybean (azolectin) was used to create lipid bilayers. PM samples of different diameters (<4 (SRM-PM4.0) and <10 μm (SRM-PM10) were purchased from The National Institute of Standards and Technology (USA) to ensure the repeatability of the measurements. Lipid membranes with incorporated gramicidin A (5 pg/mL) ion channels were used to investigate the effect of PM on ion transport. The ionic current passing through the azolectin membranes was measured in ionic gradients (50/150 mM KCl on cis/trans side). In parallel, the electric membrane capacitance measurements, analysis of the conductance and reversal potential were performed. Our results have shown that PM at concentration range from 10 to 150 μg/mL reduced the basal ionic current at negative potentials while increased it at positive ones, indicating the interaction between lipids forming the membrane and PM. Additionally, PM decreased the gramicidin A channel activity. At the same time, the amplitude of channel openings as well as single channel conductance and reversal potential remained unchanged. Lastly, particulate matter at a concentration of 150 μg/mL did not affect the electric membrane capacity to any significant extent. Understanding the interaction between PM and biological membranes could aid in the search for effective cytoprotective strategies. Perhaps, by the use of an artificial system, we will learn to support the consequences of PM-induced damage. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>Human bronchial epithelial cells were incubated in cell culture medium alone and in the presence of PM4.0 and PM10 for 24 h, 48 h and 72 h, respectively. Fragments of PM particles (&lt;4 μm and PM &lt; 10 μm) deposited on HBE cells are clearly visible. Blue arrows indicate healthy cell examples and black arrows indicate chosen particulate matter (PM). The pictures were taken with the DLTX1080PCMOSHDU2SD camera (DELTA optical) placed in inverted optical microscope (Olympus).</p>
Full article ">Figure 2
<p>Black lipid membrane technique. (<b>a</b>) Scheme of the system used in black lipid membrane (BLM) experiments including antivibration table, Faraday’s cage, chamber with <span class="html-italic">cis/trans</span> side, amplifier, converter and PC. Buffers were at pH = 7.2 and <span class="html-italic">trans</span> side was grounded. (<b>b</b>) Representative recording in 50/150 mM KCl (<span class="html-italic">cis/trans</span>) gradient before and after incorporation of gramicidin A (arrow) at 0 mV—indicates the closed state. (<b>c</b>) Multi-channel recording in 50/150 mM KCl (<span class="html-italic">cis/trans</span>) gradient after incorporation of gA. The graph shows the method of ionic current area (pA*s) calculation over 20 s through lipid membrane with incorporated gramicidin A ion channel.</p>
Full article ">Figure 3
<p>Effects of different PM concentrations on lipid membranes. (<b>a</b>) Registered current–time signals of ionic current flow through lipid bilayer membranes in control (0 μg/mL) and in the presence of 50, 100, 150 μg/mL PM4.0 and PM10 at a potential of +40 mV. (<b>b</b>,<b>c</b>) Analysis of the ionic current flow through lipid membrane at +40 and −40 mV without (control) and in the presence of PM4.0 and PM10 at concentrations of 10, 30, 50, 70, 100 and 150 μg/mL. The results are presented as mean ± SD, <span class="html-italic">n</span> = 5. Statistical significance was determined at <span class="html-italic">p</span> &lt; 0.001 (***) using one-way ANOVA.</p>
Full article ">Figure 4
<p>Electrical membrane capacitance changes in the presence of PM. Time-resolved effect of PM4.0 and PM10 (150 μg/mL) on electrical membrane capacitance changes. Results are presented as mean ± SD, <span class="html-italic">n</span> = 3. Statistical significance was determined at <span class="html-italic">p</span> &lt; 0.05 (*) using one-way ANOVA.</p>
Full article ">Figure 5
<p>Effects of PM on gramicidin A channels. (<b>a</b>) Registered signals of ionic current flow through azolectin bilayer membrane with incorporated gramicidin A (5 pg/mL) channels at −120 mV with and without presence 150 μg/mL of PM. (<b>b</b>,<b>c</b>) Effect of particulate matter on ionic current area through azolectin bilayer membrane with incorporated gramicidin A (5 pg/mL) channels with and without the presence of 150 μg/mL PM4.0 (<b>b</b>) and PM10 (<b>c</b>). (<b>d</b>,<b>e</b>) Effect of the particulate matter on ionic current amplitude of single gramicidin A (5 pg/mL) channel in the presence of PM4.0 (<b>d</b>) and PM10 (<b>e</b>) at a concentration of 150 μg/mL.</p>
Full article ">
19 pages, 10882 KiB  
Article
An Electrochemistry and Computational Study at an Electrified Liquid–Liquid Interface for Studying Beta-Amyloid Aggregation
by Bongiwe Silwane, Mark Wilson and Ritu Kataky
Membranes 2023, 13(6), 584; https://doi.org/10.3390/membranes13060584 - 5 Jun 2023
Viewed by 1445
Abstract
Amphiphilic peptides, such as Aß amyloids, can adsorb at an interface between two immiscible electrolyte solutions (ITIES). Based on previous work (vide infra), a hydrophilic/hydrophobic interface is used as a simple biomimetic system for studying drug interactions. The ITIES provides a 2D interface [...] Read more.
Amphiphilic peptides, such as Aß amyloids, can adsorb at an interface between two immiscible electrolyte solutions (ITIES). Based on previous work (vide infra), a hydrophilic/hydrophobic interface is used as a simple biomimetic system for studying drug interactions. The ITIES provides a 2D interface to study ion-transfer processes associated with aggregation, as a function of Galvani potential difference. Here, the aggregation/complexation behaviour of Aβ(1-42) is studied in the presence of Cu (II) ions, together with the effect of a multifunctional peptidomimetic inhibitor (P6). Cyclic and differential pulse voltammetry proved to be particularly sensitive to the detection of the complexation and aggregation of Aβ(1-42), enabling estimations of changes in lipophilicity upon binding to Cu (II) and P6. At a 1:1 ratio of Cu (II):Aβ(1-42), fresh samples showed a single DPV (Differential Pulse Voltammetry) peak half wave transfer potential (E1/2) at 0.40 V. Upon increasing the ratio of Cu (II) two-fold, fluctuations were observed in the DPVs, indicating aggregation. The approximate stoichiometry and binding properties of Aβ(1-42) during complexation with Cu (II) were determined by performing a differential pulse voltammetry (DPV) standard addition method, which showed two binding regimes. A pKa of 8.1 was estimated, with a Cu:Aβ1-42 ratio~1:1.7. Studies using molecular dynamics simulations of peptides at the ITIES show that Aβ(1-42) strands interact through the formation of β-sheet stabilised structures. In the absence of copper, binding/unbinding is dynamic, and interactions are relatively weak, leading to the observation of parallel and anti-parallel arrangements of β-sheet stabilised aggregates. In the presence of copper ions, strong binding occurs between a copper ion and histidine residues on two peptides. This provides a convenient geometry for inducing favourable interactions between folded β-sheet structures. Circular Dichroism spectroscopy (CD spectroscopy) was used to support the aggregation behaviour of the Aβ(1-42) peptides following the addition of Cu (II) and P6 to the aqueous phase. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Snapshot of the folded structure of Aβ<sub>1-42</sub> taken from the end of a 285 ns production run in water. (<b>b</b>) Protein secondary structure during the final 60 ns of the simulation, determined using the DSSP software [<a href="#B46-membranes-13-00584" class="html-bibr">46</a>,<a href="#B47-membranes-13-00584" class="html-bibr">47</a>].</p>
Full article ">Figure 2
<p>Snapshots from a 10 ns run showing the capture of the Aβ<sub>1-42</sub> peptide by a hydrophobic–hydrophilic interface, water molecules are shown in red (dots) and DCE molecules are shown in blue/gray (dots).</p>
Full article ">Figure 3
<p>Snapshots showing aggregation of Aβ<sub>1-42</sub> peptides at a DCE–water interface. Four chains are shown in cyan, red, lime green and orange together with protein secondary structure. The dark blue box outline indicates the 2D profile of the periodic box, which shows an average dimension of ~6.25 nm × 6.15 nm.</p>
Full article ">Figure 4
<p>Association of an Aβ<sub>1-42</sub> peptide dimer bound by copper at the dicholoroethane-water. The picture shows a snapshot form the end of a 1470.9 ns molecular dynamics run with a copper (II) ion linking two chains (as shown in <a href="#app1-membranes-13-00584" class="html-app">Figures S3b and S4 (see ESI)</a>). Individual chains are colour-coded in cyan and red, the copper ion is shown in yellow and the first residue of each chain is shown in white. The area of the interface is 9.95 nm × 9.94 nm.</p>
Full article ">Figure 5
<p>(<b>A</b>): CVs of (a) background solutions (b) with 0.01 M Aβ<sub>(1-42)</sub> added to the aqueous phase (c) with Cu(II)-Aβ<sub>(1-42)</sub>, 1:1 ratio in the aqueous phase, corresponding to the cells shown in <a href="#membranes-13-00584-sch002" class="html-scheme">Scheme 2</a>, Cell 1, 2 and 3, using a scan rate 0.1 V/s. (<b>B</b>): CV of Cu-Aβ and P6, 1:4 ratio, using a scan rate 0.1 V/s at ITIES after 10 m and after 24 h.</p>
Full article ">Figure 6
<p>Background, phosphate buffer, subtracted DPVs. (<b>A</b>): Solid line (Cell 2, <a href="#membranes-13-00584-sch002" class="html-scheme">Scheme 2</a>): Aβ<sub>1-42</sub>. Dotted line, CuAβ<sub>1-42</sub> 1:1 (Cell 3, <a href="#membranes-13-00584-sch002" class="html-scheme">Scheme 2</a>). Dashed line, CuAb<sub>1-42</sub>; P6 1:4 (Cell 4, <a href="#membranes-13-00584-sch002" class="html-scheme">Scheme 2</a>). (<b>B</b>): Effect of addition of 0.5 µM P6 to the 0.5 µM CuAβ<sub>1-42</sub> complex formed freshly. The signal starts decreasing when the ratio of P6 reaches 1:3. (<b>C</b>): Effect of excess Cu<sup>2+</sup>; (<b>a</b>): 0.02 μM Cu:0.01 μM Aβ<sub>(1-42)</sub> in a ratio 2:1 showing increasing current with time. The fluctuations in current indicate aggregation phenomenon at the liquid–liquid interface. (<b>b</b>): DPV obtained upon addition of 0.03 μM P6, showing a decrease in current with time, (<b>B</b>).</p>
Full article ">Figure 7
<p>Summary of the effect of time (0–60 min) on changes in current, at a potential of 0.46 V, observed at the electrified liquid–liquid interface for Aβ<sub>(1-42)</sub>, 1:1 Aβ<sub>(1-42)</sub> Cu complex and upon addition of P<sub>6</sub> to the complex at an Aβ<sub>(1-42)</sub> Cu:P6: 1:3 ratio.</p>
Full article ">Scheme 1
<p>(<b>A</b>) Structure of amyloid-beta 1-42 and sequence showing electroactive amino acids and metal binding sites. (<b>B</b>) Possible Cu-Aβ complexes that are reported in literature. (<b>C</b>) Structure of peptidomimetic peptides (P6), P6 binds to Cu<sup>2+</sup> through N-terminal glycine-histidine-lysine (GHK) groups and prevents its redox cycling in reducing conditions.</p>
Full article ">Scheme 1 Cont.
<p>(<b>A</b>) Structure of amyloid-beta 1-42 and sequence showing electroactive amino acids and metal binding sites. (<b>B</b>) Possible Cu-Aβ complexes that are reported in literature. (<b>C</b>) Structure of peptidomimetic peptides (P6), P6 binds to Cu<sup>2+</sup> through N-terminal glycine-histidine-lysine (GHK) groups and prevents its redox cycling in reducing conditions.</p>
Full article ">Scheme 2
<p>Configuration of the cells for the ITIES studies. x is the concentration ion in the aqueous phase. The double bar shows the polarised interface.</p>
Full article ">
12 pages, 1594 KiB  
Article
Directly Using Ti3C2Tx MXene for a Solid-Contact Potentiometric pH Sensor toward Wearable Sweat pH Monitoring
by Rongfeng Liang, Lijie Zhong, Yirong Zhang, Yitian Tang, Meixue Lai, Tingting Han, Wei Wang, Yu Bao, Yingming Ma, Shiyu Gan and Li Niu
Membranes 2023, 13(4), 376; https://doi.org/10.3390/membranes13040376 - 25 Mar 2023
Cited by 4 | Viewed by 2183
Abstract
The level of hydrogen ions in sweat is one of the most important physiological indexes for the health state of the human body. As a type of two-dimensional (2D) material, MXene has the advantages of superior electrical conductivity, a large surface area, and [...] Read more.
The level of hydrogen ions in sweat is one of the most important physiological indexes for the health state of the human body. As a type of two-dimensional (2D) material, MXene has the advantages of superior electrical conductivity, a large surface area, and rich functional groups on the surface. Herein, we report a type of Ti3C2Tx-based potentiometric pH sensor for wearable sweat pH analysis. The Ti3C2Tx was prepared by two etching methods, including a mild LiF/HCl mixture and HF solution, which was directly used as the pH-sensitive materials. Both etched Ti3C2Tx showed a typical lamellar structure and exhibited enhanced potentiometric pH responses compared with a pristine precursor of Ti3AlC2. The HF-Ti3C2Tx disclosed the sensitivities of −43.51 ± 0.53 mV pH–1 (pH 1–11) and −42.73 ± 0.61 mV pH–1 (pH 11–1). A series of electrochemical tests demonstrated that HF-Ti3C2Tx exhibited better analytical performances, including sensitivity, selectivity, and reversibility, owing to deep etching. The HF-Ti3C2Tx was thus further fabricated as a flexible potentiometric pH sensor by virtue of its 2D characteristic. Upon integrating with a solid-contact Ag/AgCl reference electrode, the flexible sensor realized real-time monitoring of pH level in human sweat. The result disclosed a relatively stable pH value of ~6.5 after perspiration, which was consistent with the ex situ sweat pH test. This work offers a type of MXene-based potentiometric pH sensor for wearable sweat pH monitoring. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>Preparation of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and morphologies. (<b>a</b>) Schematic illustration of the synthesis of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> by two etching methods; (<b>b</b>) XRD patterns for the Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. (<b>c</b>–<b>e</b>) SEM images of Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. (<b>f</b>) EDS mapping analysis of HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>.</p>
Full article ">Figure 2
<p>Element compositions and valence states for Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. (<b>a</b>) XPS survey spectra of Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. (<b>b</b>) XPS spectra of Ti 2p for Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. (<b>c</b>) XPS spectra of F 1s for Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>, and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>.</p>
Full article ">Figure 3
<p>Potentiometric pH responses. (<b>a</b>–<b>c</b>) Examination of pH reversible responses for Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>. All tests have been performed on six individual electrodes as shown in different colors (<span class="html-italic">n</span> = 6). (<b>d</b>–<b>f</b>) pH response calibration curves for the three types of pH electrodes. Corresponding sensitivities for the forward (pH = 1–11) and reverse (pH = 11–1) tests are shown in the Figures.</p>
Full article ">Figure 4
<p>Selectivity evaluation. (<b>a</b>–<b>c</b>) The selectivity examination of Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>, and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> by continually adding interfering ions. (<b>d</b>–<b>f</b>) The selectivity examination of Ti<sub>3</sub>AlC<sub>2</sub>, MILD-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>, and HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> by separation solution method. All potentiometric tests were performed on six individual electrodes (<span class="html-italic">n</span> = 6). The data represent the average values. Corresponding potentiometric response curves are shown in <a href="#app1-membranes-13-00376" class="html-app">Figures S2–S4</a>.</p>
Full article ">Figure 5
<p>Flexibility pH sensor for on-body sweat pH analysis. (<b>a</b>) The schematic fabrication of flexible pH sensor including PET substrate cleaning by O<sub>2</sub> plasma, Ag patterning by sputter, insulation layer deposition by PDMS, and WE/RE electrode modification by drop casting. The final fabricated flexible pH electrode is shown on the right side. (<b>b</b>) A photograph illustrates the on-body test of sweat pH monitoring during outdoor running. (<b>c</b>) The ion anti-interference test for the solid Ag/AgCl RE. (<b>d</b>) On-body test pH analysis by the device. (<b>e</b>) A comparison of sweat pH measured by HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>-based pH sensor and pH meter. (<b>f</b>) Calibration curves of the HF-Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>-based pH sensor before and after sweat test.</p>
Full article ">
11 pages, 1666 KiB  
Article
Interactions of Surfactants with Biomimetic Membranes—2. Generation of Electric Potential with Non-Ionic Surfactants
by Nikolai M. Kocherginsky and Brajendra K. Sharma
Membranes 2023, 13(3), 353; https://doi.org/10.3390/membranes13030353 - 18 Mar 2023
Cited by 1 | Viewed by 1715
Abstract
It is discovered that noncharged surfactants lead to electric effects that interact with biomimetic membranes made of nitrocellulose filters, which are impregnated with fatty acid esters. At a surfactant concentration as low as 64 microM in one of the solutions, they lead to [...] Read more.
It is discovered that noncharged surfactants lead to electric effects that interact with biomimetic membranes made of nitrocellulose filters, which are impregnated with fatty acid esters. At a surfactant concentration as low as 64 microM in one of the solutions, they lead to the transient formation of transmembrane electric potential. Maximum changes of this potential are proportional to the log of noncharged surfactant concentrations when it changes by three orders of magnitude. We explain this new and nontrivial effect in terms of an earlier suggested physicochemical mechanics approach and noncharged surfactants transient changes induced by membrane permeability for inorganic ions. It could be used to imitate the interactions of non-ionic drugs with biological membranes. The effect may also be used in determining the concentration of these surfactants and other non-ionic chemicals of concern, such as pharmaceuticals and personal care products. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>Structures of nonionic surfactants (<b>a</b>) sorbitan monolaurate Span-20, (<b>b</b>) other sorbitans, (<b>c</b>) polyethylene glycol sorbitan monolaurate, Tween 20, (<b>d</b>) Tergitol, and (<b>e</b>) Triton X-100.</p>
Full article ">Figure 2
<p>Transmembrane potential vs. time after the addition of 0.1 <span class="html-italic">w/v</span>% of different surfactants to the measuring (K<sup>+</sup> acceptor) chamber containing 5 mM buffer. The reference (K<sup>+</sup> donor) chamber contains 5 mM buffer + 0.5 M KCl.</p>
Full article ">Figure 3
<p>Transmembrane potential vs. time after the addition of Tergitol 15-S-7 in concentrations 16, 32, 64, 97, 130, 162, 800, and 3880 μM to the measuring (acceptor) chamber, containing the 5 mM buffer. The reference (donor) chamber contains 5 mM buffer + 0.5 M KCl.</p>
Full article ">Figure 4
<p>Transmembrane potential as a function of time when Tergitol 15-S-7 was added to the reference chamber. The measuring chamber contains a 5 mM buffer, while the reference chamber contains 5 mM buffer + 0.5 M KCl, pH = 4.6.</p>
Full article ">Figure 5
<p>Net changes of the minimum value of the transmembrane potential as a function of log C (µm) for different Tergitols: ● 15-S-5, ▲ 15-S-7 and ■ 15-S-9.</p>
Full article ">Figure 6
<p>Transmembrane potential as a function of time for impregnated nitrocellulose filters with pore size 0.05 μm and 0.45 μm.</p>
Full article ">Figure 7
<p>Time dependence of current through the membrane in response to 3.5 mM Tween 20. 0.45 μm nitrocellulose filter impregnated with isopropyl myristate separated by a 5 mM buffer+ 0.5 M KCl and a 5 mM buffer.</p>
Full article ">Figure 8
<p>The half-life time of potential relaxation kinetics vs. the thickness of the membrane with 0.05-micrometer pores after the addition of 0.1 wt% Tergitol 15-S-7.</p>
Full article ">
15 pages, 2205 KiB  
Article
Novel Experimental Setup for Coulometric Signal Transduction of Ion-Selective Electrodes
by Naela Delmo, Zekra Mousavi, Tomasz Sokalski and Johan Bobacka
Membranes 2022, 12(12), 1221; https://doi.org/10.3390/membranes12121221 - 2 Dec 2022
Cited by 4 | Viewed by 2120
Abstract
In this work, a novel and versatile experimental setup for coulometric signal transduction of ion-selective electrodes (ISEs) is introduced and studied. It is based on a constant potential coulometric measurement carried out using a one-compartment three-electrode electrochemical cell. In the setup, a potassium [...] Read more.
In this work, a novel and versatile experimental setup for coulometric signal transduction of ion-selective electrodes (ISEs) is introduced and studied. It is based on a constant potential coulometric measurement carried out using a one-compartment three-electrode electrochemical cell. In the setup, a potassium ion-selective electrode (K+- ISE) is connected as the reference electrode (RE). A poly(3,4-ethylenedioxythiophene) doped with polystyrene sulfonate (PEDOT:PSS)-based electrode with a dummy membrane (DM) and a glassy carbon (GC) rod are connected as the working electrode (WE) and counter electrode (CE), respectively. Adding a non-selective dummy membrane to the structure of the WE facilitates the regulation of the measured signal and response time. The results from electrochemical impedance spectroscopy measurements carried out on the WE showed that the time constant is profoundly influenced by the dummy membrane thickness. In addition, the redox capacitance of the PEDOT:PSS film shows a better correlation with the electrode area than the film thickness. Sequential addition/dilution experiments showed the improvement of current and cumulated charge signals in the new setup studied in this work compared to the setup used in the original coulometric signal transduction method. Both conventional ISEs and solid-contact ISEs (SCISEs) were used in this work. The results showed that the coulometric response was independent of the type of ISE used as RE, confirming the versatility of the novel set-up. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>Electrochemical impedance spectra of PEDOT:PSS-based (<b>a</b>,<b>b</b>) and MWCNTs-based (<b>c</b>,<b>d</b>) electrodes recorded in 10<sup>−1</sup> M KCl at OCP, in the frequency range of 100 kHz–10 mHz, and excitation amplitude of 10 mV.</p>
Full article ">Figure 2
<p>Electrochemical impedance spectra of GC/PEDOT:PSS electrodes with <span class="html-italic">Ø</span><sub>GC</sub> = 3 mm (<b>a</b>,<b>b</b>), <span class="html-italic">Ø</span><sub>GC</sub> = 5 mm (<b>c</b>,<b>d</b>), and different <span class="html-italic">Q</span><sub>polym</sub> (as indicated in the figures) recorded in 10<sup>−1</sup> M KCl at OCP in the frequency range of 100 kHz–10 mHz and excitation amplitude of 10 mV.</p>
Full article ">Figure 3
<p>(<b>a</b>) Comparison between the potentiometric response of different K<sup>+</sup>- ISEs: GC/PEDOT:PSS/50 µL K<sup>+</sup>- ISM (black line), GC/MWCNTs/50 µL K<sup>+</sup>- ISM (red line), and conventional K<sup>+</sup>- ISE (blue line), and GC/10 mC PEDOT:PSS/25 μL DM (green line) in 10<sup>−2</sup> M to 10<sup>−7</sup> M KCl solution with 10<sup>−1</sup> M NaCl as BGE. (<b>b</b>) Gas sensitivity of electrodes: GC/10 mC PEDOT:PSS (green line) and GC/10 mC PEDOT:PSS/25 μL DM (red line) carried out in 10<sup>−3</sup> M KCl solution.</p>
Full article ">Figure 4
<p>Chronoamperometric and chronocoulometric response collected using (<b>a</b>) the original setup (WE: GC/2 mC PEDOT:PSS/50 µL K<sup>+</sup>-ISM, RE: Ag/AgCl/3 M KCl/1 M LiOAc, CE: GC rod), and (<b>b</b>) the novel experimental setup (WE: GC/10 mC PEDOT:PSS/25 µL DM, RE: GC/2 mC PEDOT:PSS/50 µL K<sup>+</sup>-ISM, CE: GC rod) for the coulometric signal transduction. The addition/dilution steps correspond to a 5% increase/decrease of K<sup>+</sup> concentration in a starting solution of 10<sup>−3</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE.</p>
Full article ">Figure 5
<p>Chronoamperometric and chronocoulometric response in the addition/dilution experiments using working electrodes with different DM thicknesses: (<b>a</b>) 3 mm GC/10 mC PEDOT:PSS/25 µL DM and (<b>b</b>) 3 mm GC/10 mC PEDOT:PSS/100 µL DM. The addition/dilution steps correspond to a 5% increase/decrease of K<sup>+</sup> concentration in a starting solution of 10<sup>−3</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE.</p>
Full article ">Figure 6
<p>Chronoamperometric (<b>a</b>) and chronocoulometric (<b>b</b>) response during the addition/dilution experiments using different PEDOT:PSS-based WEs with equivalently thin dummy membrane and a GC/2 mC PEDOT:PSS/50 µL K<sup>+</sup>-ISM connected as the RE and GC rod as the CE. The addition/dilution steps correspond to a 5% increase/decrease in K<sup>+</sup> concentration in a starting solution of 10<sup>−3</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE.</p>
Full article ">Figure 7
<p>Chronoamperometric (<b>a</b>), chronocoulometric (<b>b</b>) response, cumulated charge <span class="html-italic">Q</span> vs. loga<sub>K+</sub> (<b>c</b>) during the calibration conducted in 10<sup>−2</sup> M to 10<sup>−3.4</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE with ∆log<span class="html-italic">C</span><sub>K+</sub> = 0.2 decade step<sup>−1</sup>; 3 mm GC/10 mC PEDOT:PSS/DM (red), 3 mm GC/20 mC PEDOT:PSS/DM (blue), 5 mm GC/27.8 mC PEDOT:PSS/DM (black), or 5 mm GC/55.5 mC PEDOT:PSS/DM (green) with equivalent dummy membrane thickness were used as WEs. GC/2 mC PEDOT:PSS/50 µL K<sup>+</sup>- ISM and GC rod were connected as the RE and CE, respectively. Slope values from results in (<b>c</b>) against the polymerization charge <span class="html-italic">Q</span><sub>polym</sub> of the PEDOT:PSS film in the WE (<b>d</b>).</p>
Full article ">Figure 8
<p>Cumulated charge <span class="html-italic">Q</span> vs. log <span class="html-italic">a</span><sub>K+</sub> from the calibration conducted in 10<sup>−2</sup> M to 10<sup>−3.4</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE with ∆log<span class="html-italic">C</span><sub>K+</sub> = 0.2 decade step<sup>−1</sup>; 3 mm GC/10 mC PEDOT:PSS/DM (red), 3 mm GC/20 mC PEDOT:PSS/DM (blue), 5 mm GC/27.8 mC PEDOT:PSS/DM (black), and 5 mm GC/55.5 mC PEDOT:PSS/DM (green) with equivalent dummy membrane thickness were used as WEs. Different K<sup>+</sup>- ISEs were used as REs, as indicated in the figure.</p>
Full article ">Figure 9
<p>Chronoamperometric (<b>a</b>) and chronocoulometric (<b>b</b>) response resulted from the addition/dilution experiments using a 3 mm GC/10 mC PEDOT:PSS/25 µL DM as the WE, GC/2 mC PEDOT:PSS/50 µL K<sup>+</sup>- ISM as RE, and GC rod as CE. The addition/dilution steps correspond to 5%, 2.5%, and 1% increase/decrease in K<sup>+</sup> concentration conducted in a starting solution of 10<sup>−3</sup> M KCl + 10<sup>−1</sup> M NaCl as BGE.</p>
Full article ">
15 pages, 2530 KiB  
Article
Electrically Enhanced Sensitivity (EES) of Ion-Selective Membrane Electrodes and Membrane-Based Ion Sensors
by Jan Migdalski and Andrzej Lewenstam
Membranes 2022, 12(8), 763; https://doi.org/10.3390/membranes12080763 - 3 Aug 2022
Cited by 2 | Viewed by 2224
Abstract
The use of external electronic enforcement in ion-sensor measurements is described. The objective is to improve the open-circuit (potentiometric) sensitivity of ion sensors. The sensitivity determines the precision of analyte determination and has been of interest since the beginning of ion-sensor technology. Owing [...] Read more.
The use of external electronic enforcement in ion-sensor measurements is described. The objective is to improve the open-circuit (potentiometric) sensitivity of ion sensors. The sensitivity determines the precision of analyte determination and has been of interest since the beginning of ion-sensor technology. Owing to the theoretical interpretation founded by W.E. Nernst, the sensitivity is characterized by the slope and numerically predicted. It is empirically determined and validated during calibration by measuring an electromotive force between the ion sensor and the reference electrode. In practice, this measurement is made with commercial potentiometers that function as unaltered “black boxes”. This report demonstrates that by gaining access to a meter’s electrical systems and allowing for versatile signal summations, the empirical slope can be increased favorably. To prove the validity of the approach presented, flow-through ion-sensor blocks used in routine measurements of blood electrolytes (Na+, K+, Li+, Cl) and multielectrode probes with flat surfaces, similar to those applied previously for monitoring transmembrane fluxes of Na+, K+, Cl through living biological cells, are used. Several options to serve real-life electroanalytical challenges, including linear calibration for sensors with high-resistance membranes, responses with non-Nernstian slopes, non-linear calibration, and discrimination of nonfunctional sensors, are shown. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Flow through set containing four ISE electrodes with plastic membranes and a REF electrode, all from KONE. (<b>b</b>) Multielectrode platform with an external diameter of 10 mm containing five Au/Ag disc substrates (0.5 mm diameter). The central electrode (indicated by an arrow) is a solid-contact reference electrode. It is surrounded by four SC ISs.</p>
Full article ">Figure 2
<p>Potential changes of three highly resistive sodium electrodes with glass membranes and their summed signals (<b>a</b>). Calibration curves of three highly resistive sodium electrodes with glass membranes and of their summed signals (<b>b</b>). Measurements were performed in 0.1–0.0001 M NaCl solutions.</p>
Full article ">Figure 3
<p>Responses of the lithium electrodes Li1, Li2, and Li3 and their summed signals recorded during LiCl addition to 140 mmol NaCl (<b>a</b>). The calibration curves of the lithium electrodes Li1, Li2, and Li3 and their summed signals (<b>b</b>). Potential changes were measured versus the flow-through KONE REF (electrode block used as shown in <a href="#membranes-12-00763-f001" class="html-fig">Figure 1</a>a).</p>
Full article ">Figure 4
<p>Potential changes of three potassium-sensitive electrodes (K1, K2, and K3) and the sums of their signals recorded during calibration in 0.1–0.00001 M KCl solutions (<b>a</b>). Calibration curves for three potassium-sensitive electrodes present in the multielectrode platform and the summed signals (<b>b</b>). The multielectrode platform shown in <a href="#membranes-12-00763-f001" class="html-fig">Figure 1</a>b was used.</p>
Full article ">Figure 5
<p>Potential changes of the flow-through sodium electrode connected simultaneously to three different ion-meter channels and their summed signals recorded during calibration in 0.1–0.000001 M NaCl solutions (<b>a</b>). Calibration curves of the flow-through sodium electrodes connected simultaneously to three different ion-meter channels and their summed signals (<b>b</b>).</p>
Full article ">Figure A1
<p>Scheme of the summing amplifier built with the Ultralow Offset Voltage Operational Amplifier OP08A (Analog Device) as well as its cooperation with a multichannel ion-meter and data acquisition card (DAC). The scheme was made with the use of free EasyEDA software.</p>
Full article ">Figure A2
<p>Potential changes of the chloride electrode connected simultaneously to three ion-meter channels with different gains equal to 1.0, 2.0, and 3.2, recorded during calibration in 0.1–0.000001 M KCl solutions (<b>a</b>). Calibration curves for chloride electrode connected simultaneously to three ion-meter channels with different gains equal to 1.0, 2.0, and 3.2 (<b>b</b>). Potential changes were measured versus the silver chloride electrode REF 251 (Metrohm).</p>
Full article ">
17 pages, 7686 KiB  
Article
Plasticized PVC Membrane Modified Electrodes: Voltammetry of Highly Hydrophobic Compounds
by Ernő Lindner, Marcin Guzinski, Bradford Pendley and Edward Chaum
Membranes 2020, 10(9), 202; https://doi.org/10.3390/membranes10090202 - 27 Aug 2020
Cited by 11 | Viewed by 4846
Abstract
In the last 50 years, plasticized polyvinyl chloride (PVC) membranes have gained unique importance in chemical sensor development. Originally, these membranes separated two solutions in conventional ion-selective electrodes. Later, the same membranes were applied over a variety of supporting electrodes and used in [...] Read more.
In the last 50 years, plasticized polyvinyl chloride (PVC) membranes have gained unique importance in chemical sensor development. Originally, these membranes separated two solutions in conventional ion-selective electrodes. Later, the same membranes were applied over a variety of supporting electrodes and used in both potentiometric and voltammetric measurements of ions and electrically charged molecules. The focus of this paper is to demonstrate the utility of the plasticized PVC membrane modified working electrode for the voltammetric measurement of highly lipophilic molecules. The plasticized PVC membrane prevents electrode fouling, extends the detection limit of the voltammetric methods to sub-micromolar concentrations, and minimizes interference by electrochemically active hydrophilic analytes. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the processes during the concentration (black arrows) and stripping (red arrows) of cations using a conductive polymer/plasticized PVC membrane modified electrode. The 18-crown-6 molecule in the PVC membrane represents any selective ionophore which may be used in this arrangement.</p>
Full article ">Figure 2
<p>Linear sweep voltammograms recorded with a bare and a membrane-coated electrode in 49.8 µmol/L ascorbic acid (AA) (<b>left</b>) and 117.6 µmol/L p-acetamino phenol (APAP) (<b>right</b>) solutions. The insets show the membrane-coated sensor signal with much higher resolution in the current scale. The PVC membrane composition is 25% PVC, 50% bis(2-ethylhexyl)sebacate (DOS), 22% tetradodecylammonium tetrakis(pentafuorophenyl)borate (TDDA-TPFPhB) and 3% sodium tetrakis[3,5bis(trifluoromethyl) phenyl] borate (NaTFPhB).</p>
Full article ">Figure 3
<p>Three different variations of the processes during voltammetric measurements of propofol with the PVC membrane modified electrode. In all three examples, propofol preferentially partitions into the membrane, and during its electrochemical oxidation, propofol positively charged intermediates are generated. The electroneutrality in the membrane is sustained through (<b>a</b>) the uptake of hydrophobic counterions, e.g., (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>l</mi> <msubsup> <mi>O</mi> <mn>4</mn> <mo>−</mo> </msubsup> </mrow> </semantics></math>) from the solution; (<b>b</b>) the release of positively charged counterions of the cation exchanger (e.g., K<sup>+</sup>) incorporated in the membrane; (<b>c</b>) the generation of negatively charged species on the counter electrode. In this scenario, both the working (WE) and counter (CE) electrodes are coated with the membrane.</p>
Full article ">Figure 4
<p>Chronoamperometric response of a PVC-membrane-coated glassy carbon (GC) electrode upon spiking the phosphate buffered saline (PBS) background electrolyte with propofol standard solutions. The concentrations in the PBS solution labeled 1 to 9 are 1.25, 2.5, 4.98, 9.9, 19.6, 38.4, 56.6, 80.5 and 111.1 μmol/L. The inset in the lower right is the calibration curve constructed from the chronoamperometric currents. The inset in the upper left corner shows linear sweep voltammograms recorded in PBS background solution and in PBS with propofol concentrations between 9.9 and 111.1 μmol/L. Membrane composition 25.5% PVC, 50.9% 2-nitrophenyl octyl ether (o-NPOE), 21.2% tetradodecylammonium tetrakis(pentafluorophenyl) borate (TDDATPFPhB), 2.4% sodium tetrakis[3,5bis(trifluoromethyl) phenyl] borate (NaTFPhB).</p>
Full article ">Figure 5
<p>Linear sweep voltammograms recorded in sertraline (<b>upper left</b>), amitriptyline (<b>upper right</b>) and aripiprazone (<b>middle left</b>), sirolimus (<b>middle right</b>), everolimus (<b>lower left</b>) and citalopram (<b>lower right</b>) solutions. The concentration ranges are indicated in the individual pictures.</p>
Full article ">Figure 6
<p>Chronoamperometric transients recorded with the plasticized PVC membrane-coated GC electrode in PBS solution of different sertraline concentrations. Applied potential: 1.0 V.</p>
Full article ">Figure 7
<p>The reproducibility of the chronoamperometric measurements in a continuous flow mode of measurement of amitriptyline with the PVC membrane-coated GC electrode in PBS solution. Applied potential: 0.95 V. The standard solutions are switched between 0 and 0.25 µmol/L and 0.25 and 0.5 µmol/L in the PBS background. The blue line represents the mean value of the current in 0.25 µmol/L amitriptyline and the red dotted lines represent ±1 standard deviation range of the data around the mean.</p>
Full article ">Figure 8
<p>Repeated LSVs of a bare (<b>left</b>) and PVC membrane-coated (<b>right</b>) GC electrode in 50 µmol/L (bare, <b>left</b>) and 10 µmol/L (membrane-coated, <b>right</b>) amitriptyline containing solutions.</p>
Full article ">Scheme 1
<p>Electrochemical notation of a potentiometric cell with a conventional, liquid inner contact ion-selective membrane electrode as indicator electrode and a double junction reference electrode.</p>
Full article ">Scheme 2
<p>Electrochemical notation of solid contact ion-selective electrodes. (<b>a</b>) Coated wire electrode; (<b>b</b>) Conductive polymer (CP)-based solid contact electrode.</p>
Full article ">Scheme 3
<p>Electrochemical notation of membrane modified voltammetric working electrodes. (<b>a</b>) Membrane directly deposited on the substrate electrode surface; (<b>b</b>) A conductive polymer (CP) layer is sandwiched between the substrate electrode and the membrane; (<b>c</b>) Multiple membrane layer modified electrode.</p>
Full article ">
18 pages, 3374 KiB  
Article
Optimization of Ruthenium Dioxide Solid Contact in Ion-Selective Electrodes
by Nikola Lenar, Beata Paczosa-Bator and Robert Piech
Membranes 2020, 10(8), 182; https://doi.org/10.3390/membranes10080182 - 9 Aug 2020
Cited by 20 | Viewed by 3040
Abstract
Ruthenium dioxide occurs in two morphologically varied structures: anhydrous and hydrous form; both of them were studied in the scope of this work and applied as mediation layers in ion-selective electrodes. The differences between the electrochemical properties of those two materials underlie their [...] Read more.
Ruthenium dioxide occurs in two morphologically varied structures: anhydrous and hydrous form; both of them were studied in the scope of this work and applied as mediation layers in ion-selective electrodes. The differences between the electrochemical properties of those two materials underlie their diverse structure and hydration properties, which was demonstrated in the paper. One of the main differences is the occurrence of structural water in RuO2•xH2O, which creates a large inner surface available for ion transport and was shown to be a favorable feature in the context of designing potentiometric sensors. Both materials were examined with SEM microscope, X-ray diffractometer, and contact angle microscope, and the results revealed that the hydrous form can be characterized as a porous structure with a smaller crystallite size and more hydrophobic properties contrary to the anhydrous form. Potentiometric and electrochemical tests carried out on designed GCD/RuO2/K+-ISM and GCD/RuO2•xH2O/K+-ISM electrodes proved that the loose porous microstructure with chemically bounded water, which is characteristic for the hydrous form, ensures the high electrical capacitance of electrodes (up to 1.2 mF) with consequently more stable potential (with the potential drift of 0.0015 mV/h) and a faster response (of a few seconds). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM scans of (<b>a</b>) hydrous RuO<sub>2</sub> and (<b>b</b>) anhydrous RuO<sub>2</sub>.</p>
Full article ">Figure 2
<p>X-ray diffractogram of (<b>a</b>) hydrous RuO<sub>2</sub> and (<b>b</b>) anhydrous RuO<sub>2</sub>.</p>
Full article ">Figure 3
<p>Contact angle microscope pictures of (<b>a</b>) anhydrous RuO<sub>2</sub>, (<b>b</b>) hydrous RuO<sub>2</sub>.</p>
Full article ">Figure 4
<p>DTA/TG (Differential Thermal Analysis/Thermogravimetry) diagrams of anhydrous (violet curves) and hydrous (blue curves) RuO<sub>2</sub> (upper curves: TG analysis, bottom curves: DTA).</p>
Full article ">Figure 5
<p>Comparison of electrochemical properties of anhydrous (violet curve) and hydrous (blue curve) RuO<sub>2</sub> tested with the use of cyclic voltammetry in the potential range from 0 to 0.3 V, inset: voltammogram of hydrous RuO<sub>2</sub> registered in a broader potential window from −1.0 to 1.0 V. Remaining cyclic voltammetry (CV) parameters: scan rate: 0.1 V/s, electrolyte: 0.01 M KCl.</p>
Full article ">Figure 6
<p>Comparison of electrochemical properties of anhydrous (violet curves) and hydrous (blue curves) RuO<sub>2</sub> tested with the use of electrochemical impedance spectroscopy in the frequency range of 100 kHz and 0.01 Hz with an amplitude of 0.01 V and set potential value of 0.15 V in 0.01 M KCl as electrolyte, (<b>a</b>) electrochemical impedance spectroscopy (EIS) curves of both tested layers, inset: closer look at hydrous RuO<sub>2</sub>, (<b>b</b>) real part of capacitance (C’) change vs. frequency, (<b>c</b>) imaginary part of capacitance (C”) change vs. frequency.</p>
Full article ">Figure 7
<p>Comparison of electrochemical properties of anhydrous (violet) and hydrous (blue) RuO<sub>2</sub> tested with the use of chronopotentiometry. Arrows point the moment of the current sign change.</p>
Full article ">Figure 8
<p>Capacitance values obtained for anhydrous (violet) and hydrous (blue) ruthenium dioxide dispersed in various solvents and applied in different amounts onto the electrode’s surface.</p>
Full article ">Figure 9
<p>Part of chronopotentiograms recorded for solid contact electrodes: GCD/RuO<sub>2</sub>•xH<sub>2</sub>O/K<sup>+</sup>-ISM electrode (blue curve), GCD/RuO<sub>2</sub>/K<sup>+</sup>-ISM electrode (violet curve) and coated disc GCD/K<sup>+</sup>-ISM electrode (black curve).</p>
Full article ">Figure 10
<p>Impedance spectrum of GCD/RuO<sub>2</sub>•xH<sub>2</sub>O/K<sup>+</sup>-ISM (blue) and GCD/RuO<sub>2</sub>/K<sup>+</sup>-ISM (violet) electrode in 0.01 M KCl solution and equivalent electrical circuits. Frequency range: 100 kHz–0.01 Hz. Equivalent circuits are shown as insets (solid lines represent data fits).</p>
Full article ">Figure 11
<p>Calibration curves recorded for GCD/RuO<sub>2</sub>•xH<sub>2</sub>O/K<sup>+</sup>-ISM (<span style="color:#3B9595">■</span>), GCD/RuO<sub>2</sub>/K<sup>+</sup>-ISM (<span style="color:#7030A0">●</span>), and GCD/K<sup>+</sup>-ISM (▲) electrode.</p>
Full article ">Figure 12
<p>Potentiometric response of (<b>a</b>) GCD/RuO<sub>2</sub>•xH<sub>2</sub>O/K<sup>+</sup>-ISM electrode and (<b>b</b>) GCD/RuO<sub>2</sub>/K<sup>+</sup>-ISM electrode during first 10 min of measurement.</p>
Full article ">Figure 13
<p>Water layer test of GCD/RuO<sub>2</sub>•xH<sub>2</sub>O/K<sup>+</sup>-ISM (blue curve) and GCD/RuO<sub>2</sub>/K<sup>+</sup>-ISM (violet curve) electrode.</p>
Full article ">
29 pages, 2738 KiB  
Article
Precipitation of Inorganic Salts in Mitochondrial Matrix
by Jerzy J. Jasielec, Robert Filipek, Krzysztof Dołowy and Andrzej Lewenstam
Membranes 2020, 10(5), 81; https://doi.org/10.3390/membranes10050081 - 27 Apr 2020
Cited by 10 | Viewed by 6711
Abstract
In the mitochondrial matrix, there are insoluble, osmotically inactive complexes that maintain a constant pH and calcium concentration. In the present paper, we examine the properties of insoluble calcium and magnesium salts, such as phosphates, carbonates and polyphosphates, which might play this role. [...] Read more.
In the mitochondrial matrix, there are insoluble, osmotically inactive complexes that maintain a constant pH and calcium concentration. In the present paper, we examine the properties of insoluble calcium and magnesium salts, such as phosphates, carbonates and polyphosphates, which might play this role. We find that non-stoichiometric, magnesium-rich carbonated apatite, with very low crystallinity, precipitates in the matrix under physiological conditions. Precipitated salt acts as pH buffer, and, hence, can contribute in maintaining ATP production in ischemic conditions, which delays irreversible damage to heart and brain cells after stroke. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The calculated activity coefficients for the main inorganic ions present in mitochondria, (<b>a</b>) in logarithmic scale and (<b>b</b>) in linear scale.</p>
Full article ">Figure 2
<p>The calculated equilibrium concentrations of free calcium (<b>a</b>) as functions of pH at 37 °C and pCO<sub>2</sub> = 7 kPa, (<b>b</b>) as a function of partial pressure of carbon dioxide at 37 °C and pH = 7.8.</p>
Full article ">Figure 3
<p>The calculated equilibrium concentrations of free magnesium (<b>a</b>) as functions of pH at 35 °C and pCO<sub>2</sub> = 7 kPa, (<b>b</b>) as a function of partial pressure of carbon dioxide at 35 °C and pH = 7.8.</p>
Full article ">Figure 4
<p>The calculated equilibrium concentrations [∑P<sub>i</sub>]<sub>eq</sub> required for the formation of calcium phosphates as functions of pH. Calcium concentration set to (<b>a</b>) [Ca<sup>2+</sup>] = 0.17 µM and (<b>b</b>) [Ca<sup>2+</sup>] = 5 µM.</p>
Full article ">Figure 5
<p>The calculated equilibrium concentrations [∑P<sub>i</sub>]<sub>eq</sub> required for the formation of magnesium phosphates as functions of pH. Magnesium concentration set to (<b>a</b>) [Mg<sup>2+</sup>] = 0.35 mM and (<b>b</b>) [Mg<sup>2+</sup>] = 1.5 mM.</p>
Full article ">Scheme 1
<p>Ion channels, exchangers, pumps and oxidative phosphorylation chain of the inner mitochondrial membrane: green—uniporters, blue—exchangers, yellow—four complexes of oxidative phosphorylation chain, and red—ATP synthasome. There are more than one type of mitoBK<sub>Ca</sub>, mitoKv or mitoCl channels.</p>
Full article ">
13 pages, 6597 KiB  
Article
Measurement of Multi Ion Transport through Human Bronchial Epithelial Cell Line Provides an Insight into the Mechanism of Defective Water Transport in Cystic Fibrosis
by Miroslaw Zajac, Andrzej Lewenstam, Piotr Bednarczyk and Krzysztof Dolowy
Membranes 2020, 10(3), 43; https://doi.org/10.3390/membranes10030043 - 12 Mar 2020
Cited by 9 | Viewed by 4336
Abstract
We measured concentration changes of sodium, potassium, chloride ions, pH and the transepithelial potential difference by means of ion-selective electrodes, which were placed on both sides of a human bronchial epithelial 16HBE14σ cell line grown on a porous support in the presence of [...] Read more.
We measured concentration changes of sodium, potassium, chloride ions, pH and the transepithelial potential difference by means of ion-selective electrodes, which were placed on both sides of a human bronchial epithelial 16HBE14σ cell line grown on a porous support in the presence of ion channel blockers. We found that, in the isosmotic transepithelial concentration gradient of either sodium or chloride ions, there is an electroneutral transport of the isosmotic solution of sodium chloride in both directions across the cell monolayer. The transepithelial potential difference is below 3 mV. Potassium and pH change plays a minor role in ion transport. Based on our measurements, we hypothesize that in a healthy bronchial epithelium, there is a dynamic balance between water absorption and secretion. Water absorption is caused by the action of two exchangers, Na/H and Cl/HCO3, secreting weakly dissociated carbonic acid in exchange for well dissociated NaCl and water. The water secretion phase is triggered by an apical low volume-dependent factor opening the Cystic Fibrosis Transmembrane Regulator CFTR channel and secreting anions that are accompanied by paracellular sodium and water transport. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic model of apparatus included. (<b>A</b>) solution inlet and outlet tubes, (<b>B</b>) ion-selective and reference electrodes, (<b>C</b>) Costar Snapwell insert with epithelial cell monolayer, and (<b>D</b>) asymmetric body. The diameter of the body is 4 cm. Distance between the cell layer and ion-selective electrodes is less than 30 µm. The inner diameter of the Costar Snapwell insert is 12 mm.</p>
Full article ">Figure 2
<p>Kinetics of the ion selective electrodes response after changing the medium. The start of medium exchange is shown by the arrow. (<b>A</b>) KHS3 solution is replaced by KHS1, (<b>B</b>) KHS2 solution is replaced by KHS1, (<b>C</b>) KHS1 solution equilibrated to pH = 6.4 is replaced by KHS1 (pH = 7.4), (<b>D</b>) KHS3 solution is replaced by KHS1 (no change in potassium concentration).</p>
Full article ">Figure 3
<p>The results (Mean ± SD) of experiments performed in sodium gradient (apical side, KHS2; basolateral side, KHS1) in control conditions and after supplementation of the solution with ion channel blockers. The graph represents: sodium, chloride, potassium transport across the epithelial cell monolayer expressed as the concentration change measured 10 min after appropriate solution flow stop; final pH value and the transepithelial potential difference change (N = 8–10 for Cl<sup>−</sup>, Na<sup>+</sup> and V<sub>te</sub>, N = 6 for K<sup>+</sup> and pH, dependent sample t-test, * <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). Note different scales for apical and basolateral graphs.</p>
Full article ">Figure 4
<p>The results (Mean ± SD) of experiments performed in reversed sodium gradient (apical side, KHS1; basolateral side, KHS2) in control conditions and after supplementation of the solution with ion channel blockers. The graph represents: sodium, chloride, potassium transport across epithelial cell monolayer expressed as the concentration change measured 10 min after appropriate solution flow stop; final pH value and the transepithelial potential difference change (N = 7–12 for Cl<sup>−</sup>, Na<sup>+</sup> and V<sub>te</sub>, N = 6 for K<sup>+</sup> and pH, dependent sample t-test, * <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). Note different scales for apical and basolateral graphs.</p>
Full article ">Figure 5
<p>The results (Mean ± SD) of experiments performed in chloride gradient (apical side, KHS3; basolateral side, KHS1) in control conditions and after supplementation of the solution with ion channel blockers. The graph represents: sodium, chloride and potassium transport across epithelial cell monolayer expressed as the concentration change measured 10 min after appropriate solution flow stop; final pH value and the transepithelial potential difference change (N = 15–17 for Cl<sup>-</sup>, Na<sup>+</sup> and V<sub>te</sub>, N = 5 for K<sup>+</sup> and pH, dependent sample t-test, * <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). Note different scales for apical and basolateral graphs.</p>
Full article ">Figure 6
<p>The results (Mean ± SD) of experiments performed in reversed chloride gradient (apical side, KHS1; basolateral side, KHS3) in control conditions and after supplementation of the solution with ion channel blockers. The graph represents: sodium, chloride, potassium transport across epithelial cell monolayer expressed as the concentration change measured 10 min after appropriate solution flow stop; final pH value and the transepithelial potential difference change (N = 10–13 for Cl<sup>−</sup>, Na<sup>+</sup> and V<sub>te</sub>, N = 5 for K<sup>+</sup> and pH, dependent sample t-test, * <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). Note different scales for apical and basolateral graphs.</p>
Full article ">Figure 7
<p>Measured (Mean ± SD) and theoretically predicted concentrations of sodium and chloride ions on both sides of the cell layer in control conditions after blocking particular ion channels present on the apical side of the monolayer. The experiments with sodium gradients and isosmotic flow of <span class="html-italic">x</span> fraction of 145 mM sodium chloride across the epithelial cell monolayer are shown.</p>
Full article ">Figure 8
<p>Measured (Mean ± SD) and theoretically predicted concentrations of sodium and chloride ions on both sides of the cell layer in control conditions after blocking particular ion channels present on the apical side of the monolayer. The experiments with sodium gradients and isosmotic flow of <span class="html-italic">x</span> fraction of 145 mM sodium chloride across the epithelial cell monolayer are shown.</p>
Full article ">Figure 9
<p>‘The dynamic water balance’ hypothesis: (<b>A</b>) water absorption, (<b>B</b>) water secretion. During the absorption phase (<b>A</b>) two exchangers are responsible for sodium and chloride influx from ASL into the cytoplasm and outward proton and bicarbonate secretion. The lower osmotic pressure of partially dissociated carbonic acid in ASL causes water absorption and depletion of its volume, which increases the concentration of the ASL volume-dependent factor (e.g., ATP) triggering the opening of the CFTR channel and starting the secretion phase (<b>B</b>). The secretion of chloride and bicarbonate ions via the CFTR channel is accompanied by the transport of sodium ions via a paracellular pathway. The appearance of net 2.6 NaCl molecules on the apical side leads to osmotic water transport, ASL hydration and dilution of the triggering factor resulting in returning the epithelium to the absorption phase. In the absence of functional CFTR, the bronchial epithelium stays in the absorptive phase leading to Cl/HCO<sub>3</sub> reaching energetic equilibrium, the action of Na/H exchanger, acidification of ASL and undiluted mucus.</p>
Full article ">
22 pages, 2616 KiB  
Article
Reference Electrodes with Polymer-Based Membranes—Comprehensive Performance Characteristics
by Peter Lingenfelter, Bartosz Bartoszewicz, Jan Migdalski, Tomasz Sokalski, Mirosław M. Bućko, Robert Filipek and Andrzej Lewenstam
Membranes 2019, 9(12), 161; https://doi.org/10.3390/membranes9120161 - 29 Nov 2019
Cited by 22 | Viewed by 6615
Abstract
Several types of liquid membrane and solid-state reference electrodes based on different plastics were fabricated. In the membranes studied, equitransferent organic (QB) and inorganic salts (KCl) are dispersed in polyvinyl chloride (PVC), polyurethane (PU), urea-formaldehyde resin (UF), polyvinyl acetate (PVA), as well as [...] Read more.
Several types of liquid membrane and solid-state reference electrodes based on different plastics were fabricated. In the membranes studied, equitransferent organic (QB) and inorganic salts (KCl) are dispersed in polyvinyl chloride (PVC), polyurethane (PU), urea-formaldehyde resin (UF), polyvinyl acetate (PVA), as well as remelted KCl in order to show the matrix impact on the reference membranes’ behavior. The comparison of potentiometic performance was made using specially designed standardized testing protocols. A problem in the reference electrode research and literature has been a lack of standardized testing, which leads to difficulties in comparing different types, qualities, and properties of reference electrodes. Herein, several protocols were developed to test the electrodes’ performance with respect to stability over time, pH sensitivity, ionic strength, and various ionic species. All of the prepared reference electrodes performed well in at least some respect and would be suitable for certain applications as described in the text. Most of the reference types, however, demonstrated some weakness that had not been previously highlighted in the literature, due in large part to the lack of exhaustive and/or consistent testing protocols. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagrams of reference electrodes (REs) used. On the left side general scheme of RE, where 1) is the Ag/AgCl electrode, 2) is the internal solution or solid contact, 3) is the membrane or frit. On the right side: (<b>i</b>.) RE with QB (PVC) and QB (PU) membranes, (<b>ii</b>.) RE with remelted inorganic salts (RKCl), (<b>iii</b>.) urea-formaldehyde (UF) resin-based RE, and (<b>iv</b>.) RE with PVA membrane. More detailed characterization of (<b>i</b>.–<b>iv</b>.) type is provided in the text.</p>
Full article ">Figure 2
<p>Stability over (<b>a</b>) 1 h, (<b>b</b>) 10 h, and (<b>c</b>) 20 h in 10<sup>−4</sup> M KCl at 23 °C for 1. QB (PVC), 2. QB (PU), 3. RKCl, 4. PVA, 5. UF, 6. Sureflow, 7. REFEX, 8. Orion Ross Ultra.</p>
Full article ">Figure 2 Cont.
<p>Stability over (<b>a</b>) 1 h, (<b>b</b>) 10 h, and (<b>c</b>) 20 h in 10<sup>−4</sup> M KCl at 23 °C for 1. QB (PVC), 2. QB (PU), 3. RKCl, 4. PVA, 5. UF, 6. Sureflow, 7. REFEX, 8. Orion Ross Ultra.</p>
Full article ">Figure 3
<p>pH titrations with the various reference electrodes tested. 1(+) QB(PVC), 2(o) QB(PU), 3(●) RKCl, 4(×) PVA, 5(■) UF, 6(□) Sureflow, 7(△) REFEX, 8(◇) Orion Ross Ultra.</p>
Full article ">Figure 4
<p>pH titrations with the various reference electrodes tested. 1(+) QB(PVC), 2(o) QB(PU), 3(●) RKCl, 4(×) PVA, 5(■) UF, 6(□) Sureflow, 7(△) REFEX, 8(◇) Orion Ross Ultra.</p>
Full article ">Figure 5
<p>MSP results for the experimental and commercial electrodes tested for this study. 1(+) QB(PVC), 2(o) QB(PU), 3(●) RKCl, 4(×) PVA, 5(■) UF, 6(□) Sureflow, 7(△) REFEX, 8(◇) Orion Ross Ultra.</p>
Full article ">Figure 6
<p>Effect of aging on the 1. QB(PVC), 2. UF, and 3. PVA reference electrodes (<b>a</b>) soon after production and (<b>b</b>) after three months of semi-regular use.</p>
Full article ">

Review

Jump to: Editorial, Research

14 pages, 593 KiB  
Review
Unintended Changes of Ion-Selective Membranes Composition—Origin and Effect on Analytical Performance
by Krzysztof Maksymiuk, Emilia Stelmach and Agata Michalska
Membranes 2020, 10(10), 266; https://doi.org/10.3390/membranes10100266 - 28 Sep 2020
Cited by 28 | Viewed by 3897
Abstract
Ion-selective membranes, as used in potentiometric sensors, are mixtures of a few important constituents in a carefully balanced proportion. The changes of composition of the ion-selective membrane, both qualitative and quantitative, affect the analytical performance of sensors. Different constructions and materials applied to [...] Read more.
Ion-selective membranes, as used in potentiometric sensors, are mixtures of a few important constituents in a carefully balanced proportion. The changes of composition of the ion-selective membrane, both qualitative and quantitative, affect the analytical performance of sensors. Different constructions and materials applied to improve sensors result in specific conditions of membrane formation, in consequence, potentially can result in uncontrolled modification of the membrane composition. Clearly, these effects need to be considered, especially if preparation of miniaturized, potentially disposable internal-solution free sensors is considered. Furthermore, membrane composition changes can occur during the normal operation of sensors—accumulation of species as well as release need to be taken into account, regardless of the construction of sensors used. Issues related to spontaneous changes of membrane composition that can occur during sensor construction, pre-treatment and their operation, seem to be underestimated in the subject literature. The aim of this work is to summarize available data related to potentiometric sensors and highlight the effects that can potentially be important also for other sensors using ion-selective membranes, e.g., optodes or voltammetric sensors. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of processes related to ion-selective membrane occurring during sensor preparation as well as application.</p>
Full article ">
24 pages, 14297 KiB  
Review
Solid-Contact Ion-Selective Electrodes: Response Mechanisms, Transducer Materials and Wearable Sensors
by Yan Lyu, Shiyu Gan, Yu Bao, Lijie Zhong, Jianan Xu, Wei Wang, Zhenbang Liu, Yingming Ma, Guifu Yang and Li Niu
Membranes 2020, 10(6), 128; https://doi.org/10.3390/membranes10060128 - 23 Jun 2020
Cited by 99 | Viewed by 10374
Abstract
Wearable sensors based on solid-contact ion-selective electrodes (SC-ISEs) are currently attracting intensive attention in monitoring human health conditions through real-time and non-invasive analysis of ions in biological fluids. SC-ISEs have gone through a revolution with improvements in potential stability and reproducibility. The introduction [...] Read more.
Wearable sensors based on solid-contact ion-selective electrodes (SC-ISEs) are currently attracting intensive attention in monitoring human health conditions through real-time and non-invasive analysis of ions in biological fluids. SC-ISEs have gone through a revolution with improvements in potential stability and reproducibility. The introduction of new transducing materials, the understanding of theoretical potentiometric responses, and wearable applications greatly facilitate SC-ISEs. We review recent advances in SC-ISEs including the response mechanism (redox capacitance and electric-double-layer capacitance mechanisms) and crucial solid transducer materials (conducting polymers, carbon and other nanomaterials) and applications in wearable sensors. At the end of the review we illustrate the existing challenges and prospects for future SC-ISEs. We expect this review to provide readers with a general picture of SC-ISEs and appeal to further establishing protocols for evaluating SC-ISEs and accelerating commercial wearable sensors for clinical diagnosis and family practice. Full article
Show Figures

Figure 1

Figure 1
<p>An overview from liquid-contact ion-selective electrodes (LC-ISEs) to solid-contact ISEs (SC-ISEs) for wearable sensors. (<b>A</b>) Classic LC-ISEs (e.g., pH meter) by a liquid-contact between internal solution and ion-selective membrane (ISM). (<b>B</b>) The structure of SC-ISEs by a solid-contact between solid ion-to-electron transducer layer and ISM. (<b>C</b>) An example of the SC-ISEs for wearable sensor applications [<a href="#B9-membranes-10-00128" class="html-bibr">9</a>].</p>
Full article ">Figure 2
<p>Response mechanisms for the SC-ISEs. (<b>A</b>) Redox capacitance-based SC-ISEs with poly(3,4-ethylenedioxythiophene) (PEDOT) as an example for redox SC transducer. (<b>B</b>) Electric-double-layer (EDL) capacitance-based SC-ISEs with carbon as an example for EDL SC transducer. Both SC-ISEs contain three interfaces, GC/SC, SC/ISM and ISM/aq. GC: glass carbon electrode substrate; SC: solid contact; aq: aqueous solution; ET: electron transfer; IT: ion transfer. The corresponding phase interfacial potentials are presented on the right (detailed illustration shown in the main text and <a href="#app1-membranes-10-00128" class="html-app">Appendix A</a>).</p>
Full article ">Figure 3
<p>Structural design and functionalization of conducting polymers (CPs) for CP-based SC-ISEs. (<b>A</b>) Scanning electronic microscopy (SEM) image of the designed high-surface 3D poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT(PSS)) by nanosphere lithography and electrosynthesis. The right: the potential traces of 3D PEDOT(PSS) functionalized with (red line) and without (black line) a lipophilic redox 1,1′-dimethylferrocene. It is found that the functionalized 3D PEDOT(PSS) exhibits much better reproducibility. Reprinted with permission from [<a href="#B24-membranes-10-00128" class="html-bibr">24</a>], Copyright (2016) John Wiley and Sons publications. (<b>B</b>) The oxidized PPy by doping perfluorooctanesulfonate (PFOS<sup>−</sup>) anion (PPy-PFOS) enhanced the hydrophobicity leading to remarkably improved reproducibility. Reprinted with permission from [<a href="#B25-membranes-10-00128" class="html-bibr">25</a>], Copyright (2017) American Chemical Society. (<b>C</b>) Superhydrophobic tetrakis-(pentafluorophenyl)borate (TPFPhB<sup>−</sup>) anion doping PEDOT as an SC transducer to reduce the water-layer effect. It should be noted the TPFPhB<sup>−</sup> ion transfer (ion exchange) between SC and ISM further enhanced the potential stability (see Equation (3)). The abbreviation of tetrakis-(pentafluorophenyl)borate on the original Figure is TFAB<sup>−</sup>. Herein it is replaced by TPFPhB<sup>−</sup>. Reprinted with permission from [<a href="#B26-membranes-10-00128" class="html-bibr">26</a>], Copyright (2019) American Chemical Society. (<b>D</b>) An ultimate approach by both C<sub>14</sub>-chain functionalized PEDOT and TPFPhB<sup>−</sup> doping to improve the performances of SC-ISEs. Reprinted with permission from [<a href="#B27-membranes-10-00128" class="html-bibr">27</a>], Copyright (2017) American Chemical Society.</p>
Full article ">Figure 4
<p>Suppress of the leaking of ISM components. (<b>A</b>) A CP-based SC-ISE by coating the silicon rubber (SR) on the ISM layer. (<b>B</b>) A photograph of the SR on the poly(vinyl chloride) (PVC)-based ISM. (<b>C</b>) Fourier transform infrared (FTIR) spectra analysis of the SR layer (bottom and top) in comparison with standard ISM components. The ISM components were observed in both bottom and top SR (typical dioctyl sebacate, DOS). (<b>D</b>) The K<sup>+</sup>-response of SR-coated SC-ISEs compared with an uncoated one. Reprinted with permission from [<a href="#B49-membranes-10-00128" class="html-bibr">49</a>], Copyright (2019) American Chemical Society.</p>
Full article ">Figure 5
<p>Carbon-based SC-ISEs. (<b>A</b>) High-surface 3D-ordered microporous (3DOM) carbon as a SC transducer for SC-ISEs. Reprinted with permission from [<a href="#B50-membranes-10-00128" class="html-bibr">50</a>], Copyright (2007) American Chemical Society. (<b>B</b>) Colloid-imprinted mesoporous (CIM) carbon with higher surface area as a SC transducer for SC-ISEs. Reprinted with permission from [<a href="#B52-membranes-10-00128" class="html-bibr">52</a>], Copyright (2014) American Chemical Society. (<b>C</b>) The porous carbon sub-micrometer spheres (PC-SMSs) for superhydrophobic SC transducer with the contact angle up to 137°. Reprinted with permission from [<a href="#B53-membranes-10-00128" class="html-bibr">53</a>], Copyright (2015) Elsevier. (<b>D</b>) Single-wall carbon nanotube (SWCNT)-based SC-ISEs and the illustrated EDL capacitance response mechanism. Reprinted with permission from [<a href="#B54-membranes-10-00128" class="html-bibr">54</a>], Copyright (2008) American Chemical Society. (<b>E</b>) The chemically prepared reduced graphene oxide (RGO) as SC transducer for SC-ISEs. Reprinted with permission from [<a href="#B55-membranes-10-00128" class="html-bibr">55</a>], Copyright (2012) Royal Society of Chemistry.</p>
Full article ">Figure 6
<p>Au nanoclusters-based SC-ISEs. (<b>A</b>–<b>C</b>) Au<sub>144</sub> redox nanocluster based-SC-ISEs. (<b>A</b>) Reduced and oxidized states of thiol monolayer-protected Au clusters (MPCs) (Au<sub>144</sub>) doped with tetrakis(4-chlorophenyl) borate (TB<sup>−</sup>) anion (MPC<sup>0</sup>/MPC<sup>+</sup>TB<sup>−</sup>). (<b>B</b>) Au MPCs-based SC-ISEs with well-defined phase interfacial potential definition. (<b>C</b>) Water-layer test of Au<sub>144</sub>-based SC-ISEs. Reprinted with permission from [<a href="#B61-membranes-10-00128" class="html-bibr">61</a>], Copyright (2012) American Chemical Society. (<b>D</b>,<b>E</b>) Au<sub>25</sub> redox nanocluster based-SC-ISEs. (<b>D</b>) Transmission electronic image (TEM) of the Au<sub>25</sub> synthesized by optimized one-phase reduction procedure. (<b>E</b>) The Au<sub>25</sub>-based SC-ISEs for K<sup>+</sup>-response to examine its long-term stability. Reprinted with permission from [<a href="#B62-membranes-10-00128" class="html-bibr">62</a>], Copyright (2016) Elsevier. (<b>F</b>) Au MPCs-based single-piece SC-ISEs through mixing the SC and ISM phase. Reprinted with permission from [<a href="#B63-membranes-10-00128" class="html-bibr">63</a>], Copyright (2016) Elsevier.</p>
Full article ">Figure 7
<p>A few other representative SC materials. (<b>A</b>) Co(II)/Co(III) complex redox buffer-based SC-ISEs. Reprinted with permission from [<a href="#B64-membranes-10-00128" class="html-bibr">64</a>], Copyright (2013) American Chemical Society. (<b>B</b>,<b>C</b>) Lithium-battery materials of LiFePO<sub>4/</sub>FePO<sub>4</sub>-based SC-ISEs. Reprinted with permission from [<a href="#B70-membranes-10-00128" class="html-bibr">70</a>], Copyright (2016) John Wiley and Sons publications. (<b>D</b>,<b>E</b>) MoS<sub>2</sub> nanomaterials was used for the EDL-type solid contact. SEM image of MoS<sub>2</sub> (<b>D</b>) and water layer test (<b>E</b>). Reprinted with permission from [<a href="#B71-membranes-10-00128" class="html-bibr">71</a>], Copyright (2016) Elsevier. (<b>F</b>) Metal–organic frameworks (MOFs)-based SC-ISEs. Reprinted with permission from [<a href="#B73-membranes-10-00128" class="html-bibr">73</a>], Copyright (2018) American Chemical Society.</p>
Full article ">Figure 8
<p>CPs-based wearable SC-ISEs for skin sweat ion sensing. (<b>A</b>) An integrated multi-parameter electrochemical wearable sensor involving Na<sup>+</sup> and K<sup>+</sup>-SC-ISEs, glucose, lactate and temperature. The PEDOT was used for the SC transducer. Reprinted with permission from [<a href="#B75-membranes-10-00128" class="html-bibr">75</a>], Copyright (2016) Springer Nature. (<b>B</b>) PEDOT-based wearable SC-ISEs for pH and Ca<sup>2+</sup> sensors. Reprinted with permission from [<a href="#B76-membranes-10-00128" class="html-bibr">76</a>], Copyright (2016) American Chemical Society. (<b>C</b>–<b>E</b>) PEDOT-based self-healable SC-ISEs for wearable senor. Reprinted with permission from [<a href="#B77-membranes-10-00128" class="html-bibr">77</a>], Copyright (2019) American Chemical Society.</p>
Full article ">Figure 9
<p>Carbon-based wearable SC-ISEs for skin sweat ion sensing. (<b>A</b>) SWCNT-based flexible SC-ISEs by using cotton yard as flexible substrate. Reprinted with permission from [<a href="#B78-membranes-10-00128" class="html-bibr">78</a>], Copyright (2013) Royal Society of Chemistry. A series of preparation including SWCNT ink and ISM dipping as shown in (<b>a</b>–<b>d</b>). (<b>B</b>) Multi-wall carbon nanotube (MWCNT) textile-based SC-ISEs for Na<sup>+</sup> and K<sup>+</sup> sensing. Reprinted with permission from [<a href="#B80-membranes-10-00128" class="html-bibr">80</a>], Copyright (2016) John Wiley and Sons publications. (<b>C</b>) Carbon textile-based sensor array for multiparameter analysis. Reprinted with permission from Science Advances [<a href="#B81-membranes-10-00128" class="html-bibr">81</a>], Copyright (2019) American Association for the Advancement of Science. (<b>D</b>) High-quality graphene-based wearable SC-ISEs for multichannel ion sensing including K<sup>+</sup>, Na<sup>+</sup>, Cl<sup>−</sup> and pH [<a href="#B82-membranes-10-00128" class="html-bibr">82</a>]. (<b>E</b>) The suggested protocol for on-body measurement. It should be noted that the importance of calibration of SC-ISEs to assure the accuracy of real-time analysis. Reprinted with permission from [<a href="#B83-membranes-10-00128" class="html-bibr">83</a>], Copyright (2019) American Chemical Society.</p>
Full article ">Figure 10
<p>Au nanomaterials-based wearable SC-ISEs for skin sweat ion sensing. (<b>A</b>–<b>C</b>) AuND array-based SC-ISEs for sweat Na<sup>+</sup> sensing. (<b>A</b>) The preparation procedure for AuND-based SC-ISEs by photolithography technique. (<b>B</b>) A schematic for the SC-ISEs on-body measurement. (<b>C</b>) Real-time analysis of sweat Na<sup>+</sup> during cycling and rest states. Reprinted with permission from [<a href="#B84-membranes-10-00128" class="html-bibr">84</a>], Copyright (2017) American Chemical Society. (<b>D</b>,<b>E</b>) Gold-based vertically aligned nanowires (V-AuNWs)-based stretchable SC-ISEs for Na<sup>+</sup>, K<sup>+</sup> and pH sensing. (<b>D</b>) The preparation of V-AuNWs on PDMS film and corresponding optical images for observing the stretching. Scale bar: 200 μm. Reprinted with permission from [<a href="#B85-membranes-10-00128" class="html-bibr">85</a>], Copyright (2019) John Wiley and Sons publications. (<b>E</b>) The Na<sup>+</sup>, K<sup>+</sup> and pH sensing from 0% to 30% stretching. Reprinted with permission from [<a href="#B86-membranes-10-00128" class="html-bibr">86</a>], Copyright (2020) American Chemical Society.</p>
Full article ">Figure 11
<p>Wearable SC-ISEs for ion detection in interstitial fluid. (<b>A</b>–<b>C</b>) Microneedle-based SC-ISEs for K<sup>+</sup> analysis in skin interstitial fluid. (<b>A</b>) Illustration of microneedle patch including working electrode (WE) and reference electrode (RE). For the WE, the bare microneedle was coated carbon, f-MWCNTs and K<sup>+</sup>-ISM. For the RE, the bare microneedle was coated Ag/AgCl; and poly(vinly butyral) membrane and polyurethane. (<b>B</b>) K<sup>+</sup> response before and after insertion into the animal skin. (<b>C</b>) Ex vivo K<sup>+</sup> measurement in chicken skin with calibration. Reprinted with permission from [<a href="#B88-membranes-10-00128" class="html-bibr">88</a>], Copyright (2019) American Chemical Society. (<b>D</b>,<b>E</b>) A cotton fiber-based SC-ISEs for Li<sup>+</sup> sensing in the human plasma. (<b>D</b>) A SEM image for the cotton-based SC-ISEs. (<b>E</b>) Li<sup>+</sup>-response in aqueous solution and human plasma. The inset shows the time traces for Li<sup>+</sup> response. Reprinted with permission from [<a href="#B89-membranes-10-00128" class="html-bibr">89</a>], Copyright (2018) American Chemical Society.</p>
Full article ">
Back to TopTop