[go: up one dir, main page]

Next Issue
Volume 2, December
Previous Issue
Volume 2, June
 
 

Membranes, Volume 2, Issue 3 (September 2012) – 15 articles , Pages 346-686

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
898 KiB  
Article
Investigation of La1−xSrxCrO3− (x ~ 0.1) as Membrane for Hydrogen Production
by Yngve Larring, Camilla Vigen, Florian Ahouanto, Marie-Laure Fontaine, Thijs Peters, Jens B. Smith, Truls Norby and Rune Bredesen
Membranes 2012, 2(3), 665-686; https://doi.org/10.3390/membranes2030665 - 11 Sep 2012
Cited by 24 | Viewed by 8868
Abstract
Various inorganic membranes have demonstrated good capability to separate hydrogen from other gases at elevated temperatures. Hydrogen-permeable, dense, mixed proton-electron conducting ceramic oxides offer superior selectivity and thermal stability, but chemically robust candidates with higher ambipolar protonic and electronic conductivity are needed. In [...] Read more.
Various inorganic membranes have demonstrated good capability to separate hydrogen from other gases at elevated temperatures. Hydrogen-permeable, dense, mixed proton-electron conducting ceramic oxides offer superior selectivity and thermal stability, but chemically robust candidates with higher ambipolar protonic and electronic conductivity are needed. In this work, we present for the first time the results of various investigations of La1−xSrxCrO3− membranes for hydrogen production. We aim in particular to elucidate the material’s complex transport properties, involving co-ionic transport of oxide ions and protons, in addition to electron holes. This opens some new possibilities for efficient heat and mass transfer management in the production of hydrogen. Conductivity measurements as a function of pH2 at constant pO2 exhibit changes that reveal a significant hydration and presence of protons. The flux and production of hydrogen have been measured under different chemical gradients. In particular, the effect of water vapor in the feed and permeate gas stream sides was investigated with the aim of quantifying the ratio of hydrogen production by hydrogen flux from feed to permeate and oxygen flux the opposite way (“water splitting”). Deuterium labeling was used to unambiguously prove flux of hydrogen species. Full article
(This article belongs to the Special Issue Membrane Processes and Energy)
Show Figures

Figure 1

Figure 1
<p>PCDC process suggested by Statoil with integrated ceramic mixed conductor membrane [<a href="#B3-membranes-02-00665" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Example of defect concentrations for LSC with 1 mol% acceptor substitution as a function of <span class="html-italic">p</span>O<sub>2</sub> at constant <span class="html-italic">p</span>H<sub>2</sub>O.</p>
Full article ">Figure 3
<p>Example of defect concentrations for LSC with 1 mol% acceptor substitution as a function of <span class="html-italic">p</span>H<sub>2</sub>O at constant <span class="html-italic">p</span>O<sub>2</sub>.</p>
Full article ">Figure 4
<p>Total a.c. conductivity as a function of water vapor partial pressure for La<sub>0.9</sub>Sr<sub>0.1</sub>CrO<sub>3−<span class="html-italic">δ</span></sub> in reducing atmospheres.</p>
Full article ">Figure 5
<p>Van 't Hoff plot of hydration equilibrium constant <span class="html-italic">K<sub>hydr</sub></span> for La<sub>0.9</sub>Sr<sub>0.1</sub>CrO<sub>3−<span class="html-italic">δ</span></sub> corresponding to conductivity data of <a href="#membranes-02-00665-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>Estimated defect concentrations as a function of inverse temperature for La<sub>0.9</sub>Sr<sub>0.1</sub>CrO<sub>3−<span class="html-italic">δ</span></sub> in wet, reducing atmospheres (<span class="html-italic">p</span>H<sub>2</sub> = 1 atm, <span class="html-italic">p</span>H<sub>2</sub>O = 0.025 atm).</p>
Full article ">Figure 7
<p>Estimated total and partial conductivities as a function of inverse temperature for La<sub>0.9</sub>Sr<sub>0.1</sub>CrO<sub>3−<span class="html-italic">δ</span></sub> in wet reducing atmospheres (<span class="html-italic">p</span>H<sub>2</sub> = 1 atm, <span class="html-italic">p</span>H<sub>2</sub>O = 0.025 atm).</p>
Full article ">Figure 8
<p>Hydrogen production plotted as apparent hydrogen (H<sub>2</sub>) flux (leakage corrected) as a function of time at 1000 °C using 50% wet H<sub>2</sub> as feed gas for different LSC samples of thickness around 0.5 mm (see <a href="#membranes-02-00665-t001" class="html-table">Table 1</a>). Wet permeate gas is used for most of the experiment, except for a short period with dry permeate (DP), with a steep decrease in the hydrogen production rate.</p>
Full article ">Figure 9
<p>Hydrogen production plotted in terms of an apparent hydrogen (H<sub>2</sub>) flux as a function of inverse temperature for sample 4SA5_#2 in 4 different gradients using dry (D) or Wet (W) Feed (F) and Permeate (P) gas.</p>
Full article ">Figure 10
<p>Measured apparent H<sub>2</sub> flux, and the same corrected for leakage using the measured He flux, for sample 4B_#3 at 1000 °C, plotted as a function of (<b>a</b>) total feed pressure using H<sub>2</sub>/He/N<sub>2</sub>/steam = 20:10:50:20 on the feed side and Ar/steam = 80:20 on the permeate side,(<b>b</b>) feed hydrogen content <span class="html-italic">x</span> with H<sub>2</sub>/He/N<sub>2</sub>/steam = <span class="html-italic">x</span>:10:70 − <span class="html-italic">x</span>:20 on the feed side at 5 bar and Ar/steam = 80:20 on the permeate side at 4.9 bar, and (<b>c</b>) feed steam content <span class="html-italic">y</span> with H<sub>2</sub>/He/N<sub>2</sub>/steam = 20:10:50 − <span class="html-italic">y</span>:<span class="html-italic">y</span> on the feed side at 5 bar and Ar/steam = 80:20 on the permeate side at 4.9 bars.</p>
Full article ">Figure 11
<p>Apparent hydrogen flux and leakage of the monolith module at 20 bars pressure.</p>
Full article ">Figure 12
<p>Mass spectrometer signal from various masses <span class="html-italic">vs.</span> time, indicative of deuterium flux through LSC disk membrane at 1040 °C when changing the feed gas from He + 2.5% D<sub>2</sub>O to D<sub>2</sub> + 2.5% D<sub>2</sub>O after approximately 1 hour.</p>
Full article ">Figure 13
<p>Mass spectrometer signal from various masses <span class="html-italic">vs.</span> time, indicative of deuterium flux through LSC disk membrane at 1040 °C when changing the feed gas from He + 2.5% D<sub>2</sub>O to D<sub>2</sub> + 2.5% D<sub>2</sub>O after approximately 2 hours. Permeate gas is Ar + 2.5% H<sub>2</sub>O.</p>
Full article ">Figure 14
<p>Ambipolar proton-electron hole and oxygen vacancy-electron hole conductivity as a function of inverse temperature for La<sub>0.9</sub>Sr<sub>0.1</sub>CrO<sub>3−<span class="html-italic">δ</span></sub>, estimated from conductivity measurements, compared with conductivities extracted from measurements of hydrogen production with disk membrane in two different gradients and with the monolith module. All proton conductivities are extrapolated to 0.025 atm H<sub>2</sub>O.</p>
Full article ">
6038 KiB  
Review
A Review of RedOx Cycling of Solid Oxide Fuel Cells Anode
by Antonin Faes, Aïcha Hessler-Wyser, Amédée Zryd and Jan Van herle
Membranes 2012, 2(3), 585-664; https://doi.org/10.3390/membranes2030585 - 31 Aug 2012
Cited by 186 | Viewed by 15581
Abstract
Solid oxide fuel cells are able to convert fuels, including hydrocarbons, to electricity with an unbeatable efficiency even for small systems. One of the main limitations for long-term utilization is the reduction-oxidation cycling (RedOx cycles) of the nickel-based anodes. This paper will review [...] Read more.
Solid oxide fuel cells are able to convert fuels, including hydrocarbons, to electricity with an unbeatable efficiency even for small systems. One of the main limitations for long-term utilization is the reduction-oxidation cycling (RedOx cycles) of the nickel-based anodes. This paper will review the effects and parameters influencing RedOx cycles of the Ni-ceramic anode. Second, solutions for RedOx instability are reviewed in the patent and open scientific literature. The solutions are described from the point of view of the system, stack design, cell design, new materials and microstructure optimization. Finally, a brief synthesis on RedOx cycling of Ni-based anode supports for standard and optimized microstructures is depicted. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Microstructural changes during a RedOx process in Ni-YSZ (yttria stabilized zirconia) based anodes [<a href="#B27-membranes-02-00585" class="html-bibr">27</a>].</p>
Full article ">Figure 2
<p>Arrhenius plot of parabolic rate constant <span class="html-italic">k<sub>p</sub></span> as a function of oxidation temperature and scale thickness indicating NiO scale morphologies and microstructures [<a href="#B48-membranes-02-00585" class="html-bibr">48</a>,<a href="#B49-membranes-02-00585" class="html-bibr">49</a>].</p>
Full article ">Figure 3
<p>Open circuit voltage (OCV) or Nernst potential versus temperature for the Ni/NiO equilibrium [<a href="#B54-membranes-02-00585" class="html-bibr">54</a>,<a href="#B55-membranes-02-00585" class="html-bibr">55</a>].</p>
Full article ">Figure 4
<p>Effect of alloying on the rate constant for oxidation of nickel in air at 900 °C [<a href="#B46-membranes-02-00585" class="html-bibr">46</a>].</p>
Full article ">Figure 5
<p>Oxide map for alloys in the Ni-Cr-Al system delineating the composition ranges for formation of different types of oxide scales [<a href="#B25-membranes-02-00585" class="html-bibr">25</a>].</p>
Full article ">Figure 6
<p>Changes in stresses in 10ScSZ (scandium-stabilized zirconia) electrolyte and anode during heating Ni-3YSZ under air [<a href="#B96-membranes-02-00585" class="html-bibr">96</a>].</p>
Full article ">Figure 7
<p>Curvature change during reoxidation and re-reduction cycles (0.27 mm half-cell, 800 °C) [<a href="#B97-membranes-02-00585" class="html-bibr">97</a>]. During re-oxidation, the half-cell shows (<b>i</b>) an initial curvature towards the electrolyte (on the concave side); then (<b>ii</b>) it reverts to “zero”-curvature and finally (<b>iii</b>) it stabilizes with a curvature towards the anode (on the concave side).</p>
Full article ">Figure 8
<p>Picture of Ni-YSZ anode supported half-cell discs after 5 reduction-oxidation cycling (RedOx cycles) at 750 °C. (<b>A</b>) and (<b>C</b>): samples with 17.5% as-sintered porosity and (<b>B</b>) and (<b>D</b>): samples with 12% as-sintered porosity. The electrolyte is face-down for (<b>A</b>) and (<b>B</b>). A clear difference in curvature is observed between the two groups of samples comparing (<b>A</b>) and (<b>B</b>), (<b>B</b>) is bent towards the anode (anode face-up on the concave side). A clear difference in crack density is observed between the two groups of samples comparing (<b>C</b>) and (<b>D</b>) [<a href="#B98-membranes-02-00585" class="html-bibr">98</a>].</p>
Full article ">Figure 9
<p>Mechanical degradation in terms of relative loss of elastic modulus of NiO-YSZ composite in its oxidized state during RedOx cycle as a function of the oxidation strain (CRS: cumulative redox strain). The measurement is reproducible (<span class="html-italic">i.e.</span>, samples 1 and 2) [<a href="#B103-membranes-02-00585" class="html-bibr">103</a>].</p>
Full article ">Figure 10
<p>Strength of the C-shaped uniaxial compressed anode rings versus number of RedOx cycles. The samples were fabricated using two different powders, coarse green NiO and fine black nickel oxide [<a href="#B104-membranes-02-00585" class="html-bibr">104</a>].</p>
Full article ">Figure 11
<p>Thin electrolyte crack formation during two RedOx cycles in the anode supported cell design. (<b>a</b>) co-firing; (<b>b</b>) reduced (<b>c</b>) re-oxidized; (<b>d</b>) second reduction; (<b>e</b>) second oxidation and (f) third reduction with an additional 100 h under reducing atmosphere [<a href="#B14-membranes-02-00585" class="html-bibr">14</a>,<a href="#B107-membranes-02-00585" class="html-bibr">107</a>,<a href="#B108-membranes-02-00585" class="html-bibr">108</a>].</p>
Full article ">Figure 12
<p>Delamination of anode and anode current collection layer in case of 8YSZ electrolyte supported cells after five RedOx cycles at 950 °C and 40 min. Right: Only Ni-8YSZ active anode; left: active anode plus Ni-8YSZ current collecting layer [<a href="#B16-membranes-02-00585" class="html-bibr">16</a>].</p>
Full article ">Figure 13
<p>Survival probability of LaSrMn-oxide (LSM) cathode (60 µm) and YSZ electrolyte (10 µm) against the strain of the Ni-YSZ anode support (1 mm) [<a href="#B110-membranes-02-00585" class="html-bibr">110</a>].</p>
Full article ">Figure 14
<p>Model proposed by Klemensø <span class="html-italic">et al.</span> including the increase of Ni contact after a RedOx cycle due to breaking of the zirconia skeleton. Further RedOx cycles will create too much porosity to maintain sufficient conductivity [<a href="#B114-membranes-02-00585" class="html-bibr">114</a>].</p>
Full article ">Figure 15
<p>Comparison between Ni-YSZ and Ni-CGO composite electrical conductivity under RedOx treatments [<a href="#B81-membranes-02-00585" class="html-bibr">81</a>].</p>
Full article ">Figure 16
<p>Dilatometry measurements of oxidation for a YSZ composite bar infiltrated with 16 wt % Ni [<a href="#B117-membranes-02-00585" class="html-bibr">117</a>].</p>
Full article ">Figure 17
<p>Left: SEM of (<b>a</b>) as received Ni and (<b>b</b>) fully oxidized Ni (NiO) particles. The secondary electron images were recorded using a beam energy of 20 keV. Right: FIB cross-sectional secondary electron images of Ni particles oxidized at 800 °C for (<b>a</b>) 30 s; (<b>b</b>) 60 s; (<b>c</b>) 90 s; (<b>d</b>) 180 s; and (<b>e</b>) 300 s. Image (<b>f</b>) is the same particle as in (<b>e</b>) but obtained using the secondary ion signal [<a href="#B50-membranes-02-00585" class="html-bibr">50</a>].</p>
Full article ">Figure 18
<p>Secondary electron image from FIB cross-section from half-cells after one RedOx cycle (<b>a</b>) at 550 °C (lower magnification); (<b>b</b>) at 550 °C (higher magnification); (<b>c</b>) at 800 °C and (<b>d</b>) at 1000 °C. NiO contains small pores after a RedOx cycle at 550 and 800 °C but a single big pore after a RedOx cycle at 1000 °C. Dark grey is YSZ and light grey is NiO. The vertical lines come from the FIB milling process (“curtain effect”) [<a href="#B85-membranes-02-00585" class="html-bibr">85</a>].</p>
Full article ">Figure 19
<p>Cross-section of a transmission electron microscope (TEM) lamella after an <span class="html-italic">in situ</span> RedOx cycle, showing the hilly surface and closed porosity of the nickel oxide after reoxidation [<a href="#B129-membranes-02-00585" class="html-bibr">129</a>].</p>
Full article ">Figure 20
<p>Fine Ni-8YSZ anode before (left) and after (right) eight RedOx cycles at 950 °C (SEM, backscattered electron detector, 10 kV) [<a href="#B126-membranes-02-00585" class="html-bibr">126</a>].</p>
Full article ">Figure 21
<p>Impedance spectra at OCV during RedOx cycling at 950 °C of (<b>a</b>) Ni-40CGO (Ce<sub>0.6</sub>Gd<sub>0.4</sub>O<sub>2−d</sub>) and (<b>b</b>) Ni-8YSZ anodes with a 8YSZ electrolyte support. Top: Nyquist plot; bottom: complex impedance plot [<a href="#B81-membranes-02-00585" class="html-bibr">81</a>,<a href="#B126-membranes-02-00585" class="html-bibr">126</a>].</p>
Full article ">Figure 22
<p>Summary of the solutions for anode RedOx instability [<a href="#B129-membranes-02-00585" class="html-bibr">129</a>].</p>
Full article ">Figure 23
<p>RedOx stability of anode supported cells (ASCs) (with support thickness of 1, 0.4 and 0.5 mm for ASC-1 (from Forschungszentrum Jülich), ASC-2 and ASC-2, respectively [<a href="#B14-membranes-02-00585" class="html-bibr">14</a>]) compared to metal supported cell (MSC), under 50 RedOx cycles of 1 min and 50 RedOx cycles of 10 min at 800 °C. Top: open circuit voltage, bottom: normalized performance at 0.7 V [<a href="#B175-membranes-02-00585" class="html-bibr">175</a>].</p>
Full article ">Figure 24
<p>Voltage with time during RedOx cycles 800 °C of segmented-in-series (SIS) cells on a flattened partially stabilized zirconia tube support (under fuel, <span class="html-italic">i</span> = 0.9 A/cm<sup>2</sup>). In the first part of the test shown in (<b>a</b>), 12 SIS cells on one side of the module were tested during 7 cycles. The module was cycled to room temperature and then back to 800 °C before the second part of the test (<b>b</b>), where 9 SIS cells on the other side of the module were tested during 12 cycles. Arrows 1 and 2 indicate when the module was left overnight at 800 °C in hydrogen without cycling. Arrow 3 indicates a longer-than-usual (1 h) fuel feed [<a href="#B179-membranes-02-00585" class="html-bibr">179</a>].</p>
Full article ">Figure 25
<p>Time dependence of cell voltage of a Sr<sub>0.8</sub>La<sub>0.2</sub>TiO<sub>3</sub> supported-cell with Ni-Ce<sub>0.8</sub>Sm<sub>0.2</sub>O<sub>2</sub>, Ni-YSZ anode and YSZ thin electrolyte over 7 RedOx cycles at 800 °C (under fuel, <span class="html-italic">i</span> = 0.9 A/cm<sup>2</sup>). The cell is alternatively exposed to dry H<sub>2</sub> for 45 min and air for 30 min [<a href="#B180-membranes-02-00585" class="html-bibr">180</a>].</p>
Full article ">Figure 26
<p>Electrochemical impedance spectroscopy at open circuit voltage (OCV), of electrolyte supported cell after RedOx treatments at 950 °C of sample composition (<b>a</b>) A; (<b>b</b>) B; (<b>c</b>) C and (<b>d</b>) D described in <a href="#membranes-02-00585-t006" class="html-table">Table 6</a> [<a href="#B16-membranes-02-00585" class="html-bibr">16</a>].</p>
Full article ">Figure 27
<p>OCV, current density (<span class="html-italic">i</span>) at 0.6 V and area specific resistance (ASR) of anode-support containing 60 wt % fine NiO, 38 wt % coarse and 2 wt % fine YSZ with the number of RedOx cycles (the last cycle is done at 850 °C). Conditions: 97% H<sub>2</sub> + 3% H<sub>2</sub>O at 800 °C. Measurement done 1 h after re-reduction when not stated otherwise [<a href="#B186-membranes-02-00585" class="html-bibr">186</a>].</p>
Full article ">Figure 28
<p>Anode containing 60 wt % NiO after 300 h at 800 °C under humidified forming gas (10% H<sub>2</sub> in N<sub>2</sub>) (<b>a</b>) and (<b>c</b>): fresh as-sintered sample, and (<b>b</b>) and (<b>d</b>): tested sample from <a href="#membranes-02-00585-f027" class="html-fig">Figure 27</a>. The grey levels separate each phase (YSZ: light grey, Ni: dark grey and porosity: black). All data-bars are 2 µm in length [<a href="#B186-membranes-02-00585" class="html-bibr">186</a>].</p>
Full article ">Figure 29
<p>Maximum cumulative RedOx strain value (CRS) obtained after three isothermal cycles at 850 °C as a function of the total porosity of the Ni-YSZ composite [<a href="#B73-membranes-02-00585" class="html-bibr">73</a>].</p>
Full article ">Figure 30
<p>Expansion after one RedOx cycle at 800 °C versus porosity of 46 NiO-8YSZ anode-support samples [<a href="#B86-membranes-02-00585" class="html-bibr">86</a>].</p>
Full article ">Figure 31
<p>Percolation threshold versus (<b>a</b>) axial ratio <span class="html-italic">M</span> and (<b>b</b>) radius of conducting particles <span class="html-italic">r</span> [<a href="#B192-membranes-02-00585" class="html-bibr">192</a>].</p>
Full article ">Figure 32
<p>Anode supported cell tested over 16 RedOx cycles and 1700 h on a single repeat unit stack configuration. Test conditions: 800 °C, active surface area about 100 cm<sup>2</sup> and constant current load of 0.25 A/cm<sup>2</sup> [<a href="#B215-membranes-02-00585" class="html-bibr">215</a>].</p>
Full article ">Figure 33
<p>(La<sub>0.75</sub>Sr<sub>0.25</sub>)<sub>1−x</sub>Cr<sub>0.5</sub>Mn<sub>0.5</sub>O<sub>3</sub> (LSCM) (47.5 wt %) + CGO (47.5 wt %) + NiO (5 wt %) under CH<sub>4</sub> with 150 mW/cm<sup>2</sup> at 750 °C before and after RedOx cycles [<a href="#B248-membranes-02-00585" class="html-bibr">248</a>].</p>
Full article ">Figure 34
<p>Open circuit voltage (open circle), current density at 0.7 V (10 min in H<sub>2</sub> and 10 min in air) at 750 °C (closed circles), and current density at 800°C applying 2 h in H<sub>2</sub> and 10 min in air (closed square), all as a function of the number of RedOx cycles with SYT ceramic anodes [<a href="#B260-membranes-02-00585" class="html-bibr">260</a>,<a href="#B261-membranes-02-00585" class="html-bibr">261</a>].</p>
Full article ">Figure 35
<p>Scheme of RedOx instability of standard anode supported Ni-YSZ half-cell [<a href="#B129-membranes-02-00585" class="html-bibr">129</a>].</p>
Full article ">Figure 36
<p>Scheme of RedOx behavior of highly porous Ni-YSZ anode supported half-cell [<a href="#B129-membranes-02-00585" class="html-bibr">129</a>].</p>
Full article ">Figure 37
<p>The central role of porosity in the Ni-YSZ anode supported solid oxide fuel cells properties.</p>
Full article ">
271 KiB  
Review
Microbial Relevant Fouling in Membrane Bioreactors: Influencing Factors, Characterization, and Fouling Control
by Bing Wu and Anthony G. Fane
Membranes 2012, 2(3), 565-584; https://doi.org/10.3390/membranes2030565 - 15 Aug 2012
Cited by 50 | Viewed by 10468
Abstract
Microorganisms in membrane bioreactors (MBRs) play important roles on degradation of organic/inorganic substances in wastewaters, while microbial deposition/growth and microbial product accumulation on membranes potentially induce membrane fouling. Generally, there is a need to characterize membrane foulants and to determine their relations to [...] Read more.
Microorganisms in membrane bioreactors (MBRs) play important roles on degradation of organic/inorganic substances in wastewaters, while microbial deposition/growth and microbial product accumulation on membranes potentially induce membrane fouling. Generally, there is a need to characterize membrane foulants and to determine their relations to the evolution of membrane fouling in order to identify a suitable fouling control approach in MBRs. This review summarized the factors in MBRs that influence microbial behaviors (community compositions, physical properties, and microbial products). The state-of-the-art techniques to characterize biofoulants in MBRs were reported. The strategies for controlling microbial relevant fouling were discussed and the future studies on membrane fouling mechanisms in MBRs were proposed. Full article
(This article belongs to the Special Issue Membranes in Water Purification)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration showing the contents of this review.</p>
Full article ">
310 KiB  
Review
Polymer Electrolytes for Lithium/Sulfur Batteries
by Yan Zhao, Yongguang Zhang, Denise Gosselink, The Nam Long Doan, Mikhail Sadhu, Ho-Jae Cheang and Pu Chen
Membranes 2012, 2(3), 553-564; https://doi.org/10.3390/membranes2030553 - 9 Aug 2012
Cited by 98 | Viewed by 17369
Abstract
This review evaluates the characteristics and advantages of employing polymer electrolytes in lithium/sulfur (Li/S) batteries. The main highlights of this study constitute detailed information on the advanced developments for solid polymer electrolytes and gel polymer electrolytes, used in the lithium/sulfur battery. This includes [...] Read more.
This review evaluates the characteristics and advantages of employing polymer electrolytes in lithium/sulfur (Li/S) batteries. The main highlights of this study constitute detailed information on the advanced developments for solid polymer electrolytes and gel polymer electrolytes, used in the lithium/sulfur battery. This includes an in-depth analysis conducted on the preparation and electrochemical characteristics of the Li/S batteries based on these polymer electrolytes. Full article
Show Figures

Figure 1

Figure 1
<p>Model for morphology change of composite cathode during charge-discharge:(<b>a</b>) ideal case;(<b>b</b>) real case [<a href="#B48-membranes-02-00553" class="html-bibr">48</a>].</p>
Full article ">Figure 2
<p>Sketch of the Sn/C/CGPE/ Li<sub>2</sub>S/C polymer battery developed herein. The battery is formed by a Sn/C composite anode, a PEO based gel polymer electrolyte, and a Li<sub>2</sub>S/C cathode. PEO = poly(ethylene oxide) [<a href="#B66-membranes-02-00553" class="html-bibr">66</a>].</p>
Full article ">Figure 3
<p>Characteristics of the PEO based gel polymer membrane to be used as electrolyte separator in the lithium-sulfur battery: (<b>a</b>) Appearance of the membrane; (<b>b</b>) Time evolution of the conductivity at room temperature [<a href="#B66-membranes-02-00553" class="html-bibr">66</a>].</p>
Full article ">Figure 4
<p>The discharge and charge reaction model of lithium/sulfur cell [<a href="#B72-membranes-02-00553" class="html-bibr">72</a>].</p>
Full article ">
559 KiB  
Article
Plasma Membranes Modified by Plasma Treatment or Deposition as Solid Electrolytes for Potential Application in Solid Alkaline Fuel Cells
by Marc Reinholdt, Alina Ilie, Stéphanie Roualdès, Jérémy Frugier, Mauricio Schieda, Christophe Coutanceau, Serguei Martemianov, Valérie Flaud, Eric Beche and Jean Durand
Membranes 2012, 2(3), 529-552; https://doi.org/10.3390/membranes2030529 - 30 Jul 2012
Cited by 12 | Viewed by 8844
Abstract
In the highly competitive market of fuel cells, solid alkaline fuel cells using liquid fuel (such as cheap, non-toxic and non-valorized glycerol) and not requiring noble metal as catalyst seem quite promising. One of the main hurdles for emergence of such a technology [...] Read more.
In the highly competitive market of fuel cells, solid alkaline fuel cells using liquid fuel (such as cheap, non-toxic and non-valorized glycerol) and not requiring noble metal as catalyst seem quite promising. One of the main hurdles for emergence of such a technology is the development of a hydroxide-conducting membrane characterized by both high conductivity and low fuel permeability. Plasma treatments can enable to positively tune the main fuel cell membrane requirements. In this work, commercial ADP-Morgane® fluorinated polymer membranes and a new brand of cross-linked poly(aryl-ether) polymer membranes, named AMELI-32®, both containing quaternary ammonium functionalities, have been modified by argon plasma treatment or triallylamine-based plasma deposit. Under the concomitant etching/cross-linking/oxidation effects inherent to the plasma modification, transport properties (ionic exchange capacity, water uptake, ionic conductivity and fuel retention) of membranes have been improved. Consequently, using plasma modified ADP-Morgane® membrane as electrolyte in a solid alkaline fuel cell operating with glycerol as fuel has allowed increasing the maximum power density by a factor 3 when compared to the untreated membrane. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscopy (SEM) pictures of representative faces of (<b>a</b>) pristine ADP-Morgane<sup>®</sup>; (<b>b</b>) plasma modified ADP-Morgane<sup>®</sup> (<span class="html-italic">P<sub>w</sub></span> = 70 W, <span class="html-italic">DC</span> = 100% and <span class="html-italic">τ</span> = 10 min); (<b>c</b>) pristine AMELI-32<sup>®</sup> and (<b>d</b>) plasma modified AMELI-32<sup>®</sup> (<span class="html-italic">P<sub>w</sub></span> = 60 W, <span class="html-italic">DC</span> = 100% and <span class="html-italic">τ</span> = 20 min) membranes.</p>
Full article ">Figure 2
<p>Cross-sectional SEM pictures of a representative plasma deposited TAA polymer thin film on ADP-Morgane<sup>®</sup> membrane, performed in the following conditions: <span class="html-italic">P<sub>w</sub></span> = 40 W, <span class="html-italic">DC</span> = 10% and <span class="html-italic">τ</span> = 60 min.</p>
Full article ">Figure 3
<p>Thin films thickness as a function of deposition time (<span class="html-italic">τ</span>) both on silicon wafer and ADP-Morgane<sup>®</sup> membrane. Plasma polymerization was performed in the following conditions: (<b>a</b>) <span class="html-italic">P<sub>w</sub></span> = 150 W, <span class="html-italic">DC</span> = 100%; (<b>b</b>) <span class="html-italic">P<sub>w</sub></span> = 40 W, <span class="html-italic">DC</span> = 100%; (<b>c</b>) <span class="html-italic">P<sub>w</sub></span> = 40 W, <span class="html-italic">DC</span> = 10%. The straight dotted lines labeled as fits correspond to linear regressions of the deposit growth on silicon wafer for deposition times above 2 min (once the permanent regime reached).</p>
Full article ">Figure 4
<p>Growth rate of plasma polymers deposited on silicon wafer as a function of average input power (<span class="html-italic">P<sub>A</sub></span>, given in brackets).</p>
Full article ">Figure 5
<p>Examples of X-ray photoelectron spectroscopy (XPS) peak decomposition for (<b>a</b>) C<sub>1s</sub>; (<b>b</b>) N<sub>1s</sub>; and (<b>c</b>) O<sub>1s</sub> photoelectrons in bulk of non-quaternized films analyzed with the SIA 200 instrument. The assignment of the various components is given in <a href="#membranes-02-00529-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 6
<p>Evolution of the bonds proportion for the N<sub>1s</sub> photoelectron peaks in the bulk of non-quaternized films (analyzed with the SIA 200 instrument) as a function of the discharge power <span class="html-italic">P<sub>w</sub></span> (<span class="html-italic">DC</span> = 100%).</p>
Full article ">Figure 7
<p>Evolution of the bonds proportion for the N<sub>1s</sub> photoelectron peaks in the bulk of non-quaternized films (analyzed with the SIA 200 instrument) as a function of the duty cycle (<span class="html-italic">DC</span>) for plasma polymers prepared at (<b>a</b>) 40 W and (<b>b</b>) 150 W.</p>
Full article ">Figure 8
<p>Evolution of the different components of the N<sub>1s</sub> peak before and after quaternization, both at the surface (without erosion) and in the bulk (after erosion). Plasma deposition was performed in the following conditions: <span class="html-italic">P<sub>w</sub></span> = 40 W, <span class="html-italic">DC</span> = 10% and <span class="html-italic">τ</span> = 10 min. XPS analyses were performed on the ESCALAB 250 instrument.</p>
Full article ">Figure 9
<p>Evolution of the (<b>a</b>) water content; (<b>b</b>) ion exchange capacity and (<b>c</b>) ionic conductivity of modified ADP-Morgane<sup>®</sup> membranes as a function of the power discharge used in the preparation of plasma films (<span class="html-italic">DC</span> = 10%, <span class="html-italic">τ</span> = 10 min); (<b>d</b>) Evolution of the NaBH<sub>4</sub> diffusion coefficient of modified ADP-Morgane<sup>®</sup> membranes as a function of the deposition time of plasma films (<span class="html-italic">P<sub>w</sub></span> = 40 or 150 W, <span class="html-italic">DC</span> = 10%). Dashed and dotted lines correspond to the reference properties of unmodified ADP-Morgane<sup>®</sup> and Nafion<sup>®</sup> membranes, respectively.</p>
Full article ">Figure 10
<p>Power density curves recorded at different temperatures in SAFC fitted with MEA realized with pristine ADP-Morgane<sup>®</sup> membrane.</p>
Full article ">Figure 11
<p>Power density curves recorded in SAFCs at 80°C fitted with MEA realized with ADP-Morgane<sup>®</sup> membranes. Membrane □ is pristine ADP-Morgane<sup>®</sup>. Membranes ♦ and ▲ are one-side treated membranes at 2 and 80 W (10 min), respectively (the treated side is positioned close to the anode during fuel cell tests). Membranes ◊ and Δ are both-sides treated membranes at 2 and 80 W (10 min), respectively.</p>
Full article ">Figure 12
<p>Polarisation curves recorded at 60 °C in SAFC fitted with MEA realized with pristine ADP-Morgane<sup>®</sup> membrane (◆) and with one-side argon plasma treated membrane at 2 W for 10 min (■). <span class="html-italic">P<sub>fuel</sub></span> = <span class="html-italic">P<sub>O2</sub></span> = 1 atm, fuel flow rate = 0.8 mL min<sup>−1</sup>, O<sub>2</sub> flow rate = 40 mL min<sup>−1</sup>.</p>
Full article ">Figure 13
<p>Plasma process equipment.</p>
Full article ">
1338 KiB  
Review
Electrochemical Membrane Reactors for Sustainable Chlorine Recycling
by Tanja Vidakovic-Koch, Isai Gonzalez Martinez, Rafael Kuwertz, Ulrich Kunz, Thomas Turek and Kai Sundmacher
Membranes 2012, 2(3), 510-528; https://doi.org/10.3390/membranes2030510 - 30 Jul 2012
Cited by 24 | Viewed by 12513
Abstract
Polymer electrolyte membranes have found broad application in a number of processes, being fuel cells, due to energy concerns, the main focus of the scientific community worldwide. Relatively little attention has been paid to the use of these materials in electrochemical production and [...] Read more.
Polymer electrolyte membranes have found broad application in a number of processes, being fuel cells, due to energy concerns, the main focus of the scientific community worldwide. Relatively little attention has been paid to the use of these materials in electrochemical production and separation processes. In this review, we put emphasis upon the application of Nafion membranes in electrochemical membrane reactors for chlorine recycling. The performance of such electrochemical reactors can be influenced by a number of factors including the properties of the membrane, which play an important role in reactor optimization. This review discusses the role of Nafion as a membrane, as well as its importance in the catalyst layer for the formation of the so-called three-phase boundary. The influence of an equilibrated medium on the Nafion proton conductivity and Cl crossover, as well as the influence of the catalyst ink dispersion medium on the Nafion/catalyst self-assembly and its importance for the formation of an ionic conducting network in the catalyst layer are summarized. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic presentations of different reactor types used for HCl electrolysis: (<b>a</b>) dimensionally stable electrodes–gas diffusion electrodes (DSE–GDE) and (<b>b</b>) gas diffusion electrodes–gas diffusion electrodes (GDE–GDE) configuration.</p>
Full article ">Figure 2
<p>Influence of temperature on equilibrium cell voltage for different HCl electrolysis processes.</p>
Full article ">Figure 3
<p>Influence of equilibrated medium on Nafion conductivity in the presence of (<b>a</b>) liquid water and vapor water at different relative humidities (RH) and (<b>b</b>) hydrochloric acid at different concentrations. Based on [<a href="#B14-membranes-02-00510" class="html-bibr">14</a>,<a href="#B15-membranes-02-00510" class="html-bibr">15</a>].</p>
Full article ">Figure 4
<p>Schematic representation of the influence of an equilibrated medium concentration on the polymer electrolyte membrane structure (<b>a</b>) concentrations lower and (<b>b</b>) greater than the Donnan concentration <span class="html-italic">c<sub>D</sub></span>. Reprinted with permission from reference [<a href="#B16-membranes-02-00510" class="html-bibr">16</a>].</p>
Full article ">Figure 5
<p>Scanning electron microscopy (SEM) micrograph of a cross-section of the membrane electrode assembly, along with a profile of elemental fluorine across the layers. Reprinted with permission from [<a href="#B33-membranes-02-00510" class="html-bibr">33</a>]. Copyright 2012, The Electrochemical Society.</p>
Full article ">Figure 6
<p>Comparison of the information obtained by traditional SEM cross-section analysis and micro-computed tomography. Reprinted with permission from [<a href="#B34-membranes-02-00510" class="html-bibr">34</a>]. Copyright 2012, The Electrochemical Society.</p>
Full article ">Figure 7
<p>Influence of Nafion loading on (<b>a</b>) electrochemically active surface area (ESA) in the presence of nitrogen at 25 °C and 20 mV s<sup>−1 </sup>and (<b>b</b>) fuel cell performance at 65 °C employing GDEs comprising carbon supported platinum catalyst. Reprinted with permission from [<a href="#B37-membranes-02-00510" class="html-bibr">37</a>,<a href="#B38-membranes-02-00510" class="html-bibr">38</a>]. Copyright 2012, Elsevier.</p>
Full article ">
1338 KiB  
Article
Proton Content and Nature in Perovskite Ceramic Membranes for Medium Temperature Fuel Cells and Electrolysers
by Philippe Colomban, Oumaya Zaafrani and Aneta Slodczyk
Membranes 2012, 2(3), 493-509; https://doi.org/10.3390/membranes2030493 - 25 Jul 2012
Cited by 43 | Viewed by 8323
Abstract
Recent interest in environmentally friendly technology has promoted research on green house gas-free devices such as water steam electrolyzers, fuel cells and CO2/syngas converters. In such applications, proton conducting perovskite ceramics appear especially promising as electrolyte membranes. Prior to a successful [...] Read more.
Recent interest in environmentally friendly technology has promoted research on green house gas-free devices such as water steam electrolyzers, fuel cells and CO2/syngas converters. In such applications, proton conducting perovskite ceramics appear especially promising as electrolyte membranes. Prior to a successful industrial application, it is necessary to determine/understand their complex physical and chemical behavior, especially that related to proton incorporation mechanism, content and nature of bulk protonic species. Based on the results of quasi-elastic neutron scattering (QNS), thermogravimetric analysis (TGA), Raman and IR measurements we will show the complexity of the protonation process and the importance of differentiation between the protonic species adsorbed on a membrane surface and the bulk protons. The bulk proton content is very low, with a doping limit (~1–5 × 10−3 mole/mole), but sufficient to guarantee proton conduction below 600 °C. The bulk protons posses an ionic, covalent bond free nature and may occupy an interstitial site in the host perovskite structure. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Sketch of protonation process: (1) dissociation of water molecules at the surface of the oxygen-deficient perovskite; (2) filling of oxygen vacancies near the ceramic surface at temperature T<sub>1</sub>; a few days are required; (3) a diffusion of protons and V<sub>O</sub>/O<sup>2−</sup> species through a ceramic is required to obtain homogeneous protonation; (4) stable, quenched state after protonation and cooling at T<sub>o</sub> &lt; T<sub>1</sub>; (<b>b</b>) Sketch of protonated ceramic: homogeneous distribution of bulk protonic species between a ceramic surface and bulk accompanied by the presence of protonic moieties adsorbed to the surface.</p>
Full article ">Figure 2
<p>Surface of SZYb high dense ceramic (99%) (<b>a</b>) before; and (<b>b</b>) after protonation at 200 °C under 15 bar during 96 h; low temperature, low-pressure protonation conditions and the use of carbonate-containing water lead to the appearance of SrCO<sub>3</sub> crystals. The Raman spectra characteristic of the ceramic surface before protonation—vibrational signature of perovskite structure, and after—vibrational signature of SrCO<sub>3</sub> (<a href="#membranes-02-00493-t002" class="html-table">Table 2</a>) are given; (<b>c</b>) Surface of BZYb sample protonated during 23 days at 200 °C under 15 bar pH<sub>2</sub>O-important layer of carbonates and hydroxides is well seen [<a href="#B24-membranes-02-00493" class="html-bibr">24</a>]; (<b>d</b>) Comparison of the BZI sample [<a href="#B26-membranes-02-00493" class="html-bibr">26</a>] before protonation—ceramic piece, and after protonation—almost powder. Its low density and the presence of secondary phases lead to crumbling and chemical decomposition. The IR spectra characteristic of the ceramic surface before protonation—vibrational signature of perovskite structure, and after—vibrational signature of BaCO<sub>3</sub>/Ba(OH)<sub>2</sub>, nH<sub>2</sub>O (<a href="#membranes-02-00493-t002" class="html-table">Table 2</a>) are given.</p>
Full article ">Figure 3
<p>Schematic summarizing the H bulk content as a function of protonation conditions (temperature, water pressure) and sample parameters (densification, <span class="html-italic">i.e</span>., the active surface area/porosity).</p>
Full article ">Figure 4
<p>(<b>a</b>) Elastic neutron spectra measured at RT, 400 °C and 660 °C under high vacuum using time-of-flight spectrometer; (<b>b</b>) thermogravimetric (TG) curves recorded for the ~ 94% dense ceramic and the very high dense (~99%) ceramic (He atmosphere). Samples were preliminarily dried at 300 °C to eliminate traces of water molecules adsorbed on the ceramic surface.</p>
Full article ">Figure 5
<p>Raman profilometry method: Spectra with characteristic backgrounds of (<b>a</b>) Ln-modified BaZrO<sub>3−δ</sub> (protonation at 250 °C under 40 bar during 72 h-protons near surface only); (<b>b</b>) Ln-modified SrZrO<sub>3−δ</sub> (protonation at 500 °C under 80 bar during 96 h homogeneous distributions of protons) were collected along the ceramic section, from the surface to the centre, just after fracture [<a href="#B23-membranes-02-00493" class="html-bibr">23</a>]; (<b>c</b>) The distributions of the spectroscopic signal measured across the half section (sample thickness = 1 mm) follows Fick’s law.</p>
Full article ">Figure 6
<p>Neutronography micrographs characteristic of two Ln-modified SrZrO<sub>3−δ</sub>ceramics. Hydrogen-rich regions appear in white. (<b>a</b>) On the ceramic surface (250 °C, 40 bar, 72 h); (<b>b</b>) throughout the ceramic (310 °C, 90 bar, 96 h).</p>
Full article ">Figure 7
<p>(<b>a</b>) IR spectra recorded on highly dense (99%) polished, thin (~120 µm) pristine (non-protonated), protonated and deprotonated Ln-modified SrZrO<sub>3−δ</sub> ceramics; (<b>b</b>) Comparison of IR spectra characteristic of low dense (94%) polished, thin Ln-modified SrZrO<sub>3−δ</sub> ceramic with a Sr(OH)<sub>x</sub>(CO<sub>3</sub>)<sub>y</sub>, nH<sub>2</sub>O powder dispersed in CsI matrix (see <a href="#membranes-02-00493-t002" class="html-table">Table 2</a> for band wave numbers and assignments).</p>
Full article ">
901 KiB  
Article
Anion- or Cation-Exchange Membranes for NaBH4/H2O2 Fuel Cells?
by Biljana Šljukić, Ana L. Morais, Diogo M. F. Santos and César A. C. Sequeira
Membranes 2012, 2(3), 478-492; https://doi.org/10.3390/membranes2030478 - 19 Jul 2012
Cited by 27 | Viewed by 13002
Abstract
Direct borohydride fuel cells (DBFC), which operate on sodium borohydride (NaBH4) as the fuel, and hydrogen peroxide (H2O2) as the oxidant, are receiving increasing attention. This is due to their promising use as power sources for space [...] Read more.
Direct borohydride fuel cells (DBFC), which operate on sodium borohydride (NaBH4) as the fuel, and hydrogen peroxide (H2O2) as the oxidant, are receiving increasing attention. This is due to their promising use as power sources for space and underwater applications, where air is not available and gas storage poses obvious problems. One key factor to improve the performance of DBFCs concerns the type of separator used. Both anion- and cation-exchange membranes may be considered as potential separators for DBFC. In the present paper, the effect of the membrane type on the performance of laboratory NaBH4/H2O2 fuel cells using Pt electrodes is studied at room temperature. Two commercial ion-exchange membranes from Membranes International Inc., an anion-exchange membrane (AMI-7001S) and a cation-exchange membrane (CMI-7000S), are tested as ionic separators for the DBFC. The membranes are compared directly by the observation and analysis of the corresponding DBFC’s performance. Cell polarization, power density, stability, and durability tests are used in the membranes’ evaluation. Energy densities and specific capacities are estimated. Most tests conducted, clearly indicate a superior performance of the cation-exchange membranes over the anion-exchange membrane. The two membranes are also compared with several other previously tested commercial membranes. For long term cell operation, these membranes seem to outperform the stability of the benchmark Nafion membranes but further studies are still required to improve their instantaneous power load. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the major migrative and diffusive fluxes across (<b>a</b>) anion and (<b>b</b>) cation exchange membranes used in direct borohydride/peroxide fuel cells (DBPFCs).</p>
Full article ">Figure 2
<p>SEM micrographs (×1000) of (<b>a</b>) AMI-7001S surface; (<b>b</b>) CMI-7000S surface; (<b>c</b>)AMI-7001S cross section and (<b>d</b>) CMI-7000S cross section.</p>
Full article ">Figure 3
<p>Polarization curves for DBPFCs employing the two tested membranes, including (<b>a</b>) cell voltage and power density curves and (<b>b</b>) the corresponding electrodes’ potentials.</p>
Full article ">Figure 4
<p>Current density <span class="html-italic">versus</span> time at an imposed cell voltage of 0.6 V for DBPFCs employing AMI-7001S and CMI-7000S membrane separators.</p>
Full article ">Figure 5
<p>Performance stability of DBPFCs employing AMI-7001S and CMI-7000S membrane separators recorded at the operating current density of 50 mA·cm<sup>−2</sup>.</p>
Full article ">Figure 6
<p>Durability tests for DBPFCs employing AMI-7001S and CMI-7000S membrane separators, at zero current flow, with (<b>a</b>) the cell voltage decay and (<b>b</b>) the corresponding electrodes’ potentials.</p>
Full article ">Figure 7
<p>Full discharge of DBPFCs employing AMI-7001S and CMI-7000S membrane separators recorded at the operating current density of 30 mA·cm<sup>−2</sup>.</p>
Full article ">
806 KiB  
Review
Molecularly Imprinted Membranes
by Francesco Trotta, Miriam Biasizzo and Fabrizio Caldera
Membranes 2012, 2(3), 440-477; https://doi.org/10.3390/membranes2030440 - 19 Jul 2012
Cited by 40 | Viewed by 10143
Abstract
Although the roots of molecularly imprinted polymers lie in the beginning of 1930s in the past century, they have had an exponential growth only 40–50 years later by the works of Wulff and especially by Mosbach. More recently, it was also proved that [...] Read more.
Although the roots of molecularly imprinted polymers lie in the beginning of 1930s in the past century, they have had an exponential growth only 40–50 years later by the works of Wulff and especially by Mosbach. More recently, it was also proved that molecular imprinted membranes (i.e., polymer thin films) that show recognition properties at molecular level of the template molecule are used in their formation. Different procedures and potential application in separation processes and catalysis are reported. The influences of different parameters on the discrimination abilities are also discussed. Full article
(This article belongs to the Special Issue Responsive Polymer Membranes)
Show Figures

Figure 1

Figure 1
<p>Uptaken amount of TCH and chloramphenicol in the MIM for TCH and NMIM: (■) tetracycline hydrochloride into P(AA-co-AN) NMIM; (●) TCH in the P(AA-co-AN) MIM; (▲) chloramphenicol in the P(AA-co-AN) MIM; ( <span class="html-fig-inline" id="membranes-02-00440-i001"> <img alt="Membranes 02 00440 i001" src="/membranes/membranes-02-00440/article_deploy/html/images/membranes-02-00440-i001.png"/></span>) chloramphenicol in the NMIM.</p>
Full article ">Figure 2
<p>(<b>a</b>) pH effect on the binding capability of blank and MIM P(AN-co-AA); (<b>b</b>) retention of folic acid by PAN and P(AN-co-AA) prepared by solvent evaporation.</p>
Full article ">Figure 3
<p>Scheme of NG-MIM preparation.</p>
Full article ">Figure 4
<p>(<b>a</b>) NG retention on poly(4-VP/PVDF) NR-imprinted and its corresponding blank membrane at different modification degree; (<b>b</b>) Binding properties of NG-imprinted membrane and its corresponding blank.</p>
Full article ">Figure 5
<p>Selectivity activity of MIM and NMIM towards mixture containing lysozyme and BHb (<b>a</b>) and Cyt c (<b>b</b>).</p>
Full article ">Figure 6
<p>Frequency change of the <span class="html-small-caps">L</span>-Gln imprinted polymer-QCM sensor for <span class="html-small-caps">L</span>-Gln and <span class="html-small-caps">D</span>-Gln.</p>
Full article ">Figure 7
<p>Separation of the <span class="html-small-caps">D</span>- and <span class="html-small-caps">L</span>-serine by the MIPCM from Ser racemate as a function of operating time as the operating pressure was varied from 0.5 to 2 bars.</p>
Full article ">Figure 8
<p>(<b>a</b>) Influence of the reduced volume of MIP on the resonance frequency of the array after rebinding of 2,4-D; (<b>b</b>) Detection of the rebinding of 2,4-D and POAc at increasing concentrations on a 2,4-D MIP and a NMIP.</p>
Full article ">Figure 9
<p>4,4'-methylenedianiline, (MDA) retention and specific binding capacities of the membranes prepared using PAN and the different acrylic copolymers.</p>
Full article ">Figure 10
<p>Permeability of PSf membranes with respect to bisphenol A. PSU has degree of sulfonation = 0.00 mol/mer; PSU1 has degree of sulfonation = 0.06 mol/mer; PSU2 has degree of sulfonation = 0.26 mol/mer; 5% is the amount of template (5 wt %).</p>
Full article ">
1080 KiB  
Article
Poly(imide)/Organically-Modified Montmorillonite Nanocomposite as a Potential Membrane for Alkaline Fuel Cells
by Liliane C. Battirola, Luiz H. S. Gasparotto, Ubirajara P. Rodrigues-Filho and Germano Tremiliosi-Filho
Membranes 2012, 2(3), 430-439; https://doi.org/10.3390/membranes2030430 - 18 Jul 2012
Cited by 8 | Viewed by 6866
Abstract
In this work we evaluated the potentiality of a poly(imide) (PI)/organically-modified montmorillonite (O-MMT) nanocomposite membrane for the use in alkaline fuel cells. Both X-ray diffraction and scanning electron microscopy revealed a good dispersion of O-MMT into the PI matrix and preservation of the [...] Read more.
In this work we evaluated the potentiality of a poly(imide) (PI)/organically-modified montmorillonite (O-MMT) nanocomposite membrane for the use in alkaline fuel cells. Both X-ray diffraction and scanning electron microscopy revealed a good dispersion of O-MMT into the PI matrix and preservation of the O-MMT layered structure. When compared to the pure PI, the addition of O-MMT improved thermal stability and markedly increased the capability of absorbing electrolyte and ionic conductivity of the composite. The results show that the PI/O-MMT nanocomposite is a promising candidate for alkaline fuel cell applications. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Fourier transform infrared spectroscopy (FTIR) spectra of pure pre-polymer (PAA) and poly(imide) (PI). (<b>A</b>) Full-range spectra and (<b>B</b>) expended low-frequency region. Vertical full and dotted lines in (<b>B</b>) indicate bands that vanished and appeared, respectively, as a consequence of the polymerization process.</p>
Full article ">Figure 2
<p>FTIR spectra of pure PI (full line) and PI/organically-modified montmorillonite (O-MMT) (dashed line).</p>
Full article ">Figure 3
<p>X-ray diffraction patterns of pure O-MMT, pure PI and PI/O-MMT.</p>
Full article ">Figure 4
<p>Scanning electron microscopy (SEM) micrograph of the PI/O-MMT nanocomposite.</p>
Full article ">Figure 5
<p>Thermal gravimetric analysis (TGA) (<b>A</b>) and DTGA (<b>B</b>) curves for the materials employed in this work. The identity of the samples is indicated in the figures.</p>
Full article ">Figure 6
<p>Aqueous 1.5 mol L<sup>−1</sup> KOH uptake by PI and PI/O-MMT nanocomposite.</p>
Full article ">Figure 7
<p>Conductivity for pure PI and PI/-O-MMT nanocomposite as a function of temperature.</p>
Full article ">
1311 KiB  
Article
Effectiveness of Water Desalination by Membrane Distillation Process
by Marek Gryta
Membranes 2012, 2(3), 415-429; https://doi.org/10.3390/membranes2030415 - 17 Jul 2012
Cited by 93 | Viewed by 12366
Abstract
The membrane distillation process constitutes one of the possibilities for a new method for water desalination. Four kinds of polypropylene membranes with different diameters of capillaries and pores, as well as wall thicknesses were used in studied. The morphology of the membrane used [...] Read more.
The membrane distillation process constitutes one of the possibilities for a new method for water desalination. Four kinds of polypropylene membranes with different diameters of capillaries and pores, as well as wall thicknesses were used in studied. The morphology of the membrane used and the operating parameters significantly influenced process efficiency. It was found that the membranes with lower wall thickness and a larger pore size resulted in the higher yields. Increasing both feed flow rate and temperature increases the permeate flux and simultaneously the process efficiency. However, the use of higher flow rates also enhanced heat losses by conduction, which decreases the thermal efficiency. This efficiency also decreases when the salt concentration in the feed was enhanced. The influence of fouling on the process efficiency was considered. Full article
(This article belongs to the Special Issue Energy Efficient Membranes)
Show Figures

Figure 1

Figure 1
<p>The heat and mass transfer in the DCMD variant of MD process.</p>
Full article ">Figure 2
<p>Heat transfer in the DCMD process.</p>
Full article ">Figure 3
<p>The heat transfer in the MD module with partially wetted membranes.</p>
Full article ">Figure 4
<p>MD experimental set-up: 1: MD module, 2: distillate tank, 3: feed tank, 4: pump, 5: heat exchanger, 6: manometer, 7: balance, T: thermometer.</p>
Full article ">Figure 5
<p>SEM images of cross section of Accurel PP membranes.</p>
Full article ">Figure 6
<p>SEM images of Accurel PP membrane surfaces (bore side).</p>
Full article ">Figure 7
<p>The influence of feed temperature and kind of used membranes on the permeate flux and thermal efficiency. Accurel PP capillary polypropylene membranes: S6/2, S6/4 and Q3/2. Feed–distilled water. T<sub>D</sub>=20 °C, m<sub>F</sub>=m<sub>D</sub>=0.014 dm<sup>3</sup>/s.</p>
Full article ">Figure 8
<p>The influence of feed temperature and kind of used membranes on the permeate flux and thermal efficiency. Accurel PP capillary polypropylene membranes: S6/2 and V8/2 HF (0.58 cm-<a href="#membranes-02-00415-t001" class="html-table">Table 1</a>). Feed–distilled water. T<sub>D</sub>= 20 °C, m<sub>F</sub>=m<sub>D</sub>=0.014 dm<sup>3</sup>/s.</p>
Full article ">Figure 9
<p>The influence of feed temperature and flow side on the permeate flux and thermal efficiency. Membrane Accurel PP V8/2 HF (0.71 cm—<a href="#membranes-02-00415-t001" class="html-table">Table 1</a>). T<sub>D</sub>= 20 °C, m<sub>F</sub>=m<sub>D</sub>=0.014 dm<sup>3</sup>/s.</p>
Full article ">Figure 10
<p>The influence of feed flow rate on the permeate flux and thermal efficiency. Accurel PP S6/2 membranes. Feed—distilled water. T<sub>F</sub>= 80 °C, T<sub>D</sub>= 20 °C.</p>
Full article ">Figure 11
<p>Changes of electrical conductivity of distillate obtained during water desalination by the MD process for different Accurel PP membranes. Feed–salt solution (170 g NaCl/dm<sup>3</sup>). T<sub>F</sub>= 80 °C, T<sub>D</sub>= 20 °C.</p>
Full article ">Figure 12
<p>The influence of feed temperature and salt concentration on the permeate flux and thermal efficiency. Accurel PP S6/2. T<sub>D</sub>= 20 °C, m<sub>F</sub>=m<sub>D</sub>=0.014 dm<sup>3</sup>/s.</p>
Full article ">Figure 13
<p>The influence of feed temperature and salt concentration on water vapor pressure. Data for calculation taken from [<a href="#B14-membranes-02-00415" class="html-bibr">14</a>].</p>
Full article ">
751 KiB  
Review
A Review of Molecular-Level Mechanism of Membrane Degradation in the Polymer Electrolyte Fuel Cell
by Takayoshi Ishimoto and Michihisa Koyama
Membranes 2012, 2(3), 395-414; https://doi.org/10.3390/membranes2030395 - 10 Jul 2012
Cited by 51 | Viewed by 12656
Abstract
Chemical degradation of perfluorosulfonic acid (PFSA) membrane is one of the most serious problems for stable and long-term operations of the polymer electrolyte fuel cell (PEFC). The chemical degradation is caused by the chemical reaction between the PFSA membrane and chemical species such [...] Read more.
Chemical degradation of perfluorosulfonic acid (PFSA) membrane is one of the most serious problems for stable and long-term operations of the polymer electrolyte fuel cell (PEFC). The chemical degradation is caused by the chemical reaction between the PFSA membrane and chemical species such as free radicals. Although chemical degradation of the PFSA membrane has been studied by various experimental techniques, the mechanism of chemical degradation relies much on speculations from ex-situ observations. Recent activities applying theoretical methods such as density functional theory, in situ experimental observation, and mechanistic study by using simplified model compound systems have led to gradual clarification of the atomistic details of the chemical degradation mechanism. In this review paper, we summarize recent reports on the chemical degradation mechanism of the PFSA membrane from an atomistic point of view. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Structure of Nafion<sup>®</sup> membrane.</p>
Full article ">Figure 2
<p>Plot showing relative fluoride emission rate (FER) from Fenton’s test as a function of concentration of reactive end-groups (taken from in reference [<a href="#B82-membranes-02-00395" class="html-bibr">82</a>]).</p>
Full article ">Figure 3
<p>Molecular model of perfluorosulfonic acid (PFSA) polymer with atom labels (taken from reference [<a href="#B137-membranes-02-00395" class="html-bibr">137</a>]).</p>
Full article ">Figure 4
<p>Potential energy profile under high humidity condition, CF<sub>3</sub>(CF<sub>2</sub>)<sub>3</sub>O(CF<sub>2</sub>)<sub>2</sub>OCF<sub>2</sub>SO<sub>3</sub><sup>−</sup> + OH. The energy values (kJ/mol) are relative to reactants. Optimized structures of reactants, products, intermediate, and transition state are also shown in the potential energy profile. Important distances are shown in Å.</p>
Full article ">Figure 5
<p>Potential energy profile under low humidity condition, CF<sub>3</sub>(CF<sub>2</sub>)<sub>3</sub>O(CF<sub>2</sub>)<sub>2</sub>OCF<sub>2</sub>SO<sub>3</sub>H +·OH. The energy values (kJ/mol) are relative to reactants. Optimized structures of reactants, products, intermediate, and transition state are also shown in the potential energy profile. Important distances are shown in Å.</p>
Full article ">Figure 6
<p>Potential energy profile under high humidity condition, model structure of Nafion side chain +·OH. The energy values (kJ/mol) are relative to reactants. Optimized structures of reactants, products, intermediate, and transition state are also shown in the potential energy profile. Important distances are shown in Å.</p>
Full article ">Figure 7
<p>Optimized geometry of Nafion<sup>®</sup> and their molecular orbitals: (<b>a</b>) highest occupied molecular orbital (HOMO) and (<b>b</b>) lowest unoccupied molecular orbital (LUMO) (taken from reference [<a href="#B146-membranes-02-00395" class="html-bibr">146</a>]).</p>
Full article ">Figure 8
<p>Degradation mechanism of Nafion<sup>®</sup> proposed by Yu <span class="html-italic">et al.</span> (taken from reference [<a href="#B147-membranes-02-00395" class="html-bibr">147</a>]).</p>
Full article ">
647 KiB  
Article
A Composite Membrane of Caesium Salt of Heteropolyacids/Quaternary Diazabicyclo-Octane Polysulfone with Poly (Tetrafluoroethylene) for Intermediate Temperature Fuel Cells
by Chenxi Xu, Xu Wang, Xu Wu, Yuancheng Cao and Keith Scott
Membranes 2012, 2(3), 384-394; https://doi.org/10.3390/membranes2030384 - 10 Jul 2012
Cited by 6 | Viewed by 7011
Abstract
Inorganic-organic composite electrolyte membranes were fabricated from CsXH3−XPMo12O40 (CsPOMo) and quaternary diazabicyclo-octane polysulfone (QDPSU) using a polytetrafluoroethylene (PTFE) porous matrix for the application of intermediate temperature fuel cells. The CsPOMo/QDPSU/PTFE composite membrane was made proton conducting [...] Read more.
Inorganic-organic composite electrolyte membranes were fabricated from CsXH3−XPMo12O40 (CsPOMo) and quaternary diazabicyclo-octane polysulfone (QDPSU) using a polytetrafluoroethylene (PTFE) porous matrix for the application of intermediate temperature fuel cells. The CsPOMo/QDPSU/PTFE composite membrane was made proton conducting by using a relatively low phosphoric acid loading, which benefits the stability of the membrane conductivity and the mechanical strength. The casting method was used in order to build a thin and robust composite membrane. The resulting composite membrane films were characterised in terms of the elemental composition, membrane structure and morphology by EDX, FTIR and SEM. The proton conductivity of the membrane was 0.04 S cm−1 with a H3PO4 loading level of 1.8 PRU (amount of H3PO4 per repeat unit of polymer QDPSU). The fuel cell performance with the membrane gave a peak power density of 240 mW cm−2 at 150 °C and atmospheric pressure. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>SEM of (<b>a</b>) CsPOMo/QDPSU/PTFE/H<sub>3</sub>PO<sub>4</sub> composite membrane; and (<b>b</b>) PTFE.</p>
Full article ">Figure 2
<p>EDX analysis of CsPOMo/QDPSU/PTFE/H<sub>3</sub>PO<sub>4</sub> composite membrane. (<b>a</b>) Cs; (<b>b</b>) S; (<b>c</b>) F; (<b>d</b>) P; (<b>e</b>) Mo.</p>
Full article ">Figure 3
<p>Infrared spectra of CsPOMo-PSU-PTFE-H<sub>3</sub>PO<sub>4</sub> composite membrane.</p>
Full article ">Figure 4
<p>Conductivities of CsPOMo/PSU/PTFE composite membrane and polybenzimidazole (PBI) membrane loaded with H<sub>3</sub>PO<sub>4</sub> (PRU 1.8) under relative humidity &lt;1%. The thickness of both membranes is 30μm.</p>
Full article ">Figure 5
<p>Polarization and power density curves of a fuel cell operated at 150 °C with (<b>a</b>) H<sub>2</sub>/O<sub>2</sub>; and (<b>b</b>) H<sub>2</sub>/air atmospheric pressure. Pt loading: cathode 0.4 mg cm<sup>−2</sup>; anode 0.2 mg cm<sup>−2</sup>; no gas humidity, H<sub>3</sub>PO<sub>4</sub> PRU: 1.8, membrane thickness 28 μm. Gas rate: anode: 40 dm<sup>3</sup> min<sup>−1</sup>; cathode: 70 dm<sup>3</sup> min<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Polarization curves of CsPOMo/QDPSU/PTFE and PBI fuel cell operated at 150 °C with H<sub>2</sub>/O<sub>2</sub>. Atmospheric pressure, no gas humidity.</p>
Full article ">Figure 7
<p>(<b>a</b>) IR corrected polarization curves of CsPOMo/PSU/PTFE; (<b>b</b>) Tafel plots obtained from polarization curves (I is current density).</p>
Full article ">
1735 KiB  
Review
Membranes in Lithium Ion Batteries
by Min Yang and Junbo Hou
Membranes 2012, 2(3), 367-383; https://doi.org/10.3390/membranes2030367 - 4 Jul 2012
Cited by 159 | Viewed by 25318
Abstract
Lithium ion batteries have proven themselves the main choice of power sources for portable electronics. Besides consumer electronics, lithium ion batteries are also growing in popularity for military, electric vehicle, and aerospace applications. The present review attempts to summarize the knowledge about some [...] Read more.
Lithium ion batteries have proven themselves the main choice of power sources for portable electronics. Besides consumer electronics, lithium ion batteries are also growing in popularity for military, electric vehicle, and aerospace applications. The present review attempts to summarize the knowledge about some selected membranes in lithium ion batteries. Based on the type of electrolyte used, literature concerning ceramic-glass and polymer solid ion conductors, microporous filter type separators and polymer gel based membranes is reviewed. Full article
(This article belongs to the Special Issue Membranes for Electrochemical Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Ragone plots (power <span class="html-italic">vs</span><span class="html-italic">.</span> energy density) for different rechargeable batteries [<a href="#B3-membranes-02-00367" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Schematic of a lithium ion battery (LIB) consisting of the negative electrode (graphitic carbon) and positive electrode (Li-intercalation compound) [<a href="#B5-membranes-02-00367" class="html-bibr">5</a>].</p>
Full article ">Figure 3
<p>Charge and discharge curves of Li/PEO/LTAP/LiCoO<sub>2</sub> cells at 50 °C at different annealing temperatures: (<b>a</b>) as-deposited; (<b>b</b>) 300 °C; (<b>c</b>) 400 °C and (<b>d</b>) 500 °C [<a href="#B15-membranes-02-00367" class="html-bibr">15</a>].</p>
Full article ">Figure 4
<p>Arrhenius plot for the lithium-ion conductivity of La<sub>0.281</sub>Li<sub>0.155</sub>TaO<sub>3</sub> compared with data for Li<sub>6</sub>BaLa<sub>2</sub>Ta<sub>2</sub>O<sub>12</sub> and lithium phosphorus oxy-nitride (LiPON) [<a href="#B27-membranes-02-00367" class="html-bibr">27</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Charge-discharge curves and cycling performance at 64μAcm<sup>−2</sup> for the 500th cycle of In/LiCoO<sub>2</sub> cells with 80Li<sub>2</sub>S-20P<sub>2</sub>S<sub>5</sub> glass-ceramic [<a href="#B28-membranes-02-00367" class="html-bibr">28</a>]; (<b>b</b>) Charge-discharge curves of the all-solid-state Li-In/70Li<sub>2</sub>S-27P<sub>2</sub>S<sub>5</sub>-3P<sub>2</sub>O<sub>5</sub>/Li<sub>4</sub>Ti<sub>5</sub>O<sub>12</sub> cell (discharge always at 64μAcm<sup>−2</sup>) [<a href="#B35-membranes-02-00367" class="html-bibr">35</a>].</p>
Full article ">Figure 6
<p>The concept of material design for the lithium superionic conductor (LISICON) system, and materials belonging to the LISICON (oxides) and the thio-LISICON (sulfides) are summarized [<a href="#B36-membranes-02-00367" class="html-bibr">36</a>].</p>
Full article ">Figure 7
<p>Schematic of the segmental motion assisted diffusion of Li ions in the poly(ethylene oxide) (PEO) matrix. The circles represent the ether oxygens of PEO [<a href="#B45-membranes-02-00367" class="html-bibr">45</a>].</p>
Full article ">Figure 8
<p>Scanningelectron micrographs (SEMs) of (<b>a</b>) Celgard separator using dry process; (<b>b</b>) Asahi separator using wet process; (<b>c</b>) Entek separator using wet process; (<b>d</b>) Tonen separator using wet process [<a href="#B49-membranes-02-00367" class="html-bibr">49</a>].</p>
Full article ">Figure 9
<p>Schematic and SEMs of a Degussa composite separator [<a href="#B54-membranes-02-00367" class="html-bibr">54</a>].</p>
Full article ">Figure 10
<p>Schematic representation of (<b>a</b>) a chemical gel network with junction points; (<b>b</b>) physical gel networks having junction zones and (<b>c</b>) fringed micelles, respectively [<a href="#B55-membranes-02-00367" class="html-bibr">55</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Surface SEM of the composite nonwoven separator;the inset is a photograph of Poly(methyl methacrylate) (PMMA) nanoparticles suspension; (<b>b</b>) Cross-section SEM; (<b>c</b>) AFM photograph of the composite nonwoven separator; (<b>d</b>) Schematic illustration of nanoporous structure [<a href="#B69-membranes-02-00367" class="html-bibr">69</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) Initial charge-discharge curves for the cells with the Celgard membrane and the Poly(acrylonitrile) (PAN) nonwoven membranes; (<b>b</b>) Discharge capacities <span class="html-italic">vs.</span> cycle numbers of the test cells at the 0.5C rate [<a href="#B78-membranes-02-00367" class="html-bibr">78</a>].</p>
Full article ">
733 KiB  
Article
Pure and Modified Co-Poly(amide-12-b-ethylene oxide) Membranes for Gas Separation Studied by Molecular Investigations
by Luana De Lorenzo, Elena Tocci, Annarosa Gugliuzza and Enrico Drioli
Membranes 2012, 2(3), 346-366; https://doi.org/10.3390/membranes2030346 - 28 Jun 2012
Cited by 18 | Viewed by 7612
Abstract
This paper deals with a theoretical investigation of gas transport properties in a pure and modified PEBAX block copolymer membrane with N-ethyl-o/p-toluene sulfonamide (KET) as additive molecules. Molecular dynamics simulations using COMPASS force field, Gusev-Suter Transition State Theory (TST) and Monte Carlo [...] Read more.
This paper deals with a theoretical investigation of gas transport properties in a pure and modified PEBAX block copolymer membrane with N-ethyl-o/p-toluene sulfonamide (KET) as additive molecules. Molecular dynamics simulations using COMPASS force field, Gusev-Suter Transition State Theory (TST) and Monte Carlo methods were used. Bulk models of PEBAX and PEBAX/KET in different copolymer/additive compositions were assembled and analyzed to evaluate gas permeability and morphology to characterize structure-performance relationships. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>3D Equilibrated simulation models of (<b>a</b>) PEBAX/30 wt % KET; (<b>b</b>) PEBAX/50 wt % KET; (<b>c</b>) PEBAX/70 wt % KET. The region of free volume inside the systems is colored in yellow.</p>
Full article ">Figure 2
<p>Polymer chain mean squared displacements curves (MSD) in pure PEBAX (time = 1.5 ns) and PEBAX/KET models (time = 7 ns).</p>
Full article ">Figure 3
<p>Transition-state theory (TST) calculated and experimental permeability data in modified PEBAX/KET models.</p>
Full article ">Figure 4
<p>Selective permeability α(CO<sub>2</sub>/N<sub>2</sub>) and α(CO<sub>2</sub>/O<sub>2</sub>) in modified PEBAX/KET systems.</p>
Full article ">Figure 5
<p>TST and GCMC solubility coefficients <span class="html-italic">vs.</span> Critical Temperature (<span class="html-italic">Tc</span>) of H<sub>2</sub>, N<sub>2</sub>, O<sub>2</sub>; CH<sub>4</sub>, CO<sub>2</sub>, H<sub>2</sub>O. Experimental data are available and reported only for CO<sub>2 </sub>and H<sub>2</sub>O gases.</p>
Full article ">Figure 6
<p>Solubility selectivity of (<b>a</b>) CO<sub>2</sub> and (<b>b</b>) H<sub>2</sub>O to non-polar gases at different concentrations of KET.</p>
Full article ">Figure 7
<p>Experimental and theoretical diffusivity of H<sub>2</sub>O in PEBAX membrane with different wt % of KET.</p>
Full article ">Figure 8
<p>Radial distribution functions (RDF) between the oxygen of CO<sub>2</sub> and (<b>a</b>) the N of PA-12 amidic group; (<b>b</b>) the O of the PTMO group; (<b>c</b>) the HN of the sulphonamidic group of KET in PEBAX/KET models.</p>
Full article ">Figure 9
<p>Molecular Structures of PEBAX chain repeat unit and <span class="html-italic">N</span>-ethyl-o/p-toluene sulfonamide (KET) modifier.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop