scholarly journals Mg/Al double-metal hydroxide regeneration of anion exchange resin by electric field intensification

2016 ◽  
Vol 75 (6) ◽  
pp. 1309-1318 ◽  
Author(s):  
Ying Wang ◽  
Zhun Li ◽  
Yansheng Li ◽  
Zhigang Liu

Fouled anion exchange resins were regenerated by electric field intensification of Mg/Al double-metal hydroxides. Regenerative experiments were performed with varying voltages (10–30 V) and dosages of Mg/Al hydroxides (0.045–0.135 mol and 0.015–0.045 mol, respectively) for 1–5 h. Optimal results were obtained under the following regenerative conditions: 20 V, 4 h, and 0.09/0.03 mol of Mg/Al hydroxides. The maximum regenerative capacity of resins was increased to 41.07%. The regenerative mechanism was presented by Fourier-transform infrared spectrum of resins and Mg/Al hydroxides, and the regenerative degree was analyzed with respect to conductivity, pH value, and electric current. Mg/Al hydroxides were also recycled after the regeneration. This method was proven to be cost-effective and environmentally friendly.

2008 ◽  
Vol 26 (9) ◽  
pp. 693-703 ◽  
Author(s):  
P. Senthil Kumar ◽  
K. Kirthika ◽  
K. Sathish Kumar

The removal of hexavalent chromium, Cr(VI), from aqueous solutions under different conditions using an anion-exchange resin (AXR) as an adsorbent was investigated under batch conditions. Such studies indicated that the percentage adsorption decreased with increasing initial Cr(VI) concentration, with the maximum removal of such ions occurred at a pH value of ca. 2.0. Both the Langmuir and Freundlich isotherm models were capable of reproducing the isotherms obtained experimentally. The sorption process was rapid during the first 20 min with equilibrium being attained within 30 min. The process followed first-order kinetics. The results demonstrate that such anion-exchange resins can be used for the efficient removal of Cr(VI) ions from water and wastewater.


Weed Science ◽  
2007 ◽  
Vol 55 (2) ◽  
pp. 157-163
Author(s):  
Glenn B. Fain ◽  
Timothy L. Grey ◽  
Glenn R. Wehtje ◽  
Charles H. Gilliam ◽  
Jason A. Osborne

Two polymeric anion-exchange resins and one sorbent resin were evaluated for their propensity to adsorb, and subsequently desorb, oryzalin. The intent was to determine whether these resins could adsorb and subsequently release oryzalin in a manner that would render these resins as an option for slow-release herbicide delivery. The dinitroaniline herbicide oryzalin is weakly acidic with a dissociation constant (pKa) of 8.6. An additional objective was to determine whether altering the pH between sorption and desorption would enhance the desired performance. Maximum oryzalin sorption by the two anion-exchange resins was between 127 and 132 mg g−1ai. The sorbent resin was adsorbed at a maximum concentration of 191 mg g−1ai. Maximum sorption occurred with the pH 10 solutions with all resins. Average oryzalin desorption by the anion-exchange resin was between 0.12 and 3.84 mg g−1per desorption event. Maximum desorption occurred at pH 6.0. Results reveal that the resins evaluated may have merit for slow-release herbicide delivery.


1978 ◽  
Vol 18 (03) ◽  
pp. 207-218 ◽  
Author(s):  
D.R. Zornes ◽  
G.P. Willhite ◽  
M.J. Michnick

Abstract The separation of petroleum mono- and disulfonates on an anion-exchange column using high-pressure liquid chromatography (HPLC) was investigated as a method /or the determination of sulfonate concentrations in aqueous or hydrocarbon solutions. Quantitative analysis o/ brine (3,000 ppm) or hexane solutions containing 500 to 5,000 ppm ppm) or hexane solutions containing 500 to 5,000 ppm of unfractionated Witco TRS 10-80 Petronate was possible to a precision of 2 percent. Extending the possible to a precision of 2 percent. Extending the method to solutions containing TRS 10-80 sulfonates with altered mono- and di-ratios was accomplished by deriving ultraviolet (UV) response factors /or the average mono- and disulfonates present in the original TRS 10-80. Application of the HPLC technique with monoand disulfonate response factors was demonstrated by analysis of sulfonate concentrations in both phases in a series of two-phase hexane-brine phases in a series of two-phase hexane-brine systems. Trends in the distribution coefficients were identified readily. Preferential partitioning of the mono- and disulfonates between the hydrocarbon and brine phases was indicated by changes in the ratio of the mono- and disulfonate areas on the chromatographs. The HPLC technique gave information on the sulfonate composition that was not obtained by traditional wet-lab methods. Introduction Surfactant concentration is an important parameter in the evaluation of surfactants used in oil recovery processes. Unfortunately, quantative analysis of processes. Unfortunately, quantative analysis of surfactants commonly used in enhanced oil recovery processes is difficult because these often are processes is difficult because these often are complex mixtures of anionic surfactant molecules. Commercially available petroleum sulfonates not only contain a range of molecular weights but also vary in percentage of mono-, di- and polysulfonated molecules. Numerous analytical techniques have been presented in the literature to determine anionic presented in the literature to determine anionic surfactants in water and oil. The two-phase titration technique, introduced by Epton in 1946, was modified and used to determine petroleum sulfonates. Disadvantages of the Epton titration method includethe average equivalent weight of the anionic surfactant must be known;it cannot differentiate between mono-, di-, and polysulfonated molecules;experimental evidence polysulfonated molecules; (3) experimental evidence indicates the method is not stoichiometric for low molecular-weight sulfonates and there is limited knowledge of be stoichiometry for polysulfonated molecules; andit is not automated easily. The Epton titration method has been used because no other good analytical techniques exist. Recent developments in HPLC and in ion-exchange resin offer the possibility of improved analytical techniques to determine petroleum sulfonates. Traditionally, the chromatographic separation of aromatic sulfonates by ion-exchange has been difficult because of excessive noncoulombic adsorption on polystyrene-type, anion-exchange resins. Successful separation of low molecular-weight aromatic sulfonates on a quaternized polyalkeneamine, anion-exchange resin was reported by Stehl. Development of pellicular anion-exchange resins with a quaternary alkylamine bonded to an inert nonpermeable core further reduced the anionic adsorption of aromatic sulfonate molecules and extended the range of application. Schmit and Singh reposed separations of naphthalene-sulfonic-acid dye intermediates on pellicular anion-exchange resin. Suffridge reported pellicular anion-exchange resin. Suffridge reported separation of a petroleum sulfonate into its monoand disulfonate constituents on a pellicular anion-exchange resin using a linear ionic-strength gradient. SPEJ P. 207


Author(s):  
Miroslav Horák ◽  
Pavel Híc ◽  
Eva Tománková ◽  
Josef Balík

The aim of this experiment was to eliminate SO2 ions present in wine using the anion-exchanger resins. To compare the effectiveness, 2 following strongly basic anion-exchange resin were used. When activated, the sodium bicarbonate solution (activation solution I) is used to prevent parallel reduction of sulphites, tartates and malates, so the anion-exchange resins were activated in two-step activation. In the second step, it was immersed into a mixture of malic acid and tartaric acid (1:1). After the application of anex into wine, the content of total SO2 was reduced to 97–201 mg.L−1 (depending on the amount of anex added into the wine sample). According to our expectations, the variants with anion-exchange resin activated only with bicarbonate solution, the tartrates and malates were significantly reduced. If the anion-exchange resin was activated with a two-steps activation, the tartaric acid and malic acid were reduced in the range of ± 0.13 g.L−1. This phenomenon was strongly reflected at the anion-exchanger Aqua Osmotic 02. The changes in antioxidant content were not affected by the type of anion-exchange resin, the method of activation, or an amount of used anion-exchanger. The color parameters of wine, expressed by the L * a * b *, were not significantly affected by the effects of anion-exchange resin use.


1979 ◽  
Vol 25 (9) ◽  
pp. 1617-1621 ◽  
Author(s):  
H Miyagi ◽  
J Miura ◽  
Y Takata ◽  
S Ganno

Abstract Several liquid-chromatographic systems involving anion-exchange resins and linear acetate gradient elution have been reported previously for the analysis of ultraviolet-absorbing constituents in human urine. We describe a high-performance liquid-chromatographic system on which a macroreticular anion-exchange resin, and stepwise elution are used. Separations were more rapid with ammonium chloride/acetonitrile stepwise elution than with acetate stepwise elution, and urine constituents could be detected at 225 nm with the former system. Analytical conditions were convenient and the method has proven to be useful for routine clinical analysis.


2020 ◽  
Vol 21 (21) ◽  
pp. 7874
Author(s):  
Natalia Pismenskaya ◽  
Veronika Sarapulova ◽  
Anastasia Klevtsova ◽  
Sergey Mikhaylin ◽  
Laurent Bazinet

This study examines the mechanisms of adsorption of anthocyanins from model aqueous solutions at pH values of 3, 6, and 9 by ion-exchange resins making the main component of heterogeneous ion-exchange membranes. This is the first report demonstrating that the pH of the internal solution of a KU-2-8 aromatic cation-exchange resin is 2-3 units lower than the pH of the external bathing anthocyanin-containing solution, and the pH of the internal solution of some anion-exchange resins with an aromatic (AV-17-8, AV-17-2P) or aliphatic (EDE-10P) matrix is 2–4 units higher than the pH of the external solution. This pH shift is caused by the Donnan exclusion of hydroxyl ions (in the KU-2-8 resin) or protons (in the AV-17-8, AV-17-2P, and EDE-10P resins). The most significant pH shift is observed for the EDE-10P resin, which has the highest ion-exchange capacity causing the highest Donnan exclusion. Due to the pH shift, the electric charge of anthocyanin inside an ion-exchange resin differs from its charge in the external solution. At pH 6, the external solution contains uncharged anthocyanin molecules. However, in the AV-17-8 and AV-17-2P resins, the anthocyanins are present as singly charged anions, while in the EDE-10P resin, they are in the form of doubly charged anions. Due to the electrostatic interactions of these anions with the positively charged fixed groups of anion-exchange resins, the adsorption capacities of AV-17-8, AV-17-2P, and EDE-10P were higher than expected. It was established that the electrostatic interactions of anthocyanins with the charged fixed groups increase the adsorption capacity of the aromatic resin by a factor of 1.8–2.5 compared to the adsorption caused by the π–π (stacking) interactions. These results provide new insights into the fouling mechanism of ion-exchange materials by polyphenols; they can help develop strategies for membrane cleaning and for extracting anthocyanins from juices and wine using ion-exchange resins and membranes.


2013 ◽  
Vol 634-638 ◽  
pp. 334-337
Author(s):  
Miao Miao He ◽  
Xiao Jun Hu ◽  
Yong Biao Peng ◽  
Xin He

Through the method of the static tests, the removal rate of aqueous dichloroethylene onto a new TCAS-loaded resin was researched. This TCAS-loaded resin was made of a novel supramolecular acceptor compound named thiacalix[4]arenetetrasulfonate(TCAS) and anion exchange resin, and the adsorption mechanism was discussed preliminarily. The results of adsorption indicated that the pH value was an important factor for the removal of dichloroethylene and it would be better for the adsorption if the pH value was greater than 6. The operating temperature should be controlled in 5 to 15°C for the adsorption of dichloroethylene onto TCAS-loaded resin while the removal rate decreased with the temperature increasing and the best time for reaction was 40min. The removal rate of dichloroethylene in aqueous solution was better when 25mL aqueous solution of dichloroethylene (1.0mg/L) was adsorbed by 0.5g TCAS-loaded resin. The dichloroethylene can be resolving and TCAS-loaded resin can be reused.


Sign in / Sign up

Export Citation Format

Share Document