Richard D. Webster

Find an error

Name:
Organization: Nanyang Technological University , Singapore
Department: School of Physical and Mathematical Sciences
Title: (PhD)

TOPICS

Co-reporter:Yuexia Zhang;Xingxing Wu;Lin Hao;Zeng Rong Wong;Sherman J. L. Lauw;Song Yang;Yonggui Robin Chi
Organic Chemistry Frontiers 2017 vol. 4(Issue 3) pp:467-471
Publication Date(Web):2017/02/28
DOI:10.1039/C6QO00738D
An unusual trimerization of enone via a formal [2 + 2 + 2] process is disclosed. The reaction is initiated by a radical process enabled by NaOtBu and N-heterocyclic carbene (NHC). Molecular oxygen (air) is involved in key steps of the radical intermediate formation and carbon–carbon bond cleavage in this trimerization reaction to form highly substituted cyclohexane derivatives. In previous studies, alkali metal tert-butoxides such as NaOtBu were mainly employed to generate radical intermediates from aryl halides. Here we provide a new avenue in using NaOtBu and combined NHC/NaOtBu to generate radical intermediates from enones for further reactions.
Co-reporter:Maja Budanović, Bahareh Khezri, Sherman J.L. Lauw, Malcolm E. Tessensohn, Richard D. Webster
Analytica Chimica Acta 2017 Volume 992(Volume 992) pp:
Publication Date(Web):1 November 2017
DOI:10.1016/j.aca.2017.08.052
•Mercury is complexed by tetrathiafulvalene preventing adsorptive/volatility losses.•Statistical analysis confirms tetrathiafulvalene is superior to other preservatives.•Addition of tetrathiafulvalene greatly improves accuracy of ICP-MS method for mercury determination.•Voltammetric measurements confirm complexation reaction between Hg and S atoms in tetrathiafulvalene.The determination of mercury simultaneously with other elements via inductively coupled plasma-mass spectrometry (ICP-MS) in airborne particulate matter (PM2.5) is still challenging due to the lack of accuracy for the low level mercury concentrations as a result of its volatility and tendency to adhere to the walls of the sample introduction system. This study investigated the effect of existing (gold and methionine) and new (lithium tetrathiafulvalene carboxylate (LiCTTF)) preservation agents in order to improve the determination of trace mercury in PM2.5 samples. Statistical analysis revealed that a concentration of 10 μg mL−1 of LiCTTF was sufficient to obtain highly accurate results with t values of 0.1044–1.1239 which are considerably less than the critical t value of 1.8 and apparent recoveries of 85–100%. An evaluation of the method revealed a spiked mercury recovery of 91% and a detection limit of 0.05 ng mL−1. The method was tested for the determination of trace metals in PM2.5 from atmospheric samples and led to the detection of low elemental concentrations in Singapore's atmosphere. The mechanism for the interaction of mercury with LiCTTF and tetrathiafulvalene (TTF) was studied by conducting in situ electrochemical studies. Cyclic voltammetry and square-wave voltammetry analyses of mercury, and mercury in presence of LiCTTF and TTF revealed complexation between the metal and sulfur-containing compounds.Download high-res image (190KB)Download full-size image
Co-reporter:Shu Jun Ng, Natalie F. Sims, Esther X.Y. Tay, Richard D. Webster
Journal of Water Process Engineering 2017 Volume 18(Volume 18) pp:
Publication Date(Web):1 August 2017
DOI:10.1016/j.jwpe.2017.05.012
Two methods to remove VOCs from water using the natural food additive, lecithin, assisted by olive oil or vitamin E were explored. Lecithin combined with olive oil was utilized in a phase exchange and solvent extraction process to remove VOCs from water into the olive oil layer. The addition of 0.5 ppm of lecithin together with the olive oil were found to be sufficient to remove halogenated and aromatic VOCs completely; however, the method was less efficient against ketones and esters due to their better water solubility and smaller molecular mass in comparison to other VOCs tested. The second method involved combining lecithin with vitamin E to produce an aqueous suspension of micelles that facilitated the removal of VOCs from water into the micelles, with the optimal ratio of lecithin to vitamin E found to be 82:18 w/w. The overall removal of VOCs from water using lecithin/vitamin E micelles was not as effective as lecithin/olive oil layers due to the problems faced when removing lecithin from the samples.
Co-reporter:Ling Ying Diane Tiong;Mer Lin Doris Ho
Journal of Radioanalytical and Nuclear Chemistry 2017 Volume 314( Issue 2) pp:1347-1351
Publication Date(Web):06 October 2017
DOI:10.1007/s10967-017-5507-4
A method was developed for removing the interference of 238U from samples containing 238Pu so that the Pu isotopes could be rapidly determined by ICP-MS. The acidified sample was first passed through a commercial TEVA® resin that retained all of the plutonium and some uranium. Second, the sample was further passed through a UTEVA® resin that retained all of the remaining uranium but was free of plutonium. Elution of both of the fractions retained by the two resins and analysis by ICP-MS, allowed a simple correction factor to be determined that enabled subtraction of the interference caused by 238U.
Co-reporter:L. Y. D. Tiong;B. K. Pong;R. D. Webster
Analytical Methods (2009-Present) 2017 vol. 9(Issue 19) pp:2909-2914
Publication Date(Web):2017/05/18
DOI:10.1039/C7AY01017F
An acid assisted closed vessel microwave digestion procedure was developed for the treatment of fine airborne particulate matter that allows elemental iodine to be solubilised simultaneously with other metallic elements. This procedure is an improvement over existing methods as it allows the iodine to be safely digested from the filter media (polypropylene or cellulose) in less than one hour in up to 70–100% recoveries, without requiring an overnight predigestion step or the addition of reducing agents, with a microwave temperature of 185 °C and the use of 65% HNO3. A range of filters were investigated to determine the material that has the best properties for microwave digestion that also enabled the deposited iodine present in total suspended particles (TSP) to be rapidly fully digested and quantitatively analyzed using inductively coupled plasma-mass spectrometry (ICP-MS).
Co-reporter:Jazreen H.Q. Lee, Yu Rong Koh, Richard D. Webster
Journal of Electroanalytical Chemistry 2017 Volume 799(Volume 799) pp:
Publication Date(Web):15 August 2017
DOI:10.1016/j.jelechem.2017.05.044
•Diethylstilbestrol (DES) undergoes electrochemical oxidation in acetonitrile.•DES exhibited comparable voltammetric behavior to bisphenol A (BPA).•The oxidation process was posited to involve up to four moles of electrons.The electro-oxidation of the former drug diethylstilbestrol (DES) was examined in detail by means of cyclic voltammetry (CV) and controlled potential electrolysis (CPE) techniques. Analogous to the electrochemistry of the structurally-related compound bisphenol A (BPA), DES undergoes two successive oxidation processes at ca. 0.64 and 0.81 vs. (Fc/Fc+)/V at glassy carbon electrodes in acetonitrile. The former anodic process was determined to involve up to four moles of electrons, whereas the latter anodic peak was assigned as the oxidation of a secondary product formed after the initial oxidation of DES. Results from variable voltammetric scan rate experiments showed that the secondary process can be outrun at high scan rates, while simultaneously detecting a new reduction wave (at ca. 0.50 vs. (Fc/Fc+)/V). Additionally, at fast scan rates, this reduction process was found to be chemically reversible; whereby a corresponding anodic peak was recorded at ca. 0.53 vs. (Fc/Fc+)/V on the second cycle. Following a similar mechanistic pathway as BPA, under prolonged periods of time (e.g. slow scan rates), DES is proposed to undergo an overall four-electron oxidation, whereas at fast scan rates the oxidation of DES was found to only involve two electrons. The CV data were modeled using digital simulations that allowed an estimation of the electrochemical and kinetic parameters associated with the electrode reactions.Download high-res image (134KB)Download full-size image
Co-reporter:Jazreen H.Q. Lee, Sherman J.L. Lauw, Richard D. Webster
Electrochemistry Communications 2016 Volume 64() pp:69-73
Publication Date(Web):March 2016
DOI:10.1016/j.elecom.2016.01.016
•CO2 was electrochemically reduced in the presence of pyridoxine at a Pt electrode.•CO2-saturated solutions gave larger currents than Ar-purged solutions.•No observable cathodic processes were recorded on a glassy carbon electrode.•Preparative scale reductions of CO2 gave 5% faradic yields for methanol formation.Experiments aimed at ameliorating carbon dioxide (CO2) into methanol were explored using pyridoxine, a member of the vitamin B6 family, to enhance the reduction process. At a platinum electrode, an aqueous solution (pH ≈ 5) of pyridoxine showed a quasi-reversible redox couple with the cathodic peak detected at ca. − 0.55 V vs. Ag/AgCl (3 M KCl) in the presence of CO2 and argon. An increase in the corresponding cathodic peak current was observed following saturation of the solution with CO2 using a Pt electrode, but with no detectable reduction current recorded at a glassy carbon electrode for the same system. Confirmation of methanol formation during the pyridoxine-assisted CO2 reduction was conducted by using gas chromatography analysis of the electrolyzed solutions and faradic yields of ca. 5% were afforded. A combination of the results from the cyclic voltammetry and constant current chronopotentiometry experiments revealed an overpotential of ≤ 200 mV was required. The results indicate a potential utility of pyridoxine as an alternative reagent to the more toxic pyridine during the electrochemical reduction of CO2.
Co-reporter:Malcolm E. Tessensohn, Richard D. Webster
Electrochemistry Communications 2016 Volume 62() pp:38-43
Publication Date(Web):January 2016
DOI:10.1016/j.elecom.2015.11.006
•Review why voltammetry is very sensitive to hydrogen bonding effects.•Review how voltammetry can differentiate between strengths of hydrogen bonding donors and acceptors.•Review several systems where hydrogen bonding effects have been reported.This review briefly details some quantitative and qualitative methods to measure the hydrogen bonding abilities of electroactive and electroinactive compounds through voltammetry and provides examples of several systems that have been examined.
Co-reporter:Jazreen H.Q. Lee, Sherman J.L. Lauw, Richard D. Webster
Electrochimica Acta 2016 Volume 211() pp:533-544
Publication Date(Web):1 September 2016
DOI:10.1016/j.electacta.2016.06.054
•Vanillan undergoes a proton-coupled electron transfer oxidation.•The initially formed −2e−/-H+ intermediate undergoes hydrolysis.•The long term oxidation product is an ortho-quinone.•A hemiketal intermediate can be detected by voltammetry.•The voltammetric behaviour is affected by trace water in the acetonitrile.A detailed electrochemical study of vanillin was performed in acetonitrile using a platinum electrode. At low scan rates ≤ 1 V s−1, vanillin displayed an anodic peak at ca. 1.12 vs. (Fc/Fc+)/V, and a cathodic peak at ca. 0.09 vs. (Fc/Fc+)/V (on the reverse cyclic voltammetry (CV) scan after the initial oxidation) due to the reduction of a secondary reaction product. Both redox processes were found to involve a transfer of two electrons via controlled-potential electrolysis experiments, and it is surmised that vanillin first undergoes a −2e−/-H+ oxidation, followed by a hydrolysis reaction and the loss of its methoxy substituent to generate a corresponding 1,2-benzoquinone that can subsequently be reduced via +2e−/ + 2H+. When higher scan rates ≥2 V s−1 were employed, however, the aforementioned homogeneous reactions were partially outrun; leading to the detection of an additional cathodic peak at ca. −0.28 vs. (Fc/Fc+)/V which is ascribed to the reduction of the transient hemiketal intermediate that is formed immediately after the reaction of the phenoxonium cation with trace water. A similar trend to the fast scan rate CVs was likewise recorded at lowered temperatures (≤−10 °C). Digital simulations were used to model the cyclic voltammetry data which enabled an estimation of the electrochemical and kinetic parameters associated with the electrode reactions. Independently, vanillin can also be electrochemically reduced at ca. −1.58 vs. (Fc/Fc+)/V, which is proposed to involve the formation of molecular hydrogen.
Co-reporter:Sherman J.L. Lauw, Chiang Zhong, Richard D. Webster
Journal of Electroanalytical Chemistry 2016 Volume 779() pp:220-228
Publication Date(Web):15 October 2016
DOI:10.1016/j.jelechem.2016.02.004
•Cinnamaldehyde is reduced by two successive one-electron steps in CH3CN.•Reduced forms undergo several follow-up homogeneous reactions such as dimerization.•Two secondary products could be voltammetrically detected at high scan rates.•Digital simulations enabled the estimation of electrochemical/kinetic parameters.An in-depth investigation on the redox reactions of cinnamaldehyde in acetonitrile was carried out using cyclic, linear sweep, and rotating disk electrode voltammetry in conjunction with controlled potential electrolysis and digital simulations. Overall, it was found that cinnamaldehyde displays a similar reduction behavior to retinal (an aldehyde form of vitamin A), and can be reduced by two consecutive one-electron processes to initially form its radical anion (Epred ≈ − 1.95 vs. (Fc/Fc+)/V, where Epred refers to the cathodic peak potential and Fc = ferrocene), and then its dianion (Epred ≈ − 2.50 vs. (Fc/Fc+)/V). Both reduction processes had limited chemically reversibility even when examined at relatively fast scan rates and low temperatures. Notably, voltammetry and electrolysis experiments revealed that the electrochemical reduction critically depends on the amount of cinnamaldehyde used. At high concentrations, the radical anions have a propensity to undergo a heterodimerization reaction with the starting material to form a radical anionic dimer which could be voltammetrically detected at fast scan rates. Nevertheless, the experimental data also indicated a number of other homogeneous reactions involving the reduced species (radical anion and dianion) which were modeled by digital simulations to determine the electrochemical and kinetic parameters affiliated with all of the heterogeneous electron transfer and homogeneous reaction steps.
Co-reporter:Yuexia Zhang; Yu Du; Zhijian Huang; Jianfeng Xu; Xingxing Wu; Yuhuang Wang; Ming Wang; Song Yang; Richard D. Webster;Yonggui Robin Chi
Journal of the American Chemical Society 2015 Volume 137(Issue 7) pp:2416-2419
Publication Date(Web):February 4, 2015
DOI:10.1021/ja511371a
An N-heterocyclic carbene-catalyzed β-hydroxylation of enals is developed. The reaction goes through a pathway involving multiple radical intermediates, as supported by experimental observations. This oxidative single-electron-transfer reaction allows for highly enantioselective access to β-hydroxyl esters that are widely found in natural products and bioactive molecules.
Co-reporter:L. N. Mataranga-Popa, I. Torje, T. Ghosh, M. J. Leitl, A. Späth, M. L. Novianti, R. D. Webster and B. König  
Organic & Biomolecular Chemistry 2015 vol. 13(Issue 40) pp:10198-10204
Publication Date(Web):14 Aug 2015
DOI:10.1039/C5OB01418B
Flavin derivatives with an extended π-conjugation were synthesized in moderate to good yields from aryl bromides via a Buchwald–Hartwig palladium catalyzed amination protocol, followed by condensation of the corresponding aromatic amines with violuric acid. The electronic properties of the new compounds were investigated by absorption and emission spectroscopy, cyclic voltammetry, density functional theory (DFT) and time dependent density functional theory (TDDFT). The compounds absorb up to 550 nm and show strong luminescence. The photoluminescence quantum yields ϕPL measured in dichloromethane reach 80% and in PMMA (poly(methyl methacrylate)) 77%, respectively, at ambient temperature. The electrochemical redox behaviour of π-extended flavins follows the mechanism previously described for the parent flavin.
Co-reporter:Yanni Yue, Maria L. Novianti, Malcolm E. Tessensohn, Hajime Hirao and Richard D. Webster  
Organic & Biomolecular Chemistry 2015 vol. 13(Issue 48) pp:11732-11739
Publication Date(Web):13 Oct 2015
DOI:10.1039/C5OB01868D
Systematic synthesis of a number of new phenolic compounds with structures similar to vitamin E led to the identification of several sterically hindered compounds that when electrochemically oxidised in acetonitrile in a –2e−/–H+ process formed phenoxonium diamagnetic cations that were resistant to hydrolysis reactions. The reactivity of the phenoxonium ions was ascertained by performing cyclic voltammetric scans during the addition of carefully controlled quantities of water into acetonitrile solutions, with the data modelled using digital simulation techniques.
Co-reporter:Jazreen H. Q. Lee;Dr. Yanni Yue;Dr. Rakesh Ganguly ;Dr. Richard D. Webster
ChemElectroChem 2015 Volume 2( Issue 3) pp:412-420
Publication Date(Web):
DOI:10.1002/celc.201402340

Abstract

Cyclic voltammetry (CV) and controlled potential electrolysis (CPE) experiments revealed that at a Pt electrode, pyridoxine undergoes a one-electron chemically irreversible oxidation at approximately 0.50 V versus Fc/Fc+, likely generating a dimeric product. Fouling of the electrode surface could be detected with repeated scans during CV and during preparative scale CPE oxidation experiments. These adsorption effects were overcome by performing bulk oxidation of pyridoxine with NOSbF6 as a chemical oxidant. At Pt electrodes, a cathodic wave associated with the reduction of pyridoxine was detected at approximately −1.60 V versus Fc/Fc+. The reduction process appeared to be chemically reversible on the shorter timeframe of CV, but not under the prolonged timescale of CPE. The reduction of pyridoxine involves a direct discharge of pyridoxine at the Pt surface, generating an oxide anion, which upon treatment with iodomethane yields N-methylated pyridoxine rather than its O-methylated analogue.

Co-reporter:Dr. Gwendeline K. S. Wong;Li Zhen Lim;Marcus Jun Wen Lim;Li Lin Ong;Dr. Bahareh Khezri;Dr. Martin Pumera;Dr. Richard D. Webster
ChemPlusChem 2015 Volume 80( Issue 8) pp:
Publication Date(Web):
DOI:10.1002/cplu.201500286
Co-reporter:Dr. Gwendeline K. S. Wong;Li Zhen Lim;Marcus Jun Wen Lim;Li Lin Ong;Dr. Bahareh Khezri;Dr. Martin Pumera;Dr. Richard D. Webster
ChemPlusChem 2015 Volume 80( Issue 8) pp:1279-1287
Publication Date(Web):
DOI:10.1002/cplu.201500070

Abstract

Carbon nanotubes (CNTs) possess well-defined structural and chemical characteristics coupled with a large surface area that makes them ideal as sorbent materials for applications where adsorption processes are required. The adsorption properties of carboxylated derivatives of multiwalled carbon nanotubes (COOH-MWCNT) and singlewalled carbon nanotubes (COOH-SWCNT), together with their nonfunctionalized counterparts (MWCNT and SWCNT) for 48 common atmospheric volatile organic compounds (VOCs) were determined using thermal desorption–gas chromatography/mass spectrometry (TD-GCMS). The CNTs exhibited similar recoveries for many of the VOCs compared to the standard sorbent materials, Carbopack X and Tenax TA. However, VOCs with electron donor–acceptor (EDA) properties such as carbonyls, alkenes, and alcohols exhibited poorer recoveries on all CNTs compared to Carbopack X and Tenax TA. The poor recoveries of VOCs from the CNTs has important implications for the long term use and storage of CNTs, because it demonstrates that they will become progressively more contaminated with common atmospheric VOCs, therefore potentially affecting their surface-based properties.

Co-reporter:B. Khezri, Y. Y. Chan, L. Y. D. Tiong and R. D. Webster  
Environmental Science: Nano 2015 vol. 17(Issue 9) pp:1578-1586
Publication Date(Web):16 Jul 2015
DOI:10.1039/C5EM00312A
Burning of joss paper and incense is still a very common traditional custom in countries with a majority Chinese population. The Hungry Ghost Festival which is celebrated in the 7 month of the Chinese calendar is one of the events where joss paper and incense are burned as offerings. This study investigates the impact of the Ghost Month Festival (open burning event) on air quality by analysis of the chemical composition of particulate matter (PM) and rainwater samples collected during this event, compared with data collected throughout the year, as well as bottom ash samples from burning the original joss paper and incense. The results showed that the change in the chemical composition of the rainwater and PM2.5 (PM ≤ 2.5 μm) atmospheric samples could be correlated directly with burning events during this festival, with many elements increasing between 18% and 60% during August and September compared to the yearly mean concentrations. The order of percentage increase in elemental composition (in rain water and PM2.5) during the Hungry Ghost Festival is as follows: Zn > Ca > K > Mg > Fe > Al > Na ∼ Mn ∼ Ti ∼ V > Cu > As > Ni > Co > Cd > Cr > Pb. The chemical composition of the original source materials (joss paper and incense for combustion) and their associated bottom ash were analysed to explain the impact of burning on air quality.
Co-reporter:Malcolm E. Tessensohn;Melvyn Lee; Dr. Hajime Hirao; Dr. Richard D. Webster
ChemPhysChem 2015 Volume 16( Issue 1) pp:160-168
Publication Date(Web):
DOI:10.1002/cphc.201402693

Abstract

Voltammetric experiments with 9,10-anthraquinone and 1,4-benzoquinone performed under controlled moisture conditions indicate that the hydrogen-bond strengths of alcohols in aprotic organic solvents can be differentiated by the electrochemical parameter ΔEpred=|Epred(1)Epred(2)|, which is the potential separation between the two one-electron reduction processes. This electrochemical parameter is inversely related to the strength of the interactions and can be used to differentiate between primary, secondary, tertiary alcohols, and even diols, as it is sensitive to both their steric and electronic properties. The results are highly reproducible across two solvents with substantially different hydrogen-bonding properties (CH3CN and CH2Cl2) and are supported by density functional theory calculations. This indicates that the numerous solvent–alcohol interactions are less significant than the quinone–alcohol hydrogen-bonding interactions. The utility of ΔEpred was illustrated by comparisons between 1) 3,3,3-trifluoro-n-propanol and 1,3-difluoroisopropanol and 2) ethylene glycol and 2,2,2-trifluoroethanol.

Co-reporter:Serena L. J. Tan, Maria L. Novianti, and Richard D. Webster
The Journal of Physical Chemistry B 2015 Volume 119(Issue 44) pp:14053-14064
Publication Date(Web):October 8, 2015
DOI:10.1021/acs.jpcb.5b07534
The electrochemical reduction mechanisms of 2 synthesized flavins (Flox) were examined in detail in deoxygenated solutions of DMSO containing varying amounts of water, utilizing variable scan rate cyclic voltammetry (ν = 0.1–20 V s–1), controlled-potential bulk electrolysis, and UV–vis spectroscopy. Flavin 1, which contains a hydrogen atom at N(3), is capable of donating its proton to other reduced flavin species. After 1e– reduction, the initially formed Fl•– receives a proton from another Flox to form FlH• (and concomitantly produce the deprotonated flavin, Fl–), although the equilibrium constant for this process favors the back reaction. Any FlH• formed at the electrode surface immediately undergoes another 1e– reduction to form FlH–, which reacts with Fl– to form 2 molecules of Fl•–. Further 1e– reduction of Fl•– at more negative potentials produces the dianion, Fl2–, which can also be protonated by another Flox to form FlH– and Fl–. Flavin 2, which is methylated at N(3) (and therefore has no acidic proton), undergoes a simple chemically reversible 1e– reduction process in DMSO provided the water content is low (<100 mM). Further 1e– reduction of Fl•– (from flavin 2) at more negative potentials leads to the dianion, Fl2–, which is protonated by trace water in solution to form FlH–, similar to the mechanism of flavin 1 at high scan rates. Addition of sufficient amounts of water to nonaqueous solvents results in protonation of the anion radical species, Fl•–, for both flavins, causing an increase in the amount of FlH– in solution. This behavior contrasts with what is observed for quinones, which are also reduced in two 1e– steps in aprotic organic solvents to form the radical anions and dianions, but are able to exist in hydrogen-bonded forms (with trace or added water) without undergoing protonation.
Co-reporter:Sherman J. L. Lauw;Xiuhui Xu ;Dr. Richard D. Webster
ChemPlusChem 2015 Volume 80( Issue 8) pp:1288-1297
Publication Date(Web):
DOI:10.1002/cplu.201500247

Abstract

Ten 1,4-phenylenediamines were studied using electrochemical techniques (voltammetry and controlled potential electrolysis) and UV/Vis spectroscopy under ambient conditions. All compounds demonstrated vibrant color changes upon one-electron electrochemical oxidation in acetonitrile, with most displaying a primary color (red, green, blue, or yellow) in their oxidized state. The four electrochromes that exhibited the most intense color changes were examined by using a gold micromesh electrode laminated inside a polymer film to determine their electrochromic properties in solution (contrast ratios, chromatic efficiency, and cycle life). Their colored radical cations were also characterized by electron paramagnetic resonance spectroscopy as well as monitored for color retention over a period of 24 hours. Notably, only relatively small potentials were required to initiate the chromatic changes and the oxidized forms of the compounds were long-lived and unaffected by atmospheric oxygen or moisture.

Co-reporter:Ya Yun Chan;Alex Y. S. Eng;Dr. Martin Pumera;Dr. Richard D. Webster
ChemElectroChem 2015 Volume 2( Issue 7) pp:1003-1009
Publication Date(Web):
DOI:10.1002/celc.201500047

Abstract

The drop-casting method for the suspension of nanomaterials on conducting surfaces is a commonly used procedure for evaluating the electrochemical properties of the drop-cast materials. In this study, we pinpoint a key limitation of the method, which may lead to misinterpretation of the obtained data, especially when evaluating heterogeneous electron-transfer rates. The electrochemical responses recorded at 1 mm-diameter copper electrodes modified with porous layers of drop-cast multiwalled carbon nanotubes (MWCNTs) in 0.1 M Na2SO4 aqueous solutions were examined. Standard amounts of the MWCNTs that are typically used for the drop-casting procedure (1 mg MWCNT in 1 mL of dimethylformamide) were deposited drop wise on the surface of the copper electrodes. Layers of MWCNTs were progressively built up on the electrode surface by varying the number of drops from 0 to 10. The ability of the MWCNTs to cover and prevent diffusion to the base copper electrode was assessed by performing oxidative cyclic (CV) and linear-sweep voltammetry (LSV) experiments in the presence of aggressive SO42− supporting electrolyte, where a large oxidative current indicated the occurrence of corroding copper metal. It was demonstrated that a total of 10 drops of the coating solution (equivalent to 640 μg cm−2 of MWCNTs per unit area) was still insufficient in providing complete coverage over the underlying electrode surface (as a corrosion current was still observed), even though considerably lower CNT loadings have been applied in many literature reports. The electrochemical results indicate that, for experiments that utilize the drop-casting procedure to modify electrode surfaces, it cannot be assumed that the base electrode, nor the pore structure of the coating material, does not significantly contribute to the overall observed voltammetric response.

Co-reporter:Ding Luo, Sangsu Lee, Bin Zheng, Zhe Sun, Wangdong Zeng, Kuo-Wei Huang, Ko Furukawa, Dongho Kim, Richard D. Webster and Jishan Wu  
Chemical Science 2014 vol. 5(Issue 12) pp:4944-4952
Publication Date(Web):15 Aug 2014
DOI:10.1039/C4SC01843E
Polycyclic hydrocarbons (PHs) with a singlet biradical ground state have recently attracted extensive interest in physical organic chemistry and materials science. Replacing the carbon radical center in the open-shell PHs with a more electronegative nitrogen atom is expected to result in the more stable aminyl radical. In this work, two kinetically blocked stable/persistent derivatives (1 and 2) of indolo[2,3-b]carbazole, an isoelectronic structure of the known indeno[2,1-b]fluorene, were synthesized and showed different ground states. Based on variable-temperature NMR/ESR measurements and density functional theory calculations, it was found that the indolo[2,3-b]carbazole derivative 1 is a persistent singlet biradical in the ground state with a moderate biradical character (y0 = 0.269) and a small singlet–triplet energy gap (ΔES–T ≅ −1.78 kcal mol−1), while the more extended dibenzo-indolo[2,3-b]carbazole 2 exhibits a quinoidal closed-shell ground state. The difference can be explained by considering the number of aromatic sextet rings gained from the closed-shell to the open-shell biradical resonance form, that is to say, two for compound 1 and one for compound 2, which determines their different biradical characters. The optical and electronic properties of 2 and the corresponding aromatic precursors were investigated by one-photon absorption, transient absorption and two-photon absorption (TPA) spectroscopies and electrochemistry. Amphoteric redox behaviour, a short excited lifetime and a moderate TPA cross section were observed for 2, which can be correlated to its antiaromaticity and small biradical character. Compound 2 showed high reactivity to protic solvents due to its extremely low-lying LUMO energy level. Unusual oxidative dimerization was also observed for the unblocked dihydro-indolo[2,3-b]carbazole precursors 6 and 11. Our studies shed light on the rational design of persistent aminyl biradicals with tunable properties in the future.
Co-reporter:Yu Du, Yuhuang Wang, Xin Li, Yaling Shao, Guohui Li, Richard D. Webster, and Yonggui Robin Chi
Organic Letters 2014 Volume 16(Issue 21) pp:5678-5681
Publication Date(Web):October 24, 2014
DOI:10.1021/ol5027415
An unprecedented N-heterocyclic carbene catalytic reductive β,β-carbon coupling of α,β-nitroalkenes, by using an organic substrate to mimic the one-electron oxidation role of the pyruvate ferredoxin oxidoreductase (PFOR) in living systems, has been developed. The reaction goes through a radical anion intermediate generated under a catalytic redox process. For the first time, the presence of radical anion intermediate in NHC organocatalysis is observed and clearly verified.
Co-reporter:Ya Yun Chan, Yanni Yue, Richard D. Webster
Electrochimica Acta 2014 Volume 138() pp:400-409
Publication Date(Web):20 August 2014
DOI:10.1016/j.electacta.2014.06.133
•Vitamins D2 and D3 undergo a chemically irreversible oxidation process.•The electrochemical oxidation occurs via one-electron on short (CV) time-scales.•On long time scales (electrolysis) the oxidation occurs via two-electrons.•Chemical oxidation was performed using two molar equivalents of NO+.•Oxidation occurs at the triene moiety.The electrochemical behavior of vitamins D2 and D3 were examined by performing cyclic voltammetry (CV), rotating disk electrode voltammetry, controlled potential electrolysis and chemical oxidation in aprotic organic solvents. Both vitamins were electrochemically oxidized in dichloromethane and acetonitrile (Epox ∼ +0.8 vs. (Fc/Fc+)/V, where Epox is the anodic peak potential and Fc = ferrocene) via a one-electron chemically irreversible process on the short voltammetric time scale (≤ seconds). Varying the scan rate (0.1 V s−1 to 20 V s−1) and temperature (233 K to 293 K) did not strongly affect the voltammetric response recorded on platinum and glassy carbon electrode surfaces with the oxidation process remaining chemically irreversible over the range of scan rates and temperatures tested, indicating that the initially formed cation radical was not long-lived. Repetitive CV experiments indicated that the oxidized product partially adsorbed onto the electrode surface, resulting in diminishing peak currents with multiple scans. Bulk controlled potential electrolysis of the vitamin D compounds performed by alternating several cycles of oxidative electrolysis and reductive pulsed stripping proved to be effective in stripping the adsorbed species off the electrode surfaces. Longer time scale bulk electrolysis experiments led to the detection of a new oxidation peak appearing at less positive potentials as the electrolysis progressed, suggesting that the compounds underwent oxidation on long time scales (minutes to hours) via a two electron process. The vitamins were most likely initially oxidized via one-electron (E-step) to form a cation radical, which reacts homogeneously in two chemical steps (C-steps) where one chemical step is fast (< 1 s) and one is relatively slow (> 1 s). On the electrolysis timescale, the oxidized product then undergoes a second electron transfer (E) at less positive potentials, so the overall mechanism is an ECCE process. Chemical oxidation of vitamin D3 with 2 mol equiv of the one-electron oxidant, NO+SbF6− in acetonitrile/dichloromethane 1:4 ratio (v/v) resulted in the complete oxidation of the starting material in an overall two-electron process, with NMR analysis of the reaction mixture indicating that the triene moiety is absent from the products.
Co-reporter:Xiuhui Xu and Richard D. Webster  
RSC Advances 2014 vol. 4(Issue 35) pp:18100-18107
Publication Date(Web):16 Apr 2014
DOI:10.1039/C4RA02523G
Eleven aromatic diesters and thioic S,S′-diesters were synthesized and investigated using electrochemical (cyclic voltammetry and controlled potential electrolysis) and UV-vis spectroscopic techniques over a range of temperatures. Nine of the compounds exhibited vibrant colour changes from a colourless state in their neutral forms to brightly coloured upon one-electron electrochemical reduction in acetonitrile. The compounds were found to display either red, green or blue colours in their one-electron reduced states. The electrochromic properties of 3 of the compounds that displayed the most vibrant colour changes were examined in solution using a gold micro-mesh electrode laminated inside a polymer film.
Co-reporter:Kwok Kiong Chan;Dr. Rakesh Ganguly;Dr. Yongxin Li ;Dr. Richard D. Webster
ChemElectroChem 2014 Volume 1( Issue 9) pp:1557-1562
Publication Date(Web):
DOI:10.1002/celc.201402203

Abstract

Caffeine (CAF) undergoes a one-electron oxidation in acetonitrile to form a cation radical, with variable scan rate CV experiments indicating that the lifetime of the cation radical improves as the trace water content of the solvent is decreased. Electrochemical oxidation (and chemical oxidation with NOSbF6) of CAF in CH3CN leads to the generation of the protonated CAF cation as the long-term product in high yield, whose structure is confirmed by single-crystal X-ray crystallography and NMR spectroscopy. The protonated cation is able to be electrochemically reduced back to CAF under electrolysis conditions. The formation of the protonated cation involves the initial one-electron oxidation of CAF to form the cation radical, which undergoes a hydrogen atom abstraction reaction. Digital simulations of the CV data show that the rate and equilibrium constants for the hydrogen atom abstraction step are kf=1.0×102 L mol−1 s−1 and Keq=1.0×102.

Co-reporter:Ying Shan Tan, Dejan Urbančok, and Richard D. Webster
The Journal of Physical Chemistry B 2014 Volume 118(Issue 29) pp:8591-8600
Publication Date(Web):July 1, 2014
DOI:10.1021/jp505456q
Six of the major vitamers and provitamins comprising vitamin A (β-carotene, retinoic acid, retinol, retinyl palmitate, retinyl acetate, and retinal) were examined using voltammetric and controlled potential electrolysis techniques in the aprotic organic solvents acetonitrile and dichloromethane at glassy carbon and platinum electrodes. All of the compounds underwent oxidation and reduction processes and displayed a number of similarities and differences in terms of the number of redox processes and chemical reversibility of the voltammetric responses. The electrochemical properties of the compounds were strongly influenced by the functional groups on the unsaturated phytyl chains (carboxylic acid, alcohol, ester, or aldehyde groups), and not only on the fully conjugated hydrocarbon unit which is common to all forms of vitamin A. The compounds were reduced at potentials between approximately −1.7 and −2.6 vs (Fc/Fc+)/V (Fc = ferrocene) and oxidized at potentials between approximately +0.2 and +0.7 vs (Fc/Fc+)/V. The average number of electrons transferred per molecule under long time scale electrolysis experiments were found to vary between 0.4 and 4 electrons depending on the exact molecular structure and experimental conditions.
Co-reporter:Ya Yun Chan, Yanni Yue, Yongxin Li, Richard D. Webster
Electrochimica Acta 2013 Volume 112() pp:287-294
Publication Date(Web):1 December 2013
DOI:10.1016/j.electacta.2013.08.181
•Bisphenol A undergoes a chemically irreversible voltammetric oxidation process.•Chemical oxidation was performed to overcome adsorption effects that cause electrode fouling.•A new product was isolated from chemical oxidation with 4 mol equiv. of the one-electron oxidant, NO+.•The oxidative mechanism was proposed to be a four-electron/two-proton process.The electrochemical behavior of bisphenol A (BPA) was examined using cyclic voltammetry, bulk electrolysis and chemical oxidation in aprotic organic solvents. It was found that BPA undergoes a chemically irreversible voltammetric oxidation process to form compounds that cannot be electrochemically converted back to the starting materials on the voltammetric timescale. To overcome the effects of electrode fouling during controlled potential electrolysis experiments, NO+ was used as a one-electron chemical oxidant. A new product, hydroxylated bisdienone was isolated from the chemical oxidation of BPA with 4 mol equiv of NO+SbF6− in low water content CH3CN. The structure of the cation intermediate species was deduced and it was proposed that BPA is oxidized in a four-electron/two-proton process to form a relatively unstable dication which reacts quickly in the presence of water in acetonitrile (in a mechanism that is similar to phenols in general). However, as the water content of the solvent increased it was found that the chemical oxidation mechanism produced a nitration product in high yield. The findings from this study provide useful insights into the reactions that can occur during oxidative metabolism of BPA and highlight the possibility of the role of a bisdienone cation as a reactive metabolite in biological systems.
Co-reporter:Gwendeline K. S. Wong, Shu Jun Ng and Richard D. Webster  
Analytical Methods 2013 vol. 5(Issue 1) pp:219-230
Publication Date(Web):05 Oct 2012
DOI:10.1039/C2AY25982F
An analytical method has been developed and validated for analyzing 48 volatile organic compounds (VOCs) that were found to be present in substantial quantities in the atmosphere in Singapore. Air samples were collected by active sampling using Tenax/Carbopack X multisorbent tubes and were evaluated by Thermal Desorption Gas Chromatography Mass Spectrometry (TD-GCMS). Experiments conducted using standards demonstrated excellent repeatability with relative standard deviation (%RSD) values lesser than 10%, good linearity with R2 values of at least 0.99 for a wide range of concentrations between 0.02 and 500 ng, breakthrough values 5% or lower, tube desorption efficiencies close to 100% and good recoveries between 61% and 120%. Sampling volumes and flow rates were tested and selected by evaluating the performance of the multisorbent tubes. 30 mL min−1 was selected as the optimal flow rate for different sampling volumes depending on the individual compound's breakthrough value and reproducibility during air sampling. Most of the target analytes exhibited acceptable breakthrough of 5% or less, reproducibility within 20% deviation and method detection limits below 500 ppbv. Criteria established by the United States Environmental Protection Agency (USEPA) for sorbent tube sampling (EPA TO-17) were met for most compounds of interest.
Co-reporter:Serena L. J. Tan, Jia Min Kan, and Richard D. Webster
The Journal of Physical Chemistry B 2013 Volume 117(Issue 44) pp:13755-13766
Publication Date(Web):September 30, 2013
DOI:10.1021/jp4069619
The electrochemical reduction mechanisms of flavin mononucleotide (FMN) in buffered aqueous solutions at pH 3–11 and unbuffered aqueous solutions at pH 2–11 were examined in detail using variable-scan-rate cyclic voltammetry (ν = 0.1–20 V s–1), controlled-potential bulk electrolysis, UV–vis spectroscopy, and rotating-disk-electrode voltammetry. In buffered solutions at pH 3–5, FMN undergoes a two-electron/two-proton (2e–/2H+) reduction to form FMNH2 at all scan rates. When the buffered pH is increased to 7–9, FMN undergoes a 2e– reduction to form FMN2–, which initially undergoes hydrogen bonding with water molecules, followed by protonation to form FMNH–. At a low voltammetric scan rate of 0.1 V s–1, the protonation reaction has sufficient time to take place. However, at a higher scan rate of 20 V s–1, the proton-transfer reaction is outrun, and upon reversal of the scan direction, less of the FMNH– is available for oxidation, causing its oxidation peak to decrease in magnitude. In unbuffered aqueous solutions, three major voltammetric waves were observed in different pH ranges. At low pH in unbuffered solutions, where [H+] ≥ [FMN], (FMN)H– undergoes a 2e–/2H+ reduction to form (FMNH2)H– (wave 1), similar to the mechanism in buffered aqueous solutions at low pH. At midrange pH values (unbuffered), where pH ≤ pKa of the phosphate group and [FMN] ≥ [H+], (FMN)H– undergoes a 2e– reduction to form (FMN2–)H– (wave 2), similar to the mechanism in buffered aqueous solutions at high pH. At high pH (unbuffered), where pH ≥ pKa = 6.2 of the phosphate group, the phosphate group loses its second proton to be fully deprotonated, forming (FMN)2–, and this species undergoes a 2e– reduction to form (FMN2–)2– (wave 3).
Co-reporter:Ying Shan Tan, Yanni Yue, and Richard D. Webster
The Journal of Physical Chemistry B 2013 Volume 117(Issue 32) pp:9371-9379
Publication Date(Web):July 23, 2013
DOI:10.1021/jp4051808
Retinal (R) can be sequentially voltammetrically reduced in CH3CN in two one-electron processes to form first the anion radical (R•–) at −1.75 (±0.04) V vs Fc/Fc+ (Fc = ferrocene) then the dianion (R2–) at −2.15 (±0.04) V vs Fc/Fc+. The anion radical undergoes a reversible dimerization reaction to form the dianion (R22–) with a forward dimerization rate constant kdim = 8 × 102 L mol–1 s–1 and a reverse monomerization rate constant kmon = 2 × 10–2 s–1 at 295 K. All three anion species (anion radical, dianion, and dimer dianion) undergo hydrogen-bonding interactions with water that is present at millimolar levels in the solvent. As the water content of the solvent increases, the fate of the reduced compounds is determined by chemically irreversible hydrolysis reactions with H2O and decomposition reactions of the highly charged R2–. Bulk-controlled potential electrolysis experiments combined with NMR analysis of the reaction solutions indicate that the reduction occurs at the aldehyde group of retinal. The electrochemical data obtained under a range of experimental conditions (varying voltammetric scan rates, temperatures, H2O content of solutions, and retinal concentrations) were modeled by digital simulation techniques to determine the kinetic and thermodynamic parameters associated with all of the homogeneous reactions.
Co-reporter:Malcolm E. Tessensohn ; Hajime Hirao
The Journal of Physical Chemistry C 2013 Volume 117(Issue 2) pp:1081-1090
Publication Date(Web):December 20, 2012
DOI:10.1021/jp311007m
The electrochemical behavior of several phenols, quinones and hydroquinone in acetonitrile (CH3CN) with varying amounts of water were investigated to understand the effect of hydrogen-bonding on their voltammetric responses. Karl Fischer coulometric titrations were performed to obtain an accurate reading of the water concentrations. The solvent/electrolyte mixture was carefully dried using 3 Å molecular sieves to obtain an initial water content that was close to the substrate concentration (∼1 × 10–3 M), and higher water contents were then achieved via the addition from microliter syringes. It was found that small changes in what is often considered “trace” amounts of water were sufficient to substantially change the potential and in some cases the appearance of the voltammetric waves observed during the oxidation of the phenols/hydroquinones and reduction of the quinones. Density functional theory calculations were performed on the reduced/oxidized species in the presence of varying numbers of water molecules to better understand the hydrogen-bonding interactions at the molecular level. The results highlight the importance of accurately knowing the trace water content of organic solvents when used for voltammetric experiments.
Co-reporter:Zhen Hui Lim, Elaine Lay Khim Chng, Yanlan Hui, and Richard D. Webster
The Journal of Physical Chemistry B 2013 Volume 117(Issue 8) pp:2396-2402
Publication Date(Web):February 11, 2013
DOI:10.1021/jp4003966
When the quinone, vitamin K1 (VK1), is electrochemically reduced in aqueous-acetonitrile solutions (CH3CN with 7.22 M H2O), it undergoes a two-electron reduction to form the dianion that is hydrogen-bonded with water [VK1(H2O)y2–]. EPR and voltammetry experiments have shown that the persistent existence of the semiquinone anion radical (also hydrogen-bonded with water) [VK1(H2O)x–•] in aqueous or organic–aqueous solutions is a result of VK1(H2O)y2– undergoing a net homogeneous electron transfer reaction (comproportionation) with VK1, and not via direct one-electron reduction of VK1. When 1 mM solutions of VK1 were electrochemically reduced by two electrons in aqueous-acetonitrile solutions, quantitative EPR experiments indicated that the amount of VK1(H2O)x–• produced was up to approximately 35% of all the reduced species. In situ electrochemical ATR-FTIR experiments on sequentially one- and two-electron bulk reduced solutions of VK1 (showing strong absorbances at 1664, 1598, and 1298 cm–1) in CH3CN containing <0.05 M H2O led to the detection of VK1–• with strong absorbances at 1710, 1703, 1593, 1559, 1492, and 1466 cm–1 and VK1(H2O)y2– with strong absorbances at 1372 and 1342 cm–1.
Co-reporter:Serena L. J. Tan
Journal of the American Chemical Society 2012 Volume 134(Issue 13) pp:5954-5964
Publication Date(Web):March 5, 2012
DOI:10.1021/ja300191u
The electrochemical behavior of the naturally occurring vitamin B2, riboflavin (Flox), was examined in detail in dimethyl sulfoxide solutions using variable scan rate cyclic voltammetry (ν = 0.1 – 20 V s–1) and has been found to undergo a series of proton-coupled electron transfer reactions. At a scan rate of 0.1 V s–1, riboflavin is initially reduced by one electron to form the radical anion (Flrad•–) at E0f = −1.22 V versus Fc/Fc+ (E0f = formal reduction potential and Fc = ferrocene). Flrad•– undergoes a homogeneous proton transfer reaction with the starting material (Flox) to produce FlradH• and Flox–, which are both able to undergo further reduction at the electrode surface to form FlredH– (E0f = −1.05 V vs Fc/Fc+) and Flrad•2– (E0f = −1.62 V vs Fc/Fc+), respectively. At faster voltammetric scan rates, the homogeneous reaction between Flrad•– and Flox begins to be outrun, which leads to the detection of a voltammetric peak at more negative potentials associated with the one-electron reduction of Flrad•– to form Flred2– (E0f = −1.98 V vs Fc/Fc+). The variable scan rate voltammetric data were modeled quantitatively using digital simulation techniques based on an interconnecting “scheme of squares” mechanism, which enabled the four formal potentials as well as the equilibrium and rate constants associated with four homogeneous reactions to be determined. Extended time-scale controlled potential electrolysis (t > hours) and spectroscopic (EPR and in situ UV–vis) experiments confirmed that the chemical reactions were completely chemically reversible.
Co-reporter:Zebing Zeng ; Young Mo Sung ; Nina Bao ; Davin Tan ; Richmond Lee ; José L. Zafra ; Byung Sun Lee ; Masatoshi Ishida ; Jun Ding ; Juan T. López Navarrete ; Yuan Li ; Wangdong Zeng ; Dongho Kim ; Kuo-Wei Huang ; Richard D. Webster ; Juan Casado ;Jishan Wu
Journal of the American Chemical Society 2012 Volume 134(Issue 35) pp:14513-14525
Publication Date(Web):August 14, 2012
DOI:10.1021/ja3050579
Stable open-shell polycyclic aromatic hydrocarbons (PAHs) are of fundamental interest due to their unique electronic, optical, and magnetic properties and promising applications in materials sciences. Chichibabin’s hydrocarbon as a classical open-shell PAH has been investigated for a long time. However, most of the studies are complicated by their inherent high reactivity. In this work, two new stable benzannulated Chichibabin’s hydrocarbons 1-CS and 2-OS were prepared, and their electronic structure and geometry in the ground state were studied by various experiments (steady-state and transient absorption spectra, NMR, electron spin resonance (ESR), superconducting quantum interference device (SQUID), FT Raman, X-ray crystallographic etc.) and density function theory (DFT) calculations. 1-CS and 2-OS exhibited tunable ground states, with a closed-shell quinoidal structure for 1-CS and an open-shell biradical form for 2-OS. Their corresponding excited-state forms 1-OS and 2-CS were also chemically approached and showed different decay processes. The biradical 1-OS displayed an unusually slow decay to the ground state (1-CS) due to a large energy barrier (95 ± 2.5 kJ/mol) arising from severe steric hindrance during the transition from an orthogonal biradical form to a butterfly-like quinoidal form. The quick transition from the quinoidal 2-CS (excited state) to the orthogonal biradicaloid 2-OS (ground state) happened during the attempted synthesis of 2-CS. Compounds 1-CS and 2-OS can be oxidized into stable dications by FeCl3 and/or concentrated H2SO4. The open-shell 2-OS also exhibited a large two-photon absorption (TPA) cross section (760 GM at 1200 nm).
Co-reporter:Richard D. Webster
The Chemical Record 2012 Volume 12( Issue 1) pp:188-200
Publication Date(Web):
DOI:10.1002/tcr.201100005

Abstract

A review summarizing the voltammetric literature of the liposoluble vitamins A, D, E and K in organic solvents containing supporting electrolyte is presented. Electrochemical studies that were performed by attaching the vitamins to electrode surfaces and performing voltammetric scans in aqueous solutions are also summarized. Vitamins A (retinol and retinal) and D (cholecaliferol and ergocalciferol) undergo chemically irreversible voltammetric oxidation processes in organic solvents to form complicated or unknown compounds that cannot be electrochemically converted back to the starting materials. In contrast to vitamins A and D, vitamins E and K undergo chemically reversible electron-transfer processes that are often coupled to proton-transfer reactions. Vitamin E (a phenol) is voltammetrically oxidized in aprotic organic solvents in a −2e-/−H+ process to form a diamagnetic cation, which is unusually long-lived compared to the analogous cations produced during the oxidation of other phenols. In an aqueous environment, vitamin E is electrochemically oxidized to the hydroquinone in a chemically irreversible −2e- process. In low moisture content aprotic solvents, vitamin K (a quinone) is reduced in two one-electron chemically reversible steps to form first a radical anion (semiquinone, at E1) and then at more negative potentials a dianion is formed (at E2). The dianion is especially prone to strong hydrogen-bonding interactions with trace water present in the organic solvents, resulting in a shift in the formal reduction potential of E2 to more positive potentials as more water is added to the solvent. DOI 10.1002/tcr.201100005

Co-reporter:Yanlan Hui and Richard D. Webster
Analytical Chemistry 2011 Volume 83(Issue 3) pp:976
Publication Date(Web):January 7, 2011
DOI:10.1021/ac102734a
Many solvents used for electrochemistry can be dried to <1 × 10−3 M water content by storing the solvents over 3 Å molecular sieves in a nitrogen or argon atmosphere. However, as soon as the solvents are placed in an electrochemical cell, the water content increases significantly. Karl Fischer coulometric titrations were conducted on several predried solvents commonly used for electrochemistry (acetonitrile, dichloromethane, N,N-dimethylformamide, and dimethyl sulfoxide) in a controlled-humidity environment (30%, 50%, and 70% relative humidity) to determine the rate of moisture uptake into the organic solvents when used under typical electrochemical conditions (either in an electrochemical cell under a nitrogen atmosphere or in an electrochemical cell directly exposed to the atmosphere). The results in this study give guidelines for estimating the water content of organic solvents under conventional electrochemical operating conditions.
Co-reporter:Ying Shan Tan, Shanshan Chen, Wan Mei Hong, Jia Min Kan, Edwin Swee Hee Kwek, Shi Yu Lim, Zhen Hui Lim, Malcolm E. Tessensohn, Yinlu Zhang and Richard D. Webster  
Physical Chemistry Chemical Physics 2011 vol. 13(Issue 28) pp:12745-12754
Publication Date(Web):14 Jun 2011
DOI:10.1039/C1CP20579J
The phenol, α-tocopherol, can be electrochemically oxidised in a −2e−/−H+ process to form a diamagnetic cation that is long-lived in dry organic solvents such as acetonitrile and dichloromethane, but in the presence of water quickly reacts to form a hemiketal. Variable scan rate cyclic voltammetry experiments in acetonitrile with carefully controlled amounts of water between 0.010 M–0.6 M were performed in order to determine the rate of reaction of the diamagnetic cation with water. The water content of the solvent was accurately determined by Karl Fischer coulometric titrations and the voltammetric data were modelled using digital simulation techniques. The oxidation peak potential of α-tocopherol measured during cyclic voltammetry experiments was found to shift to less positive potentials as increasing amounts of water (0.01–0.6 M) were added to the acetonitrile, which was interpreted based on hydrogen-bonding interactions between the phenolic hydrogen atom and water. Several other phenols were examined and they displayed similar voltammetric features to α-tocopherol, suggesting that interactions of phenols with trace amounts of water were a common occurrence in acetonitrile. The H-bonding interactions of α-tocopherol with water were also examined viaNMR and UV-vis spectroscopies, with the voltammetric and spectroscopic studies extended to include other coordinating solvents (dimethyl sulfoxide and pyridine).
Co-reporter:Wei Wei Yao, Charmaine Lau, Yanlan Hui, Hwee Ling Poh, and Richard D. Webster
The Journal of Physical Chemistry C 2011 Volume 115(Issue 5) pp:2100-2113
Publication Date(Web):January 10, 2011
DOI:10.1021/jp1096037
A robust model membrane environment has been developed to enable voltammetry experiments to be performed on low molecular weight biological molecules completely incorporated inside artificial lipid bilayer (or multilayer) membranes. The artificial supported membranes were prepared by sandwiching multilayers of lecithin between layers of Nafion that were deposited on the surface of a glassy carbon electrode. The Nafion films acted as a conduit to aid proton transfer across the lecithin solution interface, and thereby balance the charge brought about by the electrochemical reactions. Vitamin E (α-tocopherol) and vitamin K1 were separately incorporated inside the Nafion|lecithin|Nafion layers and the coated electrodes were immersed in aqueous solutions between pH 3 and 13. The membranes were conductive to ion transfer, which allowed cyclic voltammetry experiments to be performed at scan rates of at least 200 V s−1. The electrode coating procedure produced multilayer membranes with solvent-like properties enabling highly reproducible diffusion controlled voltammetric processes to be observed. Vitamin E and vitamin K1 underwent multiple electron-transfer and proton-transfer reactions inside the membranes, and in the case of vitamin E, higher scan rate voltammetric experiments allowed the detection of short-lived intermediates.
Co-reporter:Ying Shan Tan and Richard D. Webster
The Journal of Physical Chemistry B 2011 Volume 115(Issue 14) pp:4244-4250
Publication Date(Web):March 18, 2011
DOI:10.1021/jp2000333
β-Carotene (β-Car) was chemically oxidized in a −2e− process using 2 mol equiv of NOSbF6 in a 4:1 ratio (v/v) of dichloromethane:acetonitrile to form the β-carotene dication (β-Car2+). Voltammetric monitoring of the chemical oxidation experiments over a range of temperatures indicated that the half-life of β-Car2+ was approximately 20 min at −60 °C, and approximately 1 min at −30 °C. α-Tocopherol (α-TOH) was chemically oxidized in a −2e−/−H+ process using 2 mol equiv of NOSbF6 to form the diamagnetic cation (α-TO+) which survives indefinitely at −60 °C in a 4:1 ratio (v/v) of dichloromethane:acetonitrile. Cyclic voltammetry experiments indicated that the oxidative peak potential for α-TOH was approximately +0.4 V more positive than the oxidative peak potential of β-Car. When solutions of α-TO+/H+ (prepared by chemical oxidation of α-TOH with 2 NO+) were reacted with solutions containing equal molar amounts of β-Car, voltammetric monitoring indicated that α-TOH was quantitatively regenerated and β-Car2+ was formed in high yield in a homogeneous two-electron transfer, according to the reaction α-TO+ + H+ + β-Car → α-TOH + β-Car2+.
Co-reporter:Shanshan Chen;Kah Yieng Tai ;Dr. Richard D. Webster
Chemistry – An Asian Journal 2011 Volume 6( Issue 6) pp:1492-1499
Publication Date(Web):
DOI:10.1002/asia.201000909

Abstract

Dopamine was electrochemically oxidized in aqueous solutions and in the organic solvents N,N-dimethyl-formamide and dimethylsulfoxide containing varying amounts of supporting electrolyte and water, to form dopamine ortho-quinone. It was found that the electrochemical oxidation mechanism in water and in organic solvents was strongly influenced by the buffering properties of the supporting electrolyte. In aqueous solutions close to pH 7, where buffers were not used, the protons released during the oxidation process were able to sufficiently change the localized pH at the electrode surface to reduce the deprotonation rate of dopamine ortho-quinone, thereby slowing the conversion into leucoaminochrome. In N,N-dimethylformamide and dimethylsulfoxide solutions, in the absence of buffers, dopamine was oxidized to dopamine ortho-quinone that survived without further reaction for several minutes at 25 °C. The voltammetric data obtained in the organic solvents were made more complicated by the presence of HCl in commercial sources of dopamine, which also underwent an oxidation process.

Co-reporter:Yanlan Hui, Elaine Lay Khim Chng, Louisa Pei-Lyn Chua, Wan Zhen Liu and Richard D. Webster
Analytical Chemistry 2010 Volume 82(Issue 5) pp:1928
Publication Date(Web):February 10, 2010
DOI:10.1021/ac9026719
Voltammetry experiments were performed on the natural quinone, vitamin K1 (VK1), in a range of organic solvents of varying dielectric constant that are commonly used for electrochemical measurements [dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF), acetonitrile (MeCN), propionitrile (EtCN), butyronitrile (PrCN), 1,2-dichloroethane (DCE), dichloromethane (DCM), and 1,1,2,2-tetrachloroethane (TCE)]. The water content of the solvents was accurately measured using Karl Fischer (KF) coulometric titrations, and the voltammetric data were used to estimate the degree of hydrogen-bonding interactions between the reduced forms of VK1 and variable levels of water, thereby allowing a ranking of water−substrate interactions in the different solvents. The voltammetric data were analyzed based on interactions that occur between reduced forms of VK1 and the water, the solvent, and the supporting electrolyte. Calibration data were obtained that are independent of the nature of the reference electrode and allow the water content of the solvents to be calculated by performing a single voltammetric scan in the presence of VK1 and 0.2 M supporting electrolyte (Bu4NPF6).
Co-reporter:Shanshan Chen, Richard D. Webster, Carmen Talotta, Francesco Troisi, Carmine Gaeta, Placido Neri
Electrochimica Acta 2010 Volume 55(Issue 23) pp:7036-7043
Publication Date(Web):30 September 2010
DOI:10.1016/j.electacta.2010.06.076
The cyclic voltammetric properties of several substituted calix[4]arenes were examined in acetonitrile and dichloromethane. The compounds that contained one phenolic group in the macrocyclic cavity were able to be electrochemically oxidised at positive potentials. In acetonitrile, cyclic voltammetry experiments indicated that the phenolic compounds were oxidised in a two-electron (one-proton) process over all measured scan rates (up to 50 V s−1), while in dichloromethane, the oxidation process occurred by one-electron at scan rates ≥5 V s−1, to most likely form the radical cations. In both solvents, longer timescale (minutes to hours) controlled potential coulometry experiments indicated that the oxidation process occurred by two-electrons per molecule, to form reactive diamagnetic cations that could not be reduced back to the starting materials under electrolysis conditions. The ion-sensing properties of the compounds were investigated in polymer membrane ion-selective electrodes and it was found that they responded reversibly in a Nernstian fashion to Groups 1 and 2 metals and had the highest selectivity to the cesium cation.
Co-reporter:Shanshan Chen, Hong Mei Peng, Richard D. Webster
Electrochimica Acta 2010 Volume 55(Issue 28) pp:8863-8869
Publication Date(Web):1 December 2010
DOI:10.1016/j.electacta.2010.07.096
Several phenols with structures similar to vitamin E were oxidised and the intermediate species produced were characterised by in situ infrared and UV–vis spectroscopies. The Fourier transform infrared (FTIR) measurements were performed by chemically oxidising the phenols with 2 mol equiv. of NO+SbF6− in CH3CN and recording the spectra between 1900 and 1300 cm−1 with an attenuated total reflectance (ATR) probe utilising a fiber conduit and a diamond composite sensor. The compounds that formed long-lived phenoxonium cations displayed two IR absorbances at 1665 (±15) cm−1 and one at 1600 (±10) cm−1 associated with the carbonyl, symmetric ring stretch and asymmetric ring stretch modes. The para-quinones are one of the long-term products of oxidation of the phenols, and displayed solution phase IR absorbances at 1650 (±10) cm−1. In situ electrochemical UV–vis experiments performed during the oxidation of the phenols led to the detection of bands due to the phenoxonium cations at 295 (±5) and 440 (±15) nm and due to the para-quinones at 260 (±10) nm. The concentration of the substrate and the water content of the solvent had a major effect on the yields of the intermediates and products that were produced during the oxidation reactions.
Co-reporter:Wei Wei Yao, Ying Shan Tan, Ying Xiu Low, Jasmine Shu Ying Yuen, Charmaine Lau and Richard D. Webster
The Journal of Physical Chemistry B 2009 Volume 113(Issue 46) pp:15263-15271
Publication Date(Web):October 28, 2009
DOI:10.1021/jp905324q
A procedure was developed for initiating electron transfer from a gold electrode to a low molecular weight electron acceptor present inside supported lipid (lecithin) bilayers, followed by further electron transfer to an electron acceptor present in an aqueous solution. The electron acceptors present in the lecithin bilayers and aqueous phase were 7,7,8,8-tetracyanoquinodimethane (TCNQ) and [FeIII(CN)6]3−, respectively. A polished planar gold disk electrode was first coated via self-assembly procedures with an alkanethiol monolayer. A phospholipid layer consisting of multiple bilayers of lecithin containing TCNQ was subsequently deposited onto the alkanethiol monolayer. The Au/alkanethiol/lecithin−TCNQ electrode was placed in an aqueous solution containing various amounts of [FeIII(CN)6]3− and [FeII(CN)6]4−, with 0.5 M KCl as the supporting electrolyte. In the absence of TCNQ inside the alkanethiol/lecithin layers, only a small background current was observed. When TCNQ was included in the alkanethiol/lecithin layers, the voltammetry showed features typical of a catalytic process, due to the TCNQ being reduced to TCNQ−• within the lecithin bilayers and then undergoing oxidation back to TCNQ via interaction with [FeIII(CN)6]3− at the lecithin−aqueous solution interface. The procedures for preparing the alkanethiol/lecithin−TCNQ coatings were optimized in order to obtain the most reproducible voltammetric response. Experiments were also performed using tetrathiafulvalene (TTF) as an electron donor in the lipid bilayer phase.
Co-reporter:Wei Wei Yao, Hong Mei Peng and Richard D. Webster
The Journal of Physical Chemistry C 2009 Volume 113(Issue 52) pp:21805-21814
Publication Date(Web):October 27, 2009
DOI:10.1021/jp9079124
Water insoluble α-tocopherol films were deposited on the surface of gold, glassy carbon and platinum electrodes and their voltammetric behavior examined in aqueous solutions between pH 3 and 13. The voltammetric mechanism involved α-tocopherol being oxidized in a −2e−/−H+ process to form a phenoxonium cation, which underwent rapid reaction with water (or −OH at pH > 7) and rearrangement to form α-tocopherol quinone in a chemically irreversible process. The identity of α-tocopherol quinone was determined by reflectance-FTIR spectroscopy of the product on the glassy carbon electrode surface and from comparison of the voltammetric data obtained with a sample of the α-tocopherol quinone model compound. α-Tocopherol quinone films could be voltammetrically reduced at negative potentials to form α-tocopherol hydroquinone in a +2e−/+2H+ chemically reversible process. Experiments were also conducted by incorporating α-tocopherol into lipid (lecithin) multilayers deposited onto the electrode surfaces and the electrochemical results compared with voltammetric data obtained from α-tocopherol films that were directly in contact with aqueous buffered solutions.
Co-reporter:Wei Wei Yao, Hong Mei Peng and Richard D. Webster, Peter M. W. Gill
The Journal of Physical Chemistry B 2008 Volume 112(Issue 22) pp:6847-6855
Publication Date(Web):May 8, 2008
DOI:10.1021/jp710995n
Variable scan rate (0.1−500 V s−1) cyclic voltammetry experiments were performed on a series of model tocopherol (vitamin E) compounds with differing degrees of methyl substitution around the aromatic (phenolic) ring. α-Tocopherol, with a fully methylated aromatic ring, produced stable phenoxonium cations upon oxidation in CH3CN, and was modeled via an ECE mechanism (where “E” represents an electron transfer and “C” a chemical step). Compounds with less methyl substitution around the aromatic ring were more reactive following oxidation, and formed additional oxidation products (hemiketals and p-quinones), and were modeled according to a more complicated ECECC mechanism. The equilibrium and rate constants associated with the chemical steps were estimated by digital simulations of the variable scan rate data over a range of temperatures (T = 253−313 K) in acetonitrile containing 0.5 M Bu4NPF6 as the supporting electrolyte. The relative lifetimes of the phenoxonium cations of tocol and the tocopherols were compared with theoretical results obtained from molecular orbital calculations.
Co-reporter:Hong Mei Peng, Becky F. Choules, Wei Wei Yao, Zhengyang Zhang, Richard D. Webster and Peter M. W. Gill
The Journal of Physical Chemistry B 2008 Volume 112(Issue 33) pp:10367-10374
Publication Date(Web):July 29, 2008
DOI:10.1021/jp804135e
Heterocyclic compounds with structures similar to vitamin E, but without the hydroxyl hydrogen atom, were synthesized and their electrochemical behavior examined in acetonitrile solutions and as solids in aqueous solutions of varying pH by attaching the compounds to the surface of a glassy carbon electrode. Compound 1, containing a fully methylated aromatic ring was found to be the most long-lived following one-electron oxidation, with its radical cation (1+•) surviving in acidic aqueous solutions and able to be isolated as a salt, 1+•(SbF6−), when reacted with NOSbF6 in CH3CN. Electrochemical, UV−vis and FTIR experiments on 1+•, in addition to the results from theoretical calculations, indicated that the electrochemical, electronic and structural properties of 1+• are very similar to those of the radical cation of vitamin E.
Co-reporter:Hong Mei Peng, Richard D. Webster and Xingwei Li
Organometallics 2008 Volume 27(Issue 17) pp:4484-4493
Publication Date(Web):August 6, 2008
DOI:10.1021/om800404q
A series of rhodium and iridium complexes of quinoline-tethered hemilabile N-heterocyclic carbenes (NHC∧N) have been synthesized via either deprotonation of imidazolium salts or silver transmetalation. Deprotonation of imidazolium ions by tBuOK in the presence of [Rh(COD)Cl]2 (COD = 1,5-cyclooctadiene) afforded both chelating [Rh(COD)(NHC∧N)]+ and monodentate [Rh(COD)(NHC)2]+ complexes, while only the chelating carbene complexes were obtained for the iridium analogues. Silver transmetalation of this type of carbene to [M(COD)Cl]2 (M = Rh, Ir) consistently afforded M(NHC)(COD)Cl, maintaining the metal chloride and pendant quinoline entity. Carbene transmetalation to (Cp∗IrCl2)2 gave an equilibrium mixture of neutral Cp*Ir(NHC)Cl2 and ionic [Cp*Ir(NHC∧N)Cl]Cl. All these rhodium and iridium cyclooctadiene complexes can undergo one-electron oxidation in cyclic voltammetry. The variable-scan-rate cyclic voltammetry experiments indicate that these compounds undergo slow heterogeneous electron transfer and that the oxidized forms are relatively short-lived. A neutral Rh(COD)(NHC)Cl complex proved to be active in catalyzing the [3 + 2] cycloaddition reactions of diphenylcyclopropenone and internal alkynes. Crystal structures of metal complexes in each category have been reported.
Co-reporter:Jazreen H.Q. Lee, Sherman J.L. Lauw, Richard D. Webster
Electrochemistry Communications (January 2017) Volume 74() pp:
Publication Date(Web):January 2017
DOI:10.1016/j.elecom.2016.12.007
Co-reporter:Ying Shan Tan, Shanshan Chen, Wan Mei Hong, Jia Min Kan, Edwin Swee Hee Kwek, Shi Yu Lim, Zhen Hui Lim, Malcolm E. Tessensohn, Yinlu Zhang and Richard D. Webster
Physical Chemistry Chemical Physics 2011 - vol. 13(Issue 28) pp:NaN12754-12754
Publication Date(Web):2011/06/14
DOI:10.1039/C1CP20579J
The phenol, α-tocopherol, can be electrochemically oxidised in a −2e−/−H+ process to form a diamagnetic cation that is long-lived in dry organic solvents such as acetonitrile and dichloromethane, but in the presence of water quickly reacts to form a hemiketal. Variable scan rate cyclic voltammetry experiments in acetonitrile with carefully controlled amounts of water between 0.010 M–0.6 M were performed in order to determine the rate of reaction of the diamagnetic cation with water. The water content of the solvent was accurately determined by Karl Fischer coulometric titrations and the voltammetric data were modelled using digital simulation techniques. The oxidation peak potential of α-tocopherol measured during cyclic voltammetry experiments was found to shift to less positive potentials as increasing amounts of water (0.01–0.6 M) were added to the acetonitrile, which was interpreted based on hydrogen-bonding interactions between the phenolic hydrogen atom and water. Several other phenols were examined and they displayed similar voltammetric features to α-tocopherol, suggesting that interactions of phenols with trace amounts of water were a common occurrence in acetonitrile. The H-bonding interactions of α-tocopherol with water were also examined viaNMR and UV-vis spectroscopies, with the voltammetric and spectroscopic studies extended to include other coordinating solvents (dimethyl sulfoxide and pyridine).
Co-reporter:Yuexia Zhang, Xingxing Wu, Lin Hao, Zeng Rong Wong, Sherman J. L. Lauw, Song Yang, Richard D. Webster and Yonggui Robin Chi
Inorganic Chemistry Frontiers 2017 - vol. 4(Issue 3) pp:
Publication Date(Web):
DOI:10.1039/C6QO00738D
Co-reporter:L. N. Mataranga-Popa, I. Torje, T. Ghosh, M. J. Leitl, A. Späth, M. L. Novianti, R. D. Webster and B. König
Organic & Biomolecular Chemistry 2015 - vol. 13(Issue 40) pp:NaN10204-10204
Publication Date(Web):2015/08/14
DOI:10.1039/C5OB01418B
Flavin derivatives with an extended π-conjugation were synthesized in moderate to good yields from aryl bromides via a Buchwald–Hartwig palladium catalyzed amination protocol, followed by condensation of the corresponding aromatic amines with violuric acid. The electronic properties of the new compounds were investigated by absorption and emission spectroscopy, cyclic voltammetry, density functional theory (DFT) and time dependent density functional theory (TDDFT). The compounds absorb up to 550 nm and show strong luminescence. The photoluminescence quantum yields ϕPL measured in dichloromethane reach 80% and in PMMA (poly(methyl methacrylate)) 77%, respectively, at ambient temperature. The electrochemical redox behaviour of π-extended flavins follows the mechanism previously described for the parent flavin.
Co-reporter:Ding Luo, Sangsu Lee, Bin Zheng, Zhe Sun, Wangdong Zeng, Kuo-Wei Huang, Ko Furukawa, Dongho Kim, Richard D. Webster and Jishan Wu
Chemical Science (2010-Present) 2014 - vol. 5(Issue 12) pp:NaN4952-4952
Publication Date(Web):2014/08/15
DOI:10.1039/C4SC01843E
Polycyclic hydrocarbons (PHs) with a singlet biradical ground state have recently attracted extensive interest in physical organic chemistry and materials science. Replacing the carbon radical center in the open-shell PHs with a more electronegative nitrogen atom is expected to result in the more stable aminyl radical. In this work, two kinetically blocked stable/persistent derivatives (1 and 2) of indolo[2,3-b]carbazole, an isoelectronic structure of the known indeno[2,1-b]fluorene, were synthesized and showed different ground states. Based on variable-temperature NMR/ESR measurements and density functional theory calculations, it was found that the indolo[2,3-b]carbazole derivative 1 is a persistent singlet biradical in the ground state with a moderate biradical character (y0 = 0.269) and a small singlet–triplet energy gap (ΔES–T ≅ −1.78 kcal mol−1), while the more extended dibenzo-indolo[2,3-b]carbazole 2 exhibits a quinoidal closed-shell ground state. The difference can be explained by considering the number of aromatic sextet rings gained from the closed-shell to the open-shell biradical resonance form, that is to say, two for compound 1 and one for compound 2, which determines their different biradical characters. The optical and electronic properties of 2 and the corresponding aromatic precursors were investigated by one-photon absorption, transient absorption and two-photon absorption (TPA) spectroscopies and electrochemistry. Amphoteric redox behaviour, a short excited lifetime and a moderate TPA cross section were observed for 2, which can be correlated to its antiaromaticity and small biradical character. Compound 2 showed high reactivity to protic solvents due to its extremely low-lying LUMO energy level. Unusual oxidative dimerization was also observed for the unblocked dihydro-indolo[2,3-b]carbazole precursors 6 and 11. Our studies shed light on the rational design of persistent aminyl biradicals with tunable properties in the future.
Co-reporter:
Analytical Methods (2009-Present) 2013 - vol. 5(Issue 1) pp:
Publication Date(Web):
DOI:10.1039/C2AY25982F
An analytical method has been developed and validated for analyzing 48 volatile organic compounds (VOCs) that were found to be present in substantial quantities in the atmosphere in Singapore. Air samples were collected by active sampling using Tenax/Carbopack X multisorbent tubes and were evaluated by Thermal Desorption Gas Chromatography Mass Spectrometry (TD-GCMS). Experiments conducted using standards demonstrated excellent repeatability with relative standard deviation (%RSD) values lesser than 10%, good linearity with R2 values of at least 0.99 for a wide range of concentrations between 0.02 and 500 ng, breakthrough values 5% or lower, tube desorption efficiencies close to 100% and good recoveries between 61% and 120%. Sampling volumes and flow rates were tested and selected by evaluating the performance of the multisorbent tubes. 30 mL min−1 was selected as the optimal flow rate for different sampling volumes depending on the individual compound's breakthrough value and reproducibility during air sampling. Most of the target analytes exhibited acceptable breakthrough of 5% or less, reproducibility within 20% deviation and method detection limits below 500 ppbv. Criteria established by the United States Environmental Protection Agency (USEPA) for sorbent tube sampling (EPA TO-17) were met for most compounds of interest.
Co-reporter:B. Khezri, Y. Y. Chan, L. Y. D. Tiong and R. D. Webster
Environmental Science: Nano 2015 - vol. 17(Issue 9) pp:NaN1586-1586
Publication Date(Web):2015/07/16
DOI:10.1039/C5EM00312A
Burning of joss paper and incense is still a very common traditional custom in countries with a majority Chinese population. The Hungry Ghost Festival which is celebrated in the 7 month of the Chinese calendar is one of the events where joss paper and incense are burned as offerings. This study investigates the impact of the Ghost Month Festival (open burning event) on air quality by analysis of the chemical composition of particulate matter (PM) and rainwater samples collected during this event, compared with data collected throughout the year, as well as bottom ash samples from burning the original joss paper and incense. The results showed that the change in the chemical composition of the rainwater and PM2.5 (PM ≤ 2.5 μm) atmospheric samples could be correlated directly with burning events during this festival, with many elements increasing between 18% and 60% during August and September compared to the yearly mean concentrations. The order of percentage increase in elemental composition (in rain water and PM2.5) during the Hungry Ghost Festival is as follows: Zn > Ca > K > Mg > Fe > Al > Na ∼ Mn ∼ Ti ∼ V > Cu > As > Ni > Co > Cd > Cr > Pb. The chemical composition of the original source materials (joss paper and incense for combustion) and their associated bottom ash were analysed to explain the impact of burning on air quality.
Co-reporter:Yanni Yue, Maria L. Novianti, Malcolm E. Tessensohn, Hajime Hirao and Richard D. Webster
Organic & Biomolecular Chemistry 2015 - vol. 13(Issue 48) pp:NaN11739-11739
Publication Date(Web):2015/10/13
DOI:10.1039/C5OB01868D
Systematic synthesis of a number of new phenolic compounds with structures similar to vitamin E led to the identification of several sterically hindered compounds that when electrochemically oxidised in acetonitrile in a –2e−/–H+ process formed phenoxonium diamagnetic cations that were resistant to hydrolysis reactions. The reactivity of the phenoxonium ions was ascertained by performing cyclic voltammetric scans during the addition of carefully controlled quantities of water into acetonitrile solutions, with the data modelled using digital simulation techniques.
5-Benzofuranol, 2,3-dihydro-2,2,4,6,7-pentamethyl-
3-Butenoic acid, 4-(2-methoxyphenyl)-2-oxo-, methyl ester, (3E)-
Pentacyclo[19.3.1.13,7.19,13.115,19]octacosa-1(25),3,6,9,11,13(27),15,17,19(26),21,23-undecaene-5,28-dione, 11,17,23-tris(1,1-dimethylethyl)-25,26,27-tripropoxy-
2-Tridecenal, 4-oxo-, (2E)-
Furo[2,3-f:4,5-f']bis[1,3]benzodioxole
Butane, dimethyl-
Benzene, [(1Z)-1-octenylsulfinyl]-
Hexanoic acid, 3-hydroxy-, methyl ester, (S)-
3-Butenoic acid, 4-(2-methoxyphenyl)-2-oxo-, (E)-
2H-1-Benzopyran-6-ol, 3,4-dihydro-2,2,5,8-tetramethyl-