首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Correction for ‘Dipyrrolyl-bis-sulfonamide chromophore based probe for anion recognition’ by Namdev V. Ghule et al., RSC Adv., 2014, 4, 27112–27115, DOI: 10.1039/C4RA04000G.

The authors regret that an incorrect version of Fig. 1 was included in the original article. The correct version of Fig. 1 is presented below.Open in a separate windowFig. 1Color changes of receptor DPBS in chloroform upon addition of 5 equiv. of F, Cl, Br, I, H2PO4, HSO4, ClO4 and AcO (tetrabutylammonium salts).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

2.
Correction for ‘Ultrafast conversion of carcinogenic 4-nitrophenol into 4-aminophenol in the dark catalyzed by surface interaction on BiPO4/g-C3N4 nanostructures in the presence of NaBH4’ by Ahmed B. Azzam et al., RSC Adv., 2021, 11, 18797–18808. DOI: 10.1039/D1RA02852A.

The authors regret that some misleading statements were included in section 3.2.1 ‘Effect of initial concentration on 4-NP’. The corrected version of section 3.2.1 is presented below. There are no changes to Fig. 8 or its caption.Open in a separate windowFig. 8Effect of initial concentration on reduction rate of 4-NP (a); corresponding linear transform ln(At/A0) = f(t) of the 4-NP reduction kinetics curves (b); comparison of the rate constant value for the reduction of the 4-NP over initial different concentration (c). ([4-NP] = 10–70 mg L−1, [catalyst] = 0.5 g L−1). 3.2.1 Effect of initial concentration of 4-NP The effect of the initial concentration of 4-NP on the catalytic efficiency rate using 50% BiPO4/g-C3N4 catalyst was carried out by varying the concentration from 10 to 70 mg L−1, and the obtained results are shown in Fig. 8a. Interestingly, 50% BiPO4/g-C3N4 was able to reduce all 4-NP solutions at concentrations from 10 to 70 mg L−1, reflecting the high efficiency of such a catalyst towards this 4-NP reduction. At lower concentrations, a superior constant rate was recorded due to the availability of a large number of catalytic sites per given amount of 4-NP moles. And vice versa, the higher the concentration, the lower the rate constant (Fig. 8b), due to the high competition of 4-NP molecules on the limited sites. In addition, the number of molecules adsorbed at the surface of the BiPO4/g-C3N4 heterojunction increases with the increase in concentration of 4-nitrophenol and hence, the surface becomes saturated by 4-nitrophenol molecules. This leads to a decrease in concentration of BH4 ions approaching the surface of the BiPO4/g-C3N4 heterojunction, hence lowering the rate of hydrogen transfer from BH4 ion to the 4-nitrophenol molecule.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

3.
Correction for ‘High-detectivity perovskite-based photodetector using a Zr-doped TiOx cathode interlayer’ by C. H. Ji et al., RSC Adv., 2018, 8, 8302–8309.

The authors regret that the names of the authors are shown incorrectly in the original article. The corrected author list is as shown above.In addition, the authors regret that an incorrect version of Fig. 4b was included in the original article. The correct version of Fig. 4 is as shown below.Open in a separate windowFig. 4Dynamic characteristics of photocurrent response times using a laser diode at a light intensity of 650 μW cm−2 at 525 nm. (a) Photocurrent response time under −0.1 V at a pulsed frequency of 1 MHz. (b) Cut-off frequency for the perovskite photodetector under −0.1 V.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Correction for ‘Split-anion solvent extraction of light rare earths from concentrated chloride aqueous solutions to nitrate organic ionic liquids’ by Mercedes Regadío et al., RSC Adv., 2018, 8, 34754–34763, DOI: 10.1039/c8ra06055j.

The authors regret that an incorrect figure caption was given for Fig. 5. The correct version is presented below.Open in a separate windowFig. 5Viscosity as a function of the temperature and the organic phase composition: (1) after loading 39 g L−1 of REE in 20 v% Cy923 in [C101][NO3], (2) pure [C101][NO3], (3) 20 v% Cy923 in [C101][NO3] and (4) pure Cy923.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

5.
Four thiophene-containing tetraphenylethene derivatives were successfully synthesized and characterized. All these highly fluorescent compounds showed typical aggregation-induced emission (AIE) characteristics and emitted different fluorescence colors including blue-green, green, yellow and orange in the aggregation state. In addition, these luminogens also exhibited various mechanofluorochromic phenomena.

Four thiophene-containing AIE-active TPE derivatives were synthesized. Furthermore, these luminogens exhibited various mechanofluorochromic phenomena.

High-efficiency organic fluorescent materials have attracted widespread attention due to their potential applications in organic light-emitting devices and fluorescent switches.1–8 Meanwhile, smart materials sensitive to environmental stimuli have also aroused substantial interest. Mechanochromic luminescent materials exhibiting color changes under the action of mechanical force (such as rubbing or grinding) are one important type of stimuli-responsive smart materials, which can be used as pressure sensors and rewritable media.9–18 Bright solid-state emission and high contrast before and after grinding are very significant for the high efficient application of mechanochromic fluorescence materials.19–28 However, a majority of traditional emissive materials usually exhibit poor emission efficiency in the solid state due to the notorious phenomenon of aggregation caused quenching (ACQ), and the best way to solve the problem is to develop a class of novel luminescent materials oppositing to the luminophoric materials with ACQ effect. Fortunately, an unusual aggregation-induced emission (AIE) phenomenon was discovered by Tang et al. in 2001.29 Indeed, the light emission of an AIE-active compound can be enhanced by aggregate formation.30–32 Obviously, it is possible that AIE-active mechanochromic fluorescent compounds can be applied to the preparation of high-efficiency mechanofluorochromic materials. Numerous luminescent materials exhibiting mechanochromic fluorescent behavior have been discovered up to now.33 Whereas, examples of fluorescent molecules simultaneously possessing AIE and mechanofluorochromic behaviors are still limited, and the exploitation of more AIE-active mechanofluorochromic luminogens is necessary. Organic solid emitters with twisted molecular conformation can effectively prevent the formation of ACQ effect, thus exhibiting strong solid-state luminescence. Tetraphenylethene is a highly twisted fluorophore. Meanwhile, it is also a typical AIE unit, which can be used to construct high emissive stimuli-responsive functional materials.34–37The design and synthesis of novel organic emitters with tunable emission color has become a promising research topic at present. Only a limited number of organic fluorescent materials with full-color emission have been reported to date.38,39 For example, in 2018, Tang et al. reported six tetraphenylpyrazine-based compounds. Interestingly, in film states, these luminogens exhibited different fluorescence colors covering the entire visible range, and this is the first example of realizing full-color emission based on the tetraphenylpyrazine unit.40 It is still an urgent challenge to develop novel organic luminophors with tunable emission color basing on the same core structure.In this study, four organic fluorophores containing tetraphenylethene unit were successfully synthesized (Scheme 1). Introducing the thiophene and carbonyl units into the molecules possibly promoted the formation of weak intermolecular interactions such as C–H⋯S or C–H⋯O interaction, which was advantageous to the exploitation of interesting stimuli-responsive fluorescent materials. Indeed, all these compounds showed obvious AIE characteristics. Furthermore, these luminogens emitted a series of different fluorescent colors involving blue-green, green, yellow and orange in the aggregation state. In addition, these luminogens also exhibited reversible mechanofluorochromic phenomena involving different fluorescent color changes.Open in a separate windowScheme 1The molecular structures of compounds 1–4.To investigate the aggregation-induced properties of compounds 1–4, the UV-vis absorption spectra of 1, 2, 3 and 4 (20 μM) in DMF–H2O mixtures of varying proportions were studied initially (Fig. S1). Obviously, level-off tails were obviously observed in the long-wavelength region as the water content increased. This interesting phenomenon is generally associated with the formation of nano-aggregates.41 Next, the photoluminescence (PL) spectra of 1–4 in DMF–H2O mixtures with various water fraction (fw) values were explored. As shown in Fig. 1, almost no PL signals were noticed when a diluted DMF solution of luminogen 1 was excited at 365 nm, and thus almost no fluorescence could be observed upon UV illumination at 365 nm, and the corresponding absolute fluorescence quantum yield (Φ) was as low as 0.04%. However, when the water content was increased to 50%, a new blue-green emission band with a λmax at 501 nm was observed, and a faint blue-green fluorescence was noticed under 365 nm UV light. As the water content was further increased to 90%, a strong blue-green emission (Φ = 30.81%) could be observed. Furthermore, as shown in Fig. S2, the nano-aggregates (fw = 90%) obtained were confirmed by dynamic light scattering (DLS). Therefore, the compound 1 with bright blue-green emission caused by aggregate formation showed typical AIE feature.Open in a separate windowFig. 1(a) Fluorescence spectra of the dilute solutions of compound 1 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different water contents (0–90%). Excitation wavelength = 365 nm. (b) Fluorescence images of 1 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different fw values under 365 nm UV light.Similarly, as can be seen in Fig. 2–4, compounds 2–4 also showed obvious aggregation-induced green emission, aggregation-induced yellow emission, and aggregation-induced orange emission, respectively. When the water content was zero, the quantum yields of compounds 2–4 were 0.04%, 0.05% and 0.46%, respectively, while as the water content increased to 90%, the corresponding quantum yields of compounds 2–4 also increased to 30.67%, 45.57% and 26.53%, respectively. Hence, luminogens 2–4 were also AIE-active species. In addition, as shown in Fig. 5, the DFT calculations for the compounds 1–4 were performed. The calculated energy gaps (ΔE) of four compounds were 3.6178416 eV (compound 1), 3.276084 eV (compound 2), 3.3073755 eV (compound 3) and 3.0766347 eV (compound 4) respectively. Therefore, the various numbers and the various kinds of the substituents had slight effects on their molecular orbital energy levels of 1–4.Open in a separate windowFig. 2(a) Fluorescence spectra of the dilute solutions of compound 2 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different water contents (0–90%). Excitation wavelength = 365 nm. (b) Fluorescence images of 2 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different fw values under 365 nm UV light.Open in a separate windowFig. 3(a) Fluorescence spectra of the dilute solutions of compound 3 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different water contents (0–90%). Excitation wavelength = 365 nm. (b) Fluorescence images of 3 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different fw values under 365 nm UV light.Open in a separate windowFig. 4(a) Fluorescence spectra of the dilute solutions of compound 4 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different water contents (0–90%). Excitation wavelength = 365 nm. (b) Fluorescence images of 4 (2.0 × 10−5 mol L−1) in DMF–water mixtures with different fw values under 365 nm UV light.Open in a separate windowFig. 5(a) HOMO and LUMO frontier molecular orbitals of molecule 1 based on DFT (B3LYP/6-31G*) calculation. (b) HOMO and LUMO frontier molecular orbitals of molecule 2 based on DFT (B3LYP/6-31G*) calculation. (c) HOMO and LUMO frontier molecular orbitals of molecule 3 based on DFT (B3LYP/6-31G*) calculation. (d) HOMO and LUMO frontier molecular orbitals of molecule 4 based on DFT (B3LYP/6-31G*) calculation.Subsequently, the mechanochromic fluorescent behaviors of compounds 1–4 were surveyed by solid-state PL spectroscopy. As shown in Fig. 6, the as-synthesized powder sample 1 exhibited an emission band with a λmax at 444 nm, corresponding to a blue fluorescence under 365 nm UV light. Intriguingly, a new blue-green light-emitting band with a λmax at 507 nm was observed after the pristine solid sample was ground. After fuming with dichloromethane solvent vapor for 1 min, the blue-green fluorescence was converted back to the original blue fluorescence. Therefore, luminogen 1 exhibited reversible mechanochromic fluorescence feature. Furthermore, this reversible mechanofluorochromic conversion was repeated many times by grinding-exposure without showing signs of fatigue (Fig. 10).Open in a separate windowFig. 6(a) Solid-state PL spectra of compound 1 before grinding, after grinding, and after treatment with dichloromethane solvent vapor. Excitation wavelength: 365 nm. Photographic images of compound 1 under 365 nm UV light: (b) the as-synthesized powder sample. (c) The ground sample. (d) The sample after treatment with dichloromethane solvent vapor.Open in a separate windowFig. 10Repetitive experiment of mechanochromic behavior for compound 1.Similarly, as evident from Fig. 7–9, luminogens 2–4 also exhibited obvious mechanofluorochromic characteristics. Moreover, the repeatabilities of their mechanochromic behaviors were also satisfactory (Fig. S3). Hence, all the compounds 1–4 showed reversible mechanofluorochromic phenomena involving different fluorescent color changes, and the various numbers of the substituents could effectively influence the mechanofluorochromic behaviors of 1–4. Obviously, luminogen 3 or 4 after grinding exhibited more red-shifted fluorescence in comparison with that of the corresponding luminogen 1 or 2 after grinding.Open in a separate windowFig. 7(a) Solid-state PL spectra of compound 2 before grinding, after grinding, and after treatment with dichloromethane solvent vapor. Excitation wavelength: 365 nm. Photographic images of compound 2 under 365 nm UV light: (b) the as-synthesized powder sample. (c) The ground sample. (d) The sample after treatment with dichloromethane solvent vapor.Open in a separate windowFig. 8(a) Solid-state PL spectra of compound 3 before grinding, after grinding, and after treatment with dichloromethane solvent vapor. Excitation wavelength: 365 nm. Photographic images of compound 3 under 365 nm UV light: (b) the as-synthesized powder sample. (c) The ground sample. (d) The sample after treatment with dichloromethane solvent vapor.Open in a separate windowFig. 9(a) Solid-state PL spectra of compound 4 before grinding, after grinding, and after treatment with dichloromethane solvent vapor. Excitation wavelength: 365 nm. Photographic images of compound 4 under 365 nm UV light: (b) the as-synthesized powder sample. (c) The ground sample. (d) The sample after treatment with dichloromethane solvent vapor.In order to further explore the possible mechanism of mechanofluorochromism of 1–4, the powder X-ray diffraction (PXRD) measurements of various solid states of 1–4 were carried out. As depicted in Fig. 11, the pristine solid powder 1 showed many clear and intense reflection peaks, suggesting its crystalline phase. However, after the pristine powder sample was ground, the sharp and intense diffraction peaks vanished, which indicated the crystalline form was converted to the amorphous form. Interestingly, when the ground solid sample was fumigated with dichloromethane solvent vapor for 1 min, the corresponding sample powder exhibited the PXRD pattern of the initial crystalline form. Meanwhile, the structural transformations of the solid samples of 2–4 were similar to that of 1 (Fig. S4–S6). Obviously, the morphological changes of solid samples of 1–4 from crystalline state to amorphous state and vice versa could be attributed to the reversible mechanical switching in compounds 1–4, and the mechanofluorochromic phenomena observed in 1–4 were related to the morphological transition involving the ordered crystalline phase and the disordered amorphous phase.Open in a separate windowFig. 11XRD patterns of compound 1: unground, ground and after treatment with dichloromethane solvent vapor.Fortunately, single crystals of compounds 1 and 2 were obtained by slow diffusion of n-hexane into a trichloromethane solution containing small amounts of 1 or 2. As shown in Fig. 12 and and13,13, the molecular structures of 1 and 2 exhibited a twisted conformation due to the existence of tetraphenylethene unit. Meanwhile, some weak intermolecular interactions, such as C–H⋯π interaction (d = 2.866 Å) for 1, π⋯π interaction (d = 3.371 Å) for 1, C–H⋯S interaction (d = 2.977 Å) for 2, and π⋯π interaction (d = 3.189 Å) for 2, were observed. These weak intermolecular interactions gave rise to a loose packing motif of 1 or 2, which indicated their ordered crystal packings might readily collapse upon exposure to external mechanical stimulus. Therefore, their solid-state fluorescence could be adjusted by mechanical force.Open in a separate windowFig. 12The structural organization of compound 1.Open in a separate windowFig. 13The structural organization of compound 2.In summary, four fluorescent molecules containing thiophene and tetraphenylethene units were successfully designed and synthesized in this study. All these compounds showed obvious AIE characteristics. Furthermore, these luminogens emitted various fluorescence colors involving blue-green, green, yellow and orange in the aggregation state. Meanwhile, these luminogens basing on the same core structure also exhibited reversible mechanofluorochromic phenomena involving different fluorescent color changes. The results of this study will be beneficial for the exploitation of novel luminophors with full-color emission.  相似文献   

6.
7.
Correction for ’Three-dimensional directional nerve guide conduits fabricated by dopamine-functionalized conductive carbon nanofibre-based nanocomposite ink printing’ by Shadi Houshyar et al., RSC Adv., 2020, 10, 40351–40364, DOI: 10.1039/D0RA06556K.

The authors regret that an incorrect version of Fig. 2 was included in the original article. The correct version of Fig. 2 is presented below.Open in a separate windowFig. 2(a) FTIR spectra of pure PCL and PCL printed with CNF and DA (40 and 100 μg mL−1), where circles emphasize the OH peak (3700 cm−1) of the carboxylated CNF and NH peak (1565 cm−1) of dopamine. (b) Shear stress of the CNF and CNF + DA nanocomposite inks versus shear rate. (c) Viscosity versus shear rate of the prepared nanocomposite inks.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

8.
Biosurfactant lipopeptide is a promising dispersant over varieties of chemical ones in oil-spill remediation. The toxicity, biodegradability and performance of the biosurfactant lipopeptide are studied in this paper.

Biosurfactant lipopeptide is a promising dispersant over varieties of chemical ones in oil-spill remediation.

Dispersants were globally applied to physico-chemically enhance the dispersion of oil in water and were assumed to stimulate oil biodegradation by indigenous microorganisms and to reduce the environmental impact of oil spills.1,2 Since the 1960s,3,4 chemical dispersants have been applied as an emergency response to oil spills in marine ecosystems,5 and have showed effectiveness at removing oil slicks from the coast.3,6,7 However, most of the chemically synthesized dispersants are inherently toxic to various aquatic species and hardly biodegradable in the natural environment.2,8 The application of chemical dispersants in the 2010 Gulf of Mexico oil spill also raised concerns regarding the toxicity and the potential environmental impact,9,10 and caused a debate about the effectiveness of chemical dispersants on the rates of oil biodegradation.11 Biosurfactants are promising dispersants in oil-spill remediation, owning to their environmentally friendly and biodegradable properties.12 Chemical surfactants could be replaced with biosurfactants and this change would diminish the environmental impact of traditional dispersants.8,13,14 Lipopeptide produced by microorganisms is one of the representative biosurfactants and has showed great potential applications in food,15 medicine,16 microbial enhanced oil recovery,17 and other fields.18 Nevertheless, the knowledge about the application of biosurfactant lipopeptide in marine oil-spill remediation is still limited.In the present work, the dispersion effectiveness, aquatic toxicity, biodegradability and environmental compatibility of the biosurfactant lipopeptide were determined using recognized standardized methods,19–23 and the biosurfactant lipopeptide used as a bio-dispersant for marine oil-spill remediation were studied, which is, to the best of our knowledge, the first report about biosurfactant lipopeptide used in oil-spill remediation.The lipopeptide samples were isolated from cell-free broth of B. subtilis HSO121 at our laboratory.24 The typical chemical structure of the lipopeptide used in the study was shown in Fig. 1 and its critical micelle concentration (CMC) was 8.69 × 10−5 mol L−1.Open in a separate windowFig. 1Typical chemical structure of lipopeptides (a) and the surface tensions of lipopeptides respect to concentrations (b).Dispersion effectiveness (DE) of lipopeptides was examined at different surfactant-to-oil ratios (SORs), temperatures, pH values, and salinities. It indicated in Fig. 2 that DE of lipopeptides reached 70.23% at SORs of 1 : 10 (w/w) at 25 °C, pH 7 and the present of 3% NaCl (w/v). It should be noticed that the DE of lipopeptides was almost kept when SORs dropped to 1 : 250 w/w. The increase in DE with increasing SORs can be attributed to the generation of emulsions with smaller droplets and lower rising velocity.2 Sharp drop off in DE was observed for lipopeptides when SORs below 1 : 500 (w/w), and DE value was 36.45% at an extreme SORs of 1 : 1250 (w/w). It had been reported that the abrupt decline for 80 : 20 lecithin : Tween 80 (w/w) surfactant happened when SORs below 1 : 100 v/v, from 77% (SORs 1 : 100 v/v) to 15% (SORs 1 : 200 v/v),25 indicating a lower SOR in lipopeptides usage could reach its maximum effectiveness. Lipopeptides exhibited >70% DE values with temperatures ranged from 15 °C to 25 °C, and an increasing DE values when pH values raised, the largest DE was 77.45% at pH 11. DE of lipopeptides increased from 56.12% to 71.14% with increase in salinity. Higher DE at higher salinity was observed for anionic biosurfactants, which can be attributed to the electrostatic repulsion between polar head groups reduced by ions, and a close-packed arrangement of surfactant molecules at the oil–water interface were formed.2Open in a separate windowFig. 2The dispersion effectiveness (DE) of lipopeptides under different SORs (), temperatures (), pH values () and salinities ().Mortalities of zebrafish under different concentrations of different surfactants were shown in Fig. 3. It was evident that the toxicity of lipopeptides was far less than those of sodium dodecyl sulfate (SDS) and 3-(N,N-dimethyl palmityl ammonio) propane sulfonate (Betaine). The 24 h median lethal concentration (LC50) values were calculated and showed in 26 in which an 96 h LC50 of 1.9 mg L−1 for Cyprinodon variegatus was reported. Low toxicities of lipopeptides on whiteleg shrimp and copepods were also evaluated that the 96 h LC50 of lipopeptides from Bacillus sp. GY19 were 1050 mg L−1 and 1174 mg L−1, respectively.27Open in a separate windowFig. 3Mortality of zebrafish (Mz) after a 24 h exposure to Betaine, SDS or lipopeptides.Ecotoxicity of tested surfactants to the zebrafish
SurfactantLC50 (mg L−1)95% confidence intervals (mg L−1) r
Lipopeptide11451090–12290.981
SDS8.257.75–8.770.998
Betaine0.8720.853–0.8900.988
Open in a separate window Fig. 4 illustrated the evolution of Pseudokirchneriella subcapitata concentrations in the algal growth inhibition test of lipopeptides and SDS. Growth rates of P. subcapitata were decreased with the increase of surfactant concentrations. 72 h median effect concentration (EC50) values were calculated using linear regression analysis based on the dose–response curves,28 and the 72 h EC50 value of lipopeptide was 1703 mg L−1, which was about 45 times higher than that of SDS, 36.51 mg L−1. EC50 was in well accordance with LC50 mentioned above, indicating that lipopeptides showed a much lower toxicity than that of SDS. The 72 h median inhibitory concentration (IC50, equivalent to EC50) of SDS on Raphidocelis subcapitata was 36.58 mg L−1,29 which was relatively close to the result in this work. However, De Oliveira et al.28 showed that the EC50 of crude surfactin on Selenastrum capricornutum (named as well as P. subcapitata) from B. subtilis ICA56 was 49.3 mg L−1. The lower toxicity of lipopeptides in our study was probably because the lipopeptides from various Bacillus sp. strains might have different activity.27 The EC50 values for 9 types of surfactants including anionic surfactants, nonionic surfactants, and zwitterionic surfactants on P. subcapitata were range from 1.5 to 4.4 mg L−1.30 Hence, according to data mentioned, lipopeptides from HSO121 in the present showed less toxicity.Open in a separate windowFig. 4Effect of lipopeptides and SDS on growth concentrations (c) of Pseudokirchneriella subcapitata during 72 hours of incubation.Biodegradabilities of lipopeptides and sodium dodecyl benzene sulfonate (SDBS) versus time were shown in Fig. 5. With the same initial concentrations, 30 mg L−1, biodegradability values of biosurfactant lipopeptides and the synthetic surfactant SDS after 7 days incubation were 100% and 98.83%, respectively. Lipopeptides degraded much faster than SDBS that the degradability was nearly 100% after 3 days. Biodegradability of lipopeptides by P. putida CECT 324 strain, around 82% after 3 days, was reported, which was higher than that of amine oxides.29 Lipopeptides showed higher biodegradability as 94.01% even increasing the initial concentration to 300 mg L−1. Lima et al.31 studied the biodegradability of surfactants and observed the lowest decrease (24.8%) in SDS, the highest decrease in lipopeptides (69.1%) and glycolipid (73.4%). Biosurfactants were seemed to be more biodegradable than synthetic surfactants.Open in a separate windowFig. 5Degradation of surfactants (DS) including lipopeptides and SDBS during 7 d test in activated sludge systems.The degradation rates of aliphatic fractions (n-C11–C29) in crude oil (Xinjiang oilfield, P. R. China) were analysis using GC-MS and the results were illustrated in Fig. 6. Significant degradation rates of n-alkanes were observed in presence of biosurfactant lipopeptides, while n-alkanes dispersed by commercial dispersant degraded similar to that in control group (in absence of any surfactant). It was well known that the degradation rates of alkanes decreased and finally vanished with increase in the chain length of hydrocarbon. In the present study, lipopeptides showed excellent activity in accelerating degradation of long-chain hydrocarbons after 2 days. The alkanes degradation rates treated by lipopeptides after 1 day and 2 days were 38.78% (Fig. 6a) and 71.45% (Fig. 6b), respectively, which were much higher than those of commercial dispersant-treated group (9.16% and 34.16% after 1 day and 2 days treatment) and control group (13.26% and 33.55% after 1 day and 2 days treatment). It was reported that commercial chemical dispersants such as Corexit 9500A and GM-2 made no enhancement to the degradation of the petroleum hydrocarbons, whereas biosurfactants such as rhamnolipids enhanced the degradation.32,33 Degradation stimulation by lipopeptides could be attributed to their good dispersion activities and biocompatible. Dispersed oil droplets formed and considerable interfacial area was available to the microorganisms followed by microbial bioremediation. In addition, microorganisms could utilize the nutrients derived from culture broth as an excellent substrate for growth.34Open in a separate windowFig. 6Degradation of n-alkanes in crude oil (DO) either non-dispersed (control) or dispersed by commercial dispersant (001#) and lipopeptides at day 1 (a), day 2 (b).  相似文献   

9.
10.
An aggregation-induced emission enhancement (AIEE)-active fluorescent sensor based on a tetraphenylethene (TPE) unit has been successfully designed and synthesized. Interestingly, the luminogen could detect Hg2+ with high selectivity in an acetonitrile solution without interference from other competitive metal ions, and the detection limit was 7.46 × 10−6 mol L−1. Furthermore, the luminogen also showed interesting solvatochromic behavior and superior cell imaging performance.

A TPE-based AIEE-active fluorescent sensor for Hg2+ was synthesized. Furthermore, it showed solvatochromism and cell imaging characteristics.

The design and synthesis of molecular sensors for the detection of metal ions, especially transition-metal ions, has attracted much attention.1–3 Among all transition-metal ions, mercury (Hg2+) is identified as one of the most dangerous and ubiquitous heavy metals. Indeed, it is not biodegradable, and can cause extreme injury to the environment as well as human health.4–8 Additionally, it can be accumulated through the food chain in the human body, consequently giving rise to several deleterious effects such as central nervous system defects, kidney damage, endocrine system disease and so on.9–12 Although many governments around the world have adopted strict regulations to limit Hg2+ emission, the global Hg2+ pollution caused by human activities is still serious.13,14 Therefore, it is highly desirable to develop a new method for the detection of Hg2+ with high selectivity and sensitivity.15–17Most traditional fluorescent sensors suffer from a detrimental phenomenon called aggregation-caused quenching (ACQ), which usually results in a poor solid-state emission efficiency. Fortunately, in 2001, Tang et al. reported a fluorescent molecule named 1-methyl-1,2,3,4,5-pentaphenylsilole. Interestingly, the fluorescence emission of this luminogen was induced by aggregation, a phenomenon referred to as aggregation-induced emission (AIE).18 Subsequently, in 2002, Park et al. reported an interesting phenomenon named aggregation-induced emission enhancement (AIEE).19 In fact, both AIE and AIEE can achieve highly efficient emission in the solid state or aggregated state.20–27 In the past years, AIE (or AIEE) phenomenon has attracted considerable research interest owing to the potential applications in a lot of fields, including bioimaging, fluorescence sensors, organic lighting emitting diode (OLED) devices and organic lasers.28–33 Meanwhile, many stimuli-responsive materials have been reported, including photochromism, mechanochromism, and solvatochromism.34–46 At present, the solvatochromism materials have been generally used in chemical and biological systems.47 To date, many fluorescent chemosensors for Hg2+ have been reported. In contrast, the corresponding chemosensors with AIE or AIEE effect are rare, not to mention solvatochromic AIE or AIEE-active fluorescent sensors for Hg2+ with good cell imaging behavior. Indeed, preparing such multifunctional sensors is challenging and significative. In this paper, we reported an AIEE-active tetraphenylethene (TPE)-based fluorescent sensor (Scheme 1) for the detection of Hg2+. Furthermore, the luminogen also showed remarkable solvatochromism and good cell imaging characteristics.Open in a separate windowScheme 1The synthetic route of luminogen 1.To investigate the AIEE phenomenon of luminogen 1, we initially studied the UV-visible absorption spectra and photoluminescence (PL) spectra in acetonitrile–water mixtures with different water fractions (fw). The results indicated that the absorption spectra exhibited level-off tails in the long wavelength region as the water content increased (Fig. S1). It is well-known that such tails are usually associated with the Mie scattering effect, which is the key signal of nano-aggregate formation.48,49 As presented in Fig. 1, luminogen 1 showed weak fluorescence and the luminescence quantum yield (Φ) was as low as 1.35%. Interestingly, when the fw in the acetonitrile solution was increased to 80%, an obvious emission band was formed, and a yellow-green fluorescence was observed. As the water content was increased to 90%, the emission intensity was further increased, and a bright yellow-green luminescence (Φ = 27.81%) with a λmax at 545 nm could be observed.Open in a separate windowFig. 1(a) Fluorescence spectra of the dilute solutions of 1 (2.0 × 10−5 mol L−1) in acetonitrile–water mixtures with different water content (0–90%). Excitation wavelength = 380 nm. (b) Changes the emission intensity of 1 at 545 nm in acetonitrile–water mixtures with different volume fractions of water (0–90%). (c) PL images of 1 (2.0 × 10−5 mol L−1) in acetonitrile–water mixtures with different fw values under 365 nm UV illumination.Clearly, water is a poor solvent of luminogen 1. As a result, the generation of the yellow-green emission can be attributed to the aggregate formation.50–52 Moreover, as shown in Fig. 2, the nano-aggregates obtained were verified by dynamic light scatterings (DLS). Therefore, 1 is a typical AIEE-active fluorescent molecule, and its AIEE behavior is caused by the restricted intramolecular rotation. As shown in Fig. S2, solid-state compound 1 showed a strong green emission (Φ = 14.50%) with a λmax at 497 nm, and the corresponding fluorescence lifetime is 2.66 ns (Fig. S3).Open in a separate windowFig. 2Size distribution curve of luminogen 1 (2.0 × 10−5 mol L−1) in acetonitrile–water mixtures with fw = 90%.On the other hand, the luminogen 1 also displayed interesting solvatochromism effect. As presented in Fig. 3, the absorption spectra and fluorescence spectra of luminogen 1 in different solvents were investigated, respectively. Obviously, the absorption spectra was barely affected by the polarity of solvents. However, the photoluminescence peaks were gradually red-shifted from 515 nm to 625 nm, and thus 1 exhibited remarkable solvatochromism behavior. Obviously, the molecular structure of luminogen 1 is distorted due to the presence of TPE unit, and the conjugation degree of molecule 1 is different in various solvents, and the intramolecular charge transfer (ICT) effect is possibly responsible for the solvatochromism behavior of 1.Open in a separate windowFig. 3(a) UV-Vis absorption spectra and (b) normalized fluorescence spectra (Excitation wavelength = 380 nm) of luminogen 1 in different solvents (2.0 × 10−5 mol L−1). (c) Photographs of 1 under 365 nm UV illumination in different solvents: tol (toluene); THF (tetrahydrofuran); EA (ethyl acetate); BT (acetone); MeCN (acetonitrile); DMSO (dimethyl sulfoxide).Subsequently, the changes in the fluorescence of luminogen 1 induced by Hg2+ were investigated in acetonitrile (2.0 × 10−5 mol L−1) at room temperature. In the fluorometric titration experiments, as shown in Fig. 4, the emission intensity significantly decreased when the Hg2+ concentration increased from 0 to 7.0 equivalents in acetonitrile. Meanwhile, the fluorescent color changed from orange-red to colorless, and followed by a plateau upon further titration (Fig. 5). Remarkably, the emission intensity of luminogen 1 was almost quenched completely. Furthermore, based on the titration experiments, the detection limit of luminogen 1 for Hg2+ on the basis of LOD = 3 × σ/B (where σ is the standard deviation of blank sample and B is the slope between the fluorescence intensity versus Hg2+ concentration) was 7.46 × 10−6 M (Fig. S4). Moreover, a good linear relationship could be obtained (R = −0.9933) and the quenching constant of luminogen 1 with Hg2+ was 1.9 × 104 M−1 (Fig. S5). On the other hand, the binding ratio of luminogen 1 for Hg2+ was established through Job''s plot and the results showed 1 : 1 stoichiometric complexation (Fig. S6). Next, the binding interactions between 1 and Hg2+ were further investigated by NMR in acetonitrile-d3. As presented in Fig. 6, the signal of Ha from 9.42 ppm shifted to 9.62 ppm and the Hb changed from 8.43 ppm to 9.24 ppm. Besides, the Hc or Hd was slightly shifted for 0.04 ppm or 0.06 ppm, respectively. These consequences revealed that the N on the pyridine and the N on the pyrazine (near the N on the pyridine) are the most probable binding with Hg2+. Mass spectra were utilized to further demonstrate the binding mode of luminogen 1 toward Hg2+. The peak located at m/z = 820.0 was coincided well with the ensemble [1 + Hg2+ +2NO3 + Cl], confirming the binding ratio of luminogen 1 for Hg2+ with 1 : 1 stoichiometry (Fig. S7).Open in a separate windowFig. 4Fluorescence titration spectra of luminogen 1 (2.0 × 10−5 mol L−1) induced by Hg2+ (0–7.0 equiv.) in an acetonitrile solution. Excitation wavelength = 380 nm.Open in a separate windowFig. 5The emission intensity changes of luminogen 1 at 625 nm with different equivalents of Hg2+.Open in a separate windowFig. 6 1H NMR (acetonitrile-d3, 400 MHz) spectra changes of luminogen 1 in the presence of Hg2+.Next, in order to study the selectivity behavior of luminogen 1 as a fluorescent sensor for Hg2+, other metal ions, such as Zn2+, Cd2+, Ba2+, Sr2+, Mn2+, Mg2+, Ca2+, Pb2+, Ni2+, Co2+, Cu2+, Al3+, Fe3+, Cr3+, Ag+ and K+ were also measured in acetonitrile under the same experimental conditions. The corresponding UV-Vis absorption spectra were shown in Fig. S8. Furthermore, as showed in Fig. 7, when these cations were added separately into the solution containing luminogen 1, no obvious fluorescence changes were observed. Indeed, as noticed in Fig. S9, no obvious interference was observed when Hg2+ (7.0 equiv.) was added with other ions (7.0 equiv.). These results indicated that luminogen 1 could be served as a highly selective fluorescent sensor for detection of Hg2+.Open in a separate windowFig. 7(a) Fluorescence spectra of luminogen 1 (2.0 × 10−5 mol L−1) towards various cations including Zn2+, Cd2+, Hg2+, Ba2+, Sr2+, Mn2+, Mg2+, Ca2+, Pb2+, Ni2+, Co2+, Cu2+, Al3+, Fe3+, Cr3+, Ag+ and K+. Excitation wavelength = 380 nm. (b) Fluorescence photographs of luminogen 1 after addition of various metal ions (7.0 equiv.) under 365 nm light.Fluorescent probe is a powerful tool for optical imaging, which allows direct visualization of biological analytes.53 Luminogen 1 was AIEE-active due to the restriction of intramolecular rotation in the aggregate state, which is beneficial for cell imaging. Indeed, the viability of HeLa cells incubated with luminogen 1 was evaluated by the standard MTT method (Fig. 8), and the result indicated that compound 1 exhibited low cytotoxicity. Next, cell imaging behavior of luminogen 1 was investigated using a confocal laser scanning microscopy (CLSM). HeLa cells were incubated with luminogen 1 (20 μM) for 30 min at 37 °C and the fluorescence images were obtained by CLSM.Open in a separate windowFig. 8The MTT assay of compound 1 for measuring cell viability.As presented in Fig. 9, an intense yellow-green fluorescence, which was consistent with the fluorescence of luminogen 1 in acetonitrile–water mixture with high water content (80% or 90%), was displayed inside the cells. This result indicated that luminogen 1 tended to aggregate inside the cells, and thus the bright yellow-green luminescence was clearly observed. Furthermore, the merged picture c demonstrated that picture a and picture b overlapped very well. These results indicated that luminogen 1 showed superior cell imaging performance.Open in a separate windowFig. 9Fluorescence images of HeLa cells incubated with luminogen 1 (20 μM) for 30 min at 37 °C: (a) bright field image; (b) fluorescence image; (c) merge image (a and b).In summary, a TPE-based fluorescent molecule was designed and synthesized. The luminogen exhibited obvious AIEE phenomenon. Moreover, luminogen 1 could be used as a highly selective fluorescence turn-off chemosensor for Hg2+, and the detection limit was 7.46 × 10−6 mol L−1. Furthermore, 1 also displayed interesting solvatochromic behavior and superior cell imaging performance. Further studies on the design of new AIE or AIEE-active fluorescent chemosensors are in progress.  相似文献   

11.
Correction for ‘iTRAQ-based quantitative proteomic analysis for identification of biomarkers associated with emodin against severe acute pancreatitis in rats’ by Hong Xiang et al., RSC Adv., 2016, 6, 72447–72457.

The authors regret that Fig. 2–4 were shown incorrectly in the original article. An incorrect section of the SAP group in the MPO-immunohistochemical staining (Fig. 2A) and HE staining (Fig. 3) experiments was used in error. In addition, Fig. 4 has been revised to show the zymogen granule, in order to better represent the ultrastructure of the pancreas. The correct versions of Fig. 2–4 are shown below.Open in a separate windowFig. 2Emodin down-regulated the MPO protein expression in pancreas of SAP rats. (A) Effect of emodin on MPO-immunopositive area (brown) staining of pancreatic tissue in SAP rats by immunohistochemical detection. (B) Effect of emodin on MPO-immunopositive area (red) staining of pancreatic tissue in SAP rats by immunofluorescence detection. Images are presented at 200× magnification. The data are presented as the mean ± SD, n = 6. **P < 0.01 versus SO; #P < 0.05 versus SAP, ##P < 0.01 versus SAP.Open in a separate windowFig. 3Emodin improved pancreatic histopathology of SAP rats. Effect of emodin on H&E staining of pancreatic tissue in SAP rats. Images are presented at 200× magnification. The data are presented as the mean ± SD, n = 6. **P < 0.01 versus SO; #P < 0.05 versus SAP, ##P < 0.01 versus SAP.Open in a separate windowFig. 4Emodin attenuated cellular structure changes in pancreas of SAP rats. Representative images of the cells’ ultrastructure in the SO (A), SAP (B), 60 mg kg−1 emodin (C), 30 mg kg−1 emodin (D) and 15 mg kg−1 emodin (E) groups. Images are presented at 25 000× magnification.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

12.
Correction for ‘Facile synthesis of CuS mesostructures with high photothermal conversion efficiency’ by Lianjiang Tan et al., RSC Adv., 2015, 5, 35317–35324. DOI: 10.1039/C5RA01835H.

The authors wish to correct Fig. 6c–e in the original version of this paper as it unfortunately contained incorrect images. The corrected version of Fig. 6c–e is provided below as Fig. 1. The authors confirm that these errors do not affect the scientific findings and conclusions of the paper, and sincerely apologize for the errors and any confusion.Open in a separate windowFig. 1Fluorescence images of HeLa cells before and after irradiation by 980 nm laser with the power density of 0.5 W cm−2 over a period of 5 min in the absence (c and d) and presence (e and f) of the CuS mesostructures with the concentration of 0.3 g L−1. The living cells were labeled by calcein AM (green emission), and the dead cells were labeled by propidium iodide (red emission). The scale bar represents 100 mm.  相似文献   

13.
Correction for ‘Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4’ by Natda Wetchakun et al., RSC Adv., 2017, 7, 13911–13918, DOI: 10.1039/C6RA27138C.

The authors regret errors in Fig. 4, ,7,7, and 9 in the previously published article. The corrections for the errors in the article are described as follows:Open in a separate windowFig. 4Kubelka–Munk absorbance spectra and band gaps (insets) of the pure InVO4 (a) and 1.0 mol% Cu-doped InVO4 (b) samples.Open in a separate windowFig. 7Schematic of the charge migration and separation on Cu-doped InVO4.(1) The diffuse reflectance spectra of pure InVO4 and 1.0 mol% Cu-doped InVO4 are shown in Fig. 4. The absorption margin of 1.0 mol% Cu-doped InVO4 was shifted to a longer wavelength, indicating a decrease in the band gap with respect to pure InVO4. The absorption margins of the pure InVO4 and 1.0 mol% Cu-doped InVO4 samples were 505 nm and 510 nm, corresponding to band gaps of 2.51 eV and 2.45 eV, respectively (Fig. 4a and b).(2) The band edge positions of the conduction band (CB) and the valence band (VB) of InVO4 can be calculated by the following equation: E0CB = χEC − 0.5Eg,1 where χ is the electronegativity of the semiconductor, EC is the energy of free electrons on the hydrogen scale of 4.5 eV, Eg is the band gap of InVO4, and the χ value of InVO4 is 5.74 eV.2 The Eg value of InVO4 evaluated from the UV-vis DRS analysis was about 2.51 eV. The valence band energy (EVB) can be calculated by the following equation:3EVB = ECB + Eg, where ECB is the conduction band energy. Based on the equation above, the calculated CB and VB edge potentials of InVO4 were −0.02 eV and 2.49 eV, respectively. Now, we are in a position to discuss the photocatalytic mechanism of Cu-doped InVO4 for MB degradation (Fig. 7). In the photocatalysis process, when the absorbed photon energy () equals or exceeds the band gap, the Cu-doped InVO4 generates electron–hole (e/h+) pairs. In that case, the generated electrons from the valence band can be transferred to the conduction band of InVO4. Since the CB edge potential of InVO4 (−0.02 eV) is higher than the standard redox potential, E0(O2/O2˙) = −0.33 V vs. NHE at pH 7, this suggests that the electrons in the CB of InVO4 cannot reduce O2 to the superoxide radical ion (O2˙). In addition, the VB of InVO4 (2.49 eV) is higher than the standard redox potential, E0(OH/OH˙) = 1.99 V vs. NHE at pH 7. This indicates that the photogenerated holes in the valence band of InVO4 can oxidize the hydroxyl ion (OH) or water (H2O) to form the hydroxyl radical (OH˙).(3) Due to the contradiction between the scavenging test and the proposed photocatalytic mechanism, Fig. 9 was removed from the original article.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

14.
3,3-Dimethyl-1-(trifluoromethyl)-1,3-dihydro-1-λ3,2-benziodoxole represents a popular reagent for trifluoromethylation. The σ hole on the hypervalent iodine atom in this “Togni reagent” is crucial for adduct formation between the reagent and a nucleophilic substrate. The electronic situation may be probed by high resolution X-ray diffraction: the experimental charge density thus derived shows that the short intermolecular contact of 3.0 Å between the iodine and a neighbouring oxygen atom is associated with a local charge depletion on the heavy halogen in the direction of the nucleophile and visible polarization of the O valence shell towards the iodine atom. In agreement with the expectation for λ3-iodanes, the intermolecular O⋯I–Caryl halogen bond deviates significantly from linearity.

The experimentally observed electron density for the “Togni reagent” explains the interaction of the hypervalent iodine atom with a nucleophile.

A “halogen bond” denotes a short contact between a nucleophile acting as electron density donor and a (mostly heavy) halogen atom as electrophile;1,2 halogen bonds are a special of σ hole interactions.3–5 Such interactions do not only play an important role in crystal engineering;6–11 rather, the concept of a nucleophile approaching the σ hole of a neighbouring atom may also prove helpful for understanding chemical reactivity.The title compound provides an example for such a σ hole based reactivity: 3,3-dimethyl-1-(trifluoromethyl)-1,3-dihydro-1-λ3,2-benziodoxole, 1, (Scheme 1) commonly known as “Togni reagent I”, is used for the electrophilic transfer of a trifluoromethyl group by reductive elimination. The original articles in which the application of 112 and other closely related “Togni reagents”13 were communicated have been and still are highly cited. Trifluoromethylation is not the only application for hypervalent iodine compounds; they have also been used as alkynylating14 or azide transfer reagents.15,16 The syntheses of hypervalent iodanes and their application in organofluorine chemistry have been reviewed,17 and a special issues of the Journal of Organic Chemistry has been dedicated to Hypervalent Iodine Catalysis and Reagents.18 Recently, Pietrasiak and Togni have expanded the concept of hypervalent reagents to tellurium.19Open in a separate windowScheme 1Lewis structure of the Togni reagent, 1.Results from theory link σ hole interactions and chemical properties and indicate that the electron density distribution associated with the hypervalent iodine atom in 1 is essential for the reactivity of the molecule in trifluoromethylation.20 Lüthi and coworkers have studied solvent effects and shown that activation entropy and volume play relevant roles for assigning the correct reaction mechanism to trifluoromethylation via1. These authors have confirmed the dominant role of reductive elimination and hence the relevance of the σ hole interaction for the reactivity of 1 in solution by ab initio molecular dynamics (AIMD) simulations.21,22 An experimental approach to the electron density may complement theoretical calculations: low temperature X-ray diffraction data of sufficient resolution allow to obtain the experimental charge density and associate it with intra- and inter-molecular interactions.23–25 Such advanced structure models based on aspherical scattering factors have also been applied in the study of halogen bonds.26–30 In this contribution, we provide direct experimental information for the electronic situation in Togni reagent I, 1; in particular, we analyze the charge distribution around the hypervalent iodine atom.Excellent single crystals of the title compound were grown by sublimation.§ The so-called independent atom model (IAM), i.e. the structure model based on conventional spherical scattering factors for neutral atoms, confirms the solid state structure reported by the original authors,12 albeit with increased accuracy. As depicted in Fig. 1, two molecules of 1 interact via a crystallographic center of inversion. The pair of short intermolecular O⋯I contacts thus generated can be perceived as red areas on the interaction-sensitive Hirshfeld surface.31Open in a separate windowFig. 1Two neighbouring molecules of 1, related by a crystallographic center of inversion. The short intermolecular I⋯O contacts show up in red on the Hirshfeld surface32 enclosing the left molecule. (90% probability ellipsoids, H atoms omitted, symmetry operator 1 − x, 1 − y, 1 − z).The high resolution of our diffraction data for 1 allowed an atom-centered multipole refinement33,34 and thus an improved model for the experimental electron density which takes features of chemical bonding and lone pairs into account. Fig. 2 shows the deformation density, i.e. the difference electron density between this advanced multipole model and the IAM in the same orientation as Fig. 1.|| The orientation of an oxygen lone pair (blue arrow) pointing towards the σ hole of the heavy halogen in the inversion-related molecule and the region of positive charge at this iodine atom (red arrow) are clearly visible. Single-bonded terminal halides are associated with one σ hole opposite to the only σ bond, thus resulting in a linear arrangement about the halogen atom. Different geometries and potentially more than a single σ hole are to be expected for λ3-iodanes such as our target molecule, and as a tendency, the resulting halogen bonds are expected to be weaker than those subtended by single-bonded iodine atoms.35 In agreement with these theoretical considerations, the closest I⋯O contacts in 1 amount to 2.9822(9) Å. This distance is significantly shorter than the sum of the van-der-Waals radii (I, 1.98 Å; O, 1.52 Å (ref. 36)) but cannot compete with the shortest halogen bonds between iodine and oxygen37,38 or iodine and nitrogen.39–41Fig. 1 and and22 show that the Caryl–I⋯O contacts are not linear; they subtend an angle C10–I1⋯O1′ of 141.23(3)° at the iodine atom. On the basis of theoretical calculations, Kirshenboim and Kozuch35 have suggested that the split σ holes should be situated in the plane of the three substituents of the hypervalent atom and that halogen and covalent bonds should be coplanar. Fig. 3 shows that the Caryl–I⋯O interaction in 1 closely matches this expectation, with the next oxygen neighbour O1′ only 0.47 Å out of the least-squares plane through the heavy halogen I1 and its three covalently bonded partners C1, O1 and C10.Open in a separate windowFig. 2Deformation density for the pair of neighbouring molecules in 1; the dashed blue and red arrows indicate regions of opposite charge. (Contour interval 0.10e Å−3; blue lines positive, red lines negative, green lines zero contours, symmetry operator 1 − x, 1 − y, 1 − z).Open in a separate windowFig. 3A molecule of 1, shown [platon] along O1⋯C1, and its halogen-bonded neighbour O1i. Symmetry operator 1 − x, 1 − y, 1 − z.The Laplacian, the scalar derivative of the gradient vector field of the electron density, emphasizes local charge accumulations and depletions and it allows to assess the character of intra- and inter-molecular interactions. A detailed analysis of all bonds in 1 according to Bader''s Atoms In Molecules theory42 is provided in the ESI. We here only mention that the electron density in the bond critical point (bcp) of the short intermolecular I⋯O contact amounts to 0.102(5)e Å−3; we are not aware of charge density studies on λ3-iodanes, but both the electron density and its small positive Laplacian match values experimentally observed for halogen bonds involving O and terminal I in the same distance range.43The crystal structure of 1 necessarily implies additional contacts beyond the short halogen bond shown in Fig. 1 and and2.2. The shortest among these secondary interactions is depicted in Fig. 4: it involved a non-classical C–H⋯F contact with a H⋯F distance of 2.55 Å.Open in a separate windowFig. 4C–H⋯F contact in 1; additional information has been compiled in the ESI. Symmetry operators i = 1 − x, 1 − y, 1 − z; ii = 1 − x, 1 − y, −z.The topological analysis of the experimental charge density reveals that this non-classical C–H⋯F hydrogen bond and all other secondary contacts are only associated with very small electron densities in the bcps. Table S8 in the ESI provides a summary of this analysis and confirms that the I⋯O halogen bond discussed in Fig. 1 and and22 represents by far the most relevant intermolecular interaction.The relevance of this halogen bond extends beyond the crystal structure of 1: Insight into the spatial disposition of electrophilic and nucleophilic regions and hence into the expected reactivity of a molecule may be gained from another electron-density derived property, the electrostatic potential (ESP). The ESP for the pair of interacting molecules in 1 is depicted in Fig. 5.Open in a separate windowFig. 5Electrostatic potential for a pair of molecules in 1 mapped on an electron density isosurface (ρ = 0.5e Å−3; program MoleCoolQt44,45). Fig. 5 underlines the complementary electrostatic interactions between the positively charged iodine and the negatively charged oxygen atoms. One can easily imagine to “replace” the inversion-related partner molecule in crystalline 1 by an incoming nucleophile.The ESP tentatively obtained for a single molecule in the structure of 1 did not differ significantly from that derived for the inversion-related pair (Fig. 5), and even the results from theoretical calculations in the gas phase for an isolated molecule20 are in good qualitative agreement with our ESP derived from the crystal structure. In the absence of very short contacts, polarization by neighbouring molecules only has a minor influence on the ESP. The experimentally observed electron density matches the proven reactivity for the title compound, and we consider it rewarding to extend our charge density studies on related hypervalent reagents.  相似文献   

15.
Correction for ‘A sensitive OFF–ON–OFF fluorescent probe for the cascade sensing of Al3+ and F ions in aqueous media and living cells’ by Lingjie Hou et al., RSC Adv., 2020, 10, 21629–21635, DOI: 10.1039/D0RA02848G.

The authors regret that an incorrect version of Fig. 4 was included in the original article. The correct version of Fig. 4 is presented below.Open in a separate windowFig. 4The ESI-MS spectrum of Al3+–HNS complex.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

16.
The development of nanoprobes for selective detection of metal ions in solution has attracted great attention due to their impact on living organisms. As a contribution to this field, this paper reports the synthesis of silver nanoparticles modified with rutin in the presence of ascorbic acid and their successful use as a chromogenic probe for the selective detection of Fe3+ in aqueous solution. Limits of detection and quantification were found to be 17 nmol L−1 and 56 nmol L−1, respectively. The sensing ability is proposed to proceed via an iron-induced nanoparticle growth/aggregation mechanism. A practical approach using image analysis for quantification of Fe3+ is also described.

The use of rutin-modified silver nanoparticles for selective detection and sensitive quantification of Fe3+ in aqueous solution is described.

Metal ions are key species in nature due to their essential functions in living organisms.1,2 On the other hand, heavy metals as well as essential metals at abnormally high levels are toxic.2 Iron, for instance, in addition to its popular use in industry and construction, is essential to the human body and active in biological processes. Although the trivalent form of iron is particularly important for oxygen transport in blood and the mitochondrial respiratory chain, high levels of this cation are associated with important pathologies.3,4 The detection of metal ions in aqueous solution is traditionally performed by methods including atomic absorption spectrometry,5 electrochemical measurements,6,7 and inductively coupled plasma techniques,8 among others. However, these techniques have important drawbacks, notably the need for sophisticated instrumentation, in addition to being time-consuming and requiring laborious procedures. To overcome these issues, the development of chromogenic and fluorogenic chemosensors for the selective detection of metal-targets has attracted great attention, especially due to the possibility of fast, sensitive and non-expensive analysis.9,10 In the last decade, nanoscaled materials have been reported as selective probes for metal ions, including Fe3+.11–24Silver nanoparticles (AgNPs) are of particular interest because of the affordable price of starting materials, ease of controlling size and morphology, possibility to functionalize their surface with organic molecules, and optical properties that enable detection of a variety of analytes via simple UV-vis spectroscopy and digital image analysis. Furthermore, applications of AgNPs are also biotechnologically relevant due to the possibility of green synthetic protocols, including the use of plant extracts,25 natural sources,26 glycerol,27 among others.Flavonoids are secondary metabolites naturally found in fruits and other vegetables with relevant roles due to their nutritional, pharmaceutical and medicinal properties.28 Because of their adequate structural features, flavonoids are candidates to be employed in the synthesis of AgNPs.26 Rutin (RU), a sugar-based flavonoid, may be employed as reducing agent in the synthesis of AgNPs along with a stabilizer such as polyvinylpyrrolidone (PVP)26 or used as crude plant extract component.29,30This paper reports the use of RU-modified AgNPs (RU-AgNPs) as a chromogenic probe for Fe3+ in aqueous medium in the presence of ascorbic acid (AA). Sensing ability of RU-AgNPs for the selective detection of Fe3+ toward other metal cations was investigated with UV-vis spectroscopy analysis. These data and transmission electronic microscopy (TEM) results allowed a mechanistic proposal involved in the selective detection of Fe3+ by RU-AgNPs. Furthermore, a practical approach based on correlation of images of solutions obtained with a conventional smartphone and chemometrics was employed for a simpler quantification of Fe3+ in aqueous medium.Initially, the order of reagent combination was investigated in the synthesis of RU-AgNPs. Concentration of RU ranged from 0.10 to 0.50 mmol L−1, while concentrations of other components were fixed at 0.20 mmol L−1 AgNO3, 0.10 mmol L−1 AA and 0.10 M NaOH. Water was used as solvent in all cases. Narrower surface plasmon resonance (SPR) bands were obtained from adding a solution of AA and NaOH to a solution containing RU and AgNO3 (Fig. 1a) against the addition of RU, AA, and NaOH to AgNO3 solution (Fig. S1a – ESI), or the addition of RU and NaOH to a solution of AA and AgNO3 (Fig. S1b). RU-AgNPs obtained from the condition presented in Fig. 1a are small (4.1 nm average diameter) and considerably polydisperse (standard deviation of 4.7 nm), however, presenting only one population (Fig. 1b and S2). A study of the influence of RU concentration (0.10 to 0.50 mmol L−1) on the stability of RU-AgNPs over time indicated that 0.10 mmol L−1 RU generates more stable RU-AgNPs (Fig. S3). Next, a study on the influence of pH indicated that RU-AgNPs are only stable under strong alkaline conditions (pH 12.5 or higher) (Fig. S4).Open in a separate windowFig. 1(a) UV-vis analysis of RU-AgNPs under strong alkaline condition (pH > 12.5) based on the order of adding reagents (RU, 0.10 to 0.50 mmol L−1; AgNO3, 0.20 mmol L−1; AA, 0.10 mmol L−1; NaOH, 0.10 mmol L−1); (b) TEM image of RU-AgNPs under the selected condition.The ability of RU-AgNPs to sense metal cations was investigated by both naked-eye and UV-vis spectroscopy analysis (Fig. 2). The separate addition of 10 μmol L−1 of several metal cations (Fe3+, Co2+, Zn2+, Sr2+, Cu2+. Al3+, Ba2+, Cd2+, Pb2+, Ni2+, Mg2+, Hg2+, Cu+ and Cr3+) to solutions of RU-AgNPs (prepared according to Fig. 1a) indicated that only Fe3+ induces a significant colorimetric change in the final aspect of solution after 50 minutes (Fig. 2a).Open in a separate windowFig. 2Naked-eye (a) and UV-vis (b) analysis of RU-AgNPs in absence (control, C) and presence of 10 μmol L−1 of selected metal cations (Fe3+, Co2+, Zn2+, Sr2+, Cu2+. Al3+, Ba2+, Cd2+, Pb2+, Ni2+, Mg2+, Hg2+, Cu+ and Cr3+) after 50 min; (c) calibration curve obtained by the addition of different amounts (0 to 10 μmol L−1) of Fe3+ to solutions of RU-AgNPs; (d) TEM image of RU-AgNPs after addition of Fe3+ (10 μmol L−1). In all experiments, RU-AgNPs were prepared in the presence of ascorbic acid.The results presented in Fig. 2a are consistent with the UV-vis spectroscopy analysis (Fig. 2b). Co2+ ions also induce some change in the system, however at a considerably smaller extension than Fe3+. In this study, AA played a crucial role in the selective detection of Fe3+ by the referred nanoprobe. In the absence of AA, Co2+ (mainly) as well as other cations induce stronger colorimetric and spectral changes (Fig. S5a and b, respectively) in the analysis of solutions of RU-AgNPs when compared to the system containing AA. This selectivity may arise from two possible reasons: (i) preservation of RU by avoiding its oxidation in the reduction of Ag+ ions; (ii) coordination of ascorbate anion to cations other than Fe3+.Interaction of RU-AgNPs and Fe3+ (10 μmol L−1) stabilizes after approximately 40 minutes (Fig. S6). Next, a calibration curve was built from the direct relationship between the absorbance at 396 nm and the concentration of Fe3+ (Fig. 2c), presenting a good correlation (R2 = 0.9929). The limits of detection and quantification were found to be 17 nmol L−1 and 56 nmol L−1, respectively, which is very satisfactory.13,14 The influence of other cations in the detection of Fe3+ was investigated by UV-vis spectroscopy. Fig. S7 clearly demonstrates that there are only small changes when a second cation (30 μmol L−1) is added together with Fe3+ to the RU-AgNPs solution.Mechanistically, the detection of Fe3+ by RU-AgNPs in aqueous medium proceeds via a growth/aggregation-combined process. This proposal is first evidenced by UV-vis analysis due the suppression of SPR band (Fig. 2b), a behavior consistent with the literature.31,32 Interestingly, TEM analysis clearly shows a growth in AgNPs size after addition of Fe3+ to the solution (Fig. 2d), resulting in a final single AgNPs population with average diameter of 14.7 nm ± 8.9 nm. Due to the strong alkaline medium, the main specie responsible for the behavior of NPs is likely to be Fe(OH)3. A schematic illustration of the mechanism involving aggregation of RU-AgNPs induced by the addition of Fe3+ is presented in Fig. 3. AgNPs are initially formed by adsorption of anionic RU to the silver surface via deprotonated 5-hydroxychromen-4-one moiety. This is supported by the literature33 and confirmed by alteration in the 1800–1500 cm−1 region of the RU infrared spectra before and after coordination with silver (Fig. S9). Afterwards, the addition of Fe3+ induces the formation of a coordination complex through an anionic catechol group, in which at least 2 : 1 ligand–Fe3+ stoichiometry is required for an aggregated effect.Open in a separate windowFig. 3Mechanistic proposal for growth/aggregation of RU-AgNPs in the presence of Fe3+. Insert: binding model for RU-AgNPs.Due to increasing interest in image processing as an analytical tool for many purposes,34,35 Multiple Linear Regression was employed to verify the capacity of the RU-AgNPs to probe Fe3+ standards at distinct concentrations, as presented in Fig. 4. The curve was obtained by plotting the color absorbances RGB-based values versus the concentrations of Fe3+ standards after RU-AgNPs interaction. Predicted iron is a vector based on RGB values that were then extracted from the filtered images and inserted in the equation described by Beer–Lambert law in order to generate the absorbances for the construction of the analytical curve. A linear behavior between the predicted response and the measured concentrations was observed (R2 = 0.9806).Open in a separate windowFig. 4Calibration curve for Fe3+ analysis showing the predicted iron (RGB) vs. iron(iii) concentration (1 to 8 μmol L−1). Adjusted R2 = 0.9806.Regression coefficients of the calibration model (using the R, G and B channels simultaneously) obtained by MLR method are shown in eqn (1):[Fe3+] = 44.5R − 4.9G − 16.4B + 5.11where Fe3+ concentration is the dependent variable (Predicted Iron), 5.1 is the intercept (β0), 44.5, −4.9 and −16.4 are the regression coefficients of the independent variables (R, G and B channels, respectively).It is possible to observe an interesting performance of the method using the three RGB color channels allied with MLR to quantify the Fe3+ content, which presented reasonable deviations in its responses considering that it is simple and low-cost. The good linearity is similar to other colorimetric methods such as sodium determination in seawater and coconut water (R2 > 0.91) by Moraes and coauthors,36 or iron(ii) in simulated seawater (R2 = 0.9993) by Gasparotto et al.37Second order regression was applied to the dataset to obtain a better adjustment, resulting in an adjusted R-squared of 0.9955. RGB values were then inserted in eqn (2) in order to generate the construction of the correlation:[Fe3+] = 14.4R + 30.7G − 23.4B − 398RG − 348RB − 0.8BG − 978R2 + 695G2 + 26.9B2 + 4.52This paper reports the use of AgNPs functionalized with RU as nanoprobes for selective detection and sensitive quantification of Fe3+ in aqueous solution. The synthesis of RU-AgNPs is reproducible, easily performed and requires no stabilizer agent other than RU. AA has a crucial role in the selectivity by either the avoidance of oxidation of RU by silver and/or coordination of ascorbate with other cations. The literature brings relevant examples of chromogenic and fluorogenic chemosensors for selective detection of Fe3+ in solution. Many of these artificial organic receptors present high selectivity and relevant limits of detection, requiring, however, very specific reagents and laborious synthetic procedures.38–41 Metal-based nanoparticles have emerged as potential probes for detection of Fe3+.11–16 Although effective in Fe3+ sensing, the synthesis of these nanoprobes require some toxic reagents, such as NaBH4 or PVP, use plant extracts, which may lead to some drawbacks, such as the understanding of the sensing mechanism. In contrast, our method is based on commercially available, nontoxic, low-cost reagents. Fe3+ sensing performed satisfactorily in the 1–10 μmol L−1 range, and the limit of detection obtained with this method (17 nmol L−1) is comparable to the most sensitive methods reported in literature. A mechanism for the detection of Fe3+ by RU-AgNPs involves a combined growth/aggregation of the NPs. There is a still limited number of nanoscaled systems reported as being selective and sensitive in the detection of Fe3+, which reinforces the relevance of the method reported herein. The linearity range obtained by both UV-vis spectroscopy and image analysis comprises the maximum of residual Fe3+ in drinking water according to the European and US legislations.15,42  相似文献   

17.
Jian Ma  Juan Wu  Bo Wu 《RSC advances》2020,10(19):11263
Correction for ‘Composition design, electrical properties, and temperature stability in (1 − x)K0.44Na0.56Nb0.96Sb0.04O3-xBi0.45La0.05Na0.5ZrO3 lead-free ceramics’ by Jian Ma et al., RSC Adv., 2018, 8, 29871–29878.

The authors regret that an incorrect version of Fig. 4 was included in the original article. The correct version of Fig. 4 is presented below.Open in a separate windowFig. 1FE-SEM surface images of (1 − x)K0.44Na0.56Nb0.96Sb0.04O3-xBi0.45La0.05Na0.5ZrO3 ceramics with (a) x = 0, (b) x = 0.020, (c) x = 0.040, (d) x = 0.060.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

18.
The exploration of highly efficient, stable and cheap water oxidation electrocatalysts using earth-abundant elements is still a great challenge. Herein, alkaline-stable cationic Ni(ii) coordination polymers (Ni-CPs) were successfully obtained under hydrothermal conditions, which could stabilize the incorporation of Fe(iii) to form Fe-immobilized Fe@Ni-CPs. The newly developed Ni-based CPs were used for the first time as an effective electrocatalyst for the oxygen evolution reaction in strong alkaline media.

An alkaline-stable cationic Ni(ii) coordination polymer showed remarkable oxygen evolution reaction (OER) catalytic activity due to capturing Fe ions.

The oxygen evolution reaction (OER) plays a vital role in energy storage and conversion applications due to energy issues and the need for sustainable development.1–4 Because of the sluggish kinetics of the OER, excellent electrocatalysts are required to work in acidic or strong alkaline environments.5,6 Noble metal-based catalysts like IrO2 and RuO2 with high efficiency showed excellent OER catalytic activity.7 However, noble metal with high-cost and scarcity are impractical for scale-up applications. Currently, to substitute these precious metal-based materials, transition-metal-based (Fe, Co, Ni and so on) and metal-free (e.g., N, P and S) hybrid materials have been extensively developed.8–10 For example, metal–organic frameworks (MOFs) such as ZIF-8/67 derived metal–carbon composite materials exhibit promising electrocatalytic performance.11,12 Unfortunately, the pyrolysis process destroys the framework completely and causes agglomeration of metals, resulting in a decreased number of active sites. Therefore, to explore highly efficient and low-cost OER catalysts that can be directly used in the OER without calcination are desired, including complexes and MOFs.Recently, much of transition bimetallic materials showed excellent electrocatalytic activity.13–15 However, a handful of examples such as Fe–Co-MOFs or Co–Ni-MOFs have been explored due to instability, poor conductivity and harsh synthesis conditions.16–20 Notably, the disadvantages of MOFs have limited their usage in the potential OER. Therefore, it is urgent to develop cheap, stable and active OER catalysts to replace the precious metals. However, this is still a great challenge. Coordination polymers (CPs) with a low-dimensional framework similar to MOFs are constructed by metal ion and organic ligands with potential active sites and functional groups, exhibiting wide applications in sensing, photoluminescence and photocatalysis.21–26 However, rare examples of CPs have been directly explored in the OER. For Fe/Ni-based bimetal electrocatalysts, a novel strategy involves doping Fe(iii) into a functional Ni-based CPs, which could enhance the electrocatalytic OER activities.Herein, we report the hydrothermal synthesis of a Ni-based CPs as a high-performance OER electrocatalyst in strong alkaline solutions (Scheme 1). The blue crystals of [Ni(bp)3·(H2O)2]·(bp)·(ClO4)2 (bp = 4,4′-biprydine) were obtained upon the reaction of bp ligands with NiClO4·6H2O under hydrothermal systems. Chemical stability tests displayed that Ni-CPs could retain their original framework in water or even a strong alkaline (pH = 14) solution after 12 hours (Fig. S1), which is rarely reported for most transition metal CPs. The thermogravimetric analysis (TGA) showed that there is a significant change at about 110 °C due to the weight loss of the partial guest (Fig. S2). Interestingly, the cationic Ni-CPs successfully captured the Mohr''s salt (ammonium iron(ii) sulfate) by taking advantage of the post-synthetic strategy. The obtained Fe@Ni-CPs exhibited a high-efficient OER activity under strong alkaline conditions.Open in a separate windowScheme 1Illustration of the synthesis process for Fe doped Ni coordination polymers for the OER.Single-crystal X-ray diffraction analysis revealed that the Ni-CPs crystallized in the C2/c space group (Table S1). The obtained Ni coordination polymer was the isostructural compound reported by Talham,27 but their packing modes were distinctly different (Fig. S3–S5). It also had a similar coordination environment to the railroad-like double chains synthesized by Yaghi.28 The parallel chains were occupied by 4,4′-bpy, perchlorate and water molecules. In the Ni-CPs, most of the phenyl rings adopted the face-to-face mode. There were evident π⋯π interactions between the adjacent 4,4′-biprydine (Fig. 1a). In addition, there were strong intermolecular hydrogen bonds between 4,4′-bpy and perchlorate (strong Cl–O⋯C amongst adjacent layers) (Fig. 1a). The weak reaction increased the high density of the framework and protected the coordination bonds against external guest attacking. These chains and guests were further packed with a three-dimensional structure along the c-axis (Fig. 1b).Open in a separate windowFig. 1(a) The weak reaction between chains in Ni CPs, showing the Cl–O⋯C (green and black dashed line) and π⋯π interactions (pink dashed line) between the adjacent perchlorate and 4,4′-biprydine, respectively; (b) the packing view of Ni CPs along the c-axis. Cl in green, Ni in pale blue, N in blue, O in red and C in black. H atoms and partial guest molecules were omitted for clarity.The unique cationic Ni-CPs framework has the potential to immobilize some counterpart ions. To demonstrate this, Mohr''s salts were investigated. Most strikingly, the color slowly changed from blue to green in an aqueous solution, given by the optical image (Fig. 2a and b), which not only indicated that Ni-CPs captured Mohr''s salts via ion-exchange, but also suggested an alteration in the valence of Fe ions. It was possible that Fe2+ may have been further oxidized to Fe3+ under the O2 and water environment when we prepared the Fe@Ni-CPs (4Fe2+ + 2H2O + O2 = 4Fe3+ + 4OH). The PXRD pattern showed that the frameworks remained unchanged after doping with Fe ions (Fig. S1). From the transmission electron microscopy (TEM) images (Fig. 2c), after immobilization, the morphology of the Fe@Ni-CPs was still level and smooth; no ring-like patterns arose corresponding to the selected area for electron diffraction (SAED) (Fig. 2d), indicating that no bulk Fe particles formed during ion-exchange. This was further demonstrated using high resolution TEM (HRTEM) (Fig. 2e), in which there was no lattice fringe of crystallized Fe. The well distribution of C, N, O, Cl, Fe, and Ni in Fe@Ni-CPs was demonstrated by elemental mapping (Fig. 2f–l). Energy-dispersive X-ray spectroscopy (EDX) also agreed well with the above mapping data (Fig. S6). In addition, the Fe3+ uptake was 4.1 wt%, as determined by inductively coupled plasma atomic emission spectroscopy (ICP). These results showed Fe ions to have been successfully immobilized by the Ni-CPs.Open in a separate windowFig. 2(a–b) The optical image of Ni-CPs and Fe@Ni-CPs; (c) TEM images of Fe@ Ni-CPs; (d–e) the corresponding SAED and HRTEM pattern. (f–l) Element mapping of C, N, O, Cl, Fe, and Ni in Fe@Ni-CPs.The X-ray photoelectron spectroscopy (XPS) survey spectrum of the Fe@Ni-CPs also showed the presence of C, N, O, Cl, Fe, and Ni elements (Fig. 3a). The Fe 2p high resolution XPS spectrum exhibited peaks at 725 eV and 711 eV (Fig. 3b), further indicating the presence of the Fe3+ oxidation state. This could be explained by the transformations of Fe2+ to Fe3+ during the ion-exchange process. Similarly, in the Ni 2p spectra (Fig. 3c), two main peaks located at 855.8 eV and 873.5 eV can be ascribed to Ni2+ 2p3/2 and Ni2+ 2p1/2, respectively. These peaks are associated with two shakeup satellite peaks, indicating that Ni still remained in a divalent state. The Cl 2p and N 1s region could be corresponded to the ClO4 and biprydine, respectively (Fig. S7). The O 1s spectrum (Fig. 3d) was divided into two peaks at 531.7 eV and 533.2 eV, which could be assigned to the OH group from filled H2O molecules and partial ClO4, respectively.Open in a separate windowFig. 3(a) XPS survey spectrum of the Ni-CPs; XPS spectra of the Ni-CPs in the (b) Fe 2p, (c) Ni 2p, and (d) O 1s regions.The above Ni-CPs with Fe doping encouraged us to investigate its electrocatalytic application in oxygen evolution reaction. To study the electrocatalytic activity of Fe@Ni-CPs for the OER, linear sweep voltammetry (LSV) was performed in a strong alkaline solution (pH = 14) for Fe@Ni-CPs@GC (fresh samples coated on glassy carbon electrode with Nafion binder). Fe@Ni-CPs@GC directly acted as working electrodes and showed good OER activity with an onset potential of 1.52 V (Fig. 4a), overpotential of 368 mV at 10 mA cm−2, and a Tafel slope of 59.3 mV dec−1 (Fig. 4b). These OER performances are close to some reported MOFs catalysts (Table S2) and even better than commercial benchmark OER catalysts like RuO2 working at the same condition (Fig. S8). In contrast, the electrocatalytic OER activities of the pristine Ni-CPs@GC without Fe incorporation displayed much worse activity. The onset potential, overpotential (at 10 mA cm−2), and the Tafel slope reached 1.62 V, 458 mV, and 96.8 mV dec−1, respectively (Fig. 4a and b). In addition, Fe@Ni-CPs showed a strong durability during the OER process. The chronoamperometric response of Fe@Ni-CPs displayed a slight anodic current attenuation within 12 h due to the peeling of samples during the evolution of a large amount of O2 gas (Fig. S9). Furthermore, LSV of Fe@Ni-CPs showed negligible changes after OER tests for 12 h (Fig. S10). These results indicate that the Fe-doped Ni-CPs with more active sites could serve as an excellent candidate for OER in strong alkaline conditions.Open in a separate windowFig. 4(a) OER polarization curves and (b) Tafel plots of various electrocatalysts in a 1 M KOH aqueous solution; (c) linear relationship of the current density at 1.1 V (vs. RHE) vs. scan rates for Fe@Ni-CPs@GC and Ni-CPs@GC; (d) EIS of Fe@Ni-CPs and Ni-CPs electrode.When Fe(iii) was introduced, the resulting Fe@Ni-CPs catalysts greatly improved OER catalytic performance. There were dynamic collisions between Fe3+ ions and Ni-CPs, which allowed for more accessible catalytic active sites compared to the Ni-CPs. Particularly, Fe(iii) doping can contribute to the adsorption and reaction of OH groups in OER process.29 As a result, Fe@Ni-CPs enhanced charge transfer under an apt electronic environment of the mixed Fe⋯Ni systems. In addition, electrochemical impedance spectrum and double-layer capacitance (Cdl) of the Fe@Ni-CP were also studied. The Cdl of Fe@Ni-CPs was confirmed to be 269.7 μF cm−2 (Fig. 4c and S11), which is higher than that of Ni-CPs (Cdl = 174.2 μF cm−2) (Fig. 4c and S12). The semicircular diameter in EIS of Fe@Ni-PCP was smaller than that of Ni-CPS (Fig. 4d). These results further showed that Fe@Ni-CPs were more effective in enlarging the catalytically active surface area, conductivity and synergistic effects between Fe and Ni in comparison to Ni-CPs coated on electrodes.In conclusion, a new alkaline-stable cationic Ni(ii) coordinated polymers was synthesized under hydrothermal conditions. The Ni CPs could quickly interact with Mohr''s salt. Interestingly, the Ni CPs could act as a unique oxidation matrix to realize the transformation of Fe2+ to Fe3+ during the ion-exchange process. Furthermore, the resulting Fe@Ni-CPs electrode, for the first time, showed an excellent electrocatalytic activity for OER in strong alkaline media. This study provides a new avenue to explore stable coordinated polymers by incorporating the low-cost and high-activity transition metal, Fe, which will substitute the rare noble metals used in energy-related research.  相似文献   

19.
Correction for ‘Highly porous core–shell chitosan beads with superb immobilization efficiency for Lactobacillus reuteri 121 inulosucrase and production of inulin-type fructooligosaccharides’ by Thanapon Charoenwongpaiboon et al., RSC Adv., 2018, 8, 17008–17016.

The authors regret that Fig. 9 in the original article was displayed incorrectly. The correct version is shown below.Open in a separate windowFig. 9Batch reusability of INU-CSBs for IFOS synthesis. Reaction condition: 10 U mL−1 of biocatalysts were incubated with 200 g L−1 sucrose in acetate buffer pH 5.5, 40 °C and 2 h per batch.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

20.
Novel Fe3C nanoparticles encapsulated with nitrogen-doped graphitic shells were synthesized by floating catalytic pyrolysis. Due to the short synthesis time and controllable pyrolytic temperature, the diameters of Fe3C core nanoparticles ranged from 5 to 15 nm (Fe3C@NGS900 prepared at 900 °C) and the average thickness of N-doped graphitic shells was ∼1.2 nm, leading to their high electrochemical performance: specific capacity of 1300 mA h g−1 at current density 0.2 A g−1, outstanding rate capability of 939 mA h g−1 at 3 A g−1, improved initial coulombic efficiency (Fe3C@NGS900: 72.1% vs. NGS900 (pure graphitic shells): 52%) for lithium ion batteries (LIBs), and impressive long-term cycle performance (1399 mA h g−1 maintained at 3 A g−1 after 500 cycles for LIBs; 214 mA h g−1 maintained at 1 A g−1 after 500 cycles for sodium ion batteries).

Novel Fe3C nanoparticles encapsulated with nitrogen-doped graphitic shells were synthesized by floating catalytic pyrolysis.

Because of the fast development of portable electronic devices and hybrid electric vehicles, lithium ion batteries (LIBs)1–8 and sodium ion batteries (NIBs)9–16 with high energy/power density, good cycling performance, and lack of memory effects are in ever-increasing need. Due to the low theoretical capacity of carbon materials (graphite: 372 mA h g−1),8,17–22 optimizing the morphology of graphitic electrode materials has been important to improve specific capacity.1–7,23,24 On the other hand, chemical doping (e.g., N, S, B, P) is an effective strategy to raise their specific capacity by increasing conductivity or active sites for Li+ or Na+ storage.3,4,25–28 Additionally, metallic compounds (e.g., Fe3C) have been also introduced into improving electrochemical performance of carbon anodes, because such materials are proposed to activate some components for reversible transformation of the solid electrolyte interface (SEI) and further benefit reversible capacity.13,16,29,30 Unfortunately, most of them have been prepared by complex methods including tedious synthetic steps or long-time annealing,6,17–19 from which, it is hard to prepare Fe3C particles with desirable small sizes.8,13,15,16 Thus, developing appropriate Fe3C/C electrode materials still requires further research.In this work, Fe3C nanoparticles encapsulated with nitrogen-doped graphitic shells (Fe3C@NGS) were in situ approached from floating catalytic pyrolysis. Due to the short annealing time of the pyrolysis, Fe3C@NGSs was prepared with controllable sizes. Furthermore, the in situ approach led the graphitic shells just grew on the surface of Fe3C core nanoparticles, which improved electron transfer between the cores and the shells. Thus, such nanoparticles might be a superb electrode material towards high performance applications of LIBs and NIBs.For preparing the Fe3C@NGSs with controllable sizes, floating catalytic pyrolysis was carried out to shorten synthetic time: the gas mixture was introduced into quartz pipe furnace, which was set at 700–1100 °C. For a typical experiment, nitrogen (flow rate: 80 L h−1), acetylene (10 mL min−1) and ammonia (100 mL min−1) gases were embedded into iron pentacarbonyl held at 10 °C to form the gas mixture. After the pyrolysis, Fe3C@NGS was collected at the other end of the quartz pipe. The details of materials characterization and electrochemical measurements can be found in ESI.As shown in transmission electron microscope (TEM) images of the prepared Fe3C@NGSs (Fig. 1a–c and S1a of ESI), the metallic cores (dark section) are encapsulated with their shell (light section). The XRD results (Fig. 1d) shows the metallic core nanoparticles are Fe3C. Thus, the sample prepared at 900 and 1100 °C have been marked with Fe3C@NGS900 and Fe3C@NGS1100, respectively. However, the sample prepared at 700 °C, which has been marked with FN@NGS700, has been oxidized in air at room temperature, because of its poor graphitic layers. Compared with XRD pattern of Fe3C@NGS900, those peaks of Fe3C@NGS1100 are much sharper indicating much bigger ferrous cores of Fe3C@NGS1100. According to the TEM and high resolution TEM (HRTEM) images for Fe3C@NGS900 (Fig. 1a–c, S1a, and b), the diameter of the core nanoparticles is ranged from 5–15 nm and the average thickness of their shells is ∼1.2 nm, respectively. Moreover, the spacing of the lattice fringes (Fig. S1a and b) is ∼0.34 nm corresponding to the characteristic (002) peak of graphite implying high graphitization of those shells.25,28,31 In the HRTEM images, every Fe3C cores is found to be a single crystal and encapsulated with the graphitic shell. The boundary between Fe3C cores and N-doped graphitic shells is continuous and distinguished, indicating Fe3C cores are tightly encapsulated with the graphitic shell.Open in a separate windowFig. 1(a) TEM images of core-shells nanoparticles (FN@NGS700) prepared at 700 °C, TEM and HRTEM (inset) images of nanoparticles (Fe3C@NGS900) prepared at 900 °C (b) and nanoparticles (Fe3C@NGS1100) prepared at 1100 °C (c) and (d) XRD patterns of core-shells nanoparticles prepared at different temperatures.Fe3C@NGS samples have been also analyzed by X-ray photoelectron spectroscopy (XPS), which suggests Fe3C@NGS samples contain Fe, C, O and N atoms (FN@NGS700: C content of 16.3 wt%, Fe content of 54.6 wt%, N content of 1.6 wt% and O content of 27.5 wt%; Fe3C@NGS900: C content of 26 wt%, Fe content of 67 wt%, N content of 2 wt% and O content of 5 wt%; Fe3C@NGS1100: C content of 33.7 wt%, Fe content of 52 wt%, N content of 2.3 wt% and O content of 7 wt%). According to XPS results of Fe3C@NGS900, the weight ratio of Fe3C cores and graphitic shells is 3.39 : 1. Additionally, the element contents of Fe3C@NGS samples are different due to their morphology and structure (Fig. 1). Higher O content of FN@NGS700 is because its ferrous cores have been oxidized, and thick-walled graphitic shells (∼3 nm) of Fe3C@NGS1100 leads to its higher C content, as XPS can only measure element contents of surface of samples.The electrochemical properties of the Fe3C@NGS-based electrodes have been shown in Fig. 2 and and3.3. Since FN@NGS700 sample has been completely oxidized, electrochemical properties of ferrous oxide has not been measured. The cyclic voltammetry (CV) curves of Fe3C@NGS900-based electrode show details of possible lithium storage process. Besides the similar peak (0.5 V) in the initial cycle for the SEI formation, two reduction peaks are located at 0.7 and ∼1.5 V, corresponding to the reduction of some SEI components (Li+ insertion). During the Li+ extraction process, the corresponding oxidation peaks is found to shift from ∼1.7 to ∼1.9 V. The almost overlapped oxidation peaks demonstrate good reversibility and cycling stability of core–shell Fe3C@NGS.15 As shown in Fig. 2b, the galvanostatic discharge–charge (GDC) profiles of Fe3C@NGS900 in a voltage range of 0.005–3 V (vs. Li+/Li) exhibits the typical shape of Fe3C@NGS-based anodes, and Fe3C@NGS900 delivers initial charge and discharge capacities of 1246.7 and 1729.1 mA h g−1, respectively. The initial columbic efficiency (CE) reaches up to 72.1% which is higher than 52% of pure graphitic shells.3 For Fe3C@NGSs electrode, the phenomenon of capacity increment is related to growing reversibly SEI film via the decomposition of electrolyte due to the catalysis of Fe3C.16,29 From the second cycle, the shape of the discharge profiles changes with respect to that of the first cycle, which may be due to the modification of the SEI film.12Open in a separate windowFig. 2Electrochemical performance of prepared nanoparticles as anodes for LIBs: CV profile (a) of Fe3C@NGS900 at a scan rate of 0.1 mV s−1 between 0.01–3 V vs. Li+/Li for the 1st–2nd charge/discharge cycles; (b) the galvanostatic charge–discharge profiles of Fe3C@NGS900; (c) charge–discharge cycling performance of Fe3C@NGS core–shell nanoparticles at different current densities from 0.2 to 3 A g−1 at room temperature; (d) cycling performance of Fe3C@NGS at 3 A g−1.Open in a separate windowFig. 3Electrochemical performance of the prepared nanoparticles as anodes for NIBs: (a) CV profile of Fe3C@NGS900 at a scan rate of 0.1 mV s−1 between 0.01–2.5 V vs. Na+/Na for the 1st–3rd charge/discharge cycles. (b) The galvanostatic charge–discharge profiles of Fe3C@NGS900; (c) cycling performance of Fe3C@NGS at a current density of 1 A g−1, and the corresponding columbic efficiency.The excellent rate capability of Fe3C@NGS-based anodes have been investigated by testing charge/discharge at current densities of 0.2, 0.5, 1 and 3 A g−1 for every 5 cycles (Fig. 2c). At the corresponding rates, the reversible capacities are 1300, 1101, 1062 and 939 mA h g−1 for Fe3C@NGS900; 925, 721, 690 and 663 mA h g−1 for Fe3C@NGS1100, indicating the smaller size of the Fe3C nanoparticles might enhance their electrochemical capability. Compared with reported pure graphitic shells (NGS900 prepared by removing Fe3C cores of Fe3C@NGS900),3 the core–shell nanoparticles (Fe3C@NGS900) exhibits impressive rate performances (NGS900: 760 mA h g−1 at 0.5 A g−1, 620 mA h g−1 at 1 A g−1 and 340 mA h g−1 at 5 A g−1), which might be caused by Fe3C, as a good conductor of electricity, can effectively improve electrical conductivity of carbon electrode material.13 Calculated from eqn (1) of ESI, specific capacity of Fe3C cores of Fe3C@NGS900 can be evaluated (1199 mA h g−1 at 0.5 A g−1 and 1193 mA h g−1 at 1 A g−1, respectively). Based on the conversion mechanism for lithium storage, if possible, Fe3C can store only 1/6 Li per unit (∼26 mA h g−1),15 which is negligible regarding to the high capacity of ∼1300 mA h g−1. The specific capacity of Fe3C is larger than what it should be, which might be resulted from the pseudocapacity on the interface between the material and the electrolyte.16 For evaluating N doping structure in the graphitic shells, Fe3C@NGS samples have been prepared with different percent of doping content at 900 °C by introducing ammonia with different flow rates (0, 30, 100 or 500 mL min−1). As a result, Fe3C@GS900 prepared without ammonia has graphitic shells without N-doping leading to its poor electrochemical performance: at current densities of 0.2, 0.5, 1 and 3 A g−1, its reversible capacities are 575, 492, 458 and 402 mA h g−1 (Fig. S4); the performances of Fe3C@NGS900 (ammonia flow rate: 100 mL min−1; content of N: 2 wt%) and Fe3C@NGS900A (ammonia flow rate: 30 mL min−1; content of N: 1.5 wt%) are similar; the sample prepared with ammonia flow rates of 500 mL min−1 (FN@NGS900B) was violent oxidized to ferrous oxide in the air. Hence, N doping structure in graphitic shells has been confirmed to enhance diffusion.The long-term cycling performance of these two electrodes also has been investigated in Fig. 2d. The Fe3C@NGS-based anode exhibits a favorable reversible capacity, which can reach 1399 mA h g−1 after 500 discharge/charge cycles at 3.0 A g−1, showing high capacity retention with CE of ∼100%. The capacities of Fe3C samples increases with cycle number rising (from 120 to 450), which might be attributed to the pseducapacity presented by the Fe3C.16To better study the kinetic properties of Fe3C@NGS900 and Fe3C@NGS1100, Fig. S3 of ESI shows the Nyquist plots and equivalent circuit obtained from electrochemical impedance spectroscopy (EIS) measurements. Here, Rs represents the ohmic resistance of the battery. Constant phase element (CPE) represents the double layer capacitive reactance between the electrode materials and the electrolyte. The semicircles and straight lines correspond to the electrochemical polarization impedance (Rp) and Warburg resistance (W), respectively.32 Both the fitted Rs value (5.526 Ω) and Rp value (20.826 Ω) for Fe3C@NGS900 electrode is much lower than that Fe3C@NGS1100 electrode (Rs: 8.501 Ω; Rp: 43.665 Ω), indicating the superior redox kinetics in the Fe3C@NGS900 composite.In order to study the electrochemical properties of the Fe3C@NGSs electrode as anodes for NIBs, CV analysis has been carried out at a scanning rate of 0.1 mV s−1 between 0.005 and 2.5 V vs. Na+/Na. As shown in Fig. 3a, there are two irreversible reduction peaks around 2.02 V and 0.83 V found during the initial cathodic scan, which could be ascribed to the interaction of Na ions with specific functional groups and the decomposition of electrolyte along with the formation of SEI film on the electrode surfaces.33,34 For the subsequent cycles, the peak at 2.02 V disappears and the peak at 0.83 V shifts to 0.60 V. For the anodic scan, the main oxidation peak ranging from about 0.47 V to 1.28 V is supposed to be the desodiation reactions.35Fig. 3b depicts the GDC curves of the Fe3C@NGS900-based electrode for the 1st–5th cycle at current density of 0.1 A g−1. The large irreversible capacity in the 1st cycle is attributed to the SEI formation and the irreversible insertion of sodium ion with a relatively large ionic radius.22,23 Following the first cycle, the charge–discharge curves become more linear which exhibits a higher and more stable CE indicating a stable SEI layer formed in the first cycle.Meanwhile, cycling performance of Fe3C@NGS-based anodes for NIBs have been shown in Fig. 3c. In the extended cycling test at 1 A g−1, a reversible capacity 214 mA h g−1 of Fe3C@NGS900 electrode is still maintained after 500 cycles, which is ∼2 times the capacity delivered by the Fe3C@NGS1100 indicating excellent cycling stability of Fe3C@NGS900. Thus, Fe3C@NGS-base anodes holds great potential as a promising candidate compared with other carbonaceous anode materials for NIBs (113 mA h g−1 at 1 A g−1 (modified PFR/C),36 188.6 mA h g−1 at 0.1 A g−1 after 300 cycles (S/C),33 150 mA h g−1 at 1 A g−1 after 200 cycles (S/graphene)35).It is noticed that the excellent electrochemical performance of the prepared Fe3C@NGSs is apparently ascribed to their novel structure. First, because of in situ growing graphitic shells on the surface of Fe3C during floating catalytic pyrolysis, contact between Fe3C cores and graphitic shells effectively increases, leading to lower contact resistance and faster electron transfer between the cores and the shells, comparing with traditional ferrous/carbon composites generated by multistep carbonization approach.8,13,15,16 Second, due to floating catalytic pyrolysis, sizes of Fe3C core nanoparticles have been under control (Fe3C@NGS900: 5–15 nm vs. Fe3C@NGS1100: 15–40 nm). Smaller size of Fe3C core nanoparticles might increase surface of Fe3C nanoparticles, leading to distinguished improvement of their electrochemical performance (Fig. 2 and and3),3), due to active sites for Li+ or Na+ storage rising. For comparison, sizes of reported Fe3C composites have been listed: Fe3C@C: 60 nm nanoparticles encapsulated with 4 nm carbon shells;8 Fe3C@PC: 29 ± 5 nm nanoparticles embedded in 300 nm porous carbon;16 Fe3C/C: 300 nm;13 Fe@Fe3C/C: 28–58 nm Fe@Fe3C nanoparticles.15 Third, introducing Fe3C to electrode material can promote the reversible formation/decomposition of the SEI film, causing improvement of initial CE (72.1% vs. 52%) for LIBs, due to the catalysis function of Fe3C.16,29 Fourth, due to in situ N-doping (∼2 wt%) during their floating catalytic pyrolysis, such defects of the graphitic shells might offer lots channels for fast diffusion of electrolyte and Li+/Na+ into those nanoparticles. Fifth, such prepared core–shell nanoparticles have ultra thin-walled graphitic shells (∼1.2 nm shown in Fig. 1d), which shorten the diffusion route of ions and electrolyte. All above confirm Fe3C@NGS900 has novel structure towards the electrochemical applications, compared with the reported works: Fe@Fe3C/C sample was prepared by sol–gel and carbonization approach, and whether its pure Fe cores was good for Li+ storage was doubtful;15 the Fe3C@C nanoparticles were prepared with no doping carbon shells;8 the graphitization of carbon structure was doubtful, when ferrous/carbon composites were prepared by polymerization-carbonization of iron phthalocyanine13 or hydrothermal method-carbonization.16 Thus, the Fe3C@NGS900 performs better (1300 mA h g−1 at current density of 0.2 A g−1; 939 mA h g−1 at 3 A g−1) than many reported ferrous/carbon composite anode materials for LIBs (0.2 A g−1: <382 (Fe@Fe3C/C),15 ∼480 (Fe3C/C),8 787.9 (Fe2O3/C),5 ∼850 (Fe3O4/C),19 873 (N,S/C),36 and 881 (Fe3O4/C)9 mA h g−1; 3 A g−1: ∼300 (Fe2O3@C),18 ∼370 (FeS@C)21 and ∼612 (Fe2O3/C)14 mA h g−1).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号