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 515 毫秒
1.
Correction for ‘Ascorbic acid/Fe0 composites as an effective persulfate activator for improving the degradation of rhodamine B’ by Xiangyu Wang et al., RSC Adv., 2018, 8, 12791–12798.

The authors regret that the unit on the x-axis of Fig. 1 was incorrectly written as “% wt” rather than “‰ wt” in the original article. The correct version of Fig. 1 is presented below.Open in a separate windowFig. 1(a) Comparison of removal efficiency of RhB in different systems (C0 = 50 mg L−1, PS dosage = 1.4 g L−1, Fe0 dosage = 1 g L−1, H2A/Fe0 dosage = 1 g L−1, H2A dosage = 1.6 g L−1 and T = 298 K); (b) effect of H2A concentration on removal efficiency of RhB in the H2A/Fe0–PS system (C0 = 50 mg L−1, Fe0 dosage = 0.8 g L−1, T = 298 K and the solution volume is 50 mL).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

2.
A screen-printed electrode (SPGPUE) was prepared with graphite–polyurethane composite ink containing gold nanoparticles (AuNPs), resulting in a screen-printed graphite–polyurethane composite electrode modified with gold nanoparticles (SPGPUE–AuNPs). Gold nanoparticles were prepared by the citrate method and extracted from the water medium since polyurethane is not compatible with humidity. After extraction to chloroform, they were characterized via transmission electron microscopy (TEM). The presence of gold on the SPGPUE–AuNP surface was confirmed via SEM and EDX analyses, while thermogravimetry revealed the presence of approximately 3.0% (m/m) gold in the composite. An electrochemical pretreatment in 0.10 mol L−1 phosphate buffer (pH 7.0) with successive cycling between −1.0 V and 1.0 V (vs. pseudo-Ag/AgCl) under a scan rate of 200 mV s−1 and 150 cycles was required in order to provide a suitable electrochemical response for the voltammetric determination of dopamine. After the optimization of the parameters of differential pulse voltammetry (DPV), an analytical curve was obtained within a linear dynamic range of 0.40–60.0 μmol L−1 and detection limit (LOD) of 1.55 ×10−8 mol L−1 for dopamine at the SPGPUE–AuNP. A non-modified SPGPUE was used for comparison and a linear range was obtained between 2.0 and 10 μmol L−1 with an LOD of 2.94 × 10−7 mol L−1. During the dopamine determination in cerebrospinal synthetic fluid (CSF), recoveries between 89.3 and 103% were achieved. There were no significant interferences from ascorbic acid and uric acid, but some from epinephrine due to the structural similarity.

A screen-printed modified composite electrode (SPGPUE) was prepared with graphite–polyurethane ink containing gold nanoparticles (AuNPs), resulting in a sensor with improved sensitivity regarding the unmodified device in dopamine determination.  相似文献   

3.
Two types of magnetic microspheres (Fe3O4@MIL-100 and Fe3O4@SiO2@polythiophene) were prepared and characterized as mixed sorbents for magnetic solid-phase extraction (MSPE) of six phthalic acid esters (PAEs), including dimethyl phthalate (DMP), diethyl phthalate (DEP), di-n-butyl phthalate (DBP), benzyl butyl phthalate (BBP), di-2-ethylhexyl phthalate (DEHP), and di-n-octyl phthalate (DnOP) from water samples prior to gas chromatography-mass spectrometry (GC-MS) analysis. The synthetic magnetic nanocomposites exhibited good repeatability and chemical stability, and improved extraction efficiency for the tested PAEs. The mixture of the two types of nanoparticles substantially improved the extraction efficiency of both DMP and DEP. The key parameters affecting the extraction efficiency, such as the type and the amount of sorbent, eluent (desorption solvent), adsorption and desorption time, pH of sample solution, and sample volume, were investigated and optimized, respectively. Under optimized conditions, the developed method showed satisfactory linearity in the range of 5–5000 μg L−1 with coefficients of determination (R2) > 0.9935. The method detection limits (MDLs) and limits of quantitation (LOQs) were between 0.35–0.91 μg L−1 and 1.1–2.9 μg L−1, respectively. At three fortification levels (1.0, 10.0, and 50.0 μg L−1), the mean recoveries ranged from 76.9–109.1% with favorable relative standard deviations (RSDs) < 9%. The feasibility of the method was evaluated by analysis of water samples from various sources (tap, drinking, and mineral water). The results show that the developed method is suitable for determination of trace level PAEs in water samples.

A mixture of Fe3O4@MIL-100 and Fe3O4@SiO2@polythiophene nanoparticles exhibit high extraction efficiency for PAEs in water.  相似文献   

4.
In the present study, a novel resource utilization method using wet magnesia flue gas desulfurization (FGD) residue for the simultaneous removal of ammonium nitrogen (NH4–N) and heavy metal pollutants from vanadium (V) industrial wastewater was proven to be viable and effective. In this process, the wet magnesia FGD residue could not only act as a reductant of hexavalent chromium [Cr(vi)] and pentavalent vanadium [V(v)], but also offered plenty of low cost magnesium ions to remove NH4–N using struvite crystallization. The optimum experimental conditions for Cr(vi) and V(v) reduction are as follows: the reduction pH = 2.5, the wet magnesia FGD residue dose is 42.5 g L−1, t = 15.0 min. The optimum experimental conditions for NH4–N and heavy metal pollutants removal are as follows: the precipitate pH = 9.5, the n(Mg2+) : n(NH4+) : n(PO43−) = 0.3 : 1.0 : 1.0, t = 20.0 min. Finally the NH4–N, V and Cr were separated from the vanadium containing industrial wastewater by forming the difficult to obtain, soluble coprecipitate containing struvite and heavy metal hydroxides. The residual pollutant concentrations in the wastewater were as follows: Cr(vi) was 0.047 mg L−1, total Cr was 0.1 mg L−1, V was 0.14 mg L−1, NH4–N was 176.2 mg L−1 (removal efficiency was about 94.5%) and phosphorus was 14.7 mg L−1.

A novel resource utilization method using wet magnesia flue gas desulfurization residue for the simultaneous removal of ammonium nitrogen and heavy metal pollutants from vanadium industrial wastewater was proven to be viable and effective.  相似文献   

5.
Four additives (Na2WO4, nano-hydroxyapatite, K2TiF6 and NaF) were added into the Na5P3O10 + NaOH + C3H8O3 base electrolyte according to the orthogonal design of four factors three levels (L9 (34)). Nine different micro-arc oxidation (MAO) coatings were fabricated on Mg–2Zn–0.5Ca alloys through orthogonal experiments. The effects of four additives on the microstructure, mechanical properties, corrosion resistance and biocompatibility of MAO coatings were investigated through X-ray diffraction (XRD), scanning electron microscopy (SEM) with energy dispersive X-ray spectroscopy (EDS), electrochemical corrosion test and in vitro degradation test. The addition of nano-hydroxyapatite and K2TiF6 showed self-sealing effects and contributed to the corrosion resistance of the samples significantly. The addition of 0.5 g L−1 Na2WO4 markedly elevated the bonding strength of the coatings with the substrate. The optimal combination of factors and levels considering both mechanical properties and corrosion resistance was: 0.5 g L−1 Na2WO4, 0 g L−1 NaF, 5 g L−1 n-HAp, 5 g L−1 K2TiF6. The growth mechanism of MAO coatings combining with the visual phenomenon was discussed as well.

Large amount of micro-pores formed in MAO coatings were interconnected and sealed.  相似文献   

6.
Bi0.4Sb1.6Te3 (BST) is known to be a unique p-type commercial thermoelectric (TE) alloy used at room temperatures, but its figure of merit (ZT) is relatively low for wide industrial applications. To improve its ZT value is vitally important. Here, we show that the incorporation of 0.5 wt% PbTe nanoparticles into BST concurrently causes a large enhancement of power factor (PF) and a significant reduction of lattice thermal conductivity κL. The increase in PF mainly benefits from the optimization of carrier concentration, maintenance of high carrier mobility and constant rise in Seebeck coefficient. The decrease in κL can be attributed to the enhanced phonon scattering by the dispersed PbTe nanoparticles and the interfaces between PbTe and the BST matrix by using the Callaway model. Specifically, an ultralow κL of 0.26 W m−1 K−1 at 429 K is achieved for the composites incorporating 0.5 wt% PbTe nanoinclusions. Consequently, an excellent ZT = 1.6 at 482 K and a high average ZTave = 1.38 at 300–500 K are achieved, indicating that incorporation of PbTe in BST is an effective approach to improve its thermoelectric performance.

We introduce 0.5 wt% PbTe nanoparticles into the Bi0.4Sb1.6Te3 matrix and possess an ultralow lattice thermal conductivity of 0.26 W m−1 K−1 at 429 K and an excellent ZT value of 1.6 at 482 K as well as a high average ZTave of 1.38 at 300–500 K.  相似文献   

7.
The use of efficient green cleaning agents, such as biosurfactants, is important in oil sludge treatment. Enhanced oil recovery from oily sludge by different rhamnolipids was comparatively evaluated. Using crude glycerol, the wild-type strain Pseudomonas aeruginosa SG and the recombinant strains P. aeruginosa PrhlAB and P. stutzeri Rhl produced 1.98 g L−1, 2.87 g L−1 and 0.87 g L−1 of rhamnolipids, respectively. The three bacterial strains produced different rhamnolipid mixtures under the same conditions. The proportions of mono-rhamnolipids in the three rhamnolipid products were 55.92%, 94.92% and 100%, respectively. These rhamnolipid products also possessed different bioactivities. Emulsifying activity became higher as the proportion of mono-rhamnolipids increased. The three rhamnolipid products were stable at temperatures lower than 121 °C, pH values from 5–11 and NaCl concentrations from 0–15%. All three rhamnolipid products could recover oil from oily sludge, but oil recovery efficiency was positively related to the proportion of mono-rhamnolipids. Mono-rhamnolipids produced by the recombinant strain Rhl exhibited the best oil recovery efficiency (53.81%). The results reveal that mono-rhamnolipids are the most promising for oil recovery from oily sludge.

Oil recovery from oily sludge is positively related to the proportion of mono-rhamnolipids.  相似文献   

8.
A novel visible-light-driven Z-scheme heterojunction, Bi2WO6/Ag2S/ZnS, was synthesized and its photocatalytic activity was evaluated for the treatment of a binary mixture of dyes, and its physicochemical properties were characterized using FT-IR, XRD, DRS and FE-SEM techniques. The Bi2WO6/Ag2S/ZnS Z-scheme heterojunctions not only facilitate the charge separation and transfer, but also maintain the redox ability of their components. The superior photocatalytic activity demonstrated by the Z-scheme Bi2WO6/Ag2S/ZnS attributes its unique properties such as the rapid generation of electron–hole pairs, slow recombination rate, and narrow bandgap. The performance of the Bi2WO6/Ag2S/ZnS was evaluated for the simultaneous degradation of methyl green (MG) and auramine-O (AO) dyes, while the influences of the initial MG concentration (4–12 mg L−1), initial AO concentration (2–6 mg L−1), pH (3–9), irradiation time (60–120 min) and photocatalyst dosage (0.008–0.016 g L−1) were investigated through the response surface methodology. The desirability function approach was applied to optimize the process and results revealed that maximum photocatalytic degradation efficiency was obtained at optimum conditions including 6.08 mg L−1 of initial MG concentration, 4.04 mg L−1 of initial AO concentration, 7.25 of pH, 90.58 min of irradiation time and 0.013 g L−1 of photocatalyst dosage. In addition, a possible photocatalytic mechanism of the Bi2WO6/Ag2S/ZnS heterojunction was proposed based on the photoinduced charge carriers.

A Z-scheme Bi2WO6/Ag2S/ZnS heterojunction was successfully synthesized as a novel visible-light-driven photocatalyst for the degradation of multiple dye pollutants.  相似文献   

9.
In this work, we report the facile hydrothermal synthesis of manganese cobaltite nanoparticles (MnCo2O4.5 NPs) which can efficiently activate peroxymonosulfate (PMS) for the generation of sulfate free radicals (SO4˙) and degradation of organic dyes. The synthesized MnCo2O4.5 NPs have a polyhedral morphology with cubic spinel structure, homogeneously distributed Mn, Co, and O elements, and an average size less than 50 nm. As demonstrated, MnCo2O4.5 NPs showed the highest catalytic activity among all tested catalysts (MnO2, CoO) and outperformed other spinel-based catalysts for Methylene Blue (MB) degradation. The MB degradation efficiency reached 100% after 25 min of reaction under initial conditions of 500 mg L−1 Oxone, 20 mg L−1 MnCo2O4.5, 20 mg L−1 MB, unadjusted pH, and T = 25 °C. MnCo2O4.5 NPs showed a great catalytic activity in a wide pH range (3.5–11), catalyst dose (10–60 mg L−1), Oxone concentration (300–1500 mg L−1), MB concentration (5–40 mg L−1), and temperature (25–55 °C). HCO3, CO32− and particularly Cl coexisting anions were found to inhibit the catalytic activity of MnCo2O4.5 NPs. Radical quenching experiments revealed that sulfate radicals are primarily responsible for MB degradation. A reaction sequence for the catalytic activation of PMS was proposed. The as-prepared MnCo2O4.5 NPs could be reused for at least three consecutive cycles with small deterioration in their performance due to low metal leaching. This study suggests a facile route for synthesizing MnCo2O4.5 NPs with high catalytic activity for PMS activation and efficient degradation of organic dyes.

Catalytic degradation of organic dyes via manganese cobaltite nanoparticles-activated peroxymonosulfate.  相似文献   

10.
A facile and green method was adopted to synthesize highly selective gum acacia-mediated silver nanoparticles as dual sensor (fluorescence turn-on and colorimetric) for Hg(ii) and fluorescence turn-off sensor for S2− and malachite green. The mechanism proposed for a dual response towards Hg(ii) is the redox reaction between Ag(0) and Hg(ii), resulting in the formation of Ag(i) and Hg(0) and electron transfer from gum acacia to Ag(i), which further leads to the formation of an Ag@Hg nanoalloy. The enhanced fluorescence signal was quenched selectively by S2− owing to the formation of Ag2S and HgS. The reported nanosensor was found to be useful for sensing malachite green via the inner filter effect. The linear ranges were 3 nmol L−1 to 13 μmol L−1 for Hg(ii), 3–170 μmol L−1 for S2− and 7–80 μmol L−1 for malachite green, and the corresponding detection limits were 2.1 nmol L−1 for Hg(ii), 1.3 μmol L−1 for S2− and 1.6 μmol L−1 for malachite green.

Gum acacia-stabilized silver nanoparticles for the detection of Hg(ii), S2− and malachite green.  相似文献   

11.
Core–shell magnetic Fe3O4@PVBC–TMT (Fe3O4@polyvinylbenzyl chloride–trithiocyanuric acid) nanoparticles containing trithiocyanuric acid groups were fabricated and employed for the fast removal of heavy metals from an aquatic environment. The morphology, structure and properties of Fe3O4@PVBC–TMT nanoparticles were characterized by a series of modern analytical tools. The adsorption behavior of the Fe3O4@PVBC–TMT nanoparticles for heavy metals ions in aqueous solutions was investigated by batch experiments. The maximum removal capacities of the Fe3O4@PVBC–TMT nanoparticles toward Mn2+, Ni2+, Cu2+, Cd2+ and Pb2+ ions were 127.4, 146.6, 180.5, 311.5, and 528.8 mg g−1, respectively. Importantly, it is found that Pb2+ ions can be completely and quickly removed by the Fe3O4@PVBC–TMT nanoparticles. The equilibrium was established within 6 min, and the removal efficiencies were found to be 99.9%, 99.8% and 99.5% for Pb2+ ions at the initial concentrations of 100 mg L−1, 200 mg L−1 and 300 mg L−1, respectively. It is hoped that the core–shell magnetic Fe3O4@PVBC–TMT nanoparticles may find application in wastewater treatment.

Core–shell Fe3O4@PVBC–TMT nanoparticles were fabricated and served as a valid magnetic adsorbent for the removal of heavy metals ions.  相似文献   

12.
Immobilized TiO2 nanoparticles modified by nanoscale CuS (CuS@TiO2NPs) were successfully synthesized and used as fibers for solid-phase microextraction (SPME) for the determination of some polycyclic aromatic hydrocarbons (PAHs) in water samples. A novel fiber has been developed by postprecipitation of CuS coated the titania nanoparticles in situ grown on a titanium wire annealed at 550 °C in a nitrogen ambient atmosphere. Its morphology and surface properties were characterized by scanning electron microscopy and energy dispersive X-ray spectrometry. It was connected to high performance liquid chromatography-ultraviolet detector (HPLC-UV) equipment by replacing the sample loop of a six-port injection valve, building the online SPME-HPLC-UV system. Variables affecting extraction procedures, including desorption time, stirring speed, extraction temperature, extraction time and ionic strength were investigated and the parameters were optimized. The SPME fiber exhibits high selectivity for the five PAHs studied. The linear ranges varied between 0.15 μg L−1 and 200 μg L−1 with correlation coefficients ranging from 0.9913 to 0.9985. LODs and LOQs ranged from 0.02–0.04 μg L−1 and 0.07–0.13 μg L−1. RSDs for one fiber and fiber-to-fiber were in the range of 3.2–4.3% and 4.6–6.8%, respectively. Additionally, the fiber possessed advantages such as resistance to organic solvent, high mechanical strength and difficult breakage, making it have strong potential applications in the selective extraction of PAHs from complex water samples at trace levels.

Immobilized TiO2 nanoparticles modified by nanoscale CuS (CuS@TiO2NPs) were successfully synthesized and used as fibers for solid-phase microextraction (SPME) for the determination of some polycyclic aromatic hydrocarbons (PAHs) in water samples.  相似文献   

13.
The effects of formic acid, acetic acid and levulinic acid on acetone–butanol–ethanol (ABE) fermentation under different pH adjustment conditions were investigated using Clostridium acetobutylicum as the fermentation strain. CaCO3 supplementation can alleviate the inhibitory effect of formic acid on ABE production. The ABE titers from the medium containing 0.5 g L−1 formic acid with pH adjusted by CaCO3 and KOH were 11.08 g L−1 and 1.04 g L−1, which reached 64.8% and 6.3% of the control group, respectively. Compared with CaCO3 pH adjustment, fermentation results with higher ABE titers and yields were obtained from the medium containing acetic acid or levulinic acid, when the pH was adjusted by KOH. When formic acid, acetic acid, and levulinic acid co-existed in the medium, better fermentation result was achieved by adjusting the pH by CaCO3. Moreover, 12.50 g L−1 ABE was obtained from the medium containing 2.0 g L−1 acetic acid, 0.4 g L−1 formic acid, and 1.0 g L−1 levulinic acid as compared to 3.98 g L−1 ABE obtained from the same medium when the pH was adjusted by KOH. CaCO3 supplementation is a more favorable pH adjustment method for ABE medium preparation from lignocellulosic hydrolysate.

The effects of formic acid, acetic acid and levulinic acid on acetone–butanol–ethanol (ABE) fermentation under different pH adjustment conditions were investigated using Clostridium acetobutylicum as the fermentation strain.  相似文献   

14.
We report the larvicidal effects of four different morphologies of zinc oxide nanoparticles (ZnO NPs) [star-shaped (S), needle-like (N), plate-like (P) and cubical (C)] on larvae of Aedes albopictus and Anopheles vagus; the mosquitoes causing dengue fever and malaria, respectively. The nanoparticles were characterized by several analytical techniques, and their sizes and shapes were determined. Second instar larvae of the two types of mosquitoes were exposed to several concentrations of nanoparticles (25 mg L−1, 50 mg L−1, 75 mg L−1, 100 mg L−1) at 25 ± 2 °C and 84 ± 5% R.H, separately, for each morphology. Larval mortality was reported at 24 h intervals up to 21 days. The resulting LC50 for Aedes albopictus were, respectively, 38.90 mg L−1, 47.53 mg L−1, 68.38 mg L−1, 50.24 mg L−1 for S-, N-, P- and C-shaped nanoparticles. The LC50 of Anopheles vagus is lower (LC50 4.78 mg L−1, 6.51 mg L−1, 13.64 mg L−1, 10.47 mg L−1), respectively, for S-, N-, P- and C-shaped nanoparticles indicating that the nanoparticles are more toxic to Anopheles vagus larvae. The highest larvicidal effect was obtained from star-shaped nanoparticles [Aedes albopictus (38.90 mg L−1) on Anopheles vagus (4.78 mg L−1)], and the lowest was shown by the plate-like nanoparticles [Aedes albopictus (68.38 mg L−1), Anopheles vagus (13.64 mg L−1)]. The rate of development of surviving mosquito larvae was retarded when exposed to ZnO nanoparticles suggesting the possibility for these nanoparticles to kill and delay the growth of Aedes albopictus and Anopheles vagus larvae.

We report the larvicidal impacts of four different morphologies of zinc oxide nanoparticles (ZnO NPs) [star-shaped (S), needle-like (N), plate-like (P), and cubical (C)] on mosquito larvae of Aedes albopictus and Anopheles vagus.  相似文献   

15.
Nickel cobalt sulfide nanoparticles (NCS) embedded onto a nitrogen and sulfur dual doped graphene (NS-G) surface are successfully synthesized via a two-step facile hydrothermal process. The electrical double-layer capacitor of NS-G acts as a supporting host for the growth of pseudocapacitance NCS nanoparticles, thus enhancing the synergistic electrochemical performance. The specific capacitance values of 1420.2 F g−1 at 10 mV s−1 and 630.6 F g−1 at 1 A g−1 are achieved with an impressive capability rate of 76.6% preservation at 10 A g−1. Furthermore, the integrating NiCo2S4 nanoparticles embedding onto the NS-G surface also present a surprising improvement in the cycle performance, maintaining 110% retention after 10 000 cycles. Owing to the unique morphology an impressive energy density of 19.35 W h kg−1 at a power density of 235.0 W kg−1 suggests its potential application in high-performance supercapacitors.

Newly developed in situ hydrothermal synthesis governs morphology of Ni–Co–S embedded on N–S doped graphene thus providing exceptional capacitive behavior.  相似文献   

16.
Herein, silica-coated iron oxide nanoparticles modified with imidazolium-based polymeric ionic liquid (Fe3O4@SiO2@PIL) were fabricated as a sustainable sorbent for magnetic solid-phase extraction (MSPE) and simultaneous determination of trace antidiabetic drugs in human plasma by high-performance liquid chromatography-ultraviolet detection (HPLC-UV). The Fe3O4 core was functionalized by silica (SiO2) and vinyl layers where the ionic liquid 1-vinyl-3-octylimidazolium bromide (VOIM-Br) was attached through a free radical copolymerization process. In order to achieve hydrophobic magnetic nanoparticles and increase the merits of the sorbent, Br anions were synthetically replaced with PF6. The properties and morphology of the sorbent were characterized by various techniques and all the results illustrated the prosperous synthesis of Fe3O4@SiO2@PIL. A comprehensive study was carried out to investigate and optimize various parameters affecting the extraction efficiency. The limit of detection (LOD, S/N = 3) for empagliflozin, metformin and canagliflozin was 1.3, 6.0 and 0.8 ng mL−1, respectively. Linearity (0.997 ≥ r2 ≥ 0.993) and linear concentration ranges of 5.0–1200.0, 20.0–1800.0 and 5.0–1000.0 ng mL−1 were obtained for empagliflozin, metformin and canagliflozin, respectively. Intra-assay (3.8–7.5%, n = 9) and inter-assay (3.2–8.5%, n = 12) precisions as well as accuracies (≤9.1%) displayed good efficiency of the method. Finally, the method was applied for the quantitation of antidiabetic drugs in human plasma after oral administration and main pharmacokinetic data including Tmax (h), Cmax (ng mL−1), AUC0–24 (ng h mL−1), AUC0–∞ (ng h mL−1), and T1/2 (h) were evaluated.

A sustainable nanoscale core–shell modified with hydrophobic polymeric ionic liquid was fabricated for simultaneous extraction and determination of antidiabetic drugs.  相似文献   

17.
To explore the interactive molecules of squid ink polysaccharides (SIP) for further understanding the action mechanisms of SIP bio-function, this study prepared SIP binding proteins from mouse liver using superparamagnetic nanometer beads. Michaelis–Menten constant (Km) was detected from a Lineweaver–Burk double reciprocal plot to assess effect of SIP on activity of aldehyde oxidase (AOX). Results showed that three proteins, AOX-3, regucalcin (RGN) and α1-antitrypsin (A1AT3) were separated from mouse liver by magnetic nanoparticles conjugated with SIP. Contents of AOX-3 were much more than RGN and A1AT3. SIP (0.5 mg mL−1) reduced Km value of aldehyde oxidase of mouse liver from 91.79 μmol L−1 to 43.70 μmol L−1.

Superparamagnetic nanometer beads bonding SIP was employed to pull down the binding protein from the liver of mouse, which was identified as aldehyde oxidase 3. By means of enzyme kinetics analysis, SIP was found to activate AOX3 enzyme activity.  相似文献   

18.
The preparation of water-dispersible hybrid nanoparticles comprising fullerene and porphyrin from cyclodextrin complexes is described. In the presence of polyethylene glycol, C60 fullerene and porphyrin were expelled from the cyclodextrin cavity to form fullerene–porphyrin hybrid nanoparticles in water. The fullerene–porphyrin hybrid nanoparticles exhibit improved singlet oxygen generation ability under photoirradiation compared with that of C60 nanoparticles.

Hybrid nanoparticles comprising fullerene and porphyrin are formed via guest exchange reaction of cyclodextrin complexes. The hybrid nanoparticles exhibit singlet oxygen generation ability under photoirradiation.

Water-dispersible colloidal fullerene assemblies, referred to as fullerene nanoparticles (NPs), have recently received increasing attention.1–3 Fullerene NPs are negatively charged and can be dispersed in water in the absence of any solubilizer. Fullerene NPs demonstrate promise within biological and medical applications, as radical scavengers and photosensitizers for photodynamic therapy. To further extend the applications of fullerene NPs, additional hybridization with desired functional molecules is required. Porphyrin and associated derivatives are highly promising candidates for hybridization with fullerenes to increase photoactivity.4 Numerous studies on the complexation of fullerenes with porphyrin molecules using synthetic organic chemistry5–7 or supramolecular chemistry8–10 have been reported. Although fullerene NPs have been intensively studied over the last decade, no reliable method to achieve the hybridization of porphyrins with fullerene NPs has been proposed.Poly(ethylene glycol) monomethyl ether (PEG) was recently observed to accelerate the decomposition of fullerene C60–γ-CD complexes in water, which leads to the rapid aggregation of C60 to form water-dispersible C60 NPs.11 In this method, C60–γ-CD complexes can exist as stable isolated molecules in water, enabling the precise size control and step-wise growth of C60 NPs.12,13 Herein, the preparation of hybrid NPs comprising C60 and hydrophobic porphyrin molecules are reported. C60–γ-CD and porphyrin-trimethyl-β-cyclodextrin (por–TMe-β-CD) complexes are mixed in water in the presence of PEG. Both complexes decompose through the interaction of PEG with the CDs, leading to the formation of C60–porphyrin hybrid NPs (denoted as C60–por NPs). The C60–por NPs are negatively charged and easily disperse in water. Additionally, the ability of C60–por NPs to generate activated oxygen is also evaluated.The C60–γ-CD complex14,15 and 1–TMe-β-CD complex (Fig. 1)16–18 were prepared according to a previously described procedure (see the ESI for details). The 1H NMR spectrum of the mixed solution comprising the C60–γ-CD complex and PEG after 1 h incubation at 80 °C shows that the peaks attributed to the C60–γ-CD complexes completely disappeared (Fig. S1). Hence, the 1H NMR data confirm the decomposition of the C60–γ-CD complexes and the formation of water-dispersible C60 NPs, as previously reported.11–13 The effect of PEG on 1–TMe-β-CD complexes was also investigated by 1H-NMR as shown in Fig. S2. After incubating the mixed solution of 1–TMe-β-CD complex and PEG ([1] = 0.1 mM, [PEG] = 5.0 g L−1) for 1 h at room temperature, peaks attributed to the 1–TMe-β-CD complex were still evident at 4.97 ppm and above 7.7 ppm (Fig. S2(i)). Hence, PEG has no influence upon the 1–TMe-β-CD structure at room temperature. Conversely, after incubating for 1 h at 80 °C, a dark purple precipitate formed and the aforementioned 1H NMR peaks completely disappeared (Fig. S2(ii)). In the absence of PEG, the 1–TMe-β-CD complex was stable in water both at room temperature (Fig. S2(iii)) and 80 °C (Fig. S2(iv)). These results suggest a decomposition route of 1–TMe-β-CD by interaction with PEG at 80 °C, with a concomitant formation of non-dispersible large aggregates.Open in a separate windowFig. 1Chemical structures of porphyrin derivatives used in this study.To obtain hybrid NPs comprising C60 and 1 (C60–1 NPs), PEG (Mw = 2000) was added to an aqueous solution containing C60–γ-CD and 1–TMe-β-CD complexes ([C60] = [1] = 0.1 mM, [PEG] = 5.0 g L−1), which were thereafter incubated at room temperature or 80 °C. The 1H NMR spectrum of the mixed solution at room temperature shows peaks attributed to γ-CD in the C60–γ-CD complex at 5.03 ppm, TMe-β-CD in the 1–TMe-β-CD complex at 4.97 ppm, and 1 in the 1–TMe-β-CD complex in the region of 7.6–8.5 ppm (Fig. 2a(i)). The data indicate that PEG fails to induce the decomposition of the C60–γ-CD and 1–TMe-β-CD complexes at room temperature. Conversely, after the mixture was heated at 80 °C for 1 h, the peaks attributed to the C60–γ-CD and 1–TMe-β-CD complexes completely disappeared (Fig. 2a(ii)). Hence, C60–γ-CD and 1–TMe-β-CD were decomposed at 80 °C, in the presence of PEG. The solution after being subjected to heat treatment at 80 °C for 1 h, was dark purple in the absence of any precipitate. The hydrodynamic diameter and ζ-potential of the reacted solution were 125 nm (polydispersity index = 0.21) and −20.2 mV, respectively. Water dispersible fullerene NPs typically exhibit negative ζ-potentials, the origin of which still requires elucidating.19,20 Hence, the formation of water-dispersible nano-composites, C60–1 NPs, is suggested.Open in a separate windowFig. 2(a) 1H NMR spectra of mixed solutions comprising fullerene C60–γ-cyclodextrin (CD) and 1–trimethyl (TMe)-β-CD complexes ([C60] = [1] = 0.1 mM) (i) before and (ii) after heating at 80 °C for 1 h, in the presence of polyethylene glycol (PEG) (5 g L−1). Open circles: free γ-CD, filled circles: C60–γ-CD, open diamonds: free TMe-β-CD, and filled diamonds: porphyrin–TMe-β-CD (por–TMe-β-CD) complex. The spectra at 7.6–8.5 ppm, are amplified five-fold. (b) Ultraviolet-visible (UV/Vis) absorption spectra of the mixed solution comprising C60–γ-CD and 1–TMe-β-CD complexes ([C60] = [1] = 0.1 mM) with PEG (5 g L−1), before (dashed line) and after (solid line) heating at 80 °C for 1 h. (c) UV/Vis absorption spectra of the mixed solution comprising the C60–γ-CD complex as a function of the 1–TMe-β-CD complex concentration ([C60] = 0.1 mM, [1] = 0–0.2 mM) with PEG (5 g L−1) after heating at 80 °C for 1 h.The por–TMe-β-CD complexes using 2–6 (Fig. 1), were also prepared adopting the same procedure as that for the 1–TMe-β-CD complex. Each por–TMe-β-CD complex solution was mixed with C60–γ-CD and PEG ([C60] = [por] = 0.1 mM, [PEG] = 5.0 g L−1). The 1H NMR spectrum of each individual mixed solution after being incubated for 1 h at 80 °C, is shown in Fig. S3. In the 1H NMR spectrum of the mixture comprising C60–γ-CD and 2–TMe-β-CD, the peaks attributed to these complexes at 5.03, 4.98, and 7.60–8.50 ppm, completely disappeared after being subjected to incubation for 1 h at 80 °C, without precipitation (Fig. S3a). Similar changes in the 1H NMR spectrum of the mixture comprising C60–γ-CD and 3–TMe-β-CD complexes are observed, as shown in Fig. S3b. The data suggest that the 2–TMe-β-CD and 3–TMe-β-CD complexes can be decomposed in a similar manner as the C60–γ-CD complexes, and imply the formation of C60–2 and C60–3 NPs.The 1H NMR spectra of the mixed solutions comprising C60–γ-CD and 4, 5, or 6–TMe-β-CD complexes failed to show peaks attributed to the C60–γ-CD complex, and peaks associated with the respective por–TMe-β-CD complexes were observed after incubation for 1 h at 80 °C (Fig. S3c–e, respectively). Hence, the data suggest that the C60–γ-CD complex decomposed in the presence of PEG, and the 4, 5, and 6–TMe-β-CD complexes were observed to be stable without decomposition at 80 °C. There have been reports suggesting the strong interaction of water-soluble tetraphenyl porphyrins with TMe-β-CDs.21,22 Polar substituents prompt the penetration of the polarized porphyrin rims into the TMe-β-CD cavity. Porphyrins 4–6 possess polar substituents, which are suggested to enable the formation of stable TMe-β-CD complexes. Furthermore, the size of the β-CD cavity is sufficiently narrow to prevent any strong interaction with PEG.23 Thus, PEG-induced decomposition of the 4, 5, and 6–TMe-β-CD complexes is not possible.The absorption behavior of C60–1 NPs was investigated using ultraviolet-visible (UV/Vis) spectroscopy. In the UV/Vis spectra, the characteristic peak of solvated C60–γ-CD, at 333 nm shifted to 344 nm after heating at 80 °C for 1 h (Fig. 2b). An additional broad absorption at 400–550 nm is also apparent, which is characteristic of solid-state crystalline C60 and arises from the electronic interactions between adjacent C60 molecules.24,25 The characteristic peak of the solvated 1–TMe-β-CD complex at 415 nm, shifted to 432 nm, with induced broadening after being subjected to heat treatment at 80 °C for 1 h (Fig. 2b). In the absence of C60–γ-CD complexes, the characteristic absorption peak attributed to the solvated 1–TMe-β-CD complex completely disappeared after heating for 1 h at 80 °C, in the presence of PEG (Fig. S4). The data show that 1, when expelled from the TMe-β-CD cavities, forms non-dispersible precipitates in the absence of C60. For 1 dispelled from the TMe-β-CD cavities to be stably dispersed in water, formation of co-aggregates with C60 may be a prerequisite. C60–2 and C60–3 NPs also show similar UV-Vis absorption spectra after being subjected to heating at 80 °C for 1 h, as shown in Fig. S5a and b, respectively.To further elucidate the composite formation of C60 and 1, the influence of 1 concentration on C60–1 NP formation was investigated by UV/Vis spectroscopy (Fig. 2c). The intensity of the absorption peak at 432 nm increased as a function of 1 concentration from 0.05 to 0.1 mM. Conversely, the absorption peak at 345 nm, derived from the formation of fullerene NPs, shifted to 338 nm, with increasing concentration of 1. This absorption peak derives from the fullerene nanoparticle size, and as the size decreased (i.e., the NPs became smaller), the peak blue-shifted.11 The results suggest that in the presence of 1, the fullerene interaction might be disturbed, or smaller C60 NPs might form. For C60–1 NPs formulated with 0.2 mM of 1 ([C60] = 0.1 mM, [1] = 0.2 mM), the absorption peak derived from the Soret band of 1 split into two peaks (Fig. 2c). The absorption peak at the shorter wavelength of 415 nm is consistent with that of the 1–TMe-β-CD complex. The absorption peak at the longer wavelength of 431 nm is almost consistent with the absorption peaks in the UV/Vis spectra of C60–1 NPs fabricated with 0.05 and 0.1 mM of 1. The findings indicate that in the sample comprising 0.2 mM of 1, a portion of the 1–TMe-β-CD complexes remained in the complex state after heating for 1 h at 80 °C, in the presence of PEG. The absorption peak at 338 nm, which reflects the state of fullerene NPs, is similar to that of C60–1 NPs fabricated with 0.1 mM of 1. When C60 and 1 form co-aggregated NPs, the ratio of 1 to C60 is thought to be limited to ∼1 : 1.Morphological observations of the hybrid NPs were also undertaken. In the absence of the por–TMe-β-CD complex, C60 NPs possessing fairly monodisperse size distributions were observed (Fig. S6a). The average diameter of the individual NPs, determined from the transmission electron microscopy (TEM) images, is 82 nm. C60 NPs have been previously reported to exhibit lattice fringes and diffraction patterns, which suggests that C60 NPs maintain the face-centered cubic (fcc) crystalline structure.11 C60–1 NPs prepared with 0.05 mM C60 and 0.1 mM 1, possessed irregular shapes (Fig. 3a and b, respectively). The average diameter of C60–1 NPs, determined by TEM, is 119 nm (Fig. 3a and S6b), demonstrating the larger C60–1 NP size than that of the C60 NPs (82 nm). Increasing the concentration of 1 to 0.1 mM results in the average diameter of C60–1 NPs to increase to 131 nm (Fig. 3b and S6c). Similar morphology is observed from the TEM micrographs of C60–2 and C60–3 NPs having average diameters of 109 nm and 144 nm, respectively (Fig. 3c and d, respectively). A lower PEG molecular weight or a lower reaction temperature during C60 NP formation via C60–γ-CD complexes have been reported to induce an increase in the diameter of the C60 NPs.11,12 Thus, the findings suggest that slower reaction conditions result in less nucleation and a larger NP formation. Porphyrin 3 possesses a methoxy substituent at the para position of the phenyl group and is more polar than 1 or 2. Previous reports have demonstrated that the higher the polarity of the phenyl group, the more stable the complex with TMe-β-CD,21,22 which indicates that 3–TMe-β-CD is more stable than 1– or 2–TMe-β-CD in water. Thus, the aforementioned decomposition, which results from the interaction with PEG, is slower in 3 with a concomitant increase in the NP size.Open in a separate windowFig. 3Transmission electron microscopy (TEM) images of C60–1 nanoparticles (NPs) prepared with (a) 0.05 and (b) 0.1 mM of the 1–TMe-β-CD complex. TEM images of (c) C60–2 and (d) C60–3 NPs. Scale bars in images (a–d) are 100 nm. (e) High resolution TEM micrograph and selected-area electron diffraction pattern of C60–1 NP. Scale bar is 10 nm. (f) 13C NMR spectra of (i) C60 NPs and (ii) C60–1 NPs. (g) Illustration of a C60–por NP. A portion of C60 form crystalline structures, while a portion of the porphyrin molecules interact with C60 at the molecular level.In the high-resolution TEM micrograph (Fig. 3e), the C60–1 NPs only exhibited partial lattice fringes, and hence did not show clear diffraction patterns compared with the C60 NPs (inset in Fig. 3e).11 The findings demonstrate the highly amorphous nature of the C60–1 NPs, and that the C60 crystal structure was only retained in part. 13C NMR spectra also provide important information about the structure of the C60–1 NPs. A characteristic C60 cluster signal at 142.4 ppm was detected in both C60 NPs and C60–1 NPs (Fig. 3f).26 The C60–1 NP dispersions also exhibited several small new signals at 141.5, 143.3, and 143.7 ppm, as shown in Fig. 3f(ii). When a fullerene and a porphyrin molecule form a stable complex in solution, the C60 signal shifts depending on the interaction type between the fullerene and the porphyrin molecule.27 Thus, the porphyrin molecule interacted with the aggregate of C60 in C60–1, as illustrated in Fig. 3g.Some porphyrin molecules can form a co-crystal with fullerene C60.28 To form a crystal structure, not only the interaction between molecules but also the relationship with the solvent, such as gradually changing the polarity of the solvent or removing the solvent, are important. In our system, porphyrin molecules that are pseudo-dissolved by TMe-β-CDs are added to water, which is a poor solvent for porphyrins, through the interaction of PEG with TMe-β-CDs. The porphyrin molecule should immediately aggregate and have difficulty forming a crystal structure. Furthermore, because water is also a poor solvent for fullerene C60, C60 also immediately aggregates in water. Thus, it should be extremely difficult for C60 and porphyrin molecules to regularly associate to form a co-crystal structure.The concentration of singlet oxygen molecules (1O2, Type-II energy transfer pathway) generated by photoirradiation was measured according to a chemical method using 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA)15,29 as a marker to determine the biological activities of C60 NPs, C60–1 NPs, C60–2 NPs, and C60–3 NPs. The absorption of ABDA at the absorption maximum (380 nm) was monitored as a function of irradiation time ([C60] = 0.1 mM, [por] = 0 or 0.1 mM). Under visible-light irradiation at wavelengths > 620 nm, C60–1 NPs, C60–2 NPs, and C60–3 NPs generated higher levels of 1O2 than C60 (Fig. 4a). These results show that the 1O2 photoproduction abilities of the C60–por NPs were higher than that of the C60 NPs. There are no significant differences in the 1O2 photoproduction abilities of C60–1 NPs, C60–2 NPs, and C60–3 NPs, which suggests that the structure of the porphyrin has an insignificant influence on the ability of the hybrid NPs. The generation of formazan, via the reduction of nitroblue tetrazolium (NBT) by oxygen radicals (O2˙), is observed as an increase of absorption intensity at 560 nm.30 The reduction of NBT by O2˙ was scarcely detected in solutions containing C60–1 NPs, C60–2 NPs, and C60–3 NPs under photoirradiation, even though formazan was readily detected in the positive control sample in the presence of reduced nicotinamide adenine dinucleotide (NADH) (Fig. 4b). The results suggest that the reactive oxygen species produced by C60–1 NPs, C60–2 NPs, and C60–3 NPs are predominantly 1O2 generated by a Type-II reaction.18Open in a separate windowFig. 4(a) 1O2 generation by NPs. Bleaching of 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA) was monitored as a function of the decrease in the absorbance at 380 nm, for C60 NPs (black circles and solid line), C60–1 NPs (red circles and solid line), C60–2 NPs (blue circles and solid line), and C60–3 NPs (green circles and solid line) ([C60] = 15 μM, [por] = 0 or 15 μM, [ABDA] = 25 μM). (b) O2˙ generation by NPs. The amount of formazan generated by the reduction of nitroblue tetrazolium (NBT) in the presence of O2˙ was analyzed by the absorbance at 560 nm, of C60–1 NPs (red circles) and C60–2 NPs (blue circles) in the absence (solid lines) and presence (dashed lines) of NADH ([C60] = 15 μM, [por] = 15 μM, [NBT] = 200 μM, [NADH] = 0 or 625 μM). All samples were photoirradiated at >620 nm, in O2-saturated aqueous solutions.In summary, the preparation of hybrid C60–porphyrin NPs was achieved via a guest exchange reaction comprising porphyrin CD complexes and C60. Seven C60–por NP derivatives with various moieties were prepared. CD porphyrin complexes possessing phenyl and methoxyphenyl moieties were decomposed in the presence of PEG at the same time as C60–γ-CD complexes and formed NPs with C60. Porphyrins containing a hydrophilic moiety form stable complexes with TMe-β-CD and fail to co-aggregate with C60. The C60–por NPs are negatively charged and are easily dispersed and stable in water. The 1O2 generation ability of C60–por NPs under photoirradiation (>620 nm) is greater than that of C60 NPs. The findings herein demonstrate a new method to fabricate fullerene–porphyrin composite materials, which provides a route to highly functional fullerene-based materials.  相似文献   

19.
In this article, we present a novel synthesis of mesoporous SiO2/Ag nanostructures for dye (methylene blue) adsorption and surface plasmon mediated photocatalysis. Mesoporous SiO2 nanoparticles with a pore size of 3.2 nm were synthesized using cetyltrimethylammonium bromide as a structure directing agent and functionalized with (3-aminopropyl)trimethoxysilane to introduce amine groups. The adsorption behavior of non-porous SiO2 nanoparticles was compared with that of the mesoporous silica nanoparticles. The large surface area and higher porosity of mesoporous SiO2 facilitated better adsorption of the dye as compared to the non-porous silica. Ag decorated SiO2 nanoparticles were synthesized by attaching silver (Ag) nanoparticles of different morphologies, i.e. spherical and triangular, on amine functionalized silica. The photocatalytic activity of the mesoporous SiO2/Ag was compared with that of non-porous SiO2/Ag nanoparticles and pristine Ag nanoparticles. Mesoporous SiO2 nanoparticles (kd = 31.3 × 10−3 g mg−1 min−1) showed remarkable improvement in the rate of degradation of methylene blue as compared to non-porous SiO2 (kd = 25.1 × 10−3 g mg−1 min−1) and pristine Ag nanoparticles (kd = 19.3 × 10−3 g mg−1 min−1). Blue Ag nanoparticles, owing to their better charge carrier generation and enhanced surface plasmon resonance, exhibited superior photocatalysis performance as compared to yellow Ag nanoparticles in all nanostructures.

In this article, we present a novel synthesis of mesoporous SiO2/Ag nanostructures for dye (methylene blue) adsorption and surface plasmon mediated photocatalysis.  相似文献   

20.
In this work, a sensitive and efficient voltammetric biosensor was introduced for differential pulse voltammetric (DPV) determination of some phthalic acid esters (PAEs) including dibutyl phthalate (DBP), dimethyl phthalate (DMP), di(2-ethylhexyl)phthalate (DEHP) and dicyclohexyl phthalate (DCHP) in aqueous solutions. Briefly, the surface of a copper electrode was modified by azolla paste prepared using azolla powder and electroencephalography gel (EEG). The modified surface was characterized by electrochemical impedance spectroscopy (EIS), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), Fourier transform infrared spectroscopy (FT-IR), Brunauer–Emmett–Teller (BET) analysis and energy dispersive X-ray (EDX) methods. Determination of PAEs was conducted based on their blocking effect on the electrode surface for ferrous ion oxidation. The central composite design (CCD) was conducted to optimize the effects of four experimental parameters including the concentration of Fe2+ ions (CFe2+) and supporting electrolyte (Csup. elec), solution pH and modifier/gel mass ratio on the decrease in the anodic peak current of ferrous ions as the response. Predicted optimal conditions (CFe2+= 319 μM, Csup. elec= 0.125 M, pH = 7.52 and modifier/gel mass ratio = 0.19) were validated by experimental checking which resulted in an error of 1.453%. At the optimum conditions, linear relationships were found between the DPV responses and PAEs concentrations and the limit of detection (LOD) and limit of quantification (LOQ) values were in the ranges of 0.2–0.4 μg L−1 and 0.5–1.0 μg L−1, respectively. Good recovery percentages ranging from 97.3 to 100.3% with RSD < 3.2% suggested the proposed method for efficient, accurate and quick determination of PAEs in real water samples.

In this work, a sensitive and efficient voltammetric biosensor was introduced for differential pulse voltammetric (DPV) determination of dibutyl phthalate, dimethyl phthalate, di(2-ethylhexyl)phthalate and dicyclohexyl phthalate in aqueous solutions.  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