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 15 毫秒
1.
2.
3.
Correction for ‘Efficient removal of cobalt from aqueous solution using β-cyclodextrin modified graphene oxide’ by Wencheng Song et al., RSC Adv., 2013, 3, 9514–9521.

The authors regret that Fig. 1 and and33 were incorrect in the original article. The SEM images of both GO and β-CD, and the Raman spectra of both, were confused with other samples. The correct versions of Fig. 1 and and33 are presented below.Open in a separate windowFig. 1SEM images of (a) GO and (b) β-CD-GO.Open in a separate windowFig. 2Raman spectra of GO and β-CD-GO.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Correction for ‘Hybrid cellulose nanocrystal/alginate/gelatin scaffold with improved mechanical properties and guided wound healing’ by Yue Shan et al., RSC Adv., 2019, 9, 22966–22979, https://doi.org/10.1039/C9RA04026A.

The authors regret that incorrect versions of Fig. 7 and and88 were included in the original article. The correct versions of Fig. 7 and and88 are presented below.Open in a separate windowFig. 7H&E staining images in control, SA/Ge, and SA/Ge/CNC groups at 7 days and 14 days after surgery. The bar corresponds to 50 μm.Open in a separate windowFig. 8Masson’s trichrome staining images in control, SA/Ge, and SA/Ge/CNC groups at 7 and 14 days after surgery. The bar corresponds to 50 μm.An independent expert has viewed the corrected images/data and has concluded that they are consistent with the discussions and conclusions presented.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

5.
Correction for ‘Chrysomycins A–C, antileukemic naphthocoumarins from Streptomyces sporoverrucosus’ by Shreyans K. Jain et al., RSC Adv., 2013, 3, 21046–21053, https://doi.org/10.1039/c3ra42884b.

The authors regret that incorrect versions of Fig. 6 and Fig. 7 were included in the original article. The correct versions of Fig. 6 and and77 are presented below.Open in a separate windowFig. 6Influence of compounds 1–3 on the nuclear morphology of human leukaemia HL-60 cells. The cells were treated with 1, 3 and 5 μM concentrations of these compounds for 24 h and stained with Hoechst 33258 for 40 min. The altered nuclear morphology and apoptotic bodies indicated by white arrows are seen in treated cells while the nuclei of the untreated cells were round and intact.Open in a separate windowFig. 7Phase contrast microscopy of compound-treated leukaemia HL-60 cells. Cells were treated with compounds 1–3 at 1, 3 and 5 μM for 24 h and visualized using a phase contrast microscope (Olympus1X72). The morphology of treated cells altered in a concentration-dependent manner, while the untreated cells remained healthy.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

6.
Correction for ‘Green-synthesised cerium oxide nanostructures (CeO2-NS) show excellent biocompatibility for phyto-cultures as compared to silver nanostructures (Ag-NS)’ by Qaisar Maqbool, RSC Adv., 2017, 7, 56575–56585, https://doi.org/10.1039/c7ra12082f.

The author regrets that Fig. 4 and and55 of the original article did not appropriately represent the findings.Open in a separate windowFig. 4Comparative TGA analysis of CeO2-NS and Ag-NS.Open in a separate windowFig. 5FTIR spectrum of CeO2-NS and Ag-NS.The correct version of Fig. 4 is shown below. In addition, the associated text on page 56578 “Experimental findings show total mass loss…” should be changed to “Experimental findings show total mass loss of 57.53% by CeO2-NS and 61.12% by Ag-NS.” Fig. 5 of the original article shows only the plot of selected data points. In order to provide clarity to readers, it should be replaced with the following original FTIR plots (complete scan).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

7.
Correction for ‘Variation in surface properties, metabolic capping, and antibacterial activity of biosynthesized silver nanoparticles: comparison of bio-fabrication potential in phytohormone-regulated cell cultures and naturally grown plants’ by Tariq Khan et al., RSC Adv., 2020, 10, 38831–38840, DOI: 10.1039/D0RA08419K.

The authors regret that an incorrect version of Fig. 7 was included in the original article. The correct version of Fig. 7 is presented below.Open in a separate windowFig. 7Venn diagram for the comparative analysis of compounds detected through LC-MS/MS.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

8.
9.
Correction for ‘A p-type multi-wall carbon nanotube/Te nanorod composite with enhanced thermoelectric performance’ by Dabin Park et al., RSC Adv., 2018, 8, 8739–8746.

The authors regret that an incorrect version of Fig. 8 was included in the original article. The correct version of Fig. 8 is presented below.Open in a separate windowFig. 1FE-SEM images of MWCNT/Te nanorod composites with various MWCNT contents (a) 1 wt%, (b) 3 wt%, and (c) 5 wt%.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

10.
Correction for ‘Improving the inhibitory effect of CXCR4 peptide antagonist in tumor metastasis with an acetylated PAMAM dendrimer’ by Changliang Liu et al., RSC Adv., 2018, 8, 39948–39956.

The authors regret that the term “CXCL12” was incorrectly displayed as “CXCR12” in Scheme 1 and Fig. 6(a)–(c) in the original article. The correct versions of Scheme 1 and Fig. 6(a)–(c) are presented below.Open in a separate windowScheme 1Schematic illustration of the preparation of the PAC80–E5 complex and the process of anti-tumor metastasis of the E5 peptide in the presence of PAC80.Open in a separate windowFig. 6(a–c) The inhibitory effect of E5 and PAC80–E5 on: (a) MCF-7; (b) MDA-MB-231; and (c) 4T1 cells detected by transwell assay. The CXCL12 supplemented sample without E5 or PAC80–E5 was set as 100% as the control. Error bars represent the standard deviation (n = 3).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

11.
Correction for ‘Synthesis of a well-dispersed CaFe2O4/g-C3N4/CNT composite towards the degradation of toxic water pollutants under visible light’ by Fei Liu et al., RSC Adv., 2019, 9, 25750–25761.

The authors regret that mistakes were made during the preparation of Fig. 1 in the published article. In the original article, Fig. 1 presented XRD data for CaFe2O4/CNT, which inadvertently duplicated the data for CaFe2O4. In addition, the spectra in the original figure do not match with the relevant discussion of Fig. 1 (on page 25753–25754) which referred to XRD patterns for g-C3N4, CaFe2O4 and CaFe2O4/g-C3N4/CNT composite. The correct image for Fig. 1 is shown below but no changes are required to the associated discussion of Fig. 1.Open in a separate windowFig. 1XRD patterns of the as-prepared CaFe2O4 and CaFe2O4/g-C3N4/CNT composite.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

12.
Correction for ‘Preparation of a novel curcumin nanoemulsion by ultrasonication and its comparative effects in wound healing and the treatment of inflammation’ by Niyaz Ahmad et al., RSC Adv., 2019, 9, 20192–20206, DOI: 10.1039/C9RA03102B.

The authors regret errors in Fig. 7 in the original article. The corrected Fig. 7 is shown below where the panels at 8 and 12 days for fusidic acid and 16 and 20 days for Cur-NE have been replaced.Open in a separate windowFig. 7Wound healing effects of optimized nanoemulsion without Cur loaded; pure clove oil; pure Cur-S; optimized nanoemulsion and marketed preparation of antibiotic fusidic acid (Fusidin; positive control) in comparison with the control after 0, 4, 8, 12, 16, 20 and 24 days of inducing wound healing.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

13.
Correction for ‘Structure evolution, amorphization and nucleation studies of carbon-lean to -rich SiBCN powder blends prepared by mechanical alloying’ by Daxin Li et al., RSC Adv., 2016, 6, 48255–48271.

The authors regret that Fig. 13 was displayed incorrectly in the original article. Due to a data processing error, partially repetitive data was displayed for the entry for 10 h. The correct version of Fig. 13 is shown below.Open in a separate windowFig. 13Solid-state 29Si NMR spectra of carbon-lean C2 (a) and carbon-rich C9 (b) powder blends subjected to different hours of milling.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

14.
Correction for ‘Photodynamic antimicrobial chemotherapy with cationic phthalocyanines against Escherichia coli planktonic and biofilm cultures’ by Min Li et al., RSC Adv., 2017, 7, 40734–40744, https://doi.org/10.1039/C7RA06073D.

The authors regret that incorrect versions of Fig. 7F (Control) and Fig. 8A (Light-alone) were included in the original article. The corrected versions are shown below. The correction does not change any results or conclusions of the original paper.Open in a separate windowFig. 7Membrane integrity detected by PI staining. (F) Images taken by fluorescence microscope of E. coli treated with 5 μM ZnPc2 in different groups.Open in a separate windowFig. 8SEM images of PACT-subjected E. coli biofilms. (A) Images of E. coli treated with 20 μM ZnPc1-PACT in different groups.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

15.
Four triphenylamine or carbazole-based benzothiadiazole fluorescent molecules have been successfully synthesized and characterized. Interestingly, the donor–acceptor (D–A) type luminogens 1, 2, 3 and 4 showed different solid-state fluorescence. Furthermore, the four compounds exhibited reversible high-contrast mechanochromism characteristics.

Four triphenylamine or carbazole-based benzothiadiazole dyes were synthesized. Interestingly, the four dyes exhibited high-contrast mechanochromism characteristics.

Stimuli-responsive materials receive much attention currently due to their academic importance and potential applications in optoelectronic devices and fluorescent sensors,1–7 especially organic smart materials whose solid-state luminescence can be tuned by external stimuli.8–13 Mechanochromic fluorescence materials, as a class of smart materials, are also receiving increasing attention.14–20 To date, a number of mechanofluorochromic organic molecules have been reported.21–24 In contrast, examples of high-contrast mechanochromic luminescence materials are still inadequate. Indeed, many traditional organic materials are aggregation caused quenching (ACQ)-active, and these materials are weakly emissive or nonluminescent in the solid state due to the presence of strong intermolecular electronic interactions in their aggregated state, which promotes the formation of exciplexes and excimers.25–27 Obviously, the ACQ effect is unbeneficial to gain high-contrast mechanofluorochromic materials.28–30 It is no doubt that mechanochromic molecules with a bright solid-state fluorescence emission are easier to achieve high-contrast mechanofluorochromic phenomenon. Therefore, the corresponding highly emissive smart luminophors have attracted considerable attention.31,32In general, the emission characteristics of mechanochromic luminescence materials depend strongly on their molecular structures and intermolecular interactions.33–35 Therefore, it is an effective method for the realization of mechanofluorochromic materials to change the morphological structures by means of external mechanical stimulus.36Benzothiadiazole-based derivatives are regarded as attractive candidates for the organic π-conjugated fluorescent dyes owing to their strongly electron-withdrawing feature.37–41 Meanwhile, the benzothiadiazole unit is also advantageous to the construction of donor–acceptor (D–A) type molecules, which have emerged as a significative class of optical materials finding potential value in some areas such as in fluorescent sensors and displays.42,43 Motivated by the fact that triphenylamine or carbazole fluorogen has been broadly applied in the field of emissive materials,44,45 we attempted to link one triphenylamine or carbazole group to one benzothiadiazole moiety. As a result, we have obtained four D–A type fluorescent molecules on the basis of a combination of the electron-donating triphenylamine or carbazole unit and the electron-accepting benzothiadiazole unit (Fig. 1). Compound 1, 2, 3 or 4 contains rotatable aromatic rings, and thus their molecular structures are nonplanar, which is advantageous to the radiative decay in the aggregated state. Indeed, compounds 1, 2, 3 and 4 showed bright solid-state fluorescence with different emission colors. In addition, we found that the D–A type luminogens 1, 2, 3 and 4 applying the triphenylamine or carbazole moiety as an electron donor and the benzothiadiazole moiety as an electron acceptor exhibited various mechanochromic fluorescence characteristics with good reversibility. Furthermore, luminogen 1 showed mechanofluorochromic behavior involving color change from orange to rare red.Open in a separate windowFig. 1Molecular structures of the compounds 1, 2, 3 and 4.To investigate the solid-state fluorescence behaviors of compounds 1, 2, 3 and 4 in detail, the corresponding solid-state emission spectra were studied initially. As shown in Fig. 2, the fluorescence spectrum of triphenylamine-containing benzothiadiazole derivative 1 exhibited one emission band with the λmax at 575 nm, and the fluorescent molecule exhibited strong orange luminescence with the fluorescence quantum yield (Φ) of 7.13%, and triphenylamine-containing compound 2 exhibited strong yellow luminescence (Φ = 7.43%) with the λmax at 567 nm. In contrast, the emission spectrum of carbazole-based benzothiadiazole derivative 3 exhibited one emission band with the λmax at 504 nm, and the luminogen exhibited bright green fluorescence with the quantum yield of 16.10%, and carbazole-based compound 4 also exhibited bright green fluorescence (Φ = 16.53%) with the λmax at 498 nm. Therefore, the photoluminescence (PL) behaviors of compounds 1, 2, 3 and 4 could be adjusted via introducing various fluorogens containing triphenylamine and carbazole. In addition, the fluorescence lifetimes of 1, 2, 3 and 4 were also measured. As shown in Fig. 3, the average lifetime of fluorescent molecule 1 was 0.82 ns, the average lifetime of 2 was 1.56 ns, the average lifetime of 3 was 3.00 ns, and the average lifetime of 4 was 1.41 ns.Open in a separate windowFig. 2Solid-state emissive spectra of the compounds 1, 2, 3 and 4, and the related fluorescence images under 365 nm UV light.Open in a separate windowFig. 3(a) Time-resolved luminescence (575 nm) of solid sample 1. Excitation wavelength: 365 nm. (b) Time-resolved luminescence (567 nm) of solid sample 2. Excitation wavelength: 365 nm. (c) Time-resolved luminescence (504 nm) of solid sample 3. Excitation wavelength: 365 nm. (d) Time-resolved luminescence (498 nm) of solid sample 4. Excitation wavelength: 365 nm.Subsequently, the mechanochromic fluorescence characteristics of compounds 1, 2, 3 and 4 were investigated. As shown in Fig. 4, the solid sample of luminogen 1 showed a bright orange fluorescence. Interestingly, the orange luminescence was changed to the red luminescence with the λmax at 593 nm upon treating with mechanical force stimulus. Furthermore, the initial orange emission could be restored after treatment of the ground compound 1 with fuming dichloromethane for 1 min. Therefore, 1 showed reversible high-contrast mechanofluorochromic behavior with color change from orange to red, which is a relatively rare color conversion among all mechanochromic fluorescence phenomena.Open in a separate windowFig. 4(a) PL spectra of solid sample 1 at different conditions. Excitation wavelength: 365 nm. (b) Fluorescence image of the unground sample 1 under 365 nm UV light. (c) Fluorescence image of the ground sample 1 under 365 nm UV light. (d) Fluorescence image of the ground sample 1 after treatment with dichloromethane under 365 nm UV light.Similarly, as shown in Fig. 5, compound 2 also showed reversible high-contrast mechanochromic fluorescence behavior. Moreover, the reversible mechanochromic fluorescence of 1 or 2 could be repeated four times between the orange or yellow and red or orange emissions without obvious changes by alternating grinding and dichloromethane treatments. To date, this mechanochromic luminescence conversion of some reported mechanochromism compounds with superior performance is also repeated three or four times,46–48 and thus the reversibility of the mechanochromic fluorescence effect of 1 or 2 is good (Fig. 6). On the other hand, as shown in Fig. 7, when sample 3 were ground in an agate mortar with a pestle, the green emission was changed to the yellow-green fluorescence with the λmax at 533 nm. Moreover, the yellow-green emission could also revert to the original green emission after a 1 min treatment of the ground powder with fuming dichloromethane vapor. Furthermore, as shown in Fig. 8, compound 4 also showed similar mechanochromic fluorescence behavior.Open in a separate windowFig. 5(a) PL spectra of solid sample 2 at different conditions. Excitation wavelength: 365 nm. (b) Fluorescence image of the unground sample 2 under 365 nm UV light. (c) Fluorescence image of the ground sample 2 under 365 nm UV light. (d) Fluorescence image of the ground sample 2 after treatment with dichloromethane under 365 nm UV light.Open in a separate windowFig. 6(a) Repetitive experiment of mechanofluorochromic effect for compound 1. (b) Repetitive experiment of mechanofluorochromic effect for compound 2.Open in a separate windowFig. 7(a) PL spectra of solid sample 3 at different conditions. Excitation wavelength: 365 nm. (b) Fluorescence image of the unground sample 3 under 365 nm UV light. (c) Fluorescence image of the ground sample 3 under 365 nm UV light. (d) Fluorescence image of the ground sample 3 after treatment with dichloromethane under 365 nm UV light.Open in a separate windowFig. 8(a) PL spectra of solid sample 4 at different conditions. Excitation wavelength: 365 nm. (b) Fluorescence image of the unground sample 4 under 365 nm UV light. (c) Fluorescence image of the ground sample 4 under 365 nm UV light. (d) Fluorescence image of the ground sample 4 after treatment with dichloromethane under 365 nm UV light.As can be seen in Fig. 9, the reversibility of the mechanofluorochromic behavior of compound 3 or 4 is also excellent. Next, the powder X-ray diffraction (XRD) patterns were studied in order to ensure the morphological characteristics. As can be seen in Fig. 10, the XRD patterns of compound 1 or 2 exhibited a number of sharp reflection peaks, suggesting that the unground compound 1 or 2 was crystalline in nature. However, the ground powder sample became amorphous, with a lack of sharp diffraction peaks. Therefore, the change in fluorescence of compound 1 or 2 could be attributed to the conversion from a crystalline state to an amorphous state. On the other hand, when the ground sample was exposed to dichloromethane vapor for 1 min, the sharp and intense peaks reappeared, indicative of the recovery of the crystalline nature. As presented in Fig. 11, the structural transition of the powder sample of compound 3 or 4 was similar to that of 1 or 2. Based on the above mentioned analysis, the powder XRD results demonstrated that the interesting mechanochromic fluorescence characteristics of compounds 1, 2, 3 and 4 were ascribed to the switchable morphology transition between the crystalline state and the amorphous state.Open in a separate windowFig. 9(a) Repetitive experiment of mechanofluorochromic effect for compound 3. (b) Repetitive experiment of mechanofluorochromic effect for compound 4.Open in a separate windowFig. 10(a) Powder XRD patterns of compound 1 in different solid states. (b) Powder XRD patterns of compound 2 in different solid states.Open in a separate windowFig. 11(a) Powder XRD patterns of compound 3 in different solid states. (b) Powder XRD patterns of compound 4 in different solid states.In conclusion, in this work, four triphenylamine or carbazole-based benzothiadiazole fluorescent molecules were successfully synthesized. The compounds 1, 2, 3 and 4 belonged to the highly solid-state emissive donor–acceptor (D–A) type luminescent molecules. It is noteworthy that the four D–A type luminogens exhibited high-contrast mechanofluorochromic characteristics. Furthermore, the reversibility of their mechanochromic phenomena is good. The results of powder XRD experiments confirmed that this switchable morphology transformation is responsible for the reversible mechanochromic fluorescence characteristics of 1, 2, 3 and 4. This work is valuable for designing high-contrast mechanochromic materials involving red light-emitting feature.  相似文献   

16.
Correction for ‘Effect of molybdenum carbide concentration on the Ni/ZrO2 catalysts for steam-CO2 bi-reforming of methane’ by Weizuo Li et al., RSC Adv., 2015, 5, 100865–100872.

The authors regret that incorrect TEM images were shown in Fig. 7 of the original article. The authors would like to use the following correct TEM images, shown in Fig. 7 below. The authors state that this error has no effect upon the conclusions of the article.Open in a separate windowFig. 7TEM images of coke on the spent Mo2C–Ni/ZrO2 (a) and Ni/ZrO2 (b) catalysts. Inset in (b) is magnified shell-like carbon.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

17.
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   

18.
19.
Correction for ‘An indenocarbazole-based host material for solution processable green phosphorescent organic light emitting diodes’ by Eun Young Park et al., RSC Adv., 2021, 11, 29115–29123. DOI: 10.1039/D1RA04855D.

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. 1HOMO, LUMO distributions and energy level of PCIC predicted through DFT and TD-DFT calculations.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

20.
Correction for ‘Nano- and micro-structural control of WO3 photoelectrode films through aqueous synthesis of WO3·H2O and (NH4)0.33WO3 precursors’ by Hiroaki Uchiyama et al., RSC Adv., 2020, 10, 11444–11449.

The Royal Society of Chemistry regrets that an incorrect version of Fig. 6 was included in the original article. The correct version of Fig. 6 is presented below.Open in a separate windowFig. 6Cross-section SEM images of the WO3 heat-treated films obtained from WO3·H2O (a) and (NH4)0.33WO3 (b) precursor layers on silica glass substrates.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  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