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 125 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.
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.
5.
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   

6.
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   

7.
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   

8.
9.
Formation of a new metastable fcc-MgH2 nanocrystalline phase upon mechanically-induced plastic deformation of MgH2 powders is reported. Our results have shown that cold rolling of mechanically reacted MgH2 powders for 200 passes introduced severe plastic deformation of the powders and led to formation of micro-lathes consisting of γ- and β-MgH2 phases. The cold rolled powders were subjected to different types of defects, exemplified by dislocations, stacking faults, and twinning upon high-energy ball milling. Long term ball milling (50 hours) destabilized β-MgH2 (the most stable phase) and γ-MgH2 (the metastable phase), leading to the formation of a new phase of face centered cubic structure (fcc). The lattice parameter of fcc-MgH2 phase was calculated and found to be 0.4436 nm. This discovered phase possessed high hydrogen storage capacity (6.6 wt%) and revealed excellent desorption kinetics (7 min) at 275 °C. We also demonstrated a cyclic-phase-transformation conducted between these three phases upon changing the ball milling time to 200 hours.

Effect of mechanically-induced cold-rolling followed by high energy ball milling on cyclic phase transformation.

The worldwide interest in MgH2 is attributed to the natural abundance of Mg metal, and its capability to store hydrogen up to 7.60 wt% (0.11 kg H2 L−1).1–3 Among the metal hydride family, MgH2 has been considered as a promising candidate for solid-state hydrogen storage.4–6 Despite the attractive properties of MgH2, and the simplicity for producing the compound on an industrial scale at ambient temperature via a reactive ball milling (RBM) technique,7,8 MgH2 is a very stable compound, and possesses slow kinetics of hydrogenation and dehydrogenation under 300 °C.9 Since the 1990s efforts have been made in order to destabilize MgH2 and improve its hydrogenation and dehydrogenation kinetics upon doping with a long list of catalysts.10–15 Most of the used catalysts showed significant enhancement of the kinetics behavior for MgH2, indexed by a decreasing decomposition temperature and a speed-up of its kinetics behavior.16–20 In spite of the beneficial effects obtained upon adding such foreign catalytic agents, they always lead to a dramatic decrease of the hydrogen storage capacity of MgH2.21,22Apart from doping MgH2 powders with catalysts, it has been experimentally demonstrated by some authors that changing the crystal structure of stable β-tetragonal MgH2 phase to a less stable phase of γ-orthorhombic MgH2 led to improve the gas uptake/release kinetics and decrease the hydrogenation temperature without drastic decreasing of the storage capacity.23–26 The β-to-γ phase transformations can be attained via severe plastic deformation (SPD)27 at ambient temperature by different approaches such as high-energy ball milling (HEBM),28 cold rolling (CRing),29 equal channel angular pressing (ECAP),30 and high pressure torsion (HPT).22,31 A common result of these employed techniques is the formation of nanocrystalline phase along with introducing high intensity defects, leading to increase of grain boundaries density. Presence of these defects in the lattice leads to create nucleation points for hydrogenation, where existence of large number of grain boundaries assists fast diffusion pathways for hydrogen.The present work has been addressed in part to study the effect of HEBM on the structure and decomposition properties of CRed MgH2 powders. Moreover, we aimed to investigate experimentally the possibility of formation a new metastable MgH2 phase rather than the reported g-phase and theoretically calculated δ and ε phases32 upon long term of milling.For the purpose of the present study, 5 g Mg (∼80 μm, 99.8 wt%) powder was balanced inside a He gas atmosphere-glove box and sealed together with fifty hardened steel-balls (11 mm in diameter), using ball-to-powder weight ratio as 40 : 1. The vial was then evacuated to the level of 10−3 bar and then filled with 50 bar of H2. The RBM process was carried out at room temperature, using planetary-type HEBM. After 25 hours (h) of RBM, the vial was open inside the glove box to discharge the powders. The powders were charged and sealed in a stainless steel (SUS304) tube (0.8 cm diameter and 20 cm length) inside the glove box. The tube contained MgH2 powders were severely CRed for different number of passes (1 to 200 passes), using two-drum type manual cold roller (11 cm wide × 5.5 cm rollers diameter). The as-CRed powders for 200 passes were then HEBMed under hydrogen gas for different milling time, in the range between 3 h to 50 h. All the samples were characterized by means of X-ray diffraction (XRD) with Cu radiation, field-emission high-resolution transmission electron microscope (FE-HRTEM) equipped with energy-dispersive X-ray spectroscopy (EDS), field emission scanning electron microscope (FE-SEM)/EDS, and differential scanning calorimeter (DSC). The absorption/desorption kinetics were investigated via Sievert''s method in different temperatures under hydrogen gas pressure in the range between 200 mbar to 8 bar.The XRD patterns of MgH2 powders obtained after 25 h of RBM is shown in Fig. 1a. Bragg peaks corresponding to starting hcp-Mg powders were hardly seen and replaced by sharp diffracted lines related to γ- and β-MgH2 phases, as elucidated in Fig. 1a. The SEM observations indicated the powders tendency to agglomerate, forming large aggregates upon CRing for 25 passes (Fig. 2a). Increasing the CRing passes to 200 times led to grain refinement, as implied by the broadening in the Bragg-peaks presented in Fig. 1b. At this stage, micro-bands with thickness of 143 μm were developed as a result of cold working generated during CRing process, as shown in Fig. 2b. The FE-HRTEM image of as-CRed powders for 200 passes indicated the development of lattice imperfections (e.g. stacking faults and deformation twins). This is implying the mechanically-induced SPD, as elucidated in ESI, Fig. S1.Open in a separate windowFig. 1XRD patterns of MgH2 powders obtained after 25 h of RBMing (a), and then CRed for 200 passes (b). XRD pattern of CRed powders for 200 passes and then HEBMEd for 50 h is presented in (c).Open in a separate windowFig. 2FE-SEM micrographs taken at accelerated voltage of 1 kV of MgH2 powders obtained after 25 h of RBMing and then CRed for (a) 25, and (b) 200 passes.In order to study the effect of HEBM on the stability of cold-rolled MgH2 aggregates, the powders were charged into tool steel vial and ball milled under 50 bar of H2 for different milling time. The XRD pattern of CRed MgH2 powders obtained after 200 CR passes and then HEBMed for 50 h is displayed in Fig. 1c. Obviously, the Bragg peaks corresponding to γ- and β-MgH2 phases were completely vanished and replaced by new Bragg peaks of unreported phase, appeared at scattering angles (2θ) of, 35.034, 40.584, 58.758, 70.115, and 73.868 (Fig. 1c). XRD analysis indicated that this discovered MgH2 phase has face centered cubic structure (fcc) of space group, Fm3̄m(225). The lattice parameter (ao) of this phase, calculated from (111) was 0.44361 nm. The as-synthesized fcc-MgH2 powders had fine spherical particles with sizes distributed in the range between 0.25 μm to 1 μm, as displayed in Fig. 3.Open in a separate windowFig. 3FE-SEM micrograph of MgH2 powders obtained after 25 h of RBM, CRed for 200 passes and finally HEBMed for 50 h.
Peak position, 2θ (degree)Interplanar spacing, d (nm)Miller indexes (h,k,l)Lattice parameter, ao (nm)
35.0340.256121110.44361
40.5840.222292000.44458
58.7580.157132200.44443
70.1150.134213110.44512
73.8680.128292220.44441
Open in a separate windowThe FE-HRTEM image of the as-CRed powders for 200 passes and then HEBMed for 50 h is shown in Fig. 4. The image revealed Moiré fringe image for 3 intimated nanograins. Filtered-fringe images corresponding to Zone I is presented in Fig. 4a. The d-spacing related to fcc-MgH2 (111) was calculated and found to be 0.2559 nm, where the corresponding ao was calculated and found to be 0.4433 nm. These values matches well with the XRD analysis (Fig. 4c, was taken from Zone I and oriented to [001] axis, where the corresponding FFT showed spot-electron diffraction pattern related to fcc-MgH2 (111).Open in a separate windowFig. 4(a) FE-HRTEM image, (b) filtered atomic resolution image of Zone I, and (c) FFT image of MgH2 powders obtained after 25 h of RBM, consequently CRed for 200 passes and finally HEBMed for 50 h. Fig. 5 presents DSC thermograms of MgH2 powders obtained after 25 h of RBM, consequently cold-rolled for 200 passes and then HEBM for different times, as indexed in Fig. 5. The DSC curves for the powders obtained after 25 h of RBM and then cold-rolled only without milling for different passes are shown together in ESI (Fig. S2).Open in a separate windowFig. 5DSC thermograms of MgH2 powders obtained after 25 h of RBM, consequently CRed for 200 passes and final HEBMed for different milling times (6 h to 50 h). The values related to the peak temperatures for each milling time is elucidated in each curve.The decomposition temperature of the powders obtained after 25 h of RBM (before applying CRing process) was 415 °C (Fig. S2), and tended to decrease sharply to 361 °C upon powders CRing for 200 passes, as displayed in Fig. S2. This significant decreasing on the decomposition temperature upon increasing the number of CRing passes was attributed to destabilization of MgH2, yielded from introducing severe plastic deformation to the powders.The XRD pattern for MgH2 powders obtained after 200 CRing passes and then heated in DSC up to 500 °C showed the presence of a single hcp-Mg phase (Fig. S3). The as-HEBMed of the powders CRed for 200 passes revealed single endothermic events related to the decomposition reaction (Fig. 5).Obviously, the peak temperature of the samples obtained after different milling times shifted monotonically to the low temperature side with increasing the HEBMing time (Fig. 5). After 50 h of milling (single metastable fcc-MgH2 phase), the decomposition temperature was reordered to be 222 °C (Fig. 5). This temperature was far below the corresponding decomposition temperature (361 °C, Fig. S3) corresponding to as-CRed powders (γ- + β-MgH2 phases) obtained after 200 passes (Fig. S3).The powders obtained after 50 h of HEBM was annealed in the DSC at 220 °C (just beyond decomposition onset temperature) and at 375 °C (well above the onset decomposition temperature) for 7 min. XRD pattern of the first sample annealed at 220 °C indicated metastable fcc-MgH2-to-stable β-MgH2 phase transformation, as implied by disappearance of fcc-Bragg peaks and the presence of pronounced sharp Bragg lines corresponding to β-MgH2 coexisted with small volume fraction of γ-MgH2 (Fig. 6a).Open in a separate windowFig. 6XRD patterns of the samples obtained after 50 h and then annealed in the DSC for 7 min at (a) 220 °C, and (b) 375 °C for 7 min.FE-HRTEM (Fig. 7a) and NBD (Fig. 7b) analysis of the annealed sample at 220 °C confirmed the presence of γ- and β-MgH2 phases and a complete disappearance of metastable fcc-MgH2, as displayed in Fig. 7a and b, respectively. In contrast, XRD pattern of the sample annealed at 375 °C showed the Bragg peaks of hcp-Mg coexisted with small volume fraction of undecomposed β-MgH2 crystals (Fig. 6b). It can be concluded from the structural and thermal analysis that γ- and β-MgH2 phases tended to transform to a new metastable phase upon CRing for 200 passes followed by HEBMing for 50 h. The obtained metastable fcc-MgH2 phase transformed into more stable phases of γ- and β-MgH2 phases when annealed at 220 °C for 7 min. A complete phase transformation to hcp-Mg was attained upon annealing fcc-MgH2 at 375 °C.Open in a separate windowFig. 7(a) FE-HRTEM, and (b) NBDP of the sample milled for 50 h and then annealed in the DSC at 220 °C for 7 min.Further experiments were conducted to examine the stability of fcc-MgH2 phase against the mechanical deformation generated by HEBMing. The powders obtained after 50 h were continuously milled for 100 h, 150 h, and 200 h. The XRD pattern of the milled powders for 100 h indicated the disappearance of Bragg peaks related fcc-MgH2 phase, which were replaced by diffracted lines related to γ- and β-MgH2 phases as shown in Fig. 8a. This implies the disability of fcc-MgH2 phase to withstand against the shear and impact forces generated by the milling media. Surprisingly, increasing the HEBMing to 150 h led to clear γ- , β-MgH2 to fcc-MgH2 (less stable phase) phase transformation (Fig. 8b). This is indicated by the presence of Bragg peaks related to fcc-MgH2 phase and the disappearance of the diffracted lines corresponding to the MgH2 phases, as displayed in Fig. 8b.Open in a separate windowFig. 8XRD patterns of as-CRed MgH2 powders for 200 passes and then HEBMed for (a) 100 h, (b) 150 h, and (c) 200 h. Fig. 9 presents a schematic illustration of free energy changes of β-γ-fcc MgH2 phases upon increasing HEBM time. Based on the results motivated of the present study, mechanically-induced fcc-MgH2 to γ- and β-MgH2 phase transformation (point 3-to-points 2 + 3) was achieved upon increasing the milling time to 100 h. Introducing severe lattice imperfections to the fcc-phase is proposed to be responsible to insist this metastable phase to gain free energy and hence transformed into a more stable phase, as illustrated in Fig. 9. This phase transformation can be taken place upon annealing the powders via thermally induced phase transformation, as indexed by point 3-to-points 2 + 1 (Fig. 9).Open in a separate windowFig. 9Schematic presentation of free energy changes conducted upon HEBMing CRed MgH2 powders for long term of milling.The obtained γ- and β-MgH2 composite phases were subjected plastic deformation and defects, leading to lose free energy and transformed to the same metastable fcc-MgH2 phase with further milling time (150 h), as denoted by points 1 + 2 to point 3 (Fig. 9). Further increasing the milling time (200 h) led fcc-MgH2 to gain energy and failed into more stable phases of γ- and β-MgH2.Since high-energy ball milling process introduces intensive vacancies, lattice defects, grain boundaries and surfaces, the ball-milled powders can store a large amount of mechanical-strain energy.33,34 Introducing such defects to the crystalline lattice destroys the periodical structure of the stable tetragonal-MgH2 phase, leading to the formation of a less stable phase (fcc-MgH2). Moreover, the present result suggests that the formation enthalpy of the metastable fcc-MgH2 phase is comparable to the γ- and β-MgH2 phases and the energy barrier between these three phases is probably rather low to allow such cyclic-phase transformations. We should emphases that all the results obtained from XRD analysis were used to determine the weight percentage of each phase, using the approaches described by Guan et al.,35 Ma et al.,36 Yu et al.,37 and Cheng et al.38 Fig. 10 displays the dehydrogenation kinetics behaviors of the CRed powder obtained after 200 passes and then HEBMed for 50 h (fcc-MgH2 phase). The measurements were conducted in the temperature range of 175 °C to 275 °C under 200 mbar.Open in a separate windowFig. 10Dehydrogenation kinetics measured for 10 individual samples of MgH2 powders obtained after 25 h of RBM, consequently CRed for 200 passes and then HEBMed for 50 h. The measurements were conducted under 200 mbar H2 (a) without powders activation, and (b) after activated the powders at 300 °C under 10 bar H2.In order to maintain the crystal structure of the powders (fcc-MgH2 phase) obtained after this stage of milling, the desorption kinetics measurements measured without activation for 5 individual samples (Fig. 10a). All samples were successfully desorbed their hydrogen storage with different time scale. Generally, the fcc-MgH2 powders showed advanced dehydrogenation characteristics over γ- and β-MgH2 phases. At 175 °C, the sample desorbed ∼4 wt% H2 within 40 min (Fig. 10a). However, the powders desorbed about 6.5 wt% H2 within 22 min at 200 °C (Fig. 6a). Increasing the applied temperature to 225 °C and 250 °C, improved the dehydrogenation kinetics of fcc-MgH2, indicated by the short time required to desorb 6.6 wt% H2 in 17 min and 10 min, respectively (Fig. 10a).Outstanding desorption kinetics was conducted at 275 °C, when the sample desorbed 6.6 wt% H2 within 7 min, as elucidated in Fig. 10a. We should emphases that the XRD patterns for the examined samples at temperature range laid between 175 °C to 225 °C revealed presence of Bragg peaks corresponding to γ- and β-MgH2 phases. However, the XRD pattern for those samples examined at higher temperatures (250 °C and 275 °C) revealed diffracted lines of hcp-Mg phase. Fig. 10b displays the dehydrogenation kinetics conducted for the same samples shown in 10a, however, they were activated firstly at 300 °C (far above the transformation temperature of fcc-MgH2 to γ- and β-MgH2 phases) under 25 bar H2. Based on this activation step condition, all the samples lost their fcc-structure and completely transformed into γ- and β-MgH2 phases. This led to a significant deduction of dehydrogenation kinetics that become very slow when compared with the same samples shown in Fig. 10a. For example, the sample examined at 175 °C required more than 650 min to desorb less than 5.5 wt% H2 (Fig. 10b). The two samples examined at 200 °C and 225 °C desorbed about 6.4 and 6.6 wt% H2 within 600 and 390 min, respectively (Fig. 10b). Comparing the necessary desorption time (100 min) for the sample examined at 275 °C to release 6.6 wt% H2 (Fig. 10b) with that time required (7 min) for fcc-MgH2 sample (Fig. 10a) to desorb same hydrogen amount at the same temperature, we can realize that the kinetics desorption of fcc-MgH2 phase is 14 time faster than γ- and β-MgH2 phases.  相似文献   

10.
Correction: Flexible,micro-porous chitosan–gelatin hydrogel/nanofibrin composite bandages for treating burn wounds     
P. T. Sudheesh Kumar  G. Praveen  Mincy Raj  K. P. Chennazhi  R. Jayakumar 《RSC advances》2019,9(27):15571
  相似文献   

11.
Correction: Nano- and micro-structural control of WO3 photoelectrode films through aqueous synthesis of WO3·H2O and (NH4)0.33WO3 precursors     
Hiroaki Uchiyama  Yuki Nagayasu 《RSC advances》2020,10(24):14217
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   

12.
Correction: Three-dimensional directional nerve guide conduits fabricated by dopamine-functionalized conductive carbon nanofibre-based nanocomposite ink printing     
Shadi Houshyar  Mamatha M. Pillai  Tanushree Saha  G. Sathish-Kumar  Chaitali Dekiwadia  Satya Ranjan Sarker  R. Sivasubramanian  Robert A. Shanks  Amitava Bhattacharyya 《RSC advances》2020,10(72):43960
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   

13.
High-contrast mechanochromic benzothiadiazole derivatives based on a triphenylamine or a carbazole unit     
Yuan-yuan Zhu  Hong-ying Xia  Li-feng Yao  Dan-ping Huang  Jun-yan Song  Hai-feng He  Liang Shen  Feng Zhao 《RSC advances》2019,9(13):7176
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   

14.
Correction: Green-synthesised cerium oxide nanostructures (CeO2-NS) show excellent biocompatibility for phyto-cultures as compared to silver nanostructures (Ag-NS)     
Qaisar Maqbool 《RSC advances》2022,12(31):19953
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   

15.
Correction: Chemoselective and one-pot synthesis of novel coumarin-based cyclopenta[c]pyrans via base-mediated reaction of α,β-unsaturated coumarins and β-ketodinitriles     
Behnaz Farajpour  Abdolali Alizadeh 《RSC advances》2022,12(14):8547
Correction for ‘Chemoselective and one-pot synthesis of novel coumarin-based cyclopenta[c]pyrans via base-mediated reaction of α,β-unsaturated coumarins and β-ketodinitriles’ by Behnaz Farajpour et al., RSC Adv., 2022, 12, 7262–7267, DOI: 10.1039/D2RA00594H.

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. The CCDC number was also incorrectly cited for this same compound (compound 3d) and should instead be cited as 1970237.Open in a separate windowFig. 1ORTEP diagram of 3d (CCDC 1970237).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

16.
Correction: Noninvasive target CT detection and anti-inflammation of MRSA pneumonia with theranostic silver loaded mesoporous silica     
Hao Zhang  Qingqing Ding  Jing Ding 《RSC advances》2020,10(26):15034
Correction for ‘Noninvasive target CT detection and anti-inflammation of MRSA pneumonia with theranostic silver loaded mesoporous silica’ by Hao Zhang et al., RSC Adv., 2016, 6, 5049–5056.

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. 1(A) SEM image of PEGylated SLS NPs; inset: high-resolution TEM image highlighting the anchored Ag NPs. (B) XPS result of the SLS NPs and silver element. (C) DLS and zeta-potential profiles of the SLS NPs pre- and post-PEGylation.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

17.
Correction: High iodine adsorption performances under off-gas conditions by bismuth-modified ZnAl-LDH layered double hydroxide     
Trinh Dinh Dinh  Dongxiang Zhang  Vu Ngoc Tuan 《RSC advances》2020,10(57):34428
Correction for ‘High iodine adsorption performances under off-gas conditions by bismuth-modified ZnAl-LDH layered double hydroxide’ by Trinh Dinh Dinh et al., RSC Adv., 2020, 10, 14360–14367, DOI: 10.1039/D0RA00501K.

The authors regret that the reference for Fig. 1 was omitted. The reference has been added below.Open in a separate windowFig. 1Schematic of the device for the iodine adsorption experiments in static air.1The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

18.
Correction: A novel biocompatible,simvastatin-loaded,bone-targeting lipid nanocarrier for treating osteoporosis more effectively     
Shan Tao  Shao-qing Chen  Wen-tao Zhou  Fang-ying Yu  Lu Bao  Guo-xi Qiu  Qing Qiao  Fu-qiang Hu  Jian-wei Wang  Hong Yuan 《RSC advances》2020,10(45):26699
Correction for ‘A novel biocompatible, simvastatin-loaded, bone-targeting lipid nanocarrier for treating osteoporosis more effectively’ by Shan Tao et al., RSC Adv., 2020, 10, 20445–20459, DOI: 10.1039/D0RA00685H.

The authors regret that incorrect versions of Fig. 7, ,99 and and1010 were included in the original article. The correct versions of Fig. 7, ,99 and and1010 are presented below.Open in a separate windowFig. 7Histological analysis of organs from all experimental groups. H&E staining of heart, liver, spleen, lung, kidney, indicating the carrier has good biocompatibility. Scale bar = 50 μm.Open in a separate windowFig. 9Alkaline phosphatase (ALP) activity (arrows) and tartrate-resistant acid phosphatase (TRAP) assay results (arrowheads) of bone tissue sections. Scale bar = 50 μm. The ALP activity is much more high in SIM/LNPs and SIM/ASP6-LNPs groups, while the TRAP activity is the opposite.Open in a separate windowFig. 10Histological assessment of bone formation in all experimental groups. (A) HE staining of femur bone. Scale bar = 50 μm. Histology of bone in the all experimental groups shows all ovariectomized groups had a higher amount of adipose tissue than Sham group. The trabecular bone is much more prominent in SIM/LNPs and SIM/ASP6-LNPs groups. (B) Immunohistochemical staining for BMP-2 in typical newly-formed bone tissue (red arrows) and immunohistochemical staining for the osteogenic markers osteopontin (OPN, arrows) and osteocalcin (OCN, arrowheads). Scale bar = 50 μm. The BMP-2, OPN, OCN are much more prominent in SIM/LNPs and SIM/ASP6-LNPs groups.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

19.
Correction: Effect of molybdenum carbide concentration on the Ni/ZrO2 catalysts for steam-CO2 bi-reforming of methane     
Weizuo Li  Zhongkui Zhao  Panpan Ren  Guiru Wang 《RSC advances》2020,10(6):3625
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   

20.
Correction: The corrosion behaviors of multilayer diamond-like carbon coatings: influence of deposition periods and corrosive medium     
Mingjun Cui  Jibin Pu  Guangan Zhang  Liping Wang  Qunji Xue 《RSC advances》2021,11(58):36587
Correction for ‘The corrosion behaviors of multilayer diamond-like carbon coatings: influence of deposition periods and corrosive medium’ by Mingjun Cui et al., RSC Adv., 2016, 6, 28570–28578, DOI: 10.1039/C6RA05527C.

The authors regret that an incorrect version of Fig. 3 was included in the original article. The correct version of Fig. 3 is shown below.Open in a separate windowFig. 3SEM images of surface morphology of multilayer DLC coatings. (a) 5 deposition periods, (b) 12 deposition periods, (c) 15 deposition periods, (d) 20 deposition periods.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