首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Several salicylidene-based colorimetric and fluorimetric anion sensors are known in the literature. However, our 1H-NMR experimental results (in DMSO-d6) showed hydrolysis of imine (–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CH–) bonds in salicylidene-based receptors (SL, CL1 and CL2) in the presence of quaternary ammonium salts (n-Bu4N+) of halides (Cl and Br) and oxo-anions (H2PO4, HSO4 and CH3COO). The mono-salicylidene compound CL1 showed the most extensive –N Created by potrace 1.16, written by Peter Selinger 2001-2019 CH– bond hydrolysis in the presence of anions. In contrast, the di-salicylidene compound CL2 and the tris-salicylidene compound SL showed comparatively slow hydrolysis of –N Created by potrace 1.16, written by Peter Selinger 2001-2019 CH– bonds in the presence of anions. Anion-induced imine bond cleavage in salicylidene compounds could easily be detected in 1H-NMR due to the appearance of the salicylaldehyde –CHO peak at 10.3 ppm which eventually became more intense over time, and the –N Created by potrace 1.16, written by Peter Selinger 2001-2019 CH– peak at 8.9–9.0 ppm became considerably weaker. Furthermore, the formation of the salicylidene O–H⋯X (X = Cl/Br) hydrogen-bonded complex, peak broadening due to proton-exchange processes and keto–enol tautomerism have also been clearly observed in the 1H-NMR experiments. Control 1H-NMR experiments revealed that the presence of moisture in the organic solvents could result in gradual hydrolysis of the salicylidene compounds, and the rate of hydrolysis has further been enhanced significantly in the presence of an anion. Based on 1H-NMR results, we have proposed a general mechanism for the anion-induced hydrolysis of imine bonds in salicylidene-based receptors.

Salicylidene Schiff bases undergo imine bond hydrolysis in the presence of halides and oxo-anions in aprotic media, raising fundamental questions on the applicability of salicylidene-based receptors as anion sensors.  相似文献   

2.
We designed four series of energetic anions by replacing nitro group (NO2) with trinitromethyl group (C(NO2)3) or by inserting N-bridging groups (–NH–, –NH–NH–, –N Created by potrace 1.16, written by Peter Selinger 2001-2019 N–, –N Created by potrace 1.16, written by Peter Selinger 2001-2019 N(O)–) into the bistriazole frameworks. The properties of 40 energetic salts, based on the bistriazole-derived anions and hydroxylammonium cation, were studied by density functional theory (DFT) and volume-based thermodynamics calculations (VBT). It is found that the newly designed energetic salts have good detonation properties due to their larger nitrogen content and better oxygen balance. And one of their corresponding hydroxylammonium salts exhibits better detonation performance (D = 10.06 km s−1 and P = 48.58 GPa) than CL-20 (D = 9.54 km s−1 and P = 43.36 GPa). Moreover, 10 energetic salts not only exhibit excellent energetic properties superior to CL-20, but also have lower sensitivity than CL-20 (h50 = 13.81 cm). In addition, we rationally selected salt B6 from the 10 salts to predict its crystal structure under pressures. By converting energetic molecules with excellent detonation properties into energetic ions, some highly bistriazole-derived energetic salts with both excellent performance and low sensitivity could be developed strategically.

We designed four series of energetic anions by replacing nitro group (NO2) with trinitromethyl group (C(NO2)3) or by inserting N-bridging groups (–NH–, –NH–NH–, –N Created by potrace 1.16, written by Peter Selinger 2001-2019 N–, –N Created by potrace 1.16, written by Peter Selinger 2001-2019 N(O)–) into the bistriazole frameworks.  相似文献   

3.
Correction for ‘Mid-infrared spectroscopy and microscopy of subcellular structures in eukaryotic cells with atomic force microscopy – infrared spectroscopy’ by Luca Quaroni et al., RSC Adv., 2018, 8, 2786–2794.

In the article the assignment of two IR absorption bands, at 3010 and at 3070 cm−1 has been confused by us in the text, resulting in two incorrect statements. The misstatement does not change any of the conclusions of the work, and can be corrected by restating the following sentences.The following sentence (p. 2790, column 1, line 10):A weak but sharp band can be seen at 3010 cm−1, corresponding to the stretching mode of C–H bonds on unsaturated C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bonds in acyl chains.must be corrected to:A weak but sharp band can be seen at 3010 cm−1, corresponding to the stretching mode of C–H3 bonds on choline headgroups.In addition, the following sentence (p. 2792, column 1, line 25):Observation of a sharp band at 3010 cm−1 indicates that at least part of the acyl chains have unsaturated C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bonds.must be changed to:Observation of a band at 3070 cm−1, corresponding to the stretching mode of C–H bonds on unsaturated C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bonds in acyl chains, indicates that at least part of the acyl chains have unsaturated C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bonds.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Bimolecular nucleophilic substitution (SN2) is a fundamental reaction that has been widely studied. So far, the nucleophiles are mainly anionic species in SN2 reactions. In this study, we use density functional theory calculations to assess the mechanisms of substitution of carbonyl, imidoyl, and vinyl compounds with a neutral nucleophile, pyridine. Charge decomposition analysis is performed to explore the main components of the transition state''s LUMO. For reactions of imidoyl or carbonyl compounds with pyridine or Cl, the LUMOs of the transition states are composed of mixed orbitals originating from the nucleophile and the substrate. Considering the unique mixed nature of the orbitals, the reaction mode is termed SNm (m means mix). Moreover, the main components of the transition state''s LUMO are pure σ*C–Cl MO in the reactions of H2C Created by potrace 1.16, written by Peter Selinger 2001-2019 CHCl with pyridine or Cl. Computations were also performed for RY Created by potrace 1.16, written by Peter Selinger 2001-2019 CHX substrates with different X and Y groups (X= Cl, Br, or F; Y = O, N, or C).

The nucleophilic substitution of carbonyl, imidoyl, and vinyl carbon centers with pyridine or halides is investigated in this paper.  相似文献   

5.
A-234, [EtO–P( Created by potrace 1.16, written by Peter Selinger 2001-2019 O)(F)–N Created by potrace 1.16, written by Peter Selinger 2001-2019 C(Me)–N(Et)2], is the suspected A-type nerve agent used in the Skripal attack on the 4th of March 2018. Studies related to the structure and reactivity of this compound are limited. We, therefore, aimed at understanding the underlying hydrolysis mechanism of A-234 within the DFT framework. The attack of the water molecule can occur at the phosphinate and acetoamidine reactive centres. Our theoretical findings indicate that the hydrolysis at the acetoamidine centre is thermodynamically favoured compared to the hydrolysis at the phosphinate centre. The hydrolysis at the acetoamidine moiety may proceed via two pathways, depending on the nitrogen atom participating in the hydrolysis. The main pathway consists of four distinct channels to reach the final product, with the concerted 1,3-proton shift favoured kinetically and thermodynamically in the gas phase and water as solvent. The results are in good agreement with the literature, although some differences in the reaction mechanism were observed.

A theoretical study of the hydrolysis mechanism of A-234 [EtO–P( Created by potrace 1.16, written by Peter Selinger 2001-2019 O)(F)–N Created by potrace 1.16, written by Peter Selinger 2001-2019 C(Me)–N(Et)2]; the suspected novichok agent in the Skripal attack.  相似文献   

6.
In this study, novel green nano-zerovalent iron (G-NZVI) is synthesized for the first time using onion peel extract for the prevention of rapid surface oxidation and the enhancement of particle dispersibility with a high reductive capacity. The results from various surface analyses revealed that the spherical shape of G-NZVI was fully covered by the onion peel extract composed of polyphenolic compounds with C Created by potrace 1.16, written by Peter Selinger 2001-2019 C–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C unsaturated carbon, C Created by potrace 1.16, written by Peter Selinger 2001-2019 C, C–O, and O–H bonds, resulting in high mobility during column chromatography. Furthermore, the obtained G-NZVI showed the complete removal of 50 mg L−1 of bromate (BrO3) in 2 min under both aerobic (k = 4.42 min−1) and anaerobic conditions (k = 4.50 min−1), showing that G-NZVI had outstanding oxidation resistance compared to that of bare NZVI. Moreover, the observed performance of G-NZVI showed that it was much more reactive than other well-known reductants (e.g., Fe and Co metal organic frameworks), regardless of whether aerobic or anaerobic conditions were used. The effects of G-NZVI loading, the BrO3 concentration, and pH on the BrO3 removal kinetics using G-NZVI were also investigated in this study. The results provide the novel insight that organic onion peel waste can be reused to synthesize highly reactive anti-oxidative nanoparticles for the treatment of inorganic chemical species and heavy metals in water and wastewater.

In this study, novel green nano-zerovalent iron (G-NZVI) is synthesized for the first time using onion peel extract for the prevention of rapid surface oxidation and the enhancement of particle dispersibility with a high reductive capacity.  相似文献   

7.
Recently, inexpensive and readily available tBuOK has seen widespread use in transition-metal-free reactions. Herein, we report the use of tBuOK for S–S, S–Se, N Created by potrace 1.16, written by Peter Selinger 2001-2019 N and C Created by potrace 1.16, written by Peter Selinger 2001-2019 N bond formations, which significantly extends the scope of tBuOK in chemical synthesis. Compared with traditional methods, we have realized mild and general methods for disulfide, azobenzenes imine etc. synthesis.

Inexpensive and readily available tBuOK can trigger a series of bond formation reactions, including S–S, S–Se, Se–Se, and N Created by potrace 1.16, written by Peter Selinger 2001-2019 N and C Created by potrace 1.16, written by Peter Selinger 2001-2019 N bonds.  相似文献   

8.
A novel DPyDB-C Created by potrace 1.16, written by Peter Selinger 2001-2019 N-18C6 compound was synthesised by linking a pyrene moiety to each phenyl group of dibenzo-18-crown-6-ether, the crown ether, through –HC Created by potrace 1.16, written by Peter Selinger 2001-2019 N– bonds and characterized by FTIR, 1H-NMR, 13C-NMR, TGA, and DSC techniques. The quantitative 13C-NMR analysis revealed the presence of two position isomers. The electronic structure of the DPyDB-C Created by potrace 1.16, written by Peter Selinger 2001-2019 N-18C6 molecule was characterized by UV-vis and fluorescence spectroscopies in four solvents with different polarities to observe particular behavior of isomers, as well as to demonstrate a possible non-bonding chemical association (such as ground- and excited-state associations, namely, to probe if there were forming dimers/excimers). The interpretation of the electronic structure was realized through QM calculations. The TD-CAM-B3LYP functional, at the 6-311+G(d,p) basis set, indicated the presence of predominant π → π* and mixed π → π* + n → π* transitions, in line with the UV-vis experimental data. Even though DPyDB-C Created by potrace 1.16, written by Peter Selinger 2001-2019 N-18C6 computational studies revealed a π-extended conjugation effect with predominantly π → π* transitions, thorough fluorescence analysis was observed a weak emission, as an effect of PET and ACQ. In particular, the WAXD analysis of powder and thin films obtained from n-hexane, 1,2-dichloroethane, and ethanol indicated an amorphous organization, whereas from toluene a smectic ordering was obtained. These results were correlated with MD simulation, and it was observed that the molecular geometry of DPyDB-C Created by potrace 1.16, written by Peter Selinger 2001-2019 N-18C6 molecule played a defining role in the pyrene stacking arrangement.

Herein, we report the formation of a potential supramolecular arrangement mediated by inter- and intra-molecular interactions between di-iminopyrene-dibenzo-18-crown-6-ether molecules.  相似文献   

9.
The formation of polycyclic aromatic hydrocarbons (PAHs) on the C11H11 potential energy surface involved in the reactions of a phenyl radical (C6H5) with cis-3-penten-1-yne (cis-C1H Created by potrace 1.16, written by Peter Selinger 2001-2019 C2–C3H Created by potrace 1.16, written by Peter Selinger 2001-2019 C4H–C5H3, referred to as C5H6) and its three radicals (CH Created by potrace 1.16, written by Peter Selinger 2001-2019 C–Ċ Created by potrace 1.16, written by Peter Selinger 2001-2019 CH–CH3, CH Created by potrace 1.16, written by Peter Selinger 2001-2019 C–CH Created by potrace 1.16, written by Peter Selinger 2001-2019 Ċ–CH3, and cis-CH Created by potrace 1.16, written by Peter Selinger 2001-2019 C–CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CH–ĊH2, referred to as the C3-, C4-, and C5-radicals with the same chemical components, C5H5) assisted by H atoms is investigated by performing combined density functional theory (DFT) and ab initio calculations. Five potential pathways for the formation of PAHs have been explored in detail: Pathways I–II correspond to the reaction of C6H5 with C5H6 at the C1 and C2 position, and Pathways III–V involve the reaction of C6H5 with the C3-, C4-, and C5-radicals with the assistance of H atoms. The initial association of C6H5 with C5H6 or C5H5 is found to be highly exothermic with only minor barriers (1.4–7.1 kcal mol−1), which provides a large driving force for the formation of PAHs. The hydrogen atom is beneficial for the ring enlargement and ring formation processes. The present calculations predict 9 potential PAHs, six (CS6, CS10, CS13, CS26, CS28 and CS29) of which are indicated to be energetically more favorable along Pathways I, III, IV and V at low temperature. The calculated barriers for the formation of these PAHs are around 19.2–38.0 kcal mol−1. All PAHs products could be formed at flame temperature, for the medium barriers are easily overcome in various flame conditions. The theoretical results supplement the PAH formation pathway and provide help to understand PAH growth mechanism.

The formation of PAHs within 4-, 5-, 6- and 7-membered rings on the C6H5 + C5H6 potential energy surface.  相似文献   

10.
The catalytic reduction of NO with NH3 (NH3-SCR) on phosphorus-doped carbon aerogels (P-CAs) was studied in the temperature range of 100–200 °C. The P-CAs were prepared by a one-pot sol–gel method by using phosphoric acid as a phosphorus source followed by carbonization at 600–900 °C. A correlation between catalytic activity and surface P content is observed. The P-CA-800vac sample obtained via carbonization at 800 °C and vacuum treatment at 380 °C shows the highest NO conversion of 45.6–76.8% at 100–200 °C under a gas hourly space velocity of 500 h−1 for the inlet gas mixture of 500 ppm NO, 500 ppm NH3 and 5.0 vol% O2. The coexistence of NH3 and O2 is essential for the high conversion of NO on the P-CA carbon catalysts, which can decrease the spillover of NO2 and N2O. The main Brønsted acid sites derived from P-doping and contributed by the C–OH group at edges of carbon sheets are beneficial for NH3 adsorption. In addition, the C3–P Created by potrace 1.16, written by Peter Selinger 2001-2019 O configuration seems to have the most active sites for favorable adsorption and dissociation of O2 and facilitates the formation of NO2. Therefore, the simultaneous presence of acidic groups for NH3 adsorption and the C3–P Created by potrace 1.16, written by Peter Selinger 2001-2019 O active sites for NO2 generation due to the activation of O2 molecules is likely responsible for the significant increase in the NH3-SCR activity over the P–CAs. The transformation of C3–P Created by potrace 1.16, written by Peter Selinger 2001-2019 O to C–O–P functional groups after the reaction is found, which could be assigned to the oxidation of C3–P Created by potrace 1.16, written by Peter Selinger 2001-2019 O by the dissociated O*, resulting in an apparent decrease of catalytic activity for P-CAs. The C–O–P based functional groups are also active in the NH3-SCR reaction.

P species can effectively enhance the catalytic activity of carbon aerogels for NO reduction at low temperature.  相似文献   

11.
Onion-like graphitic structures are of great importance in different fields. Pentagons, heptagons, and octagons are essential features of onion-like graphitic structures that could generate important properties for diverse applications such as anodes in Li metal batteries or the oxygen reduction reaction. These carbon nanomaterials are fullerenes organized in a nested fashion. In this work, we produced graphitic nano onion-like structures containing phosphorus and nitrogen (NP-GNOs), using the aerosol assisted chemical vapor deposition method. The NP-GNOs were grown at high temperature (1020 °C) using ferrocene, trioctylphosphine oxide, benzylamine, and tetrahydrofuran precursors. The morphology, structure, composition, and surface chemistry of NP-GNOs were characterized using different techniques. The NP-GNOs showed diameters of 110–780 nm with Fe-based nanoparticles inside. Thermogravimetric analysis showed that NP-GNOs are thermally stable with an oxidation temperature of 724 °C. The surface chemistry analysis by FTIR and XPS revealed phosphorus–nitrogen codoping, and several functionalities containing C–H, N–H, P–H, P–O, P Created by potrace 1.16, written by Peter Selinger 2001-2019 O, C Created by potrace 1.16, written by Peter Selinger 2001-2019 O, and C–O bonds. We show density functional theory calculations of phosphorus–nitrogen doping and functionalized C240 fullerenes. We present the optimized structures, electronic density of states, HOMO, and LUMO wave functions for P-doped and OH-functionalized fullerenes. The P Created by potrace 1.16, written by Peter Selinger 2001-2019 O and P–O bonds attributed to phosphates or hydroxyl groups attached to phosphorus atoms doping the NP-GNOs could be useful in improving supercapacitor function.

Nitrogen–phosphorus doped graphitic nano onion-like structures.  相似文献   

12.
This study investigated the effects of partially replacing wheat flour with flaxseed flour (FF) on the quality parameters of Chinese Steamed Bread (CSB). FF was utilized as a functional ingredient of CSB at varying levels. The pasting properties of flour blends, the rheological and microstructural characteristics of dough, the textural and quality characteristics and functional group structure of CSB were analyzed. Results showed that FF addition influenced the pasting characteristics of wheat flour by decreasing the final viscosity, breakdown and setback values, but had little effect on the rheological properties of the dough. The microstructure of the dough indicated that the disruption degree of the gluten matrix increased with the increase of FF. Besides, FF addition increased the hardness and chewiness of CSB, while decreasing the cohesiveness and springiness. Additional characteristic peaks were observed at 1745, 2854, and 3006 cm−1 and associated with –C Created by potrace 1.16, written by Peter Selinger 2001-2019 O, –CH2, and cis-C Created by potrace 1.16, written by Peter Selinger 2001-2019 CH bond stretching vibrations of flaxseed. Results suggested 12% FF exhibited the best acceptability.

Chinese steamed bread supplemented with flaxseed flour can be recommended as a dietary product with health benefits.  相似文献   

13.
t-Butyl hydroperoxide-initiated cycloterpolymerization of diallylaminoaspartic acid hydrochloride [(CH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 CHCH2)2NH+CH(CO2H)CH2CO2H Cl] (I), maleic acid (HO2CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CHCO2H) (II) and cross-linker tetraallylhexane-1,6-diamine dihydrochloride [(CH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 CHCH2)2NH+(CH2)6NH+ (CH2CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2)2 2Cl] (III) afforded a new pH-responsive resin (IV), loaded with four CO2H and a chelating motif of NH+⋯CO2 in each repeating unit. The removal of cationic methylene blue (MB) (3000 ppm) at pH 7.25 and Pb(ii) (200 ppm) at pH 6 by IV at 298, 313, and 328 K followed second-order kinetics with Ea of 33.4 and 40.7 kJ mol−1, respectively. Both MB and Pb(ii) were removed fast, accounting for 97.7% removal of MB within 15 min at 313 K and 94% of Pb(ii) removal within 1 min. The super-adsorbent resin gave respective qmax values of 2609 mg g−1 and 873 mg g−1 for MB and Pb(ii). IV was also found to trap anionic dyes; it removed 91% Eriochrome Black T (EBT) from its 50 ppm solutions at pH 2. The resin was found to be effective in reducing priority metal contaminants (like Cr, Hg, Pb) in industrial wastewater to sub-ppb levels. The synthesis of the recyclable resin can be easily scaled up from inexpensive starting materials. The resin has been found to be better than many recently reported sorbents.

Cycloterpolymerization of diallylaminoaspartic acid hydrochloride (I), maleic acid (II) and a cross-linker (III) afforded a new pH-responsive resin (IV), loaded with four CO2H and a chelating motif of NH+⋯CO2 in each repeating unit.  相似文献   

14.
The reaction of the trimetallic clusters [H2Os3(CO)10] and [Ru3(CO)10L2] (L = CO, MeCN) with 2-ethynylpyridine has been investigated. Treatment of [H2Os3(CO)10] with excess 2-ethynylpyridine affords [HOs3(CO)10(μ-C5H4NCH=CH)] (1), [HOs3(CO)93-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2)] (2), [HOs3(CO)93-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CCO2)] (3), and [HOs3(CO)10(μ-CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CHC5H4N)] (4) formed through either the direct addition of the Os–H bond across the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond or acetylenic C–H bond activation of the 2-ethynylpyridine substrate. In contrast, the dominant pathway for the reaction between [Ru3(CO)12] and 2-ethynylpyridine is C–C bond coupling of the alkyne moiety to furnish the triruthenium clusters [Ru3(CO)7(μ-CO){μ3-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CHC(C5H4N) Created by potrace 1.16, written by Peter Selinger 2001-2019 CH}] (5) and [Ru3(CO)7(μ-CO){μ3-C5H4NCCHC(C5H4N)CHCHC(C5H4N)}] (6). Cluster 5 contains a metalated 2-pyridyl-substituted diene while 6 exhibits a metalated 2-pyridyl-substituted triene moiety. The functionalized pyridyl ligands in 5 and 6 derive via the formal C–C bond coupling of two and three 2-ethynylpyridine molecules, respectively, and 5 and 6 provide evidence for facile alkyne insertion at ruthenium clusters. The solid-state structures of 1–3, 5, and 6 have been determined by single-crystal X-ray diffraction analyses, and the bonding in the product clusters has been investigated by DFT. In the case of 1, the computational results reveal a rare thermodynamic preference for a terminal hydride ligand as opposed to a hydride-bridged Os–Os bond (3c,2e Os–Os–H bond).

The reactivity of 2-ethynylpyridine at low-valent triosmium and triruthenium centers has been investigated.  相似文献   

15.
A product study of the reactions of (E/Z)-1,2,3,3,3-pentafluoropropene ((E/Z)-CF3CF Created by potrace 1.16, written by Peter Selinger 2001-2019 CHF) and hexafluoroisobutylene ((CF3)2C Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2) initiated by Cl atoms were developed at 298 ± 2 K and atmospheric pressure. The experiments were carried out in a 1080 L quartz-glass environmental chamber coupled via in situ FTIR spectroscopy to monitor the reactants and products. The main products observed and their yields were as follows: CF3C(O)F (106 ± 9)% with HC(O)F (100 ± 8)% as a co-product for (E/Z)-CF3CF Created by potrace 1.16, written by Peter Selinger 2001-2019 CHF, and CF3C(O)CF3 (94 ± 5)% with HC(O)Cl (90 ± 7)% as a co-product for (CF3)2C Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2. Atmospheric implications of the end-product degradation are assessed in terms of their impact on ecosystems to help environmental policymakers consider HFOs as acceptable replacements.

An experimental product distribution study and the atmospheric implications of the reactions of Cl with two fluorinated alkenes is provided.  相似文献   

16.
Boron-doped graphene samples (BGs) with tunable boron content of 0–2.90 at% were synthesized and directly used in the gas-phase oxidation of benzyl alcohol to benzaldehyde, and showed excellent performance. XPS results indicated that the graphitic sp2 B species (BC3) is the mainly boron dopant species incorporated in the graphene lattice, which could significantly improve the content of ketone carbonyl groups (C Created by potrace 1.16, written by Peter Selinger 2001-2019 O) on the graphene. For instance, the contents of C Created by potrace 1.16, written by Peter Selinger 2001-2019 O jumped from 1.93 to 4.19 at% while BC3 doped into the graphene lattice was only 0.35 at%. The C Created by potrace 1.16, written by Peter Selinger 2001-2019 O is the active site of catalytic reaction, so BG has significantly improved catalytic activity. Compared to the un-doped graphene (G), the conversion of benzyl alcohol over BGs increased 2.35 times and the selectivity of benzaldehyde increased from 77.3% to 99.2%. Aerobic–anaerobic exchange experiments revealed that the superior catalytic performance of BG was achieved only under aerobic conditions. The study of the boron-doped carbocatalyst may also provide guidance for the design of surface modified carbon-based catalysts for the selective oxidation dehydrogenation of alcohols by regulating doping elements and their types.

Boron doped graphene for the oxidative dehydrogenation reactions.  相似文献   

17.
The correlations of the 1H NMR, 13C NMR and FT-IR spectral data from the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups in the alkyl carbamates and esters of homologous alcohols reveal R-group-dependent negative charge stabilization at the carbonyl oxygen and their donation to generic acceptors at Cα of even alkyl alcohols (R), which explains several of their apparently anomalous properties.

NMR, FT-IR spectral correlations of the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups in carbamates and esters of homologous alcohols (R) reveal R-group-dependent negative charge stabilization at the carbonyl oxygen and its donation to generic acceptors at Cα of even alkyl R.

Electronic interactions at the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups in carbamates and esters are less understood than that at the R–N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups in amides. Carbamates in which both these groups are fused at the carbonyl C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond, have several ester-like rather than amide-like features1,2 including comparable C–O bond lengths in crystals,3–6 favourable interactions between the dipoles of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 O and O–R bonds in the predominant s-trans rotamers of the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups and unfavourable lone pair repulsion between the two ester oxygens in their trace s-cis rotamers7–10 (Fig. 1a). However, several anomalous properties of carbamates and esters cannot be explained by these dipole and repulsive forces alone. For example, despite the electronic resonance at the O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O group in esters and additionally at the N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O group in carbamates (both of which augment the electron density at their carbonyl oxygens), their carbonyls have remarkably lower basicities than the carbonyls of amides11 and even those of ketones, which lack such resonance effects.12 The barriers to the C–N bond rotation for carbamates are 3–4 kcal mol−1 lower than those for analogous amides.13,14 Carbamate and ester carbonyl oxygens also have poor hydrogen bond accepting propensities compared to amides.15–20Open in a separate windowFig. 1(a) Dipole–dipole, electronic and steric interactions governing the rotamer stabilities at R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O frameworks in carbamates and esters; (b) earlier report of n → π* O⋯Cα interaction in phenyl carbamates; (c) current finding of generic donor → acceptor interactions: (i) charge → HCR* (ii) charge → σ*; (d) list of carbamates (1–8), esters (9–16) and alcohols (17–20) investigated. HCR is hyperconjugative resonance along the Cα–Cβ bond of alcohol groups in carbamates and esters.Of particular current interest is the apparent anomaly that in the crystals of phenyl carbamates, the phenyl ring plane is oriented perpendicular to the carbamate plane21,22 (Fig. 1b). The presence of an extraordinary n → π* orbital overlap interaction (O⋯Cαphenyl) between the lone-pair electrons on the carbonyl (C Created by potrace 1.16, written by Peter Selinger 2001-2019 O) oxygen and the π* orbital of the phenyl ring has been proposed as the stabilizing force for this conformer based on natural bond orbital (NBO) analysis.22 This is interesting for several reasons: (i) despite a decrease in the stretching frequency of C Created by potrace 1.16, written by Peter Selinger 2001-2019 O in the FT-IR spectrum for this conformer, which indicates a decrease in the bond order and concomitant improvement in the charge density at the carbonyl oxygen, a discussion on the origin and role of such charge on this O⋯Cαalcohol interaction is lacking. Rather, the interaction is assumed to originate from the lone pair on oxygen. (ii) Our investigation of CCDC revealed that the O⋯Cαalcohol distances at the Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups are quite non-variant (2.62–2.81 Å) for both aliphatic and aromatic carbamates and esters,3–6,23–27 notwithstanding the variations in the structure resolutions. These invoked the following questions: is the O⋯Cα interaction specific to the π* acceptors at Cαalcohol? Can any antibonding orbital at Cαalcohol act as an acceptor of the electrons from carbonyl oxygen? What is the generic nature and role of this O⋯Cα interaction in the Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups? Here, we present the first spectral evidence for the generic nature of the O⋯Cα interactions even with non-π* orbital acceptors (like σ* and hyperconjugative resonance bonds) at Cα in the Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups (Fig. 1c) of a variety of model homologous aliphatic carbamates and esters (Fig. 1d). The charge at the carbonyl oxygen is interdependent on the alkoxy groups and forms charge → acceptor O⋯Cα interactions, which influence the rotational states of the O–Cα bonds in the carbamates and esters.To explore the possibility of the O⋯Cα interactions at Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O in aliphatic carbamates and esters, we investigated the 1H NMR, 13C NMR and FT-IR spectra of the secondary and tertiary carbamates (1–4 and 5–8), acetates (9–12) and benzoates (13–16) of homologous aliphatic alcohols (H–CαH2–OH, MeOH; CH3–CαH2–OH, ethanol; (CH3)2–CαH–OH, isopropanol; and (CH3)3–Cα–OH, tert-butanol, 17–20) (
13C NMR (δ ppm) ν (cm−1) 1H NMR (δ ppm)
O–CαOCα–CβO Created by potrace 1.16, written by Peter Selinger 2001-2019 CC Created by potrace 1.16, written by Peter Selinger 2001-2019 O*OCα–HαOCα–Cβ–Hβ
152.1 (50.4)NA (NA)155.516853.69 (3.68)NA (NA)
552.3NA154.117053.78NA
951.6NA171.517483.67NA
1351.5NA165.917253.92NA
260.5 (58.3)14.6 (18.4)155.016794.13 (3.72)1.26 (1.24)
661.114.5153.617014.231.31
1060.514.2171.417404.121.25
1460.814.3166.517204.331.33
367.6 (64.5)22.1 (25.1)154.716744.92 (4.03)1.23 (1.20)
768.722.1153.75.051.30
1167.621.8170.617364.991.23
1568.321.9166.117165.241.35
478.7 (69.1)28.4 (31.2)154.51683NA (NA)1.46 (1.27)
880.428.3152.71689NA1.51
1280.128.1170.41738NA1.45
1680.928.2165.8NA1.58
Open in a separate windowa*FT-IR stretching frequencies; values in parentheses are for corresponding alcohols; NA means not applicable; “—” means unambiguous data could not be obtained.The 13C NMR signals (28,29 undergo a large downfield shift (by 18.7 ppm) incrementally as the number of methyl (CβH3) substituents on Cα increases from zero to three (Fig. 2a). There is a simultaneous downfield shift in the 1H NMR signal of the corresponding Hα by 0.55 ppm from methyl to isopropyl alcohol. Such shifts are contrary to what is expected from the increased positive induction on Cα–Hα by the CβH3 substituents. They are rather due to the positive charge at Cα stemming from the polarization of the Cα–O bond, which is stabilized through the hyperconjugation effect (and increasingly so) as the number of methyl substituents on Cα increases. Hence, the Cα–Cβ σ-bonds of alcohols have additional bonding from such hyperconjugative resonance (HCR).Open in a separate windowFig. 2(a and b) Incremental downfield shifts in the 13C and 1H signals of Hα and Cα in carbamates (1–8) and esters (9–16) relative to the corresponding homologous alcohols (17–20), indicating the stabilization of positive charge polarization at Cα by hyperconjugative resonance from the CβH3 groups; (c and d) inverse correlation between 13C and 1H NMR signals of Cα and Hα and FT-IR stretching frequencies at C Created by potrace 1.16, written by Peter Selinger 2001-2019 O in carbamates (1–7) and esters (9–15) with increasing number of methyl (CβH3) substituents (0–2) on Cα (for isopropyl N-phenyl carbamate (7), the FT-IR value is not inserted); (e) upfield shifts in the Cβ signals of carbamates (1–8) and esters (9–16) compared to corresponding alcohols (17–20), indicating electronic back donation towards Cβ; (f) shifts in the Hβ signals of carbamates (1–8) and esters (9–16) compared to corresponding alcohols (17–20), indicating the dominance of charge → HCR* interactions. The lines merely indicate the trends.In the corresponding carbamates (1–4, 5–8 (ref. 30–32)) and esters (9–12, 13–16 (ref. 33–35)), there are much larger downfield shifts for Cα (by 26.6 ppm for 1–4; 28.1 ppm for 5–8; 28.5 ppm for 9–12; and 29.4 ppm for 13–16) and Hα (by 1.23 ppm for 1–3; 1.27 ppm for 5–7; 1.32 ppm 9–11; and 1.32 ppm 13–15) compared to that for alcohols28,29 (Fig. 2a and b). These shifts are also incremental on increasing the CβH3 substitution on Cα but show steeper increase compared to that for alcohols (see ESI). This further substantiates the hyperconjugative stabilization of greater positive polarization at Cα of Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O, whose oxygen exists as an oxonium ion in the bipolar resonance form (Cα–O+ Created by potrace 1.16, written by Peter Selinger 2001-2019 C–O) (Fig. 3). Note that if merely the electron-withdrawing induction effect of the carbonyl group was influencing the chemical shifts of Cα and Hα, a constant downfield shift would have been observed on increasing the CβH3 substitution at Cα in either carbamates and esters compared to that for alcohols and not such incremental (steeper) downfield shifts.Open in a separate windowFig. 3The hyperconjugation effect stabilizing the polarization at Cα of alkoxy groups (a) and the O⋯Cα electronic interactions in the s-trans conformer at the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups in (b) carbamates and (c) esters of homologous alcohols, as observed by the correlation of their 1H NMR, 13C NMR and FT-IR data in the current work.The remarkable uniformity in the trends of such increments in the Cα and Hα chemical shifts for the R–O–C(R′) Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups (R′ represents primary and secondary amine, R stands for alkyl and aryl groups) of 1–16 reveals that this stabilization of the bipolar R–O+ Created by potrace 1.16, written by Peter Selinger 2001-2019 C(R′)–O intermediates by the hyperconjugative effect from the R group is largely independent of the acyl substituent (R′) effects and is only slightly perturbed by the cross-conjugation from the nitrogen in R′. The longer C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond lengths in secondary (1.21 ± 0.01 Å)36–40 and tertiary (1.23 ± 0.02 Å)5,6,24,25,41 carbamates compared to that in acetates (1.20 ± 0.02 Å)3,4,23,27 and benzoates (1.21 ± 0.01 Å)42–46 of the homologous alcohols in the solid state (ESI), which are reflected in their stretching frequencies (1697 ± 8 cm−1, 1679 ± 3 cm−1, 1740.5 ± 7.5 cm−1 and 1723 ± 7 cm−1, respectively) in the FT-IR spectra,30,32,47–50 however, clearly indicate the additional electronic resonance along the N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O framework. As a result, the 13C nuclear resonance for C Created by potrace 1.16, written by Peter Selinger 2001-2019 O in carbamates shifts upfield compared to that for the esters. However, since the shift is constant and independent of the number of methyl substituents on Cα, the mixing of nitrogen lone pair with the carbonyl π-cloud does not perturb the oxonium charge state or the concomitant positive charge at Cα. The resonance at O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O is hence quite strong. The shortening of the O–C single bond in the carbamates and esters, noted from the diffraction data, evidences such strong resonance at O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O. In fact, the lowering of the rotational energy barrier along the C–N bond in carbamates than that in the corresponding amides by 3–4 kcal mol−1,13,14 indicates the overwhelming influence of strong resonance at O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O in diminishing the resonance at N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O.Interestingly, the FT-IR C Created by potrace 1.16, written by Peter Selinger 2001-2019 O stretching frequencies in the carbamates and esters consistently show an inverse correlation with the downfield shifts at their alcohol Cα (Fig. 2c and d), indicating interactive dependence between the negative charge at the carbonyl oxygen and the positive polarization at Cα, an O⋯Cα interaction that remarkably improves on increasing the number of CβH3 substituents on Cα.Such O⋯Cα interactions are clearly evident in the crystal structures of the carbamates5,6,24,25,36–41 and esters3,4,23,27,42–46 (1–16) (ESI), especially when there are methyl (CβH3) substituents on Cα: (i) the Cα–Cβ bond is consistently oriented antiperiplanar to C Created by potrace 1.16, written by Peter Selinger 2001-2019 O with an angle of incidence of 160 ± 5 deg for oxygen. Note that in addition to the σ-bond along Cα–Cβ, the current results have established the presence of HCR bonding. Hence, the O⋯Cα charge donation can be either to σ* or HCR* along the Cα–Cβ bond. (ii) Only in methyl carbamates and esters (which lack Cβ), such antiperiplanar orientation of oxygen (159 ± 5 deg) is observed with a Cα–Hα σ-bond. (iii) The O⋯Cα distances are comparable (2.73 ± 0.02 Å) (within errors of estimation). The distances however increase slightly from 2.62 ± 0.01 Å to 2.81 ± 0.00 Å for methyl to t-butyl carbamates and esters due to increased O⋯Cα steric repulsions. (iv) All the atoms of the O Created by potrace 1.16, written by Peter Selinger 2001-2019 C–O–Cα–Cβ group are in plane, indicating lack of distortions due to the steric clashes between the C Created by potrace 1.16, written by Peter Selinger 2001-2019 O and R groups. There is slight deviation of the C–O–Cα–Cβ torsion from planarity (by ∼25 deg) only in the isopropyl carbamates and esters where the carbonyl oxygen is asymmetrically staggered between the small Hα and the bulky CβH3 of the isopropyl group. This slight deviation is reflected in the consistent minor deviations of their NMR and FT-IR data from the trends of their corresponding homologues. (v) The bond angles at the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O framework are comparable in the carbamates and esters for all the homologous R groups.The antiperiplanar orientation of C Created by potrace 1.16, written by Peter Selinger 2001-2019 O to the Cα–Cβ bond in the carbamates and esters in the solid state and the O⋯Cα interactions suggested by spectral indicators were consistent with each other. Interestingly, the Cβ signals in the 13C NMR spectra of the carbamates and esters identically shifted upfield on increasing the CβH3 substitution at Cα (Fig. 2e) compared to that for alcohols; this was in contrast to the Cα signals, which shifted down-field. However, the Hβ signals in the 1H NMR spectra showed downfield shifts of decreasing steepness with decrease in the electronic charge at the carbonyl oxygen in the order alkyl benzoates > N-phenyl carbamate > acetates/pyrrolidine carbamates (Fig. 2f). These indicate the predominance of the charge → HCR* (O⋯Cα–Cβ) interaction (Fig. 1c(i)) over the charge → σ* (along Cα–Hα or Cα–Cβ) interaction and slightly greater electronic back donation of the charge from C Created by potrace 1.16, written by Peter Selinger 2001-2019 O to Cα–Cβ in the acetates and pyrrolidine carbamates (Fig. 1c(ii)). There were consistent slight deviations in particularly the δ ppm values of Hβ for all the isopropyl carbamate and ester analogues (3, 7, 11, 15), as observed from the trends obtained for the remaining homologues. This was consistent with the small distortions away from the ideal anti-periplanarity of the carbonyl oxygen and the Cα–Cβ bond observed in their crystal structures, which diminished the charge → HCR* donation and would not be observed unless charge → HCR* predominated over charge → σ* in O⋯Cα.The generic charge → acceptor O⋯Cα electronic back donation interactions explain the X-ray structures of phenyl carbamates,21 where the inclusive plane of the carbamate group is perpendicular to the plane of the phenyl ring.22 Current data additionally indicate that this interaction is (a) largely charge → π* in nature rather than n → π*; (b) primarily localized between O and Cα of the phenyl ring; and (c) a consequence of the general O⋯Cα charge → acceptor electronic interactions at the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups, which are observed for a variety of Cα acceptors: (1) σ* acceptor in Cα–Hα, when R is a methyl group (Fig. 1c(ii)); (2) HCR* acceptor in Cα–Cβ, when R has a CβH group (Fig. 1c(i)); and (3) π* acceptor, when R is phenyl (Fig. 1b). In other words, the O⋯Cα electronic charge → acceptor back donation interaction is observed at alkyl carbamates and esters as well as they are observed in phenyl carbamate.Note that only a charge (rather than a lone pair) donor at the carbonyl oxygen, which has a longer coulombic interaction range, is consistent with the reported observations for phenyl carbamates22 that the rest of the π*-orbitals of the phenyl ring (other than at its Cα) that are at distances longer than is conducive for any n → π* orbital overlap interactions also accept electron density from the carbonyl oxygen. Moreover, the constancy (rather than decrease) in the O⋯Cα interactions despite increasing the substitution at Cα and the concomitant upfield shift in the Cβ signals of the current analogues and the O⋯Cα–Cβ periplanarities with little perturbation in the O⋯Cα distances in crystals all substantiate an R group-dependent stabilization of the negative charge on the carbonyl oxygen, which interacts back with Cα of R and influences the rotational states of the O–Cα bond at the R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O groups. Thus, the charge → acceptor O⋯Cα interaction is generic to the Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O group and influences the biasing of the rotational states along the O–Cα bond. The incorporation of the corresponding force fields in the computational methods will benefit energy minimization of the rotational states in these molecules. The generic charge → HCR*/σ*/π* interaction model at Cα–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O is also consistent with masking the charge at the carbonyl oxygen, hence explaining the low basicities and poor hydrogen bond acceptor propensities of the carbamates and esters.Finally, it is possible that the O⋯Cα interaction has an electrostatic component as well due to the polarized charge at Cα. Although this may influence the planarity of R–O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O even when R is a methyl group, the resonance at O–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O would have a major role in such planarity. Moreover, such electrostatic interactions are insufficient by themselves to explain the observed C Created by potrace 1.16, written by Peter Selinger 2001-2019 O⋯Cα–Cβ anti-periplanarities and the relative downfield shifting of Hβ (compared to alcohols). Note that the latter data also discount the possibility of O⋯Hα–Cα or O⋯Hβ–Cβ-type hydrogen bonding interactions.  相似文献   

18.
The effects of green waste compost on soil N,P, K,and organic matter fractions in forestry soils: elemental analysis evaluation     
Xiaojie Feng  Xiangyang Sun  Wenjie Zhou  Wei Zhang  Feiwei Che  Suyan Li 《RSC advances》2021,11(51):31983
We study the effects of green waste compost on soil fertility to provide a theoretical basis for accurately improving forestry soil quality. This study aims to investigate the effects of green waste compost on soil N, P, K, and soil organic matter (SOM) fractions using elemental and FTIR analyses. Therefore, five fertilization treatments were set up for research, including mineral fertilization (M-fert), green waste compost fertilization (G-fert), standard rate of M-fert plus G-fert (GM-fert), half the standard rate of M-fert plus G-fert (1/2 GM-fert), and a control with no fertilizer addition (N-fert). The results showed that GM-fert treatment significantly increased the content of soil NH4–N, available phosphorus (AP), available potassium (AK), water soluble organic carbon (WSOC), humus (HE), and humic acid (HA), which were 8.53 ± 0.67, 76.1 ± 5.96, 168 ± 3.42, 0.152 ± 0.01, 5.64 ± 0.15, and 4.69 ± 0.21 mg kg−1, respectively. The content of HA (36.7%, F = 7.55, P = 0.01) was positively correlated with the soil N, P, K, and the HA absorption peak. The relative intensities of the alcohol –OH, aliphatic –CH and carbohydrate C–O peaks showed the largest changes, which were 18.6 ± 0.56%, 13.1 ± 0.33%, and 16.3 ± 0.49%. –CH/C Created by potrace 1.16, written by Peter Selinger 2001-2019 C (49.8%, F = 12.9, P < 0.01) was also significantly positively correlated with soil N, P, K. In conclusion, green waste compost significantly increased soil N, P, K, and HA in forestry soils, and the –CH/C Created by potrace 1.16, written by Peter Selinger 2001-2019 C of HA was the main factor related to soil nutrients.

Green waste compost significantly increased soil N, P, K, and HE fractions, and the –CH/C Created by potrace 1.16, written by Peter Selinger 2001-2019 C components of the HA structures made the biggest contribution to soil N, P, K in forestry soil.  相似文献   

19.
A mechanism of surface hardness enhancement for H+ irradiated polycarbonate     
Sunmog Yeo  Won-Je Cho  Dong-Seok Kim  Chan Young Lee  Yong Seok Hwang  Jae Kwon Suk  Chorong Kim  Jun Mok Ha 《RSC advances》2020,10(48):28603
H+ irradiation increases the surface hardness of polycarbonate. Nano indentation measurement shows that the hardness increases up to 3.7 GPa at the dose of 5 × 1016 # cm−2 and at the irradiation energy of 150 keV. In addition, the hardness increases with the dose and the energy of H+ irradiation. In accordance with the nano indentation measurement, the Fourier-transform infrared spectroscopy (FTIR) depends on the dose and energy of H+ irradiation. The peak at ∼1500 cm−1 for the aromatic ring and the peak at ∼1770 cm−1 for the C Created by potrace 1.16, written by Peter Selinger 2001-2019 O stretch decrease with increasing dose and energy, while the increase of the dose and energy develops a new C Created by potrace 1.16, written by Peter Selinger 2001-2019 O stretch vibration at ∼1700 cm−1 and forms aromatic hydrocarbons at ∼1600 cm−1. X-ray diffraction experiments are also consistent with the nano indentation measurement and FTIR spectra. Based on the experiments, we discuss a possible mechanism of the surface hardness enhancements by ion beam irradiation.

H+ irradiation increases the surface hardness of polycarbonate.  相似文献   

20.
Functionalization of MOF-5 with mono-substituents: effects on drug delivery behavior     
Mengru Cai  Liuying Qin  Longtai You  Yu Yao  Huimin Wu  Zhiqin Zhang  Lu Zhang  Xingbin Yin  Jian Ni 《RSC advances》2020,10(60):36862
Metal organic frameworks (MOFs) are widely used in drug carrier research due to their tunability. The properties of MOFs can be adjusted through incorporation of mono-substituents to obtain pharmaceutical carriers with excellent properties. In this study, different functional groups of –NH2, –CH3, –Br, –OH and –CH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 CH are connected to MOF-5 to analyse the effect of mono-substituent incorporation on drug delivery properties. The resulting MOFs have similar structures, except for Br–MOF. The pore size of this series of MOFs ranges from 1.04 nm to 1.10 nm. Using oridonin (ORI) as a model drug, introduction of the functional groups appears to have a significant effect on the drug delivery performance of the MOFs. The IRMOFs can be ranked according to drug-loading capacity: MOF-5 > HO–MOF-5 > H3C–MOF-5 = Br–MOF-5 > H2N–MOF-5 > CH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 CH–MOF-5. The ORI release from ORI @IRMOFs is explored at two different pH values: 7.4 and 5.5, and the ORI@IRMOFs are ranked according to the cumulative release percentage of ORI: ORI@MOF-5 > ORI@Br–MOF-5 > ORI@H3C–MOF-5 > ORI@H2N–MOF-5 > CH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 CH–MOF-5 > ORI@ HO–MOF-5. In particular, the release behaviour of ORI@MOFs is described through a new model. The different drug delivery performance of MOFs may be due to the complex interactions between MOFs and ORI. In addition, the introduction of single substituents does not change the biocompatibility of MOFs. MTT in vitro experiments prove that this series of MOFs has low cytotoxicity. This study shows that the incorporation of single substituents can effectively adjust the drug delivery behaviour of MOFs, which is conducive to realization of personalized drug delivery modes. The introduction of active groups can also facilitate post-synthesis modification to achieve coupling of targeting groups. MOFs incorporated with single substituents perform favorably in terms of use as biomedical drug delivery alternative carriers in effective drug payload and flexible drug release.

Metal organic frameworks (MOFs) are widely used in drug carrier research due to their tunability.  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