首页 | 官方网站   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 demonstrations of the effective biodegradation in soil of pro-oxidant activated polyethylene (PE) have been reported recently. Nevertheless a comprehensive understanding of the ultimate fate in the environment of the oxidized fragments of oxo-biodegradable polyethylene materials needs the extension of the studies to other natural environments and in particular to aqueous media (river, lake, brackish and marine waters) where accidental plastic littering and the resulting degraded fragments eventually may end up.In this respect, as part of our continuing activity in the area of oxo-biodegradable polymeric materials, in the present paper we wish to report on the results attained in an ongoing investigation on the biodegradation in a water medium of thermally pre-oxidized low density polyethylene (LDPE) film samples containing pro-oxidant additives.Thermally oxidized LDPE-film samples and corresponding acetone extractable fractions were submitted to the effect of microorganism flora present in river water. The effective biodegradation was assessed by monitoring the amount of CO2 developed over time in a respirometer apparatus. Levels of biodegradation up to 12 and 48% for the degraded fragments and corresponding fractions extracted with boiling acetone were detected on a 100-day time frame.  相似文献   

2.
郭静波  范宇莹  李宝元  张兰河  马放 《化学通报》2018,81(3):241-247,266
邻苯二甲酸二丁酯(DBP)是目前应用最为广泛的邻苯二甲酸酯之一,属于难降解有机污染物,具有较强的内分泌干扰性。本文以分离得到的DBP降解菌Enterobacter sp.S8为研究对象,系统考察了共代谢基质类型和共代谢反应条件对S8共代谢生物降解DBP的影响,同时对其反应动力学及酶的稳定性进行了研究。结果表明,在无共基质添加的条件下,S8对DBP的降解率仅为30.2%,而在温度为30℃、pH为7、甲醇投加量为20mg/L以及DBP初始浓度为300mg/L的优化共代谢降解条件下降解率达到最大,为75.6%;其中DBP初始浓度对共代谢降解效果影响最大。本文的结果为采用生物共代谢方式进行DBP污染环境的治理提供了依据。  相似文献   

3.
To achieve the Fe−N−C materials with both high activity and durability in proton exchange membrane fuel cells, the attack of free radicals on Fe−N4 sites must be overcome. Herein, we report a strategy to effectively eliminate radicals at the source to mitigate the degradation by anchoring CeO2 nanoparticles as radicals scavengers adjacent (Scaad-CeO2) to the Fe−N4 sites. Radicals such as ⋅OH and HO2⋅ that form at Fe−N4 sites can be instantaneously eliminated by adjacent CeO2, which shortens the survival time of radicals and the regional space of their damage. As a result, the CeO2 scavengers in Fe−NC/Scaad-CeO2 achieved ∼80 % elimination of the radicals generated at the Fe−N4 sites. A fuel cell prepared with the Fe−NC/Scaad-CeO2 showed a smaller peak power density decay after 30,000 cycles determined with US DOE PGM-relevant AST, increasing the decay of Fe−NCPhen from 69 % to 28 % decay.  相似文献   

4.
Single-atom M−N2 (M=Fe, Co, Ni) catalysts exhibit high activity for CO2 reduction reaction (CO2RR). However, the CO2RR mechanism and the origin of activity at the single-atom sites remain unclear, which hinders the development of single-atom M−N2 catalysts. Here, using density functional theory calculations, we reveal intermediates-induced CO2RR activity at the single-atom M−N2 sites. At the M−N2 sites, the asymmetric *O*CO configuration tends to split into *CO and *OH intermediates. Intermediates become part of the active moiety to form M−(CO)N2 or M-(OH)N2 sites, which optimizes the adsorption of intermediates on the M sites. The maximum free energy differences along the optimal CO2RR pathway are 0.30, 0.54, and 0.28 eV for Fe−(OH)N2, Co−(CO)N2, and Ni−(OH)N2 sites respectively, which is lower than those of Fe−N2 (1.03 eV), Co−N2 (1.24 eV) and Ni−N2 (0.73 eV) sites. The intermediate modification can shift the d-band center of the spin-up (minority) state downward by regulating the charge distribution at the M sites, leading to less charge being accepted by the intermediates from the M sites. This work provides new insights into the understanding of the activity of single-atom M−N2 sites.  相似文献   

5.
Reactive oxygen species (ROS) play an important role in the biochemistry of the cell and occur in degenerative processes as well as in signal transduction. Iron?sulfur proteins are particularly oxygen‐sensitive and their inorganic cofactors frequently undergo ROS‐induced decomposition reactions. As experimental knowledge about these processes is still incomplete we present here a quantum chemical study of the relative energetics for the binding of the most relevant ROS to [Fe4S4] clusters. We find that cubane clusters with one uncoordinated Fe atom (as found, for instance, in aconitase) bind all oxygen derivatives considered, whereas activation of triplet O2 to singlet O2 is required for binding to valence‐saturated iron centers in these clusters. The radicals NO and OH feature the most exothermic binding energies to Fe atoms. Direct sulfoxidation of coordinating cysteine residues is only possible by OH or H2O2 as attacking agents. The thermodynamic picture of ROS binding to iron?sulfur clusters established here can serve as a starting point for studying reactivity‐modulating effects of the cluster‐embedding protein environment on ROS‐induced decomposition of iron?sulfur proteins.  相似文献   

6.
Demonstrated here is a supramolecular approach to fabricate highly ordered monolayered hydrogen- and halogen-bonded graphyne-like two-dimensional (2D) materials from triethynyltriazine derivatives on Au(111) and Ag(111). The 2D networks are stabilized by N⋅⋅⋅H−C(sp) bonds and N⋅⋅⋅Br−C(sp) bonds to the triazine core. The structural properties and the binding energies of the supramolecular graphynes have been investigated by scanning tunneling microscopy in combination with density-functional theory calculations. It is revealed that the N⋅⋅⋅Br−C(sp) bonds lead to significantly stronger bonded networks compared to the hydrogen-bonded networks. A systematic analysis of the binding energies of triethynyltriazine and triethynylbenzene derivatives further demonstrates that the X3-synthon, which is commonly observed for bromobenzene derivatives, is weaker than the X6-synthon for our bromotriethynyl derivatives.  相似文献   

7.
[B4O5(OH)42−] is a representative borate anion with a double six-membered ring structure, but there is limited knowledge about the hydrolysis mechanisms of [B4O5(OH)42−]. Density functional theory-based calculations show that the tetraborate ion undergoes three-step hydrolysis to form [B(OH)4] and an ring intermediate, [B3O2(OH)6]. Other new structures, such as linear trimer, branched tetraborate, analogous linear tetraborate, are observed, but they are not stable in neutral systems and change to ring structures. [B3O2(OH)6] hydrolyzes to [B(OH)4] and [B(OH)3] in the last two steps. The structure of borate anion and the coordination environment of the bridge oxygen atom control the hydrolysis process. [B4O5(OH)42−] always participates in the hydrolysis reaction, even with a decrease in concentration. [B3O3(OH)4], [B(OH)4], and [B(OH)3] have different roles in “water-poor” and “water-rich” zones. Concentration and pH of solution are the key factors that affect the distribution of borate ions.  相似文献   

8.
Series of photochromic 1,2-bis(thienyl)ethenes possessing perfluorocyclopentene backbones, either hydrogen or methyl groups at the β-positions of the thiophenes, and a variety of substituents in their α'-positions were prepared, which cover the range from electron-donating to electron-withdrawing (Me, −CH2OH, −OTBS, −TMS, −Br, 1,3-dioxan-2-yl, pyridin-4-yl, −CH2OH, −COOH). As a linear free energy relationship the spectroscopic Hammett equation gives fair to excellent fits to the excitation energy of the absorption maxima of the ring-opened as well as the ring-closed forms of the BTEs, when Hammett substituent constants σp were replaced by Brown's modified substituent constants σp+ and σp. Vice versa, hitherto unknown Hammett-Brown substituent constants can be estimated from the UV spectra. Furthermore, we compared the experimentally measured absorption maxima with values which we calculated by three different methods (DFT STEOM-DLPNO-CCSD/def2-TZVPP, TD-DFT ωB97X-D3/6-31G*, TD-DFT ωB97X-D3/6-311++G**).  相似文献   

9.
Montelukast (MKT), a leukotriene receptor antagonist, degrades when it is exposed to light. The analysis of MKT content in blood plasma by high-pressure liquid chromatography requires several sample preparation steps including deproteinization. This study aimed to evaluate MKT photodegradation in blood plasma samples and optimize a deproteinization method to reduce MKT photodegradation, and thereby improve analytical quality. We evaluated the stability of MKT in water and plasma in real time using high-pressure liquid chromatography and optimized a sample deproteinization procedure by comparing the effectiveness of several deproteinization methods. When exposed to light, MKT photodegraded quickly. Although MKT photodegradation was slightly slower than that in water, a half portion (55%) of the MKT in plasma degraded within 2 h when exposed to light. The rate of MKT photodegradation was dramatically reduced by sample deproteinization using ZnSO4–Ba(OH)2, but it was accelerated by deproteinization through precipitation using methanol or acetonitrile. These results suggest that precautions should be taken when preparing plasma samples for the analysis of MKT, and that deproteinization of such samples using ZnSO4–Ba(OH)2 can reduce the risk of analytical error arising from MKT photodegradation.  相似文献   

10.
The interpretation of 1H‐NMR chemical shifts, coupling constants, and coefficients of temperature dependence (δ(OH), J(H,OH), and Δδ(OH)/ΔT values) evidences that, in (D6)DMSO solution, the signal of an OH group involved as donor in an intramolecular H‐bond to a hydroxy or alkoxy group is shifted upfield, whereas the signal of an OH group acting as acceptor of an intramolecular H‐bond and as donor in an intermolecular H‐bond to (D6)DMSO is shifted downfield. The relative strength of the intramolecular H‐bond depends on co‐operativity and on the acidity of OH groups. The acidity of OH groups is enhanced when they are in an antiparallel orientation to a C−O bond. A comparison of the 1H‐NMR spectra of alcohols in CDCl3 and (D6)DMSO allows discrimination between weak and strong intramolecular H‐bonds. Consideration of IR spectra (CHCl3 or CH2Cl2) shows that the rule according to which the downfield shift of δ(OH) for H‐bonded alcohols in CDCl3 parallels the strength of the H‐bond is valid only for alcohols forming strong intramolecular H‐bonds. The combined analysis of J(H,OH) and δ(OH) values is illustrated by the interpretation of the spectra of the epoxyalcohols 14 and 15 (Fig. 3). H‐Bonding of hexopyranoses, hexulopyranoses, alkyl hexopyranosides, alkyl 4,6‐O‐benzylidenehexopyranosides, levoglucosans, and inositols in (D6)DMSO was investigated. Fully solvated non‐anomeric equatorial OH groups lacking a vicinal axial OR group (R=H or alkyl, or (alkoxy)alkyl) show characteristic J(H,OH) values of 4.5 – 5.5 Hz and fully solvated non‐anomeric axial OH groups lacking an axial OR group in β‐position are characterized by J(H,OH) values of 4.2 – 4.4 Hz (Figs. 4 – 6). Non‐anomeric equatorial OH groups vicinal to an axial OR group are involved in a partial intramolecular H‐bond (J(H,OH)=5.4 – 7.4 Hz), whereas non‐anomeric equatorial OH groups vicinal to two axial OR form partial bifurcated H‐bonds (J(H,OH)=5.8 – 9.5 Hz). Non‐anomeric axial OH groups form partial intramolecular H‐bonds to a cis‐1.3‐diaxial alkoxy group (as in 29 and 41 : J(H,OH)=4.8 – 5.0 Hz). The persistence of such a H‐bond is enhanced when there is an additional H‐bond acceptor, such as the ring O‐atom ( 43 – 47 : J(H,OH)=5.6 – 7.6 Hz; 32 and 33 : 10.5 – 11.3 Hz). The (partial) intramolecular H‐bonds lead to an upfield shift (relative to the signal of a fully solvated OH in a similar surrounding) for the signal of the H‐donor. The shift may also be related to the signal of the fully solvated, equatorial HO−C(2), HO−C(3), and HO−C(4) of β‐D ‐glucopyranose ( 16 : 4.81 ppm) by using the following increments: −0.3 ppm for an axial OH group, 0.2 – 0.25 ppm for replacing a vicinal OH by an OR group, ca. 0.1 ppm for replacing another OH by an OR group, 0.2 ppm for an antiperiplanar C−O bond, −0.3 ppm if a vicinal OH group is (partially) H‐bonded to another OR group, and −0.4 to −0.6 for both OH groups of a vicinal diol moiety involved in (partial) divergent H‐bonds. Flip‐flop H‐bonds are observed between the diaxial HO−C(2) and HO−C(4) of the inositol 40 (J(H,OH)=6.4 Hz, δ(OH)=5.45 ppm) and levoglucosan ( 42 ; J(H,OH)=6.7 – 7.1 Hz, δ(OH)=4.76 – 4.83 ppm; bifurcated H‐bond); the former is completely persistent and the latter to ca. 40%. A persistent, unidirectional H‐bond C(1)−OH⋅⋅⋅O−C(10) is present in ginkgolide B and C, as evidenced by strongly different δ(OH) and Δδ(OH)/ΔT values for HO−C(1) and HO−C(10) (Fig. 9). In the absence of this H‐bond, HO−C(1) of 52 resonates 1.1 – 1.2 ppm downfield, while HO−C(10) of ginkgolide A and of 48 – 50 resonates 0.5 – 0.9 ppm upfield.  相似文献   

11.
Ab initio MP2/aug’-cc-pVTZ calculations have been carried out to investigate H2CO : PXH2 pnicogen-bonded complexes and HCO2H : PXH2 complexes that are stabilized by pnicogen bonds and hydrogen bonds, with X=NC, F, Cl, CN, OH, CCH, CH3, and H. The binding energies of these complexes exhibit a second-order dependence on the O−P distance. DFT-SAPT binding energies correlate linearly with MP2 binding energies. The HCO2H : PXH2 complexes are stabilized by both a pnicogen bond and a hydrogen bond, resulting in greater binding energies for the HCO2H : PXH2 complexes compared to H2CO : PXH2. Neither the O−P distance across the pnicogen bond nor the O−P distance across the hydrogen bond correlates with the binding energies of these complexes. The nonlinearity of the hydrogen bonds suggests that they are relatively weak bonds, except for complexes in which the substituent X is either CH3 or H. The pnicogen bond is the more important stabilizing interaction in the HCO2H : PXH2 complexes except when the substituent X is a more electropositive group. EOM-CCSD spin-spin coupling constants 1pJ(O−P) across pnicogen bonds in H2CO:PXH2 and HCO2H : PXH2 complexes increase as the O−P distance decreases, and exhibit a second order dependence on that distance. There is no correlation between 2hJ(O−P) and the O−P distance across the hydrogen bond in the HCO2H : PXH2 complexes. 2hJ(O−P) coupling constants for complexes with X=CH3 and H have much greater absolute values than anticipated from their O−P distances.  相似文献   

12.
The Raman and IR spectra of salts of [M{IO5(OH)}2]5− (M = Cu, Ag, Au), [M(OH)2{IO5(OH)}2]6− (M = Pd, Pt), trans-[MO2{IO5(OH)}2]6− (M = Ru, Os) and {IM6O24]5− (M = Mo, W) are reported and assignments proposed.  相似文献   

13.
14.
《中国化学快报》2021,32(8):2524-2528
To enhance the photodegradation ability of CeO2 for organic dyes, an effective strategy is to introduce oxygen vacancies (Vo). In general, the introduced Vo are simultaneously present both on the surface and in the bulk of CeO2. The surface oxygen vacancies (Vo-s) can decrease the band gap, thus enhancing light absorption to produce more photogenerated e for photodegradation. However, the bulk oxygen vacancies (Vo-b) will inhibit photocatalytic activity by increasing the recombination of photogenerated e and Vo-b. Therefore, regulating the concentrations of Vo-s to Vo-b is a breakthrough for achieving the best utilization of photogenerated e during photodegradation. We used an easy hydrothermal method to achieve tunable concentrations of Vo-s to Vo-b in CeO2 nanorods. The optimized CeO2 presents a 70.2% removal of rhodamine B after 120 min of ultraviolet−visible light irradiation, and a superior photodegradation performance of multiple organics. This tuning strategy for Vo also provides guidance for developing other advanced metal-oxide semiconductor photocatalysts for the photodegradation of organic dyes.  相似文献   

15.
Following a known synthetic procedure, the porphyrin-cyclophane 1 having a porphyrin attached by two straps to an apolar cyclophane binding site was prepared. Upon metallation, the ZnII and FeIII derivatives 2 and 3 , respectively, were obtained in good yields. Treatment of 3 with base yielded the μ-oxo dimer 4 in which the two oxo-bridged porphyrins moieties are both capped by cyclophane binding sites. All compounds 1–4 are freely soluble in protic solvents such as MeOH and CF3CH2OH, and the FeIII derivatives 3 and 4 are active cytochrome P-450 mimics in these protic environments. Strong inclusion complexation of polycyclic aromatic hydrocarbons by 1 and 3 in alcoholic solvents was observed and quantified by 1H-NMR and UV/VIS titrations. Acenaphthylene binds in an ‘equatorial’ orientation which locates its reactive 1,2-double bond near the porphyrin center, whereas phenanthrene binds ‘axially’ with the reactive 9,10-double bond oriented away from the porphyrin. The reduction potential of 3 was not significantly altered by substrate binding. In the unbound form, the FeIII center in porphyrin 3 was found by ESR and 1H-NMR to prefer a high-spin state (S = 5.2). In CF3CH2OH, using iodosylbenzene as O-transfer agent, the FeIII derivative 3 catalyzed the oxidation of acenaphthylene to acenaphthen-1-one ( 14 ). Phenanthrene inhibited the reaction, possibly as a result of strong but nonproductive binding. Under similar conditions, isotetralin ( 18 ) was aromatized with high turnover to 1,4-dihydronaphthalene. The μ-oxo dimer 4 also showed high activity in the oxidation of acenaphthylen in MeOH, a result which provides strong evidence for efficent supramolecular catalysis. Due to as yet unknown reaction channels leading to polymeric products, poor mass balances were generally obtained in the oxidations effected in MeOH and CF3CH2OH in the presence of PhIO.  相似文献   

16.
The title compounds, Na4[NpO4(OH)2]OH·2H2O and Na4[PuO4(OH)2]OH·2H2O, are isostructural and isomorphous, and contain complex [AnO4(OH)2]3− anions (Ac is an actinide) in the form of distorted tetragonal bipyramids, Na+ cations, crystallization water molecules and outer‐sphere OH groups. The complex [AnO4(OH)2]3− anions occupy general positions and the coordinated OH groups deviate significantly from a centrosymmetric relative orientation. The [AnO4(OH)2]3− anions exhibit anisotropic actinide contraction; the shortening of the An—O(hydroxide) bonds on going from Np to Pu is greater than that of the AnO4 groups.  相似文献   

17.
The set-up of reversible binding interactions between the hydrophobic region of macrocyclic GBCAs (Gadolinium Based Contrast Agents) and SO3/OH containing pyrene derivatives provides new insights for pursuing relaxivity enhancements of this class of MRI contrast agents. The strong binding affinity allows attaining relaxation enhancements up to 50% at pyrene/GBCA ratios of 3 : 1. High resolution NMR spectra of the Yb-HPDO3A/pyrene system fully support the formation of a supramolecular adduct based on the set-up of hydrophobic interactions. The relaxation enhancement may be accounted for in terms of the increase of the molecular reorientation time (τR) and the number of second sphere water molecules. This effect is maintained in blood serum and in vivo, as shown by the enhancement of contrast in T1w-MR images obtained by simultaneous injection of GBCA and pyrene derivatives in mice.

The set-up of reversible binding interactions between the hydrophobic region of macrocyclic gadolinium based contrast agents and SO3/OH containing pyrene derivatives provides new insights for pursuing relaxivity enhancement of MRI contrast agents.  相似文献   

18.
Molybdenum-95 NMR chemical shifts are reported for a series of Mo(O) compounds of the type Mo(CO)5L (L = pyridine derivatives). A good correlation is found between the δ(95Mo) values and the Hammett sigma constant of the pyridine substituent or the pKa of the substituted pyridine. The chemical shift values, which range from −1366 ppm (3-CN, σ = 0.62, pKa = 1.35) to −1433 ppm (4-NMe2, σ = −0.83, pKa = 9.61), directly reflect the electronic properties of the pyridine derivatives even though the substituent is four or five bonds away from the molybdenum atom.  相似文献   

19.
We report a porous three-dimensional anionic tetrazolium based CuI−MOF 1 , which is capable of cleaving the N−H bond of ammonia and primary amine, as well as the O−H bond of H2O along with spontaneous H2 evolution. In the gas-solid phase reaction of 1 with ammonia and water vapor, CuI−MOF 1 was gradually oxidized to NH2−CuII−MOF and OH−CuII−MOF, through single-crystal-to-single-crystal (SCSC) structural transformations, which was confirmed by XPS, PXRD and X-ray single-crystal diffraction. Density functional theory (DFT) demonstrated that CuI−MOF could lower N−H bond dissociation free energy of ammonia through coordination-induced bond weakening and promote H2 evolution by the reduction potential of 1 . To our knowledge, this is the first example of MOFs that activate ammonia and amine in gas-solid manner.  相似文献   

20.
The formidable reactivity of the oxygen-evolving center near photosystem II is largely based on its protein environment that stabilizes it during catalysis. Inspired by this concept, the water-soluble Mn12 clusters Mn12O12(O2CC6H3(OH)2)16(H2O)4 (3,5DHMn12) and Mn12O12(O2CC6H3(OH)3)16(H2O)4 (3,4,5THMn12) were developed as efficient electrocatalysts for water oxidation. In this work, the role of the −OH groups in the electrocatalytic process was explored by describing the structural and electrocatalytic properties of two new Mn12 clusters, 3,4DHMn12 and 2,3DHMn12 , having one −OH group in the meta position relative to the benzoate-Mn moiety, and one at the para or ortho position, respectively. The Mn centers in 3,4DHMn12 were discovered to have lower oxidation potential compared with those in 2,3DHMn12 , and thus, 3,4DHMn12 can catalyze water oxidation with higher rate and TON than 2,3DHMn12 . Hence, the role of the −OH groups in the electrocatalysis was established, being involved in electronic stabilization of the Mn centers or in proton shuttling.  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-23

京公网安备 11010802026262号