首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
An experimental study on protonation of simple weakly basic molecules (L) by the strongest solid superacid, H(CHB11F11), showed that basicity of SO2 is high enough (during attachment to the acidic H atoms at partial pressure of 1 atm) to break the bridged H‐bonds of the polymeric acid and to form a mixture of solid mono‐ LH+⋅⋅⋅An, and disolvates, L−H+−L. With a decrease in the basicity of L=CO (via C), N2O, and CO (via O), only proton monosolvates are formed, which approach L−H+−An species with convergence of the strengths of bridged H‐bonds. The molecules with the weakest basicity, such as CO2 and weaker, when attached to the proton, cannot break the bridged H‐bond of the polymeric superacid, and the interaction stops at stage of physical adsorption. It is shown here that under the conditions of acid monomerization, it is possible to protonate such weak bases as CO2, N2, and Xe.  相似文献   

2.
Nitrogenase cofactors can be extracted into an organic solvent to catalyze the reduction of cyanide (CN?), carbon monoxide (CO), and carbon dioxide (CO2) without using adenosine triphosphate (ATP), when samarium(II) iodide (SmI2) and 2,6‐lutidinium triflate (Lut‐H) are employed as a reductant and a proton source, respectively. Driven by SmI2, the cofactors catalytically reduce CN? or CO to C1–C4 hydrocarbons, and CO2 to CO and C1–C3 hydrocarbons. The C? C coupling from CO2 indicates a unique Fischer–Tropsch‐like reaction with an atypical carbonaceous substrate, whereas the catalytic turnover of CN?, CO, and CO2 by isolated cofactors suggests the possibility to develop nitrogenase‐based electrocatalysts for the production of hydrocarbons from these carbon‐containing compounds.  相似文献   

3.
The dissociative photoionization of molecular‐beam cooled CH2CO in a region of ?10–20 eV was investigated with photoionization mass spectrometry using a synchrotron radiation as the light source. Photoionization efficiency curves of CH2CO+ and of observed fragment ions CH2+, CHCO+, HCO+, C2O+, CO+, and C2H2+ were measured to determine their appearance energies. Relative branching ratios as a function of photon energy were determined. Energies for formation of these observed fragment ions and their neutral counterparts upon ionization of CH2CO are computed with the Gaussian‐3 method. Dissociative photoionization channels associated with six observed fragment ions are proposed based on comparison of determined appearance energies and predicted energies. The principal dissociative processes are direct breaking of C=C and C‐H bonds to form CH2+ + CO and CHCO+ + H, respectively; at greater energies, dissociation involving H migration takes place.  相似文献   

4.
We report, for the first time, utilizing a rotating ring‐disc electrode (RRDE) assembly for detecting changes in the local pH during aqueous CO2 reduction reaction (CO2RR). Using Au as a model catalyst where CO is the only product, we show that the CO oxidation peak shifts by ?86±2 mV/pH during CO2RR, which can be used to directly quantify the change in the local pH near the catalyst surface during electrolysis. We then applied this methodology to investigate the role of cations in affecting the local pH during CO2RR and find that during CO2RR to CO on Au in an MHCO3 buffer (where M is an alkali metal), the experimentally measured local basicity decreased in the order Li+ > Na+ > K+ > Cs+, which agreed with an earlier theoretical prediction by Singh et al. Our results also reveal that the formation of CO is independent of the cation. In summary, RRDE is a versatile tool for detecting local pH change over a diverse range of CO2RR catalysts. Additionally, using the product itself (i.e. CO) as the local pH probe allows us to investigate CO2RR without the interference of additional probe molecules introduced to the system. Most importantly, considering that most CO2RR products have pH‐dependent oxidation, RRDE can be a powerful tool for determining the local pH and correlating the local pH to reaction selectivity.  相似文献   

5.
CO2 activation mediated by [LTiH]+ (L=Cp2, O) is observed in the gas phase at room temperature using electrospray‐ionization mass spectrometry, and reaction details are derived from traveling wave ion‐mobility mass spectrometry. Wheresas oxygen‐atom transfer prevails in the reaction of the oxide complex [OTiH]+ with CO2, generating [OTi(OH)]+ under the elimination of CO, insertion of CO2 into the metal–hydrogen bond of the cyclopentadienyl complex, [Cp2TiH]+, gives rise to the formate complex [Cp2Ti(O2CH)]+. DFT‐based methods were employed to understand how the ligand controls the observed variation in reactivity toward CO2. Insertion of CO2 into the Ti?H bond constitutes the initial step for the reaction of both [Cp2TiH]+ and [OTiH]+, thus generating formate complexes as intermediates. In contrast to [Cp2Ti(O2CH)]+ which is kinetically stable, facile decarbonylation of [OTi(O2CH)]+ results in the hydroxo complex [OTi(OH)]+. The longer lifetime of [Cp2Ti(O2CH)]+ allows for secondary reactions with background water, as a result of which, [Cp2Ti(OH)]+ is formed. Further, computational studies reveal a good linear correlation between the hydride affinity of [LTi]2+ and the barrier for CO2 insertion into various [LTiH]+ complexes. Understanding the intrinsic ligand effects may provide insight into the selective activation of CO2.  相似文献   

6.
Reduction of CO2 by direct one‐electron activation is extraordinarily difficult because of the ?1.9 V reduction potential of CO2. Demonstrated herein is reduction of aqueous CO2 to CO with greater than 90 % product selectivity by direct one‐electron reduction to CO2.? by solvated electrons. Illumination of inexpensive diamond substrates with UV light leads to the emission of electrons directly into water, where they form solvated electrons and induce reduction of CO2 to CO2.?. Studies using diamond were supported by studies using aqueous iodide ion (I?), a chemical source of solvated electrons. Both sources produced CO with high selectivity and minimal formation of H2. The ability to initiate reduction reactions by emitting electrons directly into solution without surface adsorption enables new pathways which are not accessible using conventional electrochemical or photochemical processes.  相似文献   

7.
Infrared photodissociation spectroscopy of mass‐selected heteronuclear cluster anions in the form of OMFe(CO)5 (M=Sc, Y, La) indicates that all these anions involve an 18‐electron [Fe(CO)4]2− building block that is bonded with the M center through two bridged carbonyl ligands. The OLaFe(CO)5 anion is determined to be a CO‐tagged complex involving a [Fe(CO)4]2−[LaO]+ anion core. In contrast, the OYFe(CO)5 anion is characterized to have a [Fe(CO)4]2−[Y(η2‐CO2)]+ structure involving a side‐on bonded CO2 ligand. The CO‐tagged complex and the [Fe(CO)4]2−[Sc(η2‐CO2)]+ isomer co‐exist for the OScFe(CO)5 anion. These observations indicate that both the ScO+ and YO+ cations supported on [Fe(CO)4]2− are able to oxidize CO to CO2. Theoretical analyses show that [Fe(CO)4]2− coordination significantly weakens the MO+ bond and decreases the energy gap of the interacting valence orbitals between MO+ and CO, leading to the CO oxidation reactions being both thermodynamically exothermic and kinetically facile.  相似文献   

8.
Electroreduction of CO2 to CO powered by renewable electricity is a possible alternative to synthesizing CO from fossil fuel. However, it is very hard to achieve high current density at high faradaic efficiency (FE). Here, the first use of N,P‐co‐doped carbon aerogels (NPCA) to boost CO2 reduction to CO is presented. The FE of CO could reach 99.1 % with a partial current density of ?143.6 mA cm?2, which is one of the highest current densities to date. NPCA has higher electrochemical active area and overall electronic conductivity than that of N‐ or P‐doped carbon aerogels, which favors electron transfer from CO2 to its radical anion or other key intermediates. By control experiments and theoretical calculations, it is found that the pyridinic N was very active for CO2 reduction to CO, and co‐doping of P with N hinder the hydrogen evolution reaction (HER) significantly, and thus the both current density and FE are very high.  相似文献   

9.
Ag is a promising catalyst for the production of carbon monoxide (CO) via the electrochemical reduction of carbon dioxide (CO2ER). Herein, we study the role of the formate (HCOO?) intermediate *OCHO, aiming to resolve the discrepancy between the theoretical understanding and experimental performance of Ag. We show that the first coupled proton‐electron transfer (CPET) step in the CO pathway competes with the Volmer step for formation of *H, whereas this Volmer step is a prerequisite for the formation of *OCHO. We show that *OCHO should form readily on the Ag surface owing to solvation and favorable binding strength. In situ surface‐enhanced Raman spectroscopy (SERS) experiments give preliminary evidence of the presence of O‐bound bidentate species on polycrystalline Ag during CO2ER which we attribute to *OCHO. Lateral adsorbate interactions in the presence of *OCHO have a significant influence on the surface coverage of *H, resulting in the inhibition of HCOO? and H2 production and a higher selectivity towards CO.  相似文献   

10.
Ni,Fe‐containing CO dehydrogenases (CODHs) use a [NiFe4S4] cluster, termed cluster C, to reversibly reduce CO2 to CO with high turnover number. Binding to Ni and Fe activates CO2, but current crystal structures have insufficient resolution to analyze the geometry of bound CO2 and reveal the extent and nature of its activation. The crystal structures of CODH in complex with CO2 and the isoelectronic inhibitor NCO? are reported at true atomic resolution (dmin≤1.1 Å). Like CO2, NCO? is a μ22 ligand of the cluster and acts as a mechanism‐based inhibitor. While bound CO2 has the geometry of a carboxylate group, NCO? is transformed into a carbamoyl group, thus indicating that both molecules undergo a formal two‐electron reduction after binding and are stabilized by substantial π backbonding. The structures reveal the combination of stable μ22 coordination by Ni and Fe2 with reductive activation as the basis for both the turnover of CO2 and inhibition by NCO?.  相似文献   

11.
The chemical fixation of CO2 under mild reaction conditions is of significance from a sustainable chemistry viewpoint. Herein a CO2‐reactive protic ionic liquid (PIL), [HDBU+][TFE?], was designed by neutralization of the superbase 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) with a weak proton donor trifluoroethanol (TFE). As a bifunctional catalyst for simultaneously activating CO2 and the substrate, this PIL displayed excellent performance in catalyzing the reactions of CO2 with 2‐aminobenzonitriles at atmospheric pressure and room temperature, thus producing a series of quinazoline‐2,4(1H,3H)‐diones in excellent yields.  相似文献   

12.
The selective formation of dialkyl formamides through photochemical CO2 reduction was developed as a means of utilizing CO2 as a C1 building block. Photochemical CO2 reduction catalyzed by a [Ru(bpy)2(CO)2]2+ (bpy: 2,2′‐bipyridyl)/[Ru(bpy)3]2+/Me2NH/Me2NH2+ system in CH3CN selectively produced dimethylformamide. In this process a ruthenium carbamoyl complex ([Ru(bpy)2(CO)(CONMe2)]+) formed by the nucleophilic attack of Me2NH on [Ru(bpy)2(CO)2]2+ worked as the precursor to DMF. Thus Me2NH acted as both the sacrificial electron donor and the substrate, while Me2NH2+ functioned as the proton source. Similar photochemical CO2 reductions using R2NH and R2NH2+ (R=Et, nPr, or nBu) also afforded the corresponding dialkyl formamides (R2NCHO) together with HCOOH as a by‐product. The main product from the CO2 reduction transitioned from R2NCHO to HCOOH with increases in the alkyl chain length of the R2NH. The selectivity between R2NCHO and HCOOH was found to depend on the rate of [Ru(bpy)2(CO)(CONR2)]+ formation.  相似文献   

13.
Nitrogenase cofactors can be extracted into an organic solvent to catalyze the reduction of cyanide (CN), carbon monoxide (CO), and carbon dioxide (CO2) without using adenosine triphosphate (ATP), when samarium(II) iodide (SmI2) and 2,6‐lutidinium triflate (Lut‐H) are employed as a reductant and a proton source, respectively. Driven by SmI2, the cofactors catalytically reduce CN or CO to C1–C4 hydrocarbons, and CO2 to CO and C1–C3 hydrocarbons. The C C coupling from CO2 indicates a unique Fischer–Tropsch‐like reaction with an atypical carbonaceous substrate, whereas the catalytic turnover of CN, CO, and CO2 by isolated cofactors suggests the possibility to develop nitrogenase‐based electrocatalysts for the production of hydrocarbons from these carbon‐containing compounds.  相似文献   

14.
Reducing CO2 selectively to one of the several C1 products is challenging, as the thermodynamic reduction potentials for the different n e/n H+ reductions of CO2 are similar and so is the reduction potential for H+ reduction. Recently, Halime, Aukauloo, and co-workers have taken inspiration from the active site of nickel CO dehydrogenase (Ni-CODH) to design bimetallic iron porphyrins bridged by a urea moiety. These complexes show fast and selective reduction of CO2 to CO and the results suggest a Ni-CODH type mechanism at play where one of the two metals binds and reduces the CO2 while the other stabilizes the reduced species by forming a bridged complex, facilitating the C−O bond cleavage.  相似文献   

15.
We report the synthesis, structural characterization, and porous properties of two isomeric supramolecular complexes of ([Cd(NH2?bdc)(bphz)0.5]?DMF?H2O}n (NH2?bdc=2‐aminobenzenedicarboxylic acid, bphz=1,2‐bis(4‐pyridylmethylene)hydrazine) composed of a mixed‐ligand system. The first isomer, with a paddle‐wheel‐type Cd2(COO)4 secondary building unit (SBU), is flexible in nature, whereas the other isomer has a rigid framework based on a μ‐oxo‐bridged Cd2(μ‐OCO)2 SBU. Both frameworks are two‐fold interpenetrated and the pore surface is decorated with pendant ?NH2 and ?N?N? functional groups. Both the frameworks are nonporous to N2, revealed by the type II adsorption profiles. However, at 195 K, the first isomer shows an unusual double‐step hysteretic CO2 adsorption profile, whereas the second isomer shows a typical type I CO2 profile. Moreover, at 195 K, both frameworks show excellent selectivity for CO2 among other gases (N2, O2, H2, and Ar), which has been correlated to the specific interaction of CO2 with the ?NH2 and ?N?N? functionalized pore surface. DFT calculations for the oxo‐bridged isomer unveiled that the ?NH2 group is the primary binding site for CO2. The high heat of CO2 adsorption (ΔHads=37.7 kJ mol?1) in the oxo‐bridged isomer is realized by NH2???CO2/aromatic π???CO2 and cooperative CO2???CO2 interactions. Further, postsynthetic modification of the ?NH2 group into ?NHCOCH3 in the second isomer leads to a reduced CO2 uptake with lower binding energy, which establishes the critical role of the ?NH2 group for CO2 capture. The presence of basic ?NH2 sites in the oxo‐bridged isomer was further exploited for efficient catalytic activity in a Knoevenagel condensation reaction.  相似文献   

16.
We present herein a Cp*Co(III)‐half‐sandwich catalyst system for electrocatalytic CO2 reduction in aqueous acetonitrile solution. In addition to an electron‐donating Cp* ligand (Cp*=pentamethylcyclopentadienyl), the catalyst featured a proton‐responsive pyridyl‐benzimidazole‐based N,N‐bidentate ligand. Owing to the presence of a relatively electron‐rich Co center, the reduced Co(I)‐state was made prone to activate the electrophilic carbon center of CO2. At the same time, the proton‐responsive benzimidazole scaffold was susceptible to facilitate proton‐transfer during the subsequent reduction of CO2. The above factors rendered the present catalyst active toward producing CO as the major product over the other potential 2e/2H+ reduced product HCOOH, in contrast to the only known similar half‐sandwich CpCo(III)‐based CO2‐reduction catalysts which produced HCOOH selectively. The system exhibited a Faradaic efficiency (FE) of about 70% while the overpotential for CO production was found to be 0.78 V, as determined by controlled‐potential electrolysis.  相似文献   

17.
Ba(CO)+ and Ba(CO)? have been produced and isolated in a low‐temperature neon matrix. The observed C?O stretching wavenumber for Ba(CO)+ of 1911.2 cm?1 is the most red‐shifted value measured for any metal carbonyl cations, indicating strong π backdonation of electron density from Ba+ to CO. Quantum chemical calculations indicate that Ba(CO)+ has a 2Π reference state, which correlates with the 2D(5d1) excited state of Ba+ that comprises significant Ba+(5dπ1)→CO(π* LUMO) backbonding, letting the Ba(CO)+ complex behave like a conventional transition‐metal carbonyl. A bonding analysis shows that the π backdonation in Ba(CO)+ is much stronger than the Ba+(5dσ/6s)←CO(HOMO) σ donation. The Ba+ cation in the 2D(5d1) excited state is a donor rather than an acceptor. Covalent bonding in the radical anion Ba(CO)? takes place mainly through Ba(5dπ)←CO?(π* SOMO) π donation and Ba(5dσ/6s)←CO?(HOMO) σ donation. The most important valence functions at barium in Ba(CO)+ cation and Ba(CO)? anion are the 5d orbitals.  相似文献   

18.
High-energy collisional activation mass spectrometry of HFe(CO)5+ ions shows that Fe(CO)5 is protonated on the iron atom rather than on one of the ligands. This finding is supported by ab initio quantum chemical calculations. The value of the proton affinity of Fe(CO)5 was measured by high-pressure mass spectrometry to be 857 kJ mol?1. The Fe? CO bond dissociation energies for HFe(CO)n+ (n = 1–5) were measured by energy-variable low-energy collisional activation mass spectrometry. The Fe? H bond dissociation energies in HFe(CO)n+ ions were also determined. A synergistic effect on the strengths of the Fe? H and Fe? CO bonds in HFe(CO)+ is noticed. It is demonstrated that the electronically unsaturated species HFe(CO)n+ (n = 3, 4) formed in exothermic proton-transfer reactions with Fe(CO)5 form adducts with CH4. Adducts between C2H5+ or C3H5+ and Fe(CO)n are observed. These adducts are probably formed in direct reactions between the respective carbocations and Fe(CO)5.  相似文献   

19.
Formation of C? C bonds from CO2 is a much sought after reaction in organic synthesis. To date, other than C? H carboxylations using stoichiometric amounts of metals, base, or organometallic reagents, little is known about C? C bond formation. In fact, to the best of our knowledge no catalytic methylation of C? H bonds using CO2 and H2 has been reported. Described herein is the combination of CO2 and H2 for efficient methylation of carbon nucleophiles such as indoles, pyrroles, and electron‐rich arenes. Comparison experiments which employ paraformaldehyde show similar reactivity for the CO2/H2 system.  相似文献   

20.
《Chemphyschem》2003,4(5):466-473
The influence of potassium, in the submonolayer regime, on the adsorption and coadsorption of CO2 and H on a stepped copper surface, Cu(115), has been studied by photoelectron spectroscopy, temperature‐programmed desorption, and work‐function measurements. Based on the fast recording of C 1s and O 1s core‐level spectra, the uptake of CO2 on K/Cu(115) surfaces at 120 K has been followed in real time, and the different reaction products have been identified. The K 2p3/2 peak exhibits a chemical shift of ?0.4 eV with CO2 saturation, the C 1s peaks of the CO3 and the CO species show shifts of ?0.8 and ?0.5 eV, respectively, and the C 1s peak of the physisorbed CO2 exhibits no shift. The effects of gradually heating the CO2/K/Cu(115) surface include the desorption of physisorbed CO2 at 143 K; the desorption of CO at 193 K; the ordering of the CO3 species, and subsequently the dissociation of the carbonate with desorption at 520–700 K. Formate, HCOO?, was synthesized by the coadsorption of H and CO2 on the K/Cu(115) surface at 125 K. Formate formed exclusively for potassium coverages of less than 0.4 monolayer, whereas both formate and carbonate were formed at higher coverages. The desorption of formate‐derived CO2 took place in the temperature range 410–425 K and carbonate‐derived CO2 desorbed at 645–660 K, depending on the potassium coverage.  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