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.
The direct methanol fuel cell (DMFC) is considered as a promising power source, because of its abundant fuel source, high energy density and environmental friendliness. Among DMFC anode materials, Pt and Pt group metals are considered to be the best electrocatalysts. The combination of Pt with some specific transition metal can reduce the cost and improve the tolerance toward CO poisoning of pure Pt catalysts. In this paper, the geometric stabilities of PtFe/PdFe atoms anchored in graphene sheet and catalytic CO oxidation properties were investigated using the density functional theory method. The results show that the Pt (Pd) and Fe atoms can replace C atoms in graphene sheet. The CO oxidation reaction by molecular O2 on PtFe–graphene and PdFe–graphene was studied. The results show that the Eley–Rideal (ER) mechanism is expected over the Langmuir–Hinshelwood mechanism for CO oxidation on both PtFe–graphene and PdFe–graphene. Further, complete CO oxidation on PtFe–graphene and PdFe–graphene proceeds via a two‐step ER reaction: CO(gas) + O2(ads) → CO2(ads) + O(ads) and CO(gas) + O(ads) → CO2(ads). Our results reveal that PtFe/PdFe commonly embedded in graphene can be used as a catalyst for CO oxidation. The microscopic mechanism of the CO oxidation reaction on the atomic catalysts was explored.  相似文献   

2.
We have elucidated the mechanism of CO oxidation catalyzed by gold nanoparticles through first‐principle density‐functional theory (DFT) calculations. Calculations on selected model show that the low‐coordinated Au atoms of the Au29 nanoparticle carry slightly negative charges, which enhance the O2 binding energy compared with the corresponding bulk surfaces. Two reaction pathways of the CO oxidation were considered: the Eley–Rideal (ER) and Langmuir–Hinshelwood (LH). The overall LH reaction O2(ads) + CO(gas) → O2(ads) + CO(ads) → OOCO(ads) → O(ads) + CO2(gas) is calculated to be exothermic by 3.72 eV; the potential energies of the two transition states ( TSLH1 and TSLH2 ) are smaller than the reactants, indicating that no net activation energy is required for this process. The CO oxidation via ER reaction Au29 + O2(gas) + CO(gas) → Au29–O2(ads) + CO(gas) → Au29–CO3(ads) → Au29–O(ads) + CO2(gas) requires an overall activation barrier of 0.19 eV, and the formation of Au29–CO3(ads) intermediate possesses high exothermicity of 4.33 eV, indicating that this process may compete with the LH mechanism. Thereafter, a second CO molecule can react with the remaining O atom via the ER mechanism with a very small barrier (0.03 eV). Our calculations suggest that the CO oxidation catalyzed by the Au29 nanoparticle is likely to occur at or even below room temperature. To gain insights into high‐catalytic activity of the gold nanoparticles, the interaction nature between adsorbate and substrate is also analyzed by the detailed electronic analysis. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

3.
Abstract

Cobalt(III) complexes of the type [Co(en)2(chel)]X.nH2O where en = ethylenediamine, chel = phthalato = C6H4CO2)2? 2, maleato = (O2CCH = CHCO2)2?, succinato = (O2CCH2CH2CO2)2?, homophthalato = (O2CC6H4(CH2)CO2)2?, citraconato = (O2CC(CH3) = CHCO2)2?, itaconato = (CH2 = C(CO2)CH2CO2)2?, X = NO? 3, Br?, (O2CC6H4CO2H)?, (O2CHC = CHCO2H)?, (O2C(CH2)2CO2H)?, (O2CC6H4(CH2)CO2H)?, (O2CHC = C(CH2)-CO2H)?, and (O2C-CH2?C(= CH2)-CO2H)?, [Co(en)2(malonato)]X.2H2O (where malonato = (O2CCH2CO2)2?, X = Cl?, Br?, and NO? 3) and [Co(en)2CO3]Cl.2H2O have been investigated for their bacterial activity against Escherichia coli B growing on EMB agar and in minimal glucose media both in lag and log phases. Among the most active are where chel = phthalato and homophthalato. The effects are distinct from those known for compounds of Pt, e.g., cis?[Pt(NH3)2Cl2] and rhodium, e.g., trans?[Rh(C5H5N)4,Cl2].6H2O. Antagonisms are reported.  相似文献   

4.
The synthesis and spectroscopic characterisation of the new diborane(4) compounds B2(1,2-O2C6Cl4)2 and B2(1,2-O2C6Br4)2 are reported together with the diborane(4) bis-amine adduct [B2(calix)(NHMe2)2] (calix=Butcalix[4]arene). B–B bond oxidative addition reactions between the platinum(0) compound [Pt(PPh3)2(η-C2H4)] and the diborane(4) compounds B2(1,2-S2C6H4)2, B2(1,2-O2C6Cl4)2 and B2(1,2-O2C6Br4)2 are also described which result in the platinum(II) bis-boryl complexes cis-[Pt(PPh3)2{B(1,2-S2C6H4)}2], cis-[Pt(PPh3)2{B(1,2-O2C6Cl4)}2] and cis-[Pt(PPh3)2{B(1,2-O2C6Br4)}2] respectively, the former two having been characterised by X-ray crystallography. In addition, the platinum complex [Pt(PPh3)2(η-C2H4)] reacts with XB(1,2-O2C6H4) (X=Cl, Br) affording the mono-boryl complexes trans-[PtX(PPh3)2{B(1,2-O2C6H4)}] as a result of oxidative addition of the B–X bonds to the Pt(0) centre; the chloro derivative has been characterised by X-ray crystallography.  相似文献   

5.
Summary The following palladium(II) and platinum(ll) complexes of rhodanine (HRd) and 3-methylrhodanine (MRd) have been prepared: Pd(HRd)1.5Cl2, Pd(HRd)2Br2, Pd(HRd)2Br2 · 0.25 EtOH, M(MRd)2X2 [M = Pd, X = Cl (0.25 EtOH) or Br; M = Pt, X = Cl or Br], Pd(MRd)3Br2, and M(MRd)4(ClO4)2 (M = Pd or Pt). The ligands are coordinated to the metal through the thiocarbonylic sulphur atom. Pd(HRd)1.5Cl2 has presumably a structure such as (X = Cl or Br) complexes have a trans-planar coordination. Pd(MRd)2X2 (X = Cl or Br) complexes arecis-planar coordinated. Pd(MRd)3Br2 has presumably a square coordination with two MRd molecules and two CI ionscis-coordinated in the equatorial plane, and a MRd molecule and a Cl ion weakly bonded in apical position. The M(MRd)4(ClO4)2 complexes have square planar coordination.Author to whom all correspondence should be addressed.  相似文献   

6.
1,3‐Dipoles of the type metallo nitrile ylide and metallo nitrile imine were prepared by mono‐α‐deprotonation of CH‐acidic {[W(CO)5CHCH2PPh3]PF6, M(CO)5CNCH2CO2R (M = Cr, W; R = Me, Et), [Pt(Cl)(CNCH2CO2Et)(PPh3)2]BF4} and NH‐acidic isocyanide complexes (Cr(CO)5CNNH2) and were stabilized by coordination to a second transition metal complex fragment {Cr(CO)5, [M(CO)5]+ (M = Mn, Re), [FeCp(CO)2]+, [Pt(Cl)(PR3)2]+ (R = Et, Ph)}. All dinuclear products 1 – 7 , 10 , and 11 are neutral species except [(Ph3P)2(Cl)Pt{μ2‐CNCH(CO2Et)}Pt(Cl)(PPh3)2]BF4 ( 8 ). Complex (OC)5W{μ2‐CNCH(CO2Et)}Pt(Cl)(PEt3)2 ( 5b ) was characterized by X‐ray diffraction. Twofold deprotonation/platination to give (OC)5Cr{μ3‐CNC(Ph)}[Pt(Cl)(PPh3)2]2 ( 9 ) was achieved in the case of Cr(CO)5CNCH2Ph.  相似文献   

7.
Preparation of the Nonahalogenodiplatinates(IV), [Pt2X9]?, X ? Cl, Br Spectroscopic Characterization, Normal Coordinate Analysis, and Crystal Structure of (PPN)[Pt2Br9] On heating the tetrabutylammonium salts (TBA)2[PtX6], with trifluoroacetic acid the nonahalogenodiplatinates(IV) (TBA)[Pt2X9], with X ? Cl, Br are formed. The X-ray structure determination on (PPN)[Pt2Br9] (orthorhombic, space group Pca2, Z = 4) shows for the anions pairs of face-sharing octahedra with nearly D3h symmetry. The mean terminal and bridging Pt? Br bond lengths are determined to be 2.42 and 2.52 Å, respectively. The electrostatic interaction of the Pt atoms results in the Pt? Pt distance of 3.23 Å and an elongation as it has been forecasted by the MO scheme for d6 systems. Using the structural data a normal coordinate analysis based on a general valence force field for [Pt2Br9]? has been performed, revealing a good agreement of the calculated frequencies with the bands observed in the IR and Raman spectra. The stronger bonding of the terminal as compared to the bridging ligands is shown by the valence force constants, fa(Br1) = 1,55 > fd(Brb) = 0,93 mdyn/ Å.  相似文献   

8.
Syntheses and Structures of [ReNBr2(Me2PhP)3] and (Me2PhPH)[ fac ‐Re(NBBr3)Br3(Me2PhP)2] [ReNBr2(Me2PhP)3] ( 1 ) has been prepared by the reaction of [ReNCl2(Me2PhP)3] with Me3SiBr in dichloromethane. The bromo complex reacts with BBr3 under formation of [Re(NBBr3)Br2(Me2PhP)3] ( 2 ) or (Me2PhPH)[fac‐Re(NBBr3)Br3(Me2PhP)2] ( 3 ) depending on the experimental conditions. The formation of the nitrido bridge leads to a significant decrease of the structural trans influence of the nitrido ligand which is evident by the shortening of the Re‐(trans)Br bond from 2.795(1) Å in [ReNBr2(Me2PhP)3] to 2.620(1) Å in [fac‐Re(NBBr3)Br3(Me2PhP)2] and 2.598(1) Å in [Re(NBBr3)Br2(Me2PhP)3], respectively.  相似文献   

9.
Spectrophotometric methods were used to investigate the rate of the reaction of Br2 with HCOOH in aqueous, acidic media. The reaction products are Br? and CO2. The kinetics of this reaction are complicated by both the formation of Br3? as Br? is formed and the dissociation of HCOOH into HCOO? and H+. Previous work on this reaction was carried out at acidities lower than the highest used here and led to the conclusion that only HCOO? reacts with Br2. It is agreed that this is by far the principal reaction. However, at the highest acidity experiments, an added small component of reaction was found, and it is suggested that it results from the direct reaction of Br2 with HCOOH itself. On this assumption, values of the rate constants for both reactions are derived here. The rate constant for the reaction of HCOO? with Br2 agrees with values previously reported, within a factor of 2 on the low side. The reaction involving HCOOH is more than 2000 times slower than the reaction involving HCOO?, but it does contribute to the overall rate as [H+] approaches 1M. These derived rate constants are able to simulate quantitatively the authors' absorbance-versus-time data, demonstrating the validity of their data treatment methods, if not mechanistic assignments. Finally, activation parameters were determined for both rate constants. The values obtained are: ΔE?(HCOOH + Br2) = 13.3 ± 1.1 kcal/mol, ΔS? (HCOOH + Br2) = ?28 ± 3 cal/deg mol, ΔE? (HCOO? + Br2) = 13.1 ± 0.9 kcal/mol, and ΔS?(HCOO? + Br2) = ?12 ± 1 cal/deg mol. That the activation energies of the two reactions turn out to be essentially identical does not support the authors' suggestion that both HCOOH and HCOO? react with Br2.  相似文献   

10.
A macrocyclic tetranuclear platinum(II) complex [Pt(en)(4,4′‐bpy)]4(NO3)8 ( 1 ?(NO3)8; en=ethylenediamine, 4,4′‐bpy=4,4′‐bipyridine) and a mononuclear platinum(IV) complex [Pt(en)2Br2]Br2 ( 2 ?Br2) formed two kinds of PtII/PtIV mixed valence assemblies when reacted: a discrete host–guest complex 1 ? 2 ?Br10 ( 3 ) and an extended 1‐D zigzag sheet 1 ?( 2 )3?Br8(NO3)6 ( 4 ). Single crystal X‐ray analysis showed that the dimensions of the assemblies could be stoichiometrically controlled. Resonance Raman spectra suggested the presence of an intervalence interaction, which is typically observed for quasi‐1‐D halogen‐bridged MII/MIV complexes. The intervalence interaction indicates the presence of an isolated {PtII???X? PtIV? X???PtII} moiety in the structure of 4 . On the basis of electronic spectra and polarized reflectance measurements, we conclude that 4 exhibits intervalence charge transfer (IVCT) bands. A Kramers–Kronig transformation was carried out to obtain an optical conductivity spectrum, and two sub‐bands corresponding to slightly different PtII–PtIV distances were observed.  相似文献   

11.
The oxidation of cis-[Pt(NH3)2(OAc)2] with H2O2 yields a mixture of two isomers: ctc-[Pt(NH3)2(OH)2(OAc)2] and ctc-[Pt(NH3)2(OH)(OAc)(OH)(OAc)]. Following modification with 4-phenylbutyric (PhB) anhydride, two isomers were separated and characterized; the symmetric ctc-[Pt(NH3)2(PhB)2(OAc)2] ( 1 ) and the nonsymmetric ctc-[Pt(NH3)2(PhB)(OAc)(PhB)(OAc)] ( 2 ). They differ in their log P values and despite having similar cellular uptake and similar DNA platination levels, the symmetric ctc-[Pt(NH3)2(OH)2(OAc)2] is more than 4-fold more potent than the nonsymmetric isomer in a panel of 4 cancer cell lines.  相似文献   

12.
The polyfluorinated title compounds, [MBr2(C18H16F8N2O2)] or [4,4′‐(HCF2CF2CH2OCH2)2‐2,2′‐bpy]MBr2, ( 1 ) (M = Pd and bpy is bipyridine) and ( 2 ) (M = Pt), have –CH(α)2OCH(β)2CF2CF2H side chains with methylene H‐atom donors at the α and β sites, and methine H‐atom donors at the terminal sites, in addition to aromatic H‐atom donors. In contrast to the original expectation of isomorphous structures, ( 1 ) crystallizes in the space group C2/c and ( 2 ) in P21/n, with similar unit‐cell volumes and Z = 4. The asymmetric unit of ( 1 ) is one half of the molecule, which resides on a crystallographic twofold axis. Both ( 1 ) and ( 2 ) display stacking of the molecules, indicating a planar (bpy)MBr2 skeleton in each case. The structure of ( 1 ) exhibits columns with C—H(β)…Br hydrogen bonds between consecutive layers which conforms to a static (β,β) linkage between layers. In the molecular plane, ( 1 ) shows double C—H(α)…Br hydrogen bonds self‐repeating along the b axis, the planar molecules being connected into infinite belts. Compound ( 2 ) has no crystallographic symmetry and forms π‐dimer pairs as supermolecules, which then stack parallel to the a axis. The π‐dimer‐pair supermolecules exhibit (Pt—)Br…Br(—Pt) contacts [3.6937 (7) Å] to neighbouring π‐dimer pairs crosslinking the columns. The structure of ( 2 ) reveals many C—H…F(—C) interactions between F atoms and aromatic C—H groups, in addition to those between F atoms and methylene C—H groups.  相似文献   

13.
Ligand exchange reactions of cis‐PtCl2(PPh3)2 and [NMe4]SCF3 in different ratios were studied. Depending on the stoichiometry reactions proceeded with formation of products expected for the chosen ratio, i. e. cis‐Pt(SCF3)Cl(PPh3)2, cis‐Pt(SCF3)2(PPh3)2, and [NMe4][Pt(SCF3)3(PPh3)]. Starting from cis‐PtCl2(MeCN)2 and [NMe4]SCF3 and adding PPh3 after substitution, product mixtures were dominated by the corresponding trans‐isomers. Results of the single crystal structure analyses of cis‐Pt(SCF3)2(PPh3)2 and trans‐Pt(SCF3)Cl(PPh3)2 are discussed.  相似文献   

14.
The preparation of a series of new compoundsLnBr3(2-bipy)2·6H2O (Ln=La, Pr, Nd, Eu, Gd) andLnBr2OH(2-bipy)2·4H2O (Ln=Pr, Nd, Sm, Eu, Gd) is described. The IR spectra of these compounds are discussed. The thermal decomposition of compoundsLnBr3(2-bipy)2·6H2O has been investigated.
2,2-Bipyridylkomplexe einiger Seltenerdmetallbromide
Zusammenfassung Es wurden 2,2-Bipyridylkomplexe des TypsLnBr3(2-bipy)2·6H2O (Ln=La, Pr, Nd, Sm, Eu, Gd) undLnBr2OH(2-bipy)2·4H2O (Ln=Pr, Nd, Sm, Eu, Gd) dargestellt. Die IR-Spektren werden diskutiert. Die thermische Zersetzung von VerbindungenLnBr3(2-bipy)2·6H2O wird untersucht.
  相似文献   

15.
195Pt NMR chemical shifts of octahedral Pt(IV) complexes with general formula [Pt(NO3)n(OH)6 ? n]2?, [Pt(NO3)n(OH2)6 ? n]4 ? n (n = 1–6), and [Pt(NO3)6 ? n ? m(OH)m(OH2)n]?2 + n ? m formed by dissolution of platinic acid, H2[Pt(OH)6], in aqueous nitric acid solutions are calculated employing density functional theory methods. Particularly, the gauge‐including atomic orbitals (GIAO)‐PBE0/segmented all‐electron relativistically contracted–zeroth‐order regular approximation (SARC–ZORA)(Pt) ∪ 6–31G(d,p)(E)/Polarizable Continuum Model computational protocol performs the best. Excellent second‐order polynomial plots of δcalcd(195Pt) versus δexptl(195Pt) chemical shifts and δcalcd(195Pt) versus the natural atomic charge QPt are obtained. Despite of neglecting relativistic and spin orbit effects the good agreement of the calculated δ 195Pt chemical shifts with experimental values is probably because of the fact that the contribution of relativistic and spin orbit effects to computed σiso 195Pt magnetic shielding of Pt(IV) coordination compounds is effectively cancelled in the computed δ 195Pt chemical shifts, because the relativistic corrections are expected to be similar in the complexes and the proper reference standard used. To probe the counter‐ion effects on the 195Pt NMR chemical shifts of the anionic [Pt(NO3)n(OH)6 ? n]2? and cationic [Pt(NO3)n(OH2)6 ? n]4 ? n (n = 0–3) complexes we calculated the 195Pt NMR chemical shifts of the neutral (PyH)2[Pt(NO3)n(OH)6 ? n] (n = 1–6; PyH = pyridinium cation, C5H5NH+) and [Pt(NO3)n(H2O)6 ? n](NO3)4 ? n (n = 0–3) complexes. Counter‐anion effects are very important for the accurate prediction of the 195Pt NMR chemical shifts of the cationic [Pt(NO3)n(OH2)6 ? n]4 ? n complexes, while counter‐cation effects are less important for the anionic [Pt(NO3)n(OH)6 ? n]2? complexes. The simple computational protocol is easily implemented even by synthetic chemists in platinum coordination chemistry that dispose limited software availability, or locally existing routines and knowhow. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

16.
Substitution reactions of three dinuclear Pt(II) complexes connected by a pyridine‐bridging ligand of variable length, namely [ cis‐{PtOH2(NH3)2}2–μ–L]4+, where L = 4,4′‐bis(pyridine)sulfide ( Pt1 ), 4,4′‐bis(pyridine)disulfide ( Pt2 ), and 1,2‐bis(4‐pyridyl)ethane ( Pt3 ) with S‐donor nucleophiles (thiourea, 1,3‐dimethyl‐2‐thiourea, and 1,1,3,3‐tetramethyl‐2‐thiourea) and anionic nucleophiles (SCN?, I?, and Br?) were investigated. The substitutions were followed under pseudofirst‐order conditions as a function of the nucleophile concentration and temperature, using stopped‐flow and UV–visible spectrophotometric methods. The observed pKa values were, respectively, Pt1 (pKa1: 4.86; pKa2: 5.53), Pt2 (pKa1: 5.19; pKa2: 6.42), and Pt3 (pKa1: 5.04; pKa2: 5.45). The second‐order rate constants for the lability of aqua ligands in the first step decreased in the order Pt2 > Pt3 > Pt1 , whereas for the second step it is Pt1 > Pt2 > Pt3 . The obtained results indicate that introduction of a spacer atom(s) on the structure of the bridging ligand influences the substitution reactivity as well as acidity of the investigated dinuclear Pt(II) complexes. Also nonplanarity of the bridging ligand of Pt1 complex significantly slows down the rate of substitution due to steric hindrance, whereas release of the strain enhances the dissociation of the bridging ligand. The release of the bridging ligand in the second step was confirmed by the 1H NMR of Pt1‐Cl with thiourea in DMF‐d7. The temperature dependence of the second–order rate constants and the negative values of entropies of activation (ΔS#) support an associative mode of the substitution mechanism.  相似文献   

17.
Preparation and Vibrational Spectra of Nonahalogenodirhodates(III), [Rh2ClnBr9-n]3?, n = 0–9 The pure nonahalogenodirhodates(III), A3[Rh2ClnBr9-n], A = K, Cs, (TBA); n = 0–4, 9, have been prepared. They are formed from the monomer chlorobromorhodates(III), [RhClnBr6-n]3?, n = 0–6, which are bridged to confacial bioctahedral complexes by ligand abstraction in less polar organic solvents. From the mixtures the complexions are separated by ion exchange chromatography on DEAE-cellulose. The solid, air-stable, air-stable, K-, Cs- and (TBA)-salts of [Rh2ClnBr9-n]3?, n = 0–4, are green, of [Rh2Cl9]3? are brown. The IR and Raman spectra of [Rh2Br9]3? and [Rh2Cl9]3? are assigned according to the point group D3h. The chlorobromodirhodates exist as mixtures of geometrical and structural isomers, which belong to different point groups. The vibrational spectra exhibit bands in characteristic regions; at high wavenumbers stretching vibrations with terminal ligands v(Rh—Clt): 360–320, v(Rh—Brt): 280–250; in a middle region with bridging ligands v(Rh—Clb): 300–270, v(Rh—Brb): 210–170 cm?1; the deformation bands are observed at distinct lower frequencies. The terminal ligands are fixed very strong, and the distance between v(Rh—Xt) and v(Rh—Xb) increases with decreasing size of the cations.  相似文献   

18.
利用俄歇电子能谱(AES)和程序升温脱附谱(TDS)研究了NO2在Ag/Pt(110)双金属表面的吸附和分解.室温下NO2 在Ag/Pt(110)双金属表面发生解离吸附, 生成NO(ads)和O(ads)表面吸附物种. 在升温过程中NO(ads)物种发生脱附或者进一步分解. 500 K时NO2在Ag/Pt(110)双金属表面发生解离吸附生成O(ads)表面吸附物种. Pt 向Ag传递电子, 从而削弱Pt-O键的强度, 降低O(ads)从Pt 表面的并合脱附温度. 发现能够形成具有稳定组成的Ag/Pt(110)合金结构, 其表现出与Pt(110)-(1×2)相似的解离吸附NO2能力, 但与O(ads)的结合明显弱于Pt(110)-(1×2). 该AgPt(110)合金结构是可能的低温催化直接分解氮氧化物活性结构.  相似文献   

19.
The reaction of a mixture of cis and trans-[PtCl2(SMe2)2] with 4,7-phen (4,7-phen = 4,7-phenanthroline) in a molar ratio of 1 : 1 or 2 : 1 resulted in the formation of mono and binuclear complexes trans-[PtCl2(SMe2)(4,7-phen)] (1) and trans-[Pt2Cl4(SMe2)2(μ-4,7-phen)] (2), respectively. The products have been fully characterized by elemental analysis, 1H, 13C{1H}, HHCOSY, HSQC, HMBC, and DEPT-135 NMR spectroscopy. The crystal structure of 1 reveals that platinum has a slightly distorted square planar geometry. Both chlorides are trans with a deviation from linearity 177.66(3)°, while the N–Pt–S angle is 175.53(6)°. Similarly, the reaction of a mixture of cis and trans-[PtBr2(SMe2)2] with 4,7-phen in a 1 : 1 or 2 : 1 mole ratio afforded the mono or binuclear complexes trans-[PtBr2(SMe2)(4,7-phen)] (3) and trans-[Pt2Br4(SMe2)2(μ-4,7-phen)] (4), respectively. The crystal structure of trans-[Pt2Br4(SMe2)2(μ-4,7-phen)].C6H6 reveals that 4,7-phen bridges between two platinum centers in a slightly distorted square planar arrangement of the platinum. In this structure, both bromides are trans, while the PtBr2(SMe2) moieties are syn to each other. NMR data of mono and binuclear complexes of platinum 14 show that the binuclear complexes exist in solution as a minor product, while the mononuclear complexes are major products.  相似文献   

20.
Nanoformulations of mononuclear Pt complexes cis-PtCl2(PPh3)2 ( 1 ), [Pt(PPh3)2(L−Cys)] ⋅ H2O ( 3 , L−Cys=L-cysteinate), trans-PtCl2(PPh2PhNMe2)2 ( 4 ; PPh2PhNMe2=4-(dimethylamine)triphenylphosphine), trans-PtI2(PPh2PhNMe2)2 ( 5 ) and dinuclear Pt cluster Pt2(μ-S)2(PPh3)4 ( 2 ) have comparable cytotoxicity to cisplatin against murine melanoma cell line B16F10. Masking of these discrete molecular entities within the hydrophobic core of Pluronic® F-127 significantly boosted their solubility and stability, ensuring efficient cellular uptake, giving in vitro IC50 values in the range of 0.87–11.23 μM. These results highlight the potential therapeutic value of Pt complexes featuring stable Pt−P bonds in nanocomposite formulations with biocompatible amphiphilic polymers.  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