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 187 毫秒
1.
Single-pulse 13C NMR spectra and spin-lattice relaxation times T1(1H), detected indirectly via 13C carbons, and T1(13C) were measured at 31°C for virgin pelletized and annealed polylactic acid (PLA) samples using the magic-angle spinning technique. The structural relaxation resulting in more regular crystals with narrower conformation distribution and increase in the lamellae thickness and crystallinity brought about by annealing at 100°C was deduced from the narrowing of the 13C NMR lines and proton spin-lattice relaxation times T1(1H). The spin-lattice relaxation times T1(13C) related to the respective carbons of the α-polymorph of PLA are also discussed in the study.  相似文献   

2.
Analysis of the 13C NMR chemical shift and coupling constant data for a number of straight-chain aliphatic trialkylphosphines and their transition metal carbonyl complexes suggests that complexation leads to: (1) a deshielding of C(1) and an increase in 1J(13C31P), (2) a slight shielding of C(2) and a decrease in 2J(13C31P), and (3) little or no change in the chemical shift for C(3) and a slight increase in 3J(13C31P). Application of these rules to the assignment of the 13C NMR spectrum of P(butyl)3 led to conflict with prior work. A study of segmental motion in these derivatives via spin-lattine (T1) relaxation time measurements was therefore performed, and these data are in complete agreement with the proposed assignments. These generalizations must be applied with care, however, since the presence of either unsaturation or branching near the phosphorus can interfere with this pattern.  相似文献   

3.
Assembly of the triangular, organic radical‐bridged complexes Cp*6Ln33‐HAN) (Cp*=pentamethylcyclopentadienyl; Ln=Gd, Tb, Dy; HAN=hexaazatrinaphthylene) proceeds through the reaction of Cp*2Ln(BPh4) with HAN under strongly reducing conditions. Significantly, magnetic susceptibility measurements of these complexes support effective magnetic coupling of all three LnIII centers through the HAN3−. radical ligand. Thorough investigation of the DyIII congener through both ac susceptibility and dc magnetic relaxation measurements reveals slow relaxation of the magnetization, with an effective thermal relaxation barrier of Ueff=51 cm−1. Magnetic coupling in the DyIII complex enables a large remnant magnetization at temperatures up to 3.0 K in the magnetic hysteresis measurements and hysteresis loops that are open at zero‐field up to 3.5 K.  相似文献   

4.
In this paper we present measurements of nuclear relaxation times of31P in the polynuclear cluster compounds Au55(PPh3)12Cl6 and Ru55(P(t-Bu)3)12Cl20. Above 15 K the data can be described as a Korringa process, while below 15 K the relaxation time appears to be thermally activated.  相似文献   

5.
205Tl longitudinal relaxation rate measurements were performed on several thallium(III) complexes with the composition Tl(OH)n(H2O)6?n(3?n)+ (n = 1,2), Tl(Cl)n(H2O)m?n(3?n)+, Tl(Br)n(H2O)m?n(3?n)+ (m = 6 for n = 1–2, m = 5 for n = 3, m = 4 for n = 4), Tl(CN)n(H2O)m?n(3?n)+ (m = 6 for n = 1–2, m = 4 for n = 3–4) in aqueous solution, at different magnetic fields and temperatures. 13C and 2D isotopic labelling and 1H decoupling experiments showed that the contribution of the dipolar relaxation path is negligible. The less symmetric lower complexes (n < 4) had faster relaxation rate dominantly via chemical shift anisotropy contribution which depended on the applied magnetic field: T1 values are between 20 and 100 ms at 9.4 T and the shift anisotropy is Δσ = 1000–2000 ppm. The tetrahedral complexes, n = 4, relax slower; their T1 is longer than 1 s and the spin–rotation mechanism is probably the dominant relaxation path as showed by a temperature dependence study. In the case of the TlCl4? complex, presumably a trace amount of TlCl52? causes a large CSA contribution, 300 ppm. Since the geometry and the bond length for the complexes in solution are known from EXAFS data, it was possible to establish a correlation between the CSA parameter and the symmetry of the complexes. The relaxation behaviour of the Tl–bromo complexes is not in accordance with any known relaxation mechanism. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

6.
Two [FeLn2Fe(μ3‐OH)2(teg)2(N3)2(C6H5COO)4] compounds (where Ln=YIII and DyIII; teg=triethylene glycol anion) have been synthesized and studied using SQUID and Mössbauer spectroscopy. The magnetic measurements on both compounds indicate dominant antiferromagnetic interactions between the metal centers. Analysis of the 57Fe Mössbauer spectra complement the ac magnetic susceptibility measurements, which show how a static magnetic field can quench the slow relaxation of magnetization generated by the anisotropic DyIII ions.  相似文献   

7.
The interactions of three alcohols, namely, 2‐butanol (BuOH), 3‐methyl‐2‐butanol (MeBuOH), and 3,3‐dimethyl‐2‐butanol (Me2BuOH) with propylene oxide octamer (PO8) and the copolymers (EO)8(PO)13(EO)8 (L35) and (EO)13(PO)30(EO)13 (L64) in D2O were studied using 13C NMR spectra and relaxations and 1H PFG NMR diffusion measurements. For L64, it was shown that the temperature at which the PO chain starts to change its conformation under dehydration decreases by 6 K for each additional methyl group in the alcohol molecule (i.e. with increasing its hydrophobicity), and the analogous conformation states are attained at temperatures approximately 10 K lower compared using ketonic analogs of the alcohols under the same conditions. Also, the first signs of L64 aggregation, according to the normalized diffusion coefficients, are at temperatures 7, 10, and 13 K lower for BuOH, MeBuOH, and Me2BuOH, respectively. These effects are much weaker for (PO)13 in L35 or nonexistent for (PO)8 in PO8, thus showing the role of cooperativity in dehydration and aggregation processes. According to diffusion measurements, the molar fraction of the alcohol hydrogen bonded to L64 increases with its hydrophobicity and, in an apparent conflict with thermodynamics, with increasing temperature at which also higher NOE can be observed. Strong hydrogen bond interaction, which is in cooperation with hydrophobic interaction, does not preclude the exchange between bound and free states of the alcohol, however. Using 13C transverse relaxation, its correlation time is shown to be of the order of 10 ms. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

8.
The product of the reaction of (3-chlorophenyl)-2-naphthylamine with arsenic trichloride has been confirmed to be benzo(a)-9-chlorophenarsazine 12-chloride by 13C-nmr spectroscopy and spin-lattice (T1) relaxation time measurements. A considerable shortening of the spin-lattice relaxation times of the arsenic bearing quaternary carbons, relative to the calculated relaxation times, was observed. This discrepancy has been interpreted in terms of an 75As-13C dipolar contribution to quaternary carbon relaxation.  相似文献   

9.
Abstract

The formation of polyhydroxo aluminum(III) complexes has been investigated at 30°C and in a 3 M (K)Cl ionic medium by p[H] measurements. The uncommon “integral titration” technique employed has enabled measurements of oversaturated solutions up to OH? to Al(III) ratios as large as 2.65. This has allowed the detection of the undescribed species Al13(OH)4+ 35. The data can very satisfactorily be explained by assuming the species Al2(OH)4+ 2, Al3(OH)3+ 6, Al13(OH)7+ 32, and Al13(OH)4+ 35. The Al(III) concentration has been changed from ≈0.0025 to ≈ 0.040 M and the spacings of the titration curves at different aluminum levels are a clear and direct evidence for the formation of Al13(OH)7+ 32, which dominates the hydrolysis products. The data presented in this paper are best accounted for if the trimer Al3(OH)3+ 6 is substituted for Al3(OH)5+ 4 which is frequently reported. The formation of the “13” cations may result from the reaction of four Al3(OH)3+ 6 with a transient Al(OH)? 4 species which is formed, upon addition of a rather concentrated basic solution, owing to a local excess of OH?.  相似文献   

10.
Two novel tetranuclear, star‐shaped iron(III) clusters, [Fe4(acac)6(Br‐mp)2] and [FeIII4(acac)6(tmp)2], are described. Both have S=5 ground states resulting from antiferromagnetic nearest‐neighbour superexchange interactions, with J=?8.2 cm?1 and J=?8.5 cm?1 for 1 and 2 , respectively. Energy barriers for the relaxation of the magnetisation of approximately 12 cm?1 were derived from AC susceptibility measurements. Magnetic resonance measurements revealed a zero‐field splitting parameter D=?0.34 cm?1 for both complexes. AC susceptibility measurements in solution demonstrated that the complexes are reasonably stable in solution. Interestingly, the magnetisation relaxation slows down significantly in frozen solution, in contrast to what is generally observed for single‐molecule magnets. This was shown to result from a large increase in τ0, the prefactor in the Arrhenius equation, with the energy barrier remaining unchanged.  相似文献   

11.
The complexes of trans-[Co(III)(R,CH3-dioxH)2(py)2]I2 (R = CH3, C2H5, n-C3H7 and n-C4H9) were investigated in solution by 1H and 13C NMR spectra and 13C spin-lattice relaxation time measurements. The 1H and 13C-resonances of the R = C2H5, n-C3H7 and n-C4H9) groups were shifted to higher field than those of the free ligands by the complexation; it was attributable to the ring current shielding due to the axial pyridine ligands of the complexes. 13C spin-lattice relaxation times were interpreted as due to movement of the axial pyridine ligands as if they twist around the CoN (pyridine nitrogen) bond axis and the above R groups were moving segmentally. These segmental movements allowed the R groups to approach closely toward the axial pyridine ring plane to experience the ring current shielding.  相似文献   

12.
Values of one-bond J(C35 Cl) and T1(35Cl) were determined for small chlorine-containing molecules through 13C scalar relaxation in the rotating frame. A method is suggested for evaluating the contribution of scalar relaxation which eliminates experimental imperfections. The magnitude of J(C35Cl) seems to be influenced by the same factors as J(CF) values. The differences in 35Cl relaxation times are rationalized in terms of the molecular reorientation time.  相似文献   

13.
Carbon-13 chemical shifts, spin-lattice relaxation times and nuclear Overhauser enhancement factors at 28°C are reported for a series of polyfluoroaliphatic compounds :
, and perfluoroalkyl nonionic surfactants CmF2m+1CH2(OC2H4)nOH with m = 6, 7 and n = 3, 4, 5, 6 and C6F13CH2CH2CONH(C2H4O)nH with n = 3, 4. The influence of the perfluoroalkyl group on the 13C chemtcal shifts of the neighbouring hydrogenated carbons is discussed in terms of hyperconjugative type interactions between lone electron pairs on fluorine and the neighbouring CC or CO bond. Relaxation data show similar flexibilities of the fluorinated chains in the different molecules investigated. Nonionic surfactants exhibit segmental motions in both the hydrophobic perfluoroalkyl and the hydrophilic polyoxyethylene chains ; these motions appear to be similar to those of the analogous hydrogenated surfactants.  相似文献   

14.
Two tetranuclear cyanide‐bridged FeIII2NiII2 compounds [Ni2(L1)4Fe2(μ‐CN)4(CN)2(L2)2] · 2ClO4 · CH3OH · 4H2O ( 1 ) and [Ni2(L1)4Fe2(μ‐CN)4(CN)2(L3)2] · 2ClO4 ( 2 ) [L1 = 4,4′‐dichloro‐2,2′‐bipyridine; L2 = hydrotris(pyrazolyl)borate; L3 = tetrakis(pyrazolyl)borate] were synthesized. Magnetic measurements indicated that both two compounds showed single‐molecule magnet (SMM) behaviors with the relaxation energy barrier of Δ/kB = 68.9(8) K for 1 and 12.6(1) K for 2 . Magneto‐structural analysis indicated that the intermolecular interactions played an important part in the slow magnetic relaxation behaviors.  相似文献   

15.
1H and 13C chemical shifts in the two decoalesced rotamers of the deuteromethyl ester of N-acetyl-l-proline are established by use of NOE and relaxation measurements. This requires the previous determination of the rotational barrier ΔGc and the measurement of the relaxation times T1 of the α-proton in the two conformers. The observed results reinforce the previous structural conclusions inferred from 13C studies on several acyl l-prolines and hydroxy l-prolines. The introduction of an acyl group in the prolines does not affect the different carbon atoms of the cycle in the two rotamers in the same way. These results can be interpreted in terms of electric field effects by the examination of X-Pro dipeptides with the N-terminal amino group shifted from α (COCH2NH2) to δ position (CO(CH2)4NH2).  相似文献   

16.
The sensitivity of saturated aliphatic ketone stretching frequencies, vCO to structure and solvent effects is expressed by a four-term equationvco = 1740 + I(XXX) -0.24 G + 2.10?3G ·I(XXX)The four terms represent: the frequency of the base element (acetone) calculated in the absence of intermolecular interactions; the contribution of intramolecular environment E given by structure parameter I(itXXX) previously established in gas phase; that of the intermolecular environment related to the Allerhand and Schleyer G parameter; that of solvent-solute interactions. This equation covers an experimental range of 78 cm?1 for 192 measurements in four highly diverse solvents (C6H14, CC14, CH3CN, CHBr3) with a standard deviation of 1.6 cm?1; it expresses an overall statistical behavior, but masks individual behaviors. The latter are determined by comparing two characteristic parameters of ketones, a topological parameter I(XXX), and p expressing solvent sensitivity [slope of straight lines ν = f(G)]; they are Interpreted in terms of geometrical effect, variation of valency angles and of conformations.  相似文献   

17.
In order to determine whether accurate rotational diffusion coefficients in liquids may be determined from the bandshapes of isotopically broadened vibrational peaks, we have investigated the isotropic and anisotropic Raman spectra of the ν3(A1), CCl3 symmetric bending, vibration in CHCl3 as a function of temperature in the liquid phase. The spectral lineshapes were fitted by a model containing four Lorentzian/Gaussian summation bands with relative peak intensities equal to the relative abundances of the four isotopic combinations and frequency displacements constrained to values measured in the matrix infrared spectrum. The calculated room temperature perpendicular diffusion coefficient, D (25°C) = 8.31010 s−1, was within the range of values reported from Raman measurements on the ν1, symmetric carbon-hydrogen stretching, vibration, but was somewhat lower than published results from NMR relaxation time measurements, T1(2D), on CDCl3, and from dielectric relaxation. The activation energy, Ea(D), determined from the ν3 bandshape measurements was 30% higher than the average value from the NMR and dielectric studies. The deviation is believed to result from the sensitivity of this quantity to the fractional Lorentzian character of the fitting functions.  相似文献   

18.
1H and 13C NMR and 1H NMR relaxation spectroscopy (RS)measurements are reported for the CDCl3 and CD2Cl2 solutions of [La(NO3)3(diaza-18-crown-6)] ({bf I}), [Pr(NO3)3(diaza-18-crown-6)] ({bf II}) and [Nd(NO3)3(diaza-18-crown-6)] ({bf III}) complexes. Temperature dependencies of the 1H NMR spectra of II have been analyzed using the dynamic NMR methods for multi-site exchange. Enantiomeric isomer interconversion in II is characterized by H = 21.5 ± 4 kJ mol-1. Studies of the values of the lanthanide-induced shifts and the longitudinal relaxation rate enhancement revealed that the structure of complexes in solution is similar to that reported for the [La(NO3)3(18-crown-6)] complex in the crystal state. Nevertheless, it appears that the principal values of the molar paramagnetic susceptibility tensor (i) significantly differ in complexes II and III. The possible reasons for the different characteristics of these complexes are discussed.  相似文献   

19.
The standard enthalpies of reaction of four zinc hydroxide nitrates Zn(OH)(NO3)-H2O, Zn3(OH)4(NO3)2, Zn5(OH)8(NO3)2·2H2O et Zn5(OH)8(NO3)2 and zinc oxide with a solution of nitric acid (2N) were measured in a solution calorimeter. These results, combined with auxiliary thermochemical values from the literature, yielded values of ?429.34, ?442.41, ?897.41 and ?750.70 kcal mol?1 respectively, for the molar enthalpies of formation of these zinc hydroxide nitrates.  相似文献   

20.
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