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 562 毫秒
1.
Resonance Raman spectra were obtained for formanilide (FA) in acetonitrile solution with 239.5‐ and 245.9‐nm excitation wavelengths in resonance with the S3 state, and density functional theory (DFT) was used to elucidate the electronic transitions and resonance Raman spectra of FA. The spectra indicate that, in the Franck–Condon region, photodissociation dynamics has a multidimensional character with the motions mainly along the CO stretching υ8, the ring CC stretch υ9, the NH wag and ring CCH in‐plane bend υ11, the NH wag and ring CCH in‐plane bend υ12, ring CC stretch and ring CCH in‐plane bend υ16, the NH wag and ring CCH in‐plane bend υ17, the ring CCH in‐plane bend υ18, and the ring trigonal bend υ24. The excited‐state dynamics of the S3 state is discussed, and the results are compared with those previously reported for benzamide (BA) to examine the N‐ or C‐terminal‐substituted aromatic effect of the peptide bond. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

2.
Fourier‐transform infrared (FT‐IR), Raman (RS), and surface‐enhanced Raman scattering (SERS) spectra of β‐hydroxy‐β‐methylobutanoic acid (HMB), L ‐carnitine, and N‐methylglycocyamine (creatine) have been measured. The SERS spectra have been taken from species adsorbed on a colloidal silver surface. The respective FT‐IR and RS band assignments (solid‐state samples) based on the literature data have been proposed. The strongest absorptions in the FT‐IR spectrum of creatine are observed at 1398, 1615, and 1699 cm−1, which are due to νs(COOH) + ν(CN) + δ(CN), ρs(NH2), and ν(C O) modes, respectively, whereas those of L ‐carnitine (at 1396/1586 cm−1 and 1480 cm−1) and HMB (at 1405/1555/1585 cm−1 and 1437–1473 cm−1) are associated with carboxyl and methyl/methylene group vibrations, respectively. On the other hand, the strongest bands in the RS spectrum of HMB observed at 748/1442/1462 cm−1 and 1408 cm−1 are due to methyl/methylene deformations and carboxyl group vibrations, respectively. The strongest Raman band of creatine at 831 cm−1w(R NH2)) is accompanied by two weaker bands at 1054 and 1397 cm−1 due to ν(CN) + ν(R NH2) and νs(COOH) + ν(CN) + δ(CN) modes, respectively. In the case of L ‐carnitine, its RS spectrum is dominated by bands at 772 and 1461 cm−1 assigned to ρr(CH2) and δ(CH3), respectively. The analysis of the SERS spectra shows that HMB interacts with the silver surface mainly through the  COO, hydroxyl, and  CH2 groups, whereas L ‐carnitine binds to the surface via  COO and  N+(CH3)3 which is rarely enhanced at pH = 8.3. On the other hand, it seems that creatine binds weakly to the silver surface mainly by  NH2, and C O from the  COO group. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

3.
The strength and geometry of adsorption of substituted propenoic acids on silver surface were studied by means of surface enhanced Raman spectroscopy (SERS) using silver sol. Based on their SERS behavior, two classes of phenylpropenoic acids studied were distinguished. The first class of propenoic acids (atropic acid, (E)‐2,3‐diphenylpropenoic acid, (E)‐2‐(2‐methoxyphenyl)‐3‐phenylpropenoic acid, (E)‐2,3‐di‐(4‐methoxyphenyl)phenylpropenoic acid and (E)‐2‐(2‐methoxyphenyl)‐3‐(4‐fluorophenyl)propenoic acid) has shown strong charge transfer (CT) effect. We suggest bidentate carboxyl bonded species based on the SERS enhanced bands of νCOO around 1394 cm−1 and νC―C of the ―C―COO moiety at 951 cm−1. In these series the plane of the α‐phenyl group (γCH out‐of‐plane vibrations at 850–700 cm−1) is almost parallel to the silver surface, while the β‐phenyl group is in tilted position depending on the type and the position of substituent(s) showing strong SERS enhanced bands of νCC + βCH (in‐plane mode) at 1075 cm−1, νCC (ring breathing mode, in‐plane) at 1000 cm−1 and γCCC (out‐of‐plane mode) around 401 cm−1. The other class of propenoic acids (cinnamic acid, (E)‐2‐phenyl‐3‐(4‐methoxyphenyl)propenoic acid) has shown weak electromagnetic (EM) enhancement (CC bands is enhanced in cinnamic acid). In this case no significant carboxyl enhancement was observed, so we suggest that adsorbed species lie parallel to the surface. The two types of adsorption can be related to the dissociation ability of the carboxylic group. In the first case the carboxylic H dissociates, while in the second case it does not, as indicated also by the characteristic νCO band at 1686 cm−1 in the FT‐Raman spectra of methanolic solutions. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
The S3 radical anion is observed in several systems: non‐aqueous polysulfides solutions, doped alkali halides, ultramarine pigments (UP) for which S3 is the blue chromophore and S2 is the yellow one and pigments of zeolite 4A structure. The S3 ion has C2V symmetry, and therefore its three vibrational modes should be observed in the Raman and in IR spectra. In resonance Raman spectroscopy, only the symmetric stretching mode ν1 and the bending mode ν2 have been observed, whereas the anti‐symmetric stretching mode ν3 has never been observed whatever the system. In this work, we confirm that ν3 is not observed in solutions with resonance Raman spectroscopy. However, our investigation of several blue UP, with various concentrations of S2, shows that there is a superposition of two bands at ca 590 cm−1: the first is assigned to ν (S2) and the second to ν3 (S3). With the 457.9 nm excitation line, for which the resonance conditions are simultaneously fulfilled for S2 and S3, the band at ca 590 cm−1 is the sum of the contributions of both ν (S2) and ν3 (S3) vibrations, while, with the 647.1 nm line, which only satisfies the resonance conditions of S3, the band at ca 584 cm−1 must be assigned only to ν3 (S3). Furthermore, ν3 (S3) is observed in green UP and in pigments of zeolite structure. The ν3 vibration of S3, which is observed neither in polysulfide solutions nor in doped alkali halides in resonance Raman conditions, can therefore be observed when this species is inserted into the β‐cages of the sodalite or of the zeolite 4A structures. So, the band at ca 590 cm−1 cannot always be assigned to S2 in these systems. This implies that the concentration of S2 in UP must be reconsidered. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

5.
Tellurates are rare minerals as the tellurate anion is readily reduced to the tellurite ion. Often minerals with both tellurate and tellurite anions are found. An example of such a mineral containing tellurate and tellurite is yecoraite. Raman spectroscopy has been used to study this mineral, the exact structure of which is unknown. Two Raman bands at 796 and 808 cm−1 are assigned to the ν1(TeO4)2− symmetric and ν3(TeO3)2− antisymmetric stretching modes and Raman bands at 699 cm−1 are attributed to the ν3(TeO4)2− antisymmetric stretching mode and the band at 690 cm−1 to the ν1(TeO3)2− symmetric stretching mode. The intense band at 465 cm−1 with a shoulder at 470 cm−1 is assigned the (TeO4)2− and (TeO3)2− bending modes. Prominent Raman bands are observed at 2878, 2936, 3180 and 3400 cm−1. The band at 3936 cm−1 appears quite distinct and the observation of multiple bands indicates the water molecules in the yecoraite structure are not equivalent. The values for the OH stretching vibrations listed provide hydrogen bond distances of 2.625 Å (2878 cm−1), 2.636 Å (2936 cm−1), 2.697 Å (3180 cm−1) and 2.798 Å (3400 cm−1). This range of hydrogen bonding contributes to the stability of the mineral. A comparison of the Raman spectra of yecoraite with that of tellurate containing minerals kuranakhite, tlapallite and xocomecatlite is made. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
The excited state structural dynamics of 4‐cyanobenzaldehyde (p‐CNB) were studied by using the resonance Raman spectroscopy and the quantum mechanical calculations. The experimental A‐ and B‐band absorptions were, respectively, assigned to the major nO → π3* and π2 → π3* transitions according to the B3LYP‐TD/6‐31G(d) and CIS/6‐31G(d) computations, and the resonance Raman spectra. It was determined that the actual S22π3) state was in energy lower than S31π3), which was just opposite to the B3LYP‐TD/6‐31G(d) calculated order of the S21π3) and S32π3). The vibrational assignments were carried out for the A‐ and B‐band resonance Raman spectra. The B‐band resonance Raman intensities of p‐CNB were dominated by the C2–C3/C5–C6 symmetric stretch mode ν8, the overtones nν8 and their combination bands with the ring C–H bend mode ν17, the C9–N10 stretch mode ν6, the C7–O8 stretch mode ν7 and the remaining modes. The conical intersection between S1(nOπ3) and S22π3) states of p‐CNB was determined at complete active space self‐consistent field (CASSCF)(8,7)/6‐311G(d,p) level of theory. The B‐band short‐time structural dynamics and the corresponding decay dynamics of p‐CNB were obtained by analysis of the resonance Raman intensity pattern and CASSCF computations. The resonance Raman spectra indicated that CI[S1(nOπ3)/S21π2π3π4)] located nearby the Franck–Condon region. The excited state decay dynamics evolving from the S2, FC2π3) to the S1(nOπ3) state was proposed. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

7.
Raman spectroscopy has been used to study vanadates in the solid state. The molecular structure of the vanadate minerals vésigniéite [BaCu3(VO4)2(OH)2] and volborthite [Cu3V2O7(OH)2·2H2O] have been studied by Raman spectroscopy and infrared spectroscopy. The spectra are related to the structure of the two minerals. The Raman spectrum of vésigniéite is characterized by two intense bands at 821 and 856 cm−1 assigned to ν1 (VO4)3− symmetric stretching modes. A series of infrared bands at 755, 787 and 899 cm−1 are assigned to the ν3 (VO4)3− antisymmetric stretching vibrational mode. Raman bands at 307 and 332 cm−1 and at 466 and 511 cm−1 are assigned to the ν2 and ν4 (VO4)3− bending modes. The Raman spectrum of volborthite is characterized by the strong band at 888 cm−1, assigned to the ν1 (VO3) symmetric stretching vibrations. Raman bands at 858 and 749 cm−1 are assigned to the ν3 (VO3) antisymmetric stretching vibrations; those at 814 cm−1 to the ν3 (VOV) antisymmetric vibrations; that at 508 cm−1 to the ν1 (VOV) symmetric stretching vibration and those at 442 and 476 cm−1 and 347 and 308 cm−1 to the ν4 (VO3) and ν2 (VO3) bending vibrations, respectively. The spectra of vésigniéite and volborthite are similar, especially in the region of skeletal vibrations, even though their crystal structures differ. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

8.
The process and the formation of new minerals upon heating carbonate rocks containing clay minerals together with dolomite are determined by thermal analysis, X‐ray diffraction (XRD), infrared and Raman spectroscopy. The dolomite–calcite–calcium oxide phase transition sequences were followed up to 947 °C in a naturally occurring dolomite sample. The spectral variations of the internal modes of the carbonate trigonal (ν1, ν2, ν3 and ν4) were used to probe the structural phase transitions. A new Raman mode emerged at 1090 cm−1 in the ν1 mode region, and infrared modes emerged at 713, 874, and 1420 cm−1 in the ν4, ν2 and ν3 regions at 750 °C, indicating the onset of the dolomite phase. The calcium oxide phase, (which on reaction with atmospheric water forms portlandite) with an onset temperature of around 950 °C, was also characterized by the appearance of the infrared mode around 450 cm−1. The minerals, which were formed upon heating the dolomite, were calcite, calcium oxide and diopside. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

9.
The structures of the naturally occurring sulfite‐bearing minerals scotlandite, hannebachite and orschallite have been studied by Raman spectroscopy. Raman bands are observed for scotlandite PbSO3 at 935, 880, 622 and 474 cm−1 and are assigned to the (SO3)2−ν1(A1), ν3(E), ν2(A1) and ν4(E) vibrational modes, respectively. For hannebachite (CaSO3)2·H2O these bands are observed at 1005, 969 and 655 cm−1 with multiple bands for the ν4(E) mode at 444, 492 and 520 cm−1. The Raman spectrum of hannebachite is very different from that of the compound CaSO3·2H2O. It is proposed, on the basis of Raman spectroscopy, that in the mineral hannebachite, the sulfite anion bonds to Ca through the sulfur atom. The Raman spectrum of the mineral orschallite Ca3[SO4](SO3)2·12H2O is complex resulting from the overlap of sulfate and sulfite bands. Raman bands at 1005 cm−1, 1096 and 1215 cm−1 are assigned to the (SO4)2−ν1 symmetric and ν3 asymmetric stretching modes. The two Raman bands at 971 and 984 cm−1 are attributed to the (SO3)2−ν3(E) and ν1(A1) stretching vibrations. The formation of sulfite compounds in nature offers a potential mechanism for the removal of sulfates and sulfites from soils. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

10.
Selenites and tellurites may be subdivided according to formula and structure. There are five groups, based upon the formulae (a) A(XO3), (b) A(XO3·) xH2O, (c) A2(XO3)3·xH2O, (d) A2(X2O5) and (e) A(X3O8). Of the selenites, molybdomenite is an example of type (a); chalcomenite, clinochalcomenite, cobaltomenite and ahlfeldite are minerals of type (b); mandarinoite Fe2Se3O9·6H2O is an example of type (c). Raman spectroscopy has been used to characterise the mineral mandarinoite. The intense, sharp band at 814 cm−1 is assigned to the symmetric stretching (Se3O9)6− units. Three Raman bands observed at 695, 723 and 744 cm−1 are attributed to the ν3 (Se3O9)6− anti‐symmetric stretching modes. Raman bands at 355, 398 and 474 cm−1 are assigned to the ν4 and ν2 bending modes. Raman bands are observed at 2796, 2926, 3046, 3189 and 3507 cm−1 and are assigned to OH stretching vibrations. The observation of multiple OH stretching vibrations suggests the non‐equivalence of water in the mandarinoite structure. The use of the Libowitzky empirical function provides hydrogen bond distances of 2.633(9) Å (2926 cm−1), 2.660(0) Å (3046 cm−1), 2.700(0) Å (3189 cm−1) and 2.905(3) Å (3507 cm−1). The sharp, intense band at 3507 cm−1 may be due to hydroxyl units. It is probable that some of the selenite units have been replaced by hydroxyl units. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

11.
Raman spectroscopy has been used to characterise four natural halotrichites: halotrichite FeSO4.Al2(SO4)3. 22H2O, apjohnite MnSO4.Al2(SO4)3.22H2O, pickingerite MgSO4.Al2(SO4)3.22H2O and wupatkiite CoSO4.Al2(SO4)3.22H2O. A comparison of the Raman spectra is made with the spectra of the equivalent synthetic pseudo‐alums. Energy dispersive X‐ray analysis (EDX) was used to determine the exact composition of the minerals. The Raman spectrum of apjohnite and halotrichite display intense symmetric bands at ∼985 cm−1 assigned to the ν1(SO4)2− symmetric stretching mode. For pickingerite and wupatkiite, an intense band at ∼995 cm−1 is observed. A second band is observed for these minerals at 976 cm−1 attributed to a water librational mode The series of bands for apjohnite at 1104, 1078 and 1054 cm−1, for halotrichite at 1106, 1072 and 1049 cm−1, for pickingerite at 1106, 1070 and 1049 cm−1 and for wupatkiite at 1106, 1075 and 1049 cm−1 are attributed to the ν3(SO4)2− antisymmetric stretching modes of ν3(Bg) SO4. Raman bands at around 474, 460 and 423 cm−1 are attributed to the ν2(Ag) SO4 mode. The band at 618 cm−1 is assigned to the ν4(Bg) SO4 mode. The splitting of the ν2, ν3 and ν4 modes is attributed to the reduction of symmetry of the SO4 and it is proposed that the sulphate coordinates to water in the hydrated aluminium in bidentate chelation. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

12.
Raman spectroscopy at both 298 and 77 K has been used to study a series of selected natural smithsonites from different origins. An intense sharp band at 1092 cm−1 is assigned to the CO32− symmetric stretching vibration. Impurities of hydrozincite are identified by a band around 1060 cm−1. An additional band at 1088 cm−1 which is observed in the 298 K spectra but not in the 77 K spectra is attributed to a CO32− hot band. Raman spectra of smithsonite show a single band in the 1405–1409 cm−1 range assigned to the ν3 (CO3)2− antisymmetric stretching mode. The observation of additional bands for the ν3g modes for some smithsonites is significant in that it shows distortion of the ZnO6 octahedron. No ν2 bending modes are observed for smithsonite. A single band at 730 cm−1 is assigned to the ν4 in phase bending mode. Multiple bands be attributed to the structural distortion are observed for the carbonate ν4 in phase bending modes in the Raman spectrum of hydrozincite with bands at 733, 707 and 636 cm−1. An intense band at 304 cm−1 is attributed to the ZnO symmetric stretching vibration. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

13.
The removal of arsenate anions from aqueous media, sediments and wasted soils is of environmental significance. The reaction of gypsum with the arsenate anion results in pharmacolite mineral formation, together with related minerals. Raman and infrared (IR) spectroscopy have been used to study the mineral pharmacolite Ca(AsO3OH)· 2H2O. The mineral is characterised by an intense Raman band at 865 cm−1 assigned to the ν1 (AsO3)2− symmetric stretching mode. The equivalent IR band is found at 864 cm−1. The low‐intensity Raman bands in the range from 844 to 886 cm−1 provide evidence for ν3 (AsO3) antisymmetric stretching vibrations. A series of overlapping bands in the 300‐450 cm−1 region are attributed to ν2 and ν4 (AsO3) bending modes. Prominent Raman bands at around 3187 cm−1 are assigned to the OH stretching vibrations of hydrogen‐bonded water molecules and the two sharp bands at 3425 and 3526 cm−1 to the OH stretching vibrations of only weakly hydrogen‐bonded hydroxyls in (AsO3OH)2− units. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
Raman spectroscopy, complemented with infrared spectroscopy, was used to study the uranyl carbonate mineral voglite. The mineral has the formula Ca2Cu2+ [(UO2)(CO3)3](CO3)6H2O, and bands attributed to these vibrating units are readily identified in the Raman spectrum. Symmetric stretching modes at 836 and 1094 cm−1 are assigned to ν1(UO2)2+ and ν1(CO3)2− units, respectively. The ν3 antisymmetric stretching modes of (UO2)2+ are not observed in the Raman spectrum but may be readily observed in the infrared spectrum at 898 cm−1. The ν3 antisymmetric stretching mode of (CO3)2− is observed in the Raman spectrum at 1369 cm−1 as a low intensity band as is also the ν3(CO3)2− infrared modes at 1362, 1425, 1509 and 1566 cm−1. No ν2(CO3)2− Raman bending modes are observed for voglite. The Raman band at 749 cm−1 and the two infrared bands at 747 and 709 cm−1 are assigned to the ν4(CO3)2− bending modes. U O bond and O H…O bond lengths in the structure of voglite were inferred from the infrared and Raman spectra. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

15.
Tellurites may be subdivided according to formula and structure. There are five groups based upon the formulae (a) A(XO3), (b) A(XO3)·xH2O, (c) A2(XO3)3·xH2O, (d) A2(X2O5) and (e) A(X3O8). Raman spectroscopy has been used to study the tellurite minerals teineite and graemite; both contain water as an essential element of their stability. The tellurite ion should show a maximum of six bands. The free tellurite ion will have C3v symmetry and four modes, 2A1 and 2 E. Raman bands for teineite at 739 and 778 cm−1 and for graemite at 768 and 793 cm−1 are assigned to the ν1 (TeO3)2− symmetric stretching mode while bands at 667 and 701 cm−1 for teineite and 676 and 708 cm−1 for graemite are attributed to the ν3 (TeO3)2− antisymmetric stretching mode. The intense Raman band at 509 cm−1 for both teineite and graemite is assigned to the water librational mode. Raman bands for teineite at 318 and 347 cm−1 are assigned to the (TeO3)2−ν2(A1) bending mode and the two bands for teineite at 384 and 458 cm−1 may be assigned to the (TeO3)2−ν4(E) bending mode. Prominent Raman bands, observed at 2286, 2854, 3040 and 3495 cm−1, are attributed to OH stretching vibrations. The values for these OH stretching vibrations provide hydrogen bond distances of 2.550(6) Å (2341 cm−1), 2.610(3) Å (2796 cm−1) and 2.623(2) Å (2870 cm−1) which are comparatively short for secondary minerals. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

16.
《Molecular physics》2012,110(17):2063-2069
The high resolution infrared absorption spectrum of CH2D81Br has been recorded by Fourier transform spectroscopy in the range 550–1075?cm?1, with an unapodized resolution of 0.0025?cm?1, employing a synchrotron radiation source. This spectral region is characterized by the ν6 (593.872?cm?1), ν5 (768.710?cm?1) and ν9 (930.295?cm?1) fundamental bands. The ground state constants up to sextic centrifugal distortion terms have been obtained for the first time by ground-state combination differences from the three bands and subsequently employed for the evaluation of the excited state parameters. Watson's A-reduced Hamiltonian in the Ir representation has been used in the calculations. The ν 6?=?1 level is essentially free from perturbation whereas the ν 5?=?1 and ν 9?=?1 states are mutually interacting through a-type Coriolis coupling. Accurate spectroscopic parameters of the three excited vibrational states and a high-order coupling constant which takes into account the interaction between ν5 and ν9 have been determined.  相似文献   

17.
Raman spectroscopy has been used to study zemannite Mg0.5[Zn2+Fe3+(TeO3)3]4.5H2O and emmonsite Fe23+Te34+O9·2H2O. Raman bands for zemannite and emmonsite, observed at 740 and 650 cm−1 and at 764 and 788 cm−1, respectively, are attributed to the ν1 (TeO3)2− symmetric stretching mode. The splitting of the symmetric stretching mode for emmonsite is in harmony with the results of X‐ray crystallography which shows three non‐equivalent TeO3 units in the crystal structure. Two bands at 658 and 688 cm−1 are assigned to ν3 (TeO3)2− anti‐symmetric stretching modes. Raman bands observed at 372 and 408 cm−1 for zemannite and 397 and 414 cm−1 for emmonsite are attributed to the (TeO3)2−ν2(A1) bending mode. The two Raman bands at 400 and 440 cm−1 for emmonsite are ascribed to the ν4(E) bending modes, while the band at 326 cm−1 is due to the ν2(A1) bending vibration. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

18.
The arsenite mineral finnemanite Pb5(As3+ O3)3Cl has been studied by Raman spectroscopy. The most intense Raman band at 871 cm−1 is assigned to the ν1(AsO3)3 symmetric stretching vibration. Three Raman bands at 898, 908 and 947 cm−1 are assigned to the ν3(AsO3)3− antisymmetric stretching vibration. The observation of multiple antisymmetric stretching vibrations suggest that the (AsO3)3− units are not equivalent in the molecular structure of finnemanite. Two Raman bands at 383 and 399 cm−1are assigned to the ν2(AsO3)3− bending modes. Density functional theory enabled calculation of the position of AsO32− symmetric stretching mode at 839 cm−1, the antisymmetric stretching mode at 813 cm−1 and the deformation mode at 449 cm−1. Raman bands are observed at 115, 145, 162, 176, 192, 216 and 234 cm−1 as well. The two most intense bands are observed at 176 and 192 cm−1. These bands are assigned to PbCl stretching vibrations and result from transverse/longitudinal splitting. The bands at 145 and 162 cm−1 may be assigned to Cl Pb Cl bending modes. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
Raman and infrared spectroscopy enabled insights into the molecular structure of the sampleite group of minerals. These minerals are based upon the incorporation of either phosphate or arsenate with chloride anion into the structure, and as a consequence the spectra reflect the bands attributable to these anions, namely, phosphate or arsenate with chloride. The sampleite vibrational spectrum reflects the spectrum of the phosphate anion and consists of ν1 at 964 cm−1, ν2 at 451 cm−1, ν3 at 1016 and 1088 cm−1 and ν4 at 643, 604, 591 and 557 cm−1. The lavendulan spectrum consists of ν1 at 854 cm−1, ν2 at 345 cm−1, ν3 at 878 cm−1 and ν4 at 545 cm−1. The Raman spectrum of lemanskiite is different from that of lavendulan consistent with a different structure. Low wavenumber bands at 227 and 210 cm−1 may be assigned to CuCl transverse/longitudinal (TO/LO) optic vibrations. Raman spectroscopy identified the substitution of arsenate by phosphate in zdenekite and lavendulan. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

20.
The mineral marthozite, a uranyl selenite, has been characterised by Raman spectroscopy at 298 K. The bands at 812 and 797 cm−1 were assigned to the symmetric stretching modes of the (UO2)2+ and (SeO3)2− units, respectively. These values gave the calculated U O bond lengths in uranyl of 1.799 and/or 1.814 Å. Average U O bond length in uranyl is 1.795 Å, inferred from the X‐ray single crystal structure analysis of marthozite by Cooper and Hawthorne. The broad band at 869 cm−1 was assigned to the ν3 antisymmetric stretching mode of the (UO2)2+ (calculated U O bond length 1.808 Å). The band at 739 cm−1 was attributed to the ν3 antisymmetric stretching vibration of the (SeO3)2− units. The ν4 and the ν2 vibrational modes of the (SeO3)2− units were observed at 424 and 473 cm−1. Bands observed at 257, and 199 and 139 cm−1 were assigned to OUO bending vibrations and lattice vibrations, respectively. O H···O hydrogen bond lengths were inferred using Libowiztky's empirical relation. The infrared spectrum of marthozite was studied for complementation. Copyright © 2008 John Wiley & Sons, Ltd.  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