Sie sind auf Seite 1von 6

Vibrational Spectroscopy 52 (2010) 122–127

Contents lists available at ScienceDirect

Vibrational Spectroscopy
journal homepage: www.elsevier.com/locate/vibspec

Spectroscopic characterization of natural chrysotile


G. Anbalagan a,*, G. Sivakumar b, A.R. Prabakaran c, S. Gunasekaran c
a
P.G. and Research Department of Physics, Presidency College, Kamaraj Salai, Chepauk, Chennai 5, India
b
CISL, Department of Physics, Annamalai University, Chidambaram 2, India
c
P.G. and Research Department of Physics, Pachaiyappa’s College, Chennai 30, India

A R T I C L E I N F O A B S T R A C T

Article history: Chrysotile, a variety of serpentine mineral was investigated by vibrational and magnetic resonance
Received 30 June 2009 spectroscopic methods. The SEM photograph of the sample shows that chrysotile is made up of thin and
Received in revised form 27 November 2009 flexible fibrils, classified as asbestos. The calculated lattice parameters were a = 5.1961  0.0053 Å,
Accepted 28 November 2009
b = 9.2365  0.0107 Å, c = 14.6208  0.0039 Å; b = 90.15  0.198; Z = 4, with space group C2/m. Three bands
Available online 5 December 2009
observed in the highest wavenumber region (3689 and 3648 cm 1) originate from the stretching vibrations
of the two crystallographically different OH groups. The Raman spectrum shows bending vibrations of the
Keywords:
SiO4 tetrahedra at 390 and 348 cm 1. Substitution of the Si ions in SiO4 tetrahedra by Al ions leads to low-
Chrysotile
field shifts in the 29Si NMR spectrum and the resonance line corresponding to Q3 (1 Al) was observed at
FT-IR
FT-Raman 89 ppm. This result was supported by the resonance line at 57 ppm in 27Al NMR spectrum.
EPR ß 2009 Elsevier B.V. All rights reserved.
29
Si NMR
27
Al NMR

1. Introduction only one type of octahedral site. Fe, Ni and Cr may substitute to
some extent for Mg and Al for some Si. The substitution of Mg2+ by
Chrysotile, a fibrous variety of serpentine with chemical impurities like Fe2+, Mn2+, etc., shows that they must be also
formula 2[Mg3Si2O5(OH)4] is the most important source of present in octahedral site symmetry in the mineral in high spin
commercial asbestos. It is a hydrous secondary mineral, resulting environment. The octahedral site symmetry of these ions is further
from the alteration of magnesium silicates, especially olivine, confirmed by the data reported earlier by Heller-Kalai et al. [4],
pyroxene and amphibole. It may also occur as a replacement of the which states that in almost all serpentines Fe2+ and Fe3+ are only in
original mineral forming a pseudomorph or it may be deposited octahedral coordination. The release and retention of plant foods,
elsewhere in the mineral openings or as replacing some other the stockpiling of water in the soil from wet to dry seasons, and the
mineral. Chrysotile, a two-layered silicate, is composed of one accessibility of the soil to atmospheric gases and organism depend
tetrahedral and one octahedral layer. The first layer is a hexagonal in large part on the properties of the sheet silicates. There is a
net of silicon–oxygen tetrahedral, in which three of the four ions continuing interest in the investigation on chrysotile because of
are linked. The second layer is like a brucite octahedral layer with the health hazards associated with asbestos, of which chrysotile
Mg2+ ions in octahedral environment of 2O2 and 4(OH) ions accounts for 95% of world production [5,6]. The main purpose of
instead of 6(OH) in brucite [1]. The bonds between the packets are this work is to study the effect due to hydroxyl stretching modes
of van der Waals type [2]. In the case of chrysotile the mismatch and to determine the site symmetry of iron and other impurities
between a brucite type layer and an undistorted Si2O5 sheet with present in the mineral chrysotile.
hexagonal rings is resolved by a continuous bending of the
structure into the cylindrical tubes [3]. High magnification electron
micrographs have shown the existence of such a tube-like 2. Experimental data
structure for the layers. The lattice parameters of chrysotile are
a = 5.1961 Å, b = 9.2365 Å, c = 14.6208 Å; b = 90.15; Z = 4, with 2.1. Samples
space group C2/m. The survey of literature indicates that there is
Serpentine mineral used in the present investigation, was
collected from Salem, Tamil Nadu, India. The crystals were
* Corresponding author. Tel.: +91 44 26602331; fax: +91 44 28510732. mechanically separated from the ore and used in a powdered
E-mail address: anbu24663@yahoo.co.in (G. Anbalagan). form for vibrational measurements.

0924-2031/$ – see front matter ß 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.vibspec.2009.11.007
G. Anbalagan et al. / Vibrational Spectroscopy 52 (2010) 122–127 123

2.2. Instrumentation Table 1


Elemental analysis of natural chrysotile.

XRPD analysis was performed with a SEIFERT X-ray diffractom- Element wt.%
eter with Cu Ka radiation (l = 1.54 Å), Cu filter on secondary Mg 40.47
optics, 45 kV power and 20 mA current. The morphological study Si 56.84
was performed with a Jeol JSM-6360 SEM. The sample was glued Ti 0.67
onto aluminum tubes with colloidal graphite and then coated with Cr 0.27
Mn 0.46
a carbon film approximately 400 Å in thickness.
Fe 1.01
Perkin Elmer spectrum one FT-IR spectrometer was used, and Al 0.30
the samples were analyzed in KBr pellets. Spectra were traced in
the range 4000–400 cm 1, and the band intensities were expressed
in transmittance (%).
Bruker IFS FT-IR spectrometer with FRA Raman attachment was enhanced by the iso-orientation of the constituting crystals. The
used and Nd:YAG laser of wavelength 1064 nm was used as the calculated unit cell parameters are compiled in Table 2.
source. The spectra were traced in the range 3500–50 cm 1 with
500 scans at 0.1 cm 1 resolution. Both infrared and Raman analysis 3.2. Infrared spectral analysis
permitted the identification of the main molecular groups present
in the samples. IR spectrum of chrysotile presented in Fig. 3 exhibits three
A part of the powdered sample is taken into a quartz tube for well-defined spectral regions. The first group of bands appears in
EPR recordings on a Varian E-112 EPR spectrometer operated at X- the 3800–3600 cm 1 region, the second in the 1200–800 cm 1 and
band frequency having a 100 kHz field modulation and a phase the third region includes the bands below 800 cm 1. Two bands
sensitive detector to obtain a first derivative signal. DPPH is used as observed at 3689 cm 1 (strong) and 3648 cm 1 (weak) (Fig. 3) are
an internal standard (g = 2.0036). Approximately 2 mg of pow- assigned to in-phase outer and inner Mg–OH stretch, respectively
dered sample was used to acquire the 29Si and 27Al NMR spectra [8–11]. These two bands could be attributed to the presence of two
under magic-angle spinning (MAS) conditions. The spectrometer
used was DSX 300. Spectra were obtained at a frequency
59.62 MHz for 29Si and 78.19 MHz for 27Al. Rapid (8–9 kHz)
sample spinning at the magic angle to the external magnetic field
was used. In most cases, conventional FT NMR with 7 ms radio
frequency pulses and 5 s time intervals was used.

3. Results and discussion

3.1. Powder X-ray diffraction analysis

The morphology of the sample represented in Fig. 1 shows that


chrysotile is made up of thin and flexible fibrils, classified as
asbestos. The chemical composition of the sample obtained by
energy dispersive analysis is given in Table 1. X-ray powder
diffraction analysis shows that chrysotile is composed of bundles
of parallel fibers and each individual fiber is elongated parallel to
one crystallographic direction (Fig. 2). The peak positions of the
observed pattern match those reported on the ICDD file no. 25-
0645. The principal reflections occurring at the d-spacing 7.3110,
4.5287, 3.6548, 2.4460 and 1.5325 Å confirm the presence of
chrysotile as a major mineral phase in the sample [7]. The intensity Fig. 2. Powder X-ray diffraction pattern of natural chrysotile studied.
of peaks at the d-spacings 3.6548, 2.4460 and 1.5325 Å was

Fig. 1. SEM image of the chrysotile shows constituent fibers. Fig. 3. Infrared spectrum of chrysotile recorded at room temperature.
124 G. Anbalagan et al. / Vibrational Spectroscopy 52 (2010) 122–127

Table 2
Crystallographic data on chrysotile mineral.

Mineral Formula Space group Cell constants

a (Å) b (Å) c (Å) b (8) V (Å)3

Chrysotile 2[Mg3Si2O5(OH)4] C2/m; Z = 4 5.1961  0.0053 9.2365  0.0107 14.6208  0.0039 90.15  0.19 701.707

Table 3
Band assignments in the powder infrared spectrum of natural chrysotile.

This work Viti and Mellini [13] Farmer [16] Foresti et al. [17] Sontevska et al. [9] Tentative assignments [12,13,19]

3689 s 3688 s 3697 s 3697 s 3688 s n(OH)


3648 w 3640 w 3651 w 3650 w 3646 w n(OH)
3444 m, br – – – – H2O stretcha
1635 m – – – – H2O benda
1079 s 1072 m 1079 s 1080 s 1078 s n(Si–O–Si)
1018 m 1020 w 1022 m 1020 m 1022 sh n(Si–O–Si)
n(Si–O–Mg)a
965 s 956 vs 957 vs 965 vs 961 vs n(Si–O–Si)
– 635 sh 654 sh – – –
610 s 608 s 604 s 610 s 611 s d(OH)
556 sh 560 sh 550 sh 560 s 556 vw Out-of-plane bending mode of Si–Ob
482 vw 490 sh 481 w – – Out-of-plane bending mode of the Mg octahedrab
– – 450 sh 460 vw 455 sh –
– 437 vs 432 vs 440 vs 438 vs d(Si–O–Si)
– 425 vw – 415 w 405 w –

s, strong; w, weak; m, medium; sh, shoulder; v, very; br, broad.


a
According to Falini et al. [19].
b
According to Viti and Mellini [13] and Suquet [12].

crystallographically distinct OH groups in the chrysotile structure 3.3. Raman spectral analysis
[9]. Two broad bands observed at 3444 and 1635 cm 1 are due to
the stretching and bending vibrations of water molecules adsorbed The Raman spectrum of chrysotile is presented in Fig. 4. The
to the chrysotile tubes [11,12]. Three bands at 1079, 1018 and observed bands and their vibrational assignments given in Table 4
965 cm 1 are registered in the 1200–800 cm 1 region and are in agreement with the corresponding literature data [8–10,19–
prescribed to the vibrations of the SiO4 tetrahedra (Table 3). 24]. The Raman bands present in the frequency range at about 1082
Precisely, the first band (1079 cm 1) is attributed to the Si–O–Si and 956 cm 1 are typical for Si–O–Si stretching vibration of the silica
stretching vibrations, perpendicular to the basal plane, whereas network. The highest frequency band in the SiO4 region observed at
the shoulder at 1018 cm 1 and the most intensive band at 1105 cm 1 arise from the antisymmetric stretching vibrations of the
965 cm 1 arise from the Si–O–Si vibrations in the basal plane Si–Onb–Si groups (nb – non-bridging oxygen) [9]. The most intense
[9,13]. The occupancy of the octahedral sheet in both 1:1 or in 2:1 bands which lie usually near 690 and 390 cm 1 are assigned,
minerals strongly influences the position of the OH bending bands. respectively, to symmetric stretching modes (ns) of the Si–Ob–Si
Thus dioctahedral minerals absorb in the 950–800 cm 1 region linkages and to bending modes n5(e) of the SiO4 tetrahedra,
while the OH absorption of trioctahedral minerals is shifted to respectively [9,20,21]. According to Kloprogge et al. [21], the band
lower frequencies in the 700–600 cm 1 range [14] corresponding- around 625 cm 1 appears from the Mg–OH stretching mode. The
ly a strong band is observed at 610 cm 1 in the present study. A Raman band observed at 464 cm 1 ascribed to Si–O bending
weak band observed at 482 cm 1 being attributed to the bending vibration associated with an Mg–OH translation vibration. The
vibrations of SiO4 tetrahedral, respectively [9,13,15–17]. The bands observed near 348,325 and 304 cm 1 are in agreement with
sharpness of OH absorptions at 3648 and 610 cm 1 indicates that the literature data [20,21]. These bands arise from the SiO4 bending
chrysotile has a relatively ordered structure. Kraineva et al. [18] vibrations [20,21]. The band observed at 232 cm 1, according to
used infrared spectroscopy to distinguish chrysotile of normal Rinaudo et al. [20] is due to the vibrations of the Onb. . .H–O groups.
strength, fractured and weathered specimens. In the fractured The band at 205 cm 1 arises from the stretching Mg–O mode (with
material, there is a new band at 1045 cm 1, while the 1020 cm 1 A1g symmetry) of the distorted MgO6 octahedra [21].
band occurs as a shoulder. In the weathered specimens, these two
bands fuse into a broad shoulder on the 960 cm 1 band. The
infrared spectrum of chrysotile obtained in the present study
indicates that the specimen is not a fractured or weathered one.
Kraineva et al. [18] attributed that the change in the IR spectra is
due to the reduction in the point-group symmetry of Si2O5 from C3v
to C2v. The vibrations of H2O are seen as broad bands in the regions
3600–3200 cm 1 (stretching modes) and 1500–1650 cm 1 (defor-
mation modes). Kraineva et al. [18] further suggested that these
bands are not very prominent for the fractured and weathered
chrysotile, particularly in the region of the stretching vibrations.
The water contents of these specimens were less than those for
chrysotile of normal strength. Fig. 2 exhibits broad bands at 3444
and 1635 cm 1 which also confirms that the chrysotile studied in Fig. 4. FT-Raman spectrum of chrysotile recorded with excitation wavelength
the present work is not a fractured or weathered one. 1064 nm.
G. Anbalagan et al. / Vibrational Spectroscopy 52 (2010) 122–127 125

Table 4
1
Wavenumber position (cm ) for vibrational bands observed in the Raman spectrum: excitation wavelength 1064 nm.

This work Sontevska et al. [9] Rinaudo et al. [20] Kloprogge et al. [21] Bard et al. [22] Tentative assignment [20,21]

1105 1106 m 1105 m 1102 1105 nas(Si–Onb–Si)


692 692 vs 692 vs 692 692 n(Si–O–Si)
622 626 m 620 m 629 623 n(Mg–OH)
464 – – 466 465 Mg–OH translation + n6(e)SiO4
438 – – 432 432 Antisymmetric Mg–OH translation
390 392 vs 389 vs 388 – Symmetric n5(e)SiO4
348 348 m 345 m 345 345 Bending of SiO4
325 320 w – 318 321 Bending of SiO4
304 306 w – 304 304 Bending of SiO4
232 235 s 231 s 231 – Onb  H–O vibrations
180 206 w – 199 – n(Mg–O)
s, strong; w, weak; m, medium; sh, shoulder; v, very; br, broad.

A strong band observed at 390 cm 1 indicates that this mineral rich and Fe2+-rich phyllosilicates, respectively. In the present study
is Mg bearing phyllosilicates [24]. Rinaudo et al. [20] assigned the the combination bands observed at 2311 and 2344 nm indicate the
band observed at 622 cm 1 to OH–Mg–OH translation modes, sample under study is Mg-rich phyllosilicate. The spectral band
however, Farmer [16] attributed the strong bands observed in the near 2311 nm is strongest with a doublet present in chrysotile as
infrared spectrum in the region 611–646 cm 1 to inner and surface reported by King and Clark [27] attributed to Mg–OH [28]. The
hydroxyl groups of the layers. Most trioctahedral phyllosilicates chrysotile secondary combination bands were observed at 2395
have strong bands around 230 and 200 cm 1and hence chrysotile and 2474 nm. Further Gionis et al. [29] attributed an overtone
spectrum exhibits a strong band at 232 cm 1. In trioctahedral band at1383 nm and several combination bands are indicative of a
phyllosilicates, the occupancy of almost all the octahedral sites by trioctahedral Mg3OH component. The first overtone (2nOH) bands
divalent cations prevents the formation of an O–H–O isosceles appear at wavenumbers 1383 nm (7231 cm 1) less than twice those
triangle which changes the symmetry of the group consequently of the fundamental bands (nOH = 3689 cm 1), due to the anharmonic
only one band is observed in the 200–300 cm 1 range in the Raman character of vibrations, X = W2nOH/2 WnOH, with X being the
spectrum of the trioctahedral phyllosilicates [20] studied. anharmonicity constant [30]. Petit et al. [30] used talc samples with
various crystal chemistries to solve the equation and the experi-
3.4. Near infrared spectral analysis mental data are well fitted with X = 85.6 cm 1. Petit et al. [30]
further suggested that the anharmonicity constant remains almost
The NIR spectrum of chrysotile contains three overtones and unchanged for several types of clay samples. The calculated
combination bands near: (a) 1481 nm, (b) 1927 nm, and (c) X = 73.5 is close to that reported by Petit et al. [30]. The high
2311 nm (Fig. 5), corresponding to (a) O–H vibrations in hydroxyl, anharmonicity also accounts for the relatively high intensity of these
(b) H–O–H vibrations in crystalline structure, and (c) M–OH bands in the NIR spectrum because these overtones would not be
vibrations in clay minerals (M = metals, Mg2+ in this case). The detectable if the OH vibrations in chrysotile were harmonic [10].
absorptions at 1481, 1927, and 2311 nm can be attributed to the
29
presence of Mg-rich TOT clay mineral [25]. The sharp and well 3.5. Si NMR spectral analysis
resolved OH combination bands in the 2200–2500 nm regions
provide unique fingerprints for specific minerals. According to The room temperature 29Si NMR spectrum of chrysotile sample is
Bishop et al. [26], Al-rich phyllosilicates exhibit OH combination given in Fig. 6. The 29Si NMR spectrum shows a sharp resonance at
bands near 2200 nm, while these bands are observed near 2290–
2310, 2330–2340 nm and near 2350–2370 nm for Fe3+-rich, Mg-

29
Fig. 5. Near infrared spectrum of chrysotile. Fig. 6. Si NMR spectrum of chrysotile.
126 G. Anbalagan et al. / Vibrational Spectroscopy 52 (2010) 122–127

27
3.6. Al NMR spectral analysis

27
Al NMR spectrum of chrysotile sample shown in Fig. 8
consists in general of a strongly broad central band accompanied at
both sides by one or two symmetrically located peaks of low
intensity which represent spinning side bands. According to the
previous work [37,38] the central line at 57 ppm must assign to
octahedral Al ions. These assignments are in agreement with the
structural composition of these samples. The spectrum shows that
Al is in tetrahedral coordination and the Al appears to be
disordered in the tetrahedral coordination, as indicated by the
broad resonance at 57 ppm.

4. Conclusions
Fig. 7. Powder X-band EPR spectrum of natural chrysotile.
The SEM micrograph of the sample shows that chrysotile is
made up of thin and flexible fibrils, classified as asbestos. Infrared
89 ppm of layer structured phyllosilicate (Q3) minerals. The spectrum indicates the presence of two crystallographically
observed peak at 89 is attributed to isolated silanol groups that are distinct OH groups in the chrysotile structure. The spectrum
present in the internal tridimite layer [31]. The chemical shift of the exhibits broad bands at 3444 and 1635 cm 1 which also confirms
single Si resonance ( 89 ppm) is in reasonable agreement with the that the experimental chrysotile is not a fractured or weathered
one reported value of 93 ppm for serpentine of unspecified one. The Raman spectrum is characterized by the main bands
morphology [32]. The only crystalline phase detectable in this observed at 1105, 692 and 622 cm 1and these bands assigned
sample is chrysotile, which is also evident from the resonance at according to the description of the vibrational bands in the infrared
89 ppm. Magi et al. [32] suggested that increasing degree of spectrum of phyllosilicates. The spectra of clay minerals show Si–O
condensation of the SiO4 tetrahedra leads to high-field shifts and stretching and bending absorptions in the region 1300–400 cm 1.
substitution of the Si ions in SiO4 tetrahedra by Al ions leads to low- The three bands were observed at 1079, 1018 and 965 cm 1 and
field shifts. The second trend is observed in the present study and are attributed to Si–O stretching modes. The presence of a single
resonance line observed at 89 ppm is assigned to Q3 (1 Al). This line at 89 ppm in 29Si NMR spectrum indicates the existence of a
assignment is in agreement with that reported by Lippmaa et al. unique environment for Si Q3 (1 Al). The EPR spectrum of the
[33]. The presence of a single line indicates the existence of a unique chrysotile sample shows intense ferromagnetic absorptions. The
environment for Si. 27
Al NMR spectrum shows that Al is in tetrahedral coordination. A
The EPR spectrum of the chrysotile sample of Fig. 7 shows narrow width of the resonance observed is the characteristic of
intense ferromagnetic (or super paramagnetic) absorptions [34]. most highly siliceous system.
This result strongly indicates that the spinning sidebands observed
in many natural aluminosilicates may be due not to chemical shift References
anisotropy, but rather to the presence of large magnetic
susceptibility broadenings [35]. A narrow width of the resonance [1] A.V. Milovisky, O.V. Kononov, Mineralogy, Mir Publications, Moscow, 1985.
[2] F.J. Wicks, E.J.W. Whittaker, Can. Mineral. 13 (1975) 227.
observed is the characteristic of most highly siliceous system. [3] S. Hurlbut Cornelius Jr., K. Cornelis, Manual of Mineralogy, John Wiley & Sons,
When no aluminum is present, the line broadening is removed. New York, 1977.
Further, when aluminum is present in a symmetrical environment [4] L. Heller-Kalai, Sh. Yariv, S. Gross, Mineral. Mag. 40 (1975) 197.
[5] G. Falini, E. Foresti, G. Lesci, N. Roveri, Chem. Commun. 14 (2002) 1512.
(in which case the quadrupole moment is small) a very narrow line [6] H. Schreier, Asbestos in the Natural Environment, Elsevier, New York, 1989.
is observed. Clearly, the line broadening mechanism involves the [7] Powder Diffraction File Search Manual Minerals, Joint Committee on Powder
aluminum atoms in the lattice [36]. Diffraction Standards – International Centre for Diffraction Data, card 25-0645,
USA, 2003.
[8] E. Foresti, E. Fornero, I.G. Lescia, C. Rinaudo, T. Zuccheria, N. Roveri, J. Hazard.
Mater. 167 (2009) 1070.
[9] V. Sontevska, G. Jovanovski, P. Makreski, J. Mol. Struct. 834–836 (2007) 318.
[10] I.R. Lewis, N.C. Chaffin, M.E. Gunter, P.R. Griffiths, Spectrochim. Acta 52A (1996)
315.
[11] L. Wang, A. Lu, C. Wang, X. Zheng, D. Zhao, R. Liu, J. Colloid Interface Sci. 295 (2006)
436.
[12] H. Suquet, Clays Clay Miner. 37 (1989) 439.
[13] C. Viti, M. Mellini, Eur. J. Mineral. 9 (1997) 585.
[14] J. Madejova, Vib. Spectrosc. 31 (2003) 1.
[15] D.G. Taylor, C.M. Nenadic, J.V. Crable, Am. Ind. Hyg. Assoc. J. 31 (1970) 100.
[16] V.C. Farmer (Ed.), The Infrared Spectra of Minerals, Mineralogical Society, London,
1974, pp. 331–358.
[17] E. Foresti, M. Gazzano, A.F. Gualteiri, I.G. Lesci, B. Lunelli, G. Pecchini, E. Renna, N.
Roveri, Anal. Bioanal. Chem. 376 (2003) 653.
[18] E.P. Kraineva, V.L. Petrov, T.I. Polupanova, J. Appl. Spectrosc. 37 (1982) 1157.
[19] G. Falini, E. Foresti, M. Gazzano, A.F. Gualtieri, M. Leoni, I.G. Lesci, N. Roveri, Chem.
Eur. J. 10 (2004) 3043.
[20] C. Rinaudo, D. Gastaldi, E. Belluso, Can. Mineral. 41 (2003) 883.
[21] J.T. Kloprogge, R.L. Frost, L. Rintoul, Phys. Chem. Chem. Phys. 1 (1999) 2559.
[22] D. Bard, J. Yarwood, B. Tylee, J. Raman Spectrosc. 28 (1997) 803.
[23] C. Groppo, C. Rinaudo, S. Cairo, D. Gastaldi, R. Compagnoni, Eur. J. Mineral. 18
(2006) 319.
[24] A. Wang, J. Freeman, K.E. Kueblur, Lunar and Planetary Science, 2002 (XXXIII
1374.pdf).
[25] J.R. Michalski, M.D. Kraft, T.G. Sharp, P.R. Christensen, Earth Planet. Sci. Lett. 248
(2006) 822.
27
Fig. 8. Al NMR spectrum of chrysotile. [26] J.L. Bishop, M.D. Lane, M.D. Dyar, A.J. Brown, Clay Miner. 43 (2008) 35.
G. Anbalagan et al. / Vibrational Spectroscopy 52 (2010) 122–127 127

[27] T.V.V. King, R.N. Clark, J. Geophys. Res. 94B (1989) 13997. [33] E. Lippmaa, M. Magi, A. Samoson, M. Tarmak, G. Engelhardt, J. Am. Chem. Soc. 103
[28] G.R. Hunt, R.P. Ashley, Econ. Geol. 74 (1979) 1613. (1981) 4992.
[29] V. Gionis, G.H. Kacandes, I.D. Kastritis, G.D. Chryssikos, Clays Clay Miner. 55 [34] R.A. Weeks, A. Chatelain, J.L. Kolopus, D. Kline, J.G. Castle, Science 167 (1970) 704.
(2007) 543. [35] L.E. Drain, Proc. Phys. Soc. 80 (1962) 1380.
[30] S. Petit, A. Decarreau, F. Martin, R. Buchet, Phys. Chem. Miner. 31 (2004) 585. [36] C.A. Fyfe, G.C. Gobbi, W.J. Murphy, R.S. Ozubko, D.A. Slack, J. Am. Chem. Soc. 106
[31] B.L. Sherriff, H.D. Grundy, J.S. Hartman, Eur. J. Miner. 3 (1991) 751. (1984) 4435.
[32] M. Magi, E. Lippmaa, A. Samoson, G. Engelhardt, A.R. Grimmer, J. Phys. Chem. 88 [37] D. Muller, W. Gessner, H.J. Behrens, G. Scheler, Chem. Phys. Lett. 79 (1981) 59.
(1984) 1518. [38] J. Sanz, J.M. Serratosa, J. Am. Chem. Soc. 106 (1984) 4790.

Das könnte Ihnen auch gefallen