Absorption and Reflection Infrared Spectra of Mgo and Other Diatomic Compounds
Absorption and Reflection Infrared Spectra of Mgo and Other Diatomic Compounds
Accepted 2003 June 9. Received 2002 December 19; in original form 2002 March 19
C 2003 RAS
IR spectra of MgO compounds 17
dispersion provide spectral artefacts (Horak & Vitek 1978; Hofmeis- The methods used here are not new, as each component exists in
ter 1995). To address these problems, we are compiling a data base various scientific literatures. However, to our knowledge the cross-
of IR spectra of minerals and other inorganic solids that may ex- checks have not been used in astronomical studies, and have not
ist in space. Both reflection and absorption techniques are utilized, been taken full advantage of elsewhere. For the most part, in miner-
as advantages and disadvantages exist for each approach, and be- alogical studies large crystals are used in reflectivity measurements
cause the optical functions obtained from reflectivity are used as to avoid potential problems, and the focus has largely been on fre-
inputs to radiative transfer models in astronomy, whereas absorp- quencies of the peaks, which are less prone to artefacts than peak
tion measurements are directly compared with spectral patterns for profiles, band strengths and widths. Because all peak parameters are
the purpose of identification and for estimates of concentration. important in astronomical applications of the data, however, multi-
This paper is the first of a series, and therefore focuses on the ple measurements and cross-checks are necessary.
methodology. Our plan is to present the data base in segments, each
of which covers structurally and/or chemically related minerals. Di- 2 E X P E R I M E N TA L M E T H O D S
atomic oxides and sulphides are examined here, with the focus being
Three single crystals of synthetic, pure MgO were purchased from
on MgO. This choice is motivated by astronomy, as well as prag-
Atomergenic Chemetals Corporation.1 The large faces are perpen-
C 2003 RAS, MNRAS 345, 16–38
18 A. M. Hofmeister, E. Keppel and A. K. Speck
detector and beamsplitter combinations as above. Very thick films Imeas
were made by placing a small amount of powder in the centre of a
25 cm diameter disc of BaF2 or of KBr. A tin gasket with a thickness
of 6 µm (NSG Precision Cells, Inc., Farmingdale, NY, USA) was I0 I0R
placed at the edge of the disc. The sample was compressed to the
same thickness as the gasket by a second BaF2 (or KBr) disc using
I0R(1-R)2(1-Ω)2
a commercial holder (Spectratech, Inc.). To attain the same thick-
ness for sample and gasket we made sure that the powder did not I0(1-R) I0R(1-R)(1-Ω)2
extend to the edge of the disc, but that the area covered by powder d
occluded the entire IR beam, and was without visible cracks and was
even in appearance upon examination under a binocular microscope. I0(1-R)(1- Ω) I0R(1-R)(1- Ω)
These measurements provide useful information on the overtones
and O–H modes at frequencies above ∼500 cm−1 . Use of BaF2 lim-
its transmission to above 690 cm−1 (<14.5 µm). Absorption spectra I0(1-R)(1- Ω)(1-R) = Imeas
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 19
is a function of the TO peak parameters: ν i , i and f i .The equa- (Shankland, Nitsen & Duba 1979). For small R, the last term is close
tions for m oscillators using the circular frequency (ω = 2πν) are to log(1 − 2R). This formula assumes parallel, smooth surfaces in a
non-converging beam. If the reflectance is unknown, then the mea-
m
f j ω2j ω2j − ω2
1 = ∞ + 2 , (4) sured absorbance is obtained relative to a baseline. This approach
j=1 ω2j − ω2 + 2j ω2 is valid for weak peaks, and is necessary if the surface is rough or
if additional internal reflections are present (e.g. from fractures or
where ∞ is obtained from measurements of n 2 in the visible, and grain boundaries) or when the beam converges strongly.
m
f j ω2j j ω Parallel surfaces can create interference fringes if the spacing (b)
2 = 2 . (5) is related to an integer multiple of the wavelength. The relationship is
j=1 ω2j − ω2 + 2j ω2
1
b= , (11)
This approach allows extrapolation of the data to frequencies above 2nν
and below the range of measurements, and aids in deciphering arte- where ν is the fringe spacing. In this study, the spacing is probably
facts from overtone-combination bands which are present as weak the gap between the diamonds or between the KBr or BaF2 discs.
shoulders or structure on the main peaks. Generally, the Kramers–
C 2003 RAS, MNRAS 345, 16–38
20 A. M. Hofmeister, E. Keppel and A. K. Speck
ZnO has the hexagonal wurtzite structure, in which both atoms λ, µm
occupy tetrahedral sites. This structure is identical with the simplest 50 25 20 16.7 14.2 12.5 10
2H polytype of hexagonal α-SiC (Wyckoff 1963). From symmetry a
analysis (Fateley et al. 1972), one band is expected in each of the po- MgO 5
0.8
Reflectivity
0.6
modes. Although two IR bands exist for this structure, it is unlikely
that these can be distinguished for ZnO by thin-film or dispersive 3
methods, as these methods revealed only one strong band in the IR 0.4
spectrum of α-SiC (e.g. Speck, Hofmeister & Barlow 1999). The 2
underlying reasons for the equivalence of the IR modes from the A
two polarizations are that (1) the two cation–anion bond lengths in 0.2
the wurtzite structure are similar, and (2) the motion of the atoms 1
Dielectric Functions
modes in the IR spectra (Choyke & Patrick 1962). Such a stacking -2
disorder has not been reported for ZnO, but exists for ZnS. 60 ε2 im(1/ε)
FeS is the nominal composition for a family of structurally and -3
chemically complex minerals (see e.g. Vaughan & Craig 1978; Deer
et al. 1992; Gaines et al. 1997). As occurs in wüstite, some of the 40
-4
Fe is ferric, and charge balance is attained through cation vacancies.
Chemical compositions are reported as Fe1−x S to indicate the min- 20
-5
imum degree of defects. Compositions near Fe0.94 S (ferric troilite)
crystallize in the NiAs parent structure (Deer et al. 1992) which has -6
anions in hexagonal close-packing, whereas the cations are octahe- 0 * -7
drally coordinated. Symmetry analysis of NiAs gives mode types
that are identical to those found in the wurzite structure, suggest-
8 MgO c
ing only one IR mode would be observed for such ferric troilites.
The troilite structure, which is formed with lower ferric contents, x
< 0.06, is described as a superlattice of the NiAs structure which
Optical Functions
6
contains 12 FeS molecules in the crystallographic unit cell (Wyck-
n
off 1963). Using symmetry analysis, we predict two IR modes for
troilite when E is polarized along the c-axis and six IR modes perpen-
dicular to c (acoustic modes are not included). Because the modes 4
polarized perpendicular to c are doubly degenerate, these are more k
intense, and six bands should be seen in unpolarized measurements.
However, the relationship of the troilite structure to the NiAs parent 2
structure through multiple stackings of the small NiAs unit cell par-
allels the polytypism of SiC, suggesting that many modes are weak. *
It is possible that troilite could have fewer IR bands, or one strong 0
200 400 600 800 1000
peak and several weak ones. Wavenumbers, cm
-1
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 21
Table 1. Kramers–Kronig analysis for MgO using = 9.83 (Fontanella λ, µm
et al. 1974) to constrain the data below 125 cm−1 . Frequencies and FWHM
are in cm−1 . The weakest peaks appear as shoulders and reliable parameters 40 25 20 16.7
3 0
could not be derived. Instead these were determined by classical dispersion
– see Table 2. The 646 and 286 cm−1 positions match frequencies observed
at the Brillouin zone edge (Peckham et al. 1967; Sangster et al. 1970). The 2.5
*
-0.01
mode assigned as transverse acoustic (TA) may be an artefact. **
Dielectric Functions
2 ε2 im(1/ε)
Assignment ν TO FWHM f ν LO
-0.02
TA 286 weak 290
1.5
LA 408, 430*
TO 415.3 33.8 8.05 739.0 -0.03
2TA ∼580 weak ∼580 1
TA+LA 646 ∼50 0.023 644
C 2003 RAS, MNRAS 345, 16–38
22 A. M. Hofmeister, E. Keppel and A. K. Speck
Table 2. Classical dispersion fitting parameters using ∞ = n 2 = 3.0176 λ, µm
10 5 4 3 2.5
(Fontanella et al. 1974). For MgO, the fit was made to the reflectivity data. 0.1 0.2
Frequencies and FWHM are in cm−1 .
a
R
Phase Assign. ν TO FWHM f ν LO ε0 0.08
Reflectivity
TA+TO* 650 75 0.025 645 0
CaO TO 298 32 9 570 12.4 A
MgS TO 240 31 11 435 15.8 0.04
CaS TO 232 22 9.5 417 14.6
∗ Resonances involving edges of the first Brillouin zone. The TO, LO -0.1
and LA branches cross and resonate within the Brillouin zone so the 0.02
data. λ, µm
5 2 1.25 1 0.8 0.625
0.1 0.4
peaks in the coated mylar beamsplitter, and appear to be artefacts of
less than perfect alignment of the spectrometer and its accessories, b
R
or of the sample with respect to the reference. (Slight motions of 0.08 H O
2
the thin, mylar beamsplitter are also a likely cause.) The features 0.3
are resolved because of the high signal-to-noise ratio. We there- calc. R
SCM absorbance
fore omitted these probable artefacts from the classical dispersion 0.06
Reflectivity
analysis. Weak peaks exist near 850 cm−1 , but were not included
0.2
in the analysis, because this transparent spectral region is affected
by a secondary reflection (as sketched in Fig. 1) and because the 0.04 ascm
absorption measurements provide better constraints (see below).
Roughly 300 synthetic spectra with differing numbers of modes 0.1
and mode parameters were calculated to probe the limitations and 0.02
trade-offs in fitting, particularly regarding the threefold multiplet.
There is little leeway in the fitting because the steep reflectivity
edge at high frequency and the gradual slope in reflectivity at low fre- 0 0
5000 10000 15000
quency provide opposing constraints (Fig. 2a). This steep slope con- Wavenumbers, cm
-1
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 23
(Hofmeister, in preparation). Similar behaviour was seen in the IR λ, µm
spectra of Mg2 SiO4 and Fe2 SiO4 with the spinel structure, which 20 10 5 4 3.33 2.5
also have simple IR spectra (Hofmeister & Mao 2001). For the cur-
a
rent paper, it suffices that the measured reflectivity is reproduced MgO
with a threefold multiplet (Fig. 2a), and the dielectric and optical 2
HO
2
functions from the Kramers–Kronig and classical dispersion analy-
ses (Figs 2b, c and 3) are in excellent agreement over the fundamental
region. 1.5
Reflectivity data from 1000 to 4000 cm−1 (10–2.5 µm) contain
SCM Absorbance
some additional weak and broad peaks (Fig. 4). Their poor resolu- d =0.006 mm
tion, along with intensities <1/100th of the fundamentals, and broad 1
between KBr disks
C 2003 RAS, MNRAS 345, 16–38
24 A. M. Hofmeister, E. Keppel and A. K. Speck
Direct observation of overtones and combinations features at ν > 5 100
1000 cm−1 requires thicknesses of ∼0.1 to 0.3 mm (Fig. 5a). Thicker
(0.5 to 1 mm) crystals were studied at frequencies up to 8500 cm−1 , total a
and indicated sorbed H2 O, but were otherwise featureless (Fig. 4b). 4 coverage 80
MgO is hygroscopic and polished surfaces become dull (foggy) after
%T
% Transmission
exposure to air over hours to days. Fine powders also accumulate
Absorbance
surface hydration on a similar time-scale, as indicated by features 3 60
near 1500 and 3400 cm−1 in the thick-film spectrum (Fig. 5a). At
low frequencies, the thick-film spectrum compares reasonably well
with the single-crystal results, considering the differences in optical 2 40
path. Towards high ν, the absorbance of the film strongly increases
(Fig. 5a). This behaviour is due to scattering. Because MgO is hard
and incompressible, the grain boundaries were not eliminated by 1 20
A
pressing between KBr discs.
% Transmission
Absorbance
Absorption spectra were acquired at low frequencies (Fig. 5b) to 0.6 60
probe the possible feature at 290 cm−1 in the reflectance (Fig. 2).
The feature is of comparable strength to the noise for the thick-
nesses examined (Fig. 5b), whereas for thicker samples the steep 0.4 40
rise in absorbance as frequency increased towards the main peak
obscures this feature (not shown). The feature is too weak to be
0.2 A 20
seen in the thinner films (Fig. 5b). Although it appears that this
TA mode is real as all data are consistent (Table 1), it is not suf-
ficiently resolved over the noise level of the spectra to be used to
determine film thickness accurately. It can serve as a rough check on 0 0
200 300 400 500 600 700 800 900
Kramers–Kronig analysis. From Fig. 5(b), the true absorption coef- -1
Wavenumbers, cm
ficient (peak minus sloping baseline) is ∼ 0.12 × 2.3/0.1 mm = 3
mm−1 . The calculated absorption coefficient of ∼15 mm−1 (Fig. 2a) Figure 6. Schematic indicating how spectral artefacts are created from light
is larger, but may be enhanced (i.e. roughly doubled in strength) by leakage. For each section, the circle (top) illustrates conditions of sampling:
a back reflection, as were the weak modes at high frequency (Fig. 4). heavy dashed line = the transmission spectrum of MgO with t = 1 µm;
To a good extent, these values differ simply as a result of the low light dashed line = the corresponding absorbance. Heavy solid line = trans-
signal-to-noise ratio and the weakness of the feature. We also com- mission for a sample that is opaque near 430 cm−1 ; light solid line = the
pared the measured absorbances in the far-IR with values calculated corresponding SCM absorbance. (a) Ideal case where the aperture is covered
entirely by the sample at a uniform thickness. Low values of transmission
from the Kramers–Kronig relationships, after subtracting the con-
drastically affect the absorbance. (b) Case of uniform thickness, but par-
tribution for the reflections (equation 10). The wedge absorbs more tial coverage. Light leakage alters the spectrum from that intrinsic to the
light than expected between 100 and 330 cm−1 (Fig. 5b). Probably material. Note that absorbance is about the same regardless of whether the
this is due to internal scattering, as the comparison for the thick crys- sample partially transmits near the peak or whether the sample is opaque at
tal showed a larger discrepancy. Once the absorbance is moderate frequencies near the peak.
(>0.4), calculations and measurements are in good agreement, and
for strong absorbances (>1.5) the reflectivity correction has little
impact (Fig. 5b). Absorption measurements of even single crystals response becomes non-linear. For this reason, commercial software
are therefore an approximation of the ideal absorption coefficient, often truncates absorbance (a SCM ) at values of 2, 3 or 4, at the dis-
with some problems occurring in the more transparent regions. cretion of the manufacturer. In our experience, absorbance near 1
The spectra collected from samples lying on an aperture show an minimizes artefacts of this nature. This observation is consistent
increase in the intensity of the peak near 1000 cm−1 that is propor- with use of a > 2 to describe optically thick conditions (Siegel &
tional to thickness, as expected from the connection of peak area Howell 1972).
or height with concentration (Beer’s law). However, the intensity of A different type of non-linear response is seen in the spectrum
the peak near 850 cm−1 does not increase as much as the difference obtained from a crystal using an IR microscope. In this case, light
in thickness would suggest. Note also the high degree of noise for leaking around the edge of the sample reduces the absorbance
measured absorbance that is greater than 2. The loss of proportion- from the actual values (Fig. 5a). As shown schematically in Fig. 6,
ality is a problem of sampling. Detectors are not noise-free, and, the leakage of light reduces absorbance of the strong peaks more
as the transmission decreases, it will lie within the uncertainty of than the weak peaks (or shoulder), creating either rounded or
‘zero’ before the true absorbance value is infinite (see Fig. 6a: a flattened profiles. Placing apertures at the image points before
small amount of noise in the transmission spectrum for the mini- and after the sample minimizes the light leakage, but not all IR
mum near 425 cm−1 strongly alters the maximum value of a). Thus, microscope systems use the double-aperturing technique. Nei-
at some low value of transmission (high absorbance), the system ther does the second aperture completely remove sampling artefacts
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 25
λ, µm (Refner, Hornlein & Hellgeth 1997). When light leakage is present,
50 25 16.7 12.5 10 8.3 peaks are seen at frequencies where the sample is opaque [as shown
in Fig. 5(a) when the same sample covers the aperture]. At fre-
1 a 6
quencies above 800 cm−1 , where the sample transmits to a greater
M degree, peaks are seen at the same positions as the sample of equal
5 thickness that was masked by a physical aperture, but the relative in-
0.8
tensities of the peaks, their components and shoulders are incorrect
SCM Absorbance
MgO
(Fig. 5a). For strong absorbers, the spectral profile is a predomi-
4
nantly determined by the amount of light leakage, as indicated by
0.6 M FI Fig. 6. Note that opaque regions of a sample with light leakage (i.e.
3 a dispersion) provide finite absorbance, and that infinite absorbance
ZnO cannot be distinguished from a = 4 at 50 per cent coverage. These
0.4
18
Mg O conclusions are irrespective of the type of material sampled. Thus,
I F 2
I for dispersions, the fraction of areal coverage is a primary factor in
3 4
I The peak for Mg18 O is shifted to 469 cm−1 and is broader than that
I
of MgO (Fig. 7a). The large breadth is probably due to the enriched
0.6
sample containing a significant proportion of the normal isotope
16
O. The normal sample would have levels below 0.2 atom per cent
I of the heavy isotope (Criss 1999). Isotopic exchange should affect
0.4 the mass without altering the lattice constant, producing a decrease
I
in frequency. Instead, an increase in ν is observed, and attributed to
FeO changes in the resonances between the IR and acoustic modes.
0.2 I The superimposed structure on the high-frequency side of the
M
peaks for MgO and CaO is mostly interference fringes (Fig. 7a).
The fringes are related to the spacing between the diamond faces,
0 which is larger than the sample thickness, and no simple method
200 400 600 800 1000 1200
Wavenumber, cm-1 exists to remove them without affecting the spectra. The spacing of
the fringes is broken by some weak combination peaks, and the LO
Figure 7. Thin-film IR spectra of X–O compounds. (a) Data on MgO and modes. For MgO, its acoustic combination band at 645 cm−1 and its
ZnO, showing procedures to merge spectral segments. Thick line = Mg18 O; strongest IR overtone near 850 cm−1 are also present. The strength
long-dashed line = ZnO; all other lines are MgO. M indicates merge point. of the 850 cm−1 band suggests a 2.3-µm film from comparison
F indicates interference fringes. Short lines indicate peaks in addition to
with the single-crystal data, whereas t = 1 µm is suggested from
the fringes (e.g. the overtones in MgO near 860 cm−1 ). The left-hand axis
is for the thick film of MgO (grey line); a 6-µm film (dotted line), and
data from a crystal scaled to t = 6 µm (thin line near x-axis marked ‘xtl’). high n is due to index of refraction mismatch. Arrows indicate approximate
Comparison of the overtone intensity indicates that the thin film of MgO has extrapolation between segments and to high ν. (c) Natural Fe3 O4 (light line)
t near 1 µm thick. (b) Effect of thickness on CaO. Heavy dotted line = film and FeO (dotted and thick lines). Scattering causes the rising baseline to
between KBr discs, obtained at t = 6 µm but scaled to t = 1.5 µm. The rise to high frequency. Vertical lines show peak positions.
C 2003 RAS, MNRAS 345, 16–38
26 A. M. Hofmeister, E. Keppel and A. K. Speck
Table 3. Thin-film absorption spectra parameters (in cm−1 ) and related information. Width is twice the FWHM on the low-frequency side
of the main peak. For comparison, acoustic frequencies, reduced atomic masses and bond lengths are included. Cation–anion distances and
acoustic (Debye) frequencies are from compilations of structural (Wyckoff 1963; Gaines et al. 1997) and elastic data (e.g. Bass 1995; Knittle
1995). Acoustic frequencies of the sulphides were estimated (except for ZnS) because the shear moduli are not available. For MgS and CaS,
resonances in the IR were used to estimate some acoustic frequencies.
comparison with the thick film. The agreement is good, consider- surements. (A quantitative comparison will be made in Section 6.3.)
ing that the fringes interfere with determination of peak strengths. A thicker film (∼2 µm) is also shown. In this case, the rapid rise
For the thinnest FeO film, a shoulder appears at 570 cm−1 . From ends with a shoulder at the LA resonance (Fig. 7a). This data better
comparison with previous reflectivity data, this is the LO mode. Sim- constrains the join to the thinnest film.
ilarly, for ZnO, a shoulder exists at 580 cm−1 , which is between the Applying the same procedure to CaO indicates a rapid rise at 590
LO positions determined from Raman spectroscopy (Arguello et al. cm−1 , confirming the above assignment to the LO mode (Table 3).
(1969). For MgO, the LO position is not revealed in the thin-film The thicker film of CaO appears to be 1.5 µm, as this thickness
spectrum; instead, the resonance peak at 650 cm−1 breaks the spac- provides the best joint between the thin-film data and the spectrum
ing of the fringes. A similar shoulder is seen in the CaO spectrum, from t = 6 µm. The rise to high ν is scattering, and the intrinsic
which is assigned as the LO component, because sums involving the spectrum is approximated by a gradual decrease in absorbance above
acoustic modes do not match this position (Table 3). The resulting 750 cm−1 .
splitting between the LO and TO modes systematically decreases for
the oxides (Table 3). This marker combines the effects of oscillator
5.3.2 Sulphides
strength and damping coefficient.
Joining the two thin-film spectra results in an asymmetric peak for Far-IR absorption spectra of the cubic sulphides MgS and CaS also
all of the oxides (ZnO data are not shown for clarity). The asymme- have one main peak, with some structure (Fig. 8). Mid-IR spectra
try increases from ZnO to CaO to MgO. For FeO, strong scattering revealed no additional features and are not shown. Their peak po-
makes it difficult to obtain data except for the thinnest films. For sitions are very similar, but occur near 250 cm−1 rather than near
these low thicknesses, the LO mode becomes less important (the 400 cm−1 for the oxides, consistent with the larger mass of the sul-
ideal is approached) and the peak appears symmetric. Magnetite phur anion, and with the longer bond distances (Table 3). In addition,
exhibits the same sort of scattering baseline as FeO (Fig. 7c). From the peaks are much narrower than those of the oxides. Structure ex-
comparing Fig. 7(a) with Figs 2(a) and 5(a), the thin-film spectra ists on the main band of MgS, which is attributed to a resonance
for MgO are clearly delivering excess intensity at high frequency, with the acoustic modes. The positions of the shoulders are in rea-
which provides the asymmetry. This compound is highly reflecting sonably good agreement with estimates of the acoustic frequencies
over a wide frequency range (Fig. 2a), which creates an artificially (Table 3). The TO and LO positions are also compatible with fre-
high apparent absorbance, in accord with equation (10). Using equa- quencies observed for α-ZnS and CdS using Raman spectroscopy
tion (10) requires reflectivity data, and it is desirable to obtain the (Arguello et al. 1969). The peak widths and the LO–TO splittings
intrinsic absorption independently. We devised the following pro- are smaller than those of the oxides and decrease as the reduced
cedure. For MgO, appearance of the overtone at 860 cm−1 indicates mass increases (Table 3). The narrower widths and low LO–TO
a film thickness near 1 µm. The single-crystal absorption spectrum splittings for MgS and CaS are consistent with the reflectivity being
(Fig. 5a) was therefore scaled to t = 1 µm. Although the absorbance high over the narrow range where the sample mostly absorbs, and
values should have been adjusted for reflectance (equation 10), the with the measured absorbance reasonably representing the intrinsic
scaling reduces the values to near zero, making the correction unim- absorbance for the thin films.
portant. The single-crystal spectrum is truncated at 745 cm−1 , be- FeS has an extremely weak IR spectrum, consisting of a few
cause of the rapid rise in absorbance associated with the LO mode. narrow, weak peaks superimposed on a scattering curve. The same
The spectrum for the 6 µm thick film, scaled to t = 1 µm, rapidly type of baseline was seen for FeO, and it also has a weak peak
rises just below this frequency. This trend, extrapolated to meet the compared with the other oxides, but for FeO a strong IR mode still
thin-film measurement, represents the closest approximation to the exists, consistent with the wüstite structure being essentially that of
intrinsic absorption spectrum obtainable without reflectivity mea- rocksalt. FeS appears to lack a strong IR mode, which is attributed to
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 27
λ, µm this mineral occupying a much different structure, related to that
100 50 33.3 25 20 16.7 of NiAs. The correlation of peak positions with those of the di-
3.2
atomic sulphides listed in Table 3 is given for convenience. Our
I I a
2.8
data differ somewhat from the optical functions derived from pre-
SCM Absorbance
2
6 DISCUSSION
MgS
1.6
LA 6.1 Useful relationships between spectral parameters and
physical properties
1.2 TA + TO
For a simple harmonic oscillator
0.8
I
2.4
and the bond distance. For example, the low frequency of CaO
CaS is related to its large bond length and intermediate reduced mass
2 (Table 3, Fig. 7). Although equation (11) is not precisely adhered
to, the trend for the oxides (Fig. 9) is consistent with the above phe-
1.6 nomenological model, and suggests that the measured frequencies
are reasonably correct.
1.2 The Lydanne–Sachs–Teller (LST) relationship relates the high-
and low-frequency limits of the real part of the dielectric function
LA LO to the IR frequencies:
0.8
2TO 2 m
2
0 νLO νLOi
= = , (13)
0.4 ∞ νTO νTOi
i=1
C 2003 RAS, MNRAS 345, 16–38
28 A. M. Hofmeister, E. Keppel and A. K. Speck
Table 4. Dielectric constants and frequencies (in cm−1 ). Values calculated
NiO MgO from the LST relation are in parentheses and boldface. A value near unity in
400
last column indicates adherence to the LST relation (equation 11). Dielectric
ZnO data are from compilations (Arguello et al. 1969; Boswarva 1970; Cook &
FeO
CoO Jaffe 1979; Subramanian et al. 1989; Gaines et al. 1997). The 0 -values have
uncertainties generally near ±10 per cent, whereas n and ∞ are precise. For
FeO the other compounds, data are unavailable on 0 , but ∞ = 4.796 for MnO,
-1
FeS
Peak position, cm
300 5.617 for NiO and 6.200 for CdO, all of which crystallize in the rocksalt
CaO structure (Gaines et al. 1997). FeS is opaque, so data on n are lacking.
ZnS MgS
CaS 0 TO 2
CdS 0 ∞ ν TO ν LO ∞ ( LO )
SrO BeO⊥c 6.87 2.95 724* 1098* 1.013
200 BeO
c 7.66 3.00 680* 1083* 1.007
MgO 9.83 3.018 409 737 1.003
300 ± 5† 578 ± 2†
Figure 9. Dependence of the main IR peak position on reduced mass and ZnS⊥c 8.58 5.55 274§ 352§ 0.937
interatomic bond distance (equation 12). Dots = frequency determined from ZnS
c 8.00 5.65 274§ 352§ 0.856
thin-film absorption data. Triangles = FeS from thin-film data. Squares = CdS⊥c 8.7 5.29 243§ 307§ 1.030
frequency from single-crystal studies. The line is a least-squares fit to the CdS
c 9 5.36 234§ 305§ 1.052
single-crystal measurements of the oxides (Tables 3 and 4). The bonding in PbS (150) 16.8@ 71@ 212@
the sulphides has a metallic character and falls on a different curve. The light ∗ Loh (1968). †Jacobson & Nixon (1968); Galtier et al. (1972) also
Mg cation falls off the trends for both oxides and sulphides. analyse BaO data using classical dispersion analysis. ‡Gielisse et al.
(1965). #From classical dispersion analysis of the previous IR reflectivity
data by Henning & Mutschke (1997). § Raman data of Arguello et al.
the reflectivity measurements for CaO and MgO (Table 4), sug-
(1969).
Assumes that LO and TO values of the hexagonal polymorph
gesting that the values for CaS and MgS obtained from fitting the are valid for the cubic phase. @Classical dispersion analysis of Geick (1964).
thin-film data are reasonably accurate.
For FeS, data on the index of refraction are not available because
this mineral is opaque in the visible. Two constraints exist. First, the Table 5. Peak parameters for the main mode in FeO.
low-frequency dielectric constant does not vary much among the
sulphides (Table 4), suggesting that 0 of FeS is 8.5 to 9. Compari- ν TO FWHM f ν LO 0 Analysis
son of the oxides with the sulphides with the same cation suggests 325 71 10.9 525 ∼23 Kramers–Kronig*
that ∞ of FeS is roughly 6 to 7. Roughly, the index of refraction of 325 71 10.9 563 16.3 initial fit
FeS is about 2.6. Secondly, we use existing reflectivity values ob- 325 71 9.0 534 14.38 final fit
tained by comparison with standards in a optical microscope. Such ∗ Kramers–Kronig analysis and measurements from Henning &
reflectivity values in the visible are available from the literature on
Mutschke (1997). The raw data were fitted using classical dispersion
ore minerals (mostly opaques, meaning that 30-µm thicknesses are with ∞ = 5.38 from microscope observations (Bowen & Schairer
impenetrable to visible light). For wavelengths near 590 nm, the 1935) as an input parameter: 0 is a result. Frequencies and FWHM
compilation of Peckett (1992) provides R = 18.4 per cent for FeO, are in cm−1 .
11 per cent for α-ZnO, 17 per cent for cubic ZnS, 22 per cent for
MnS, 44 per cent for PbS, 40 per cent for FeS (troilite), 50 per cent
for NiS and 40 per cent for FeS2 . The hexagonal minerals are bire- in Table 4, which suggests that either k is significant (∼2) or 0 for
flective and the values reported are the average. Some of these values troilite is greater than 9.
√
can be checked against the index of refraction data ( ∞ in Table 4), Reflectivity measurements on polycrystalline FeS (Henning &
which indicates that R vis = 15.8 per cent for FeO, 11.4 per cent for Mutschke 1997) give R = 53.5 per cent at 1000 cm−1 , which is
ZnO and 16.5 per cent for ZnS. The match is excellent for ZnO inconsistent with microscope observations of R = 40 per cent in the
and ZnS. The discrepancy for FeO is probably due to the variable visible. The excess is attributed to back reflections from the small
structural chemistry of this mineral (Hazen & Jeanloz 1984), partic- crystals; see Section 6.4.
ularly the amount of magnetite exsolved in the wüstites specifically
examined for these two different measurements. Because the same
6.2 Re-analysis of FeO data
type of behaviour occurs for troilite, we infer that its high-frequency
reflectivity should also be within a few per cent of the microscope The wüstite sample from which IR reflectivity was obtained was
determinations of 40 per cent. Because troilite is an opaque, k is synthesized by quenching (Henning & Mutschke 1997) and is poly-
non-zero. For this case, equation (1) gives an upper limit of n = crystalline (Mutschke, private communication). Grain sizes were
4.4 for FeS in the visible (with k = 0). This value is significantly not reported. At ambient conditions R of FeO has a minimum near
larger than that estimated from trends with chemistry and structure 125 cm−1 and rises to lower frequencies (Fig. 10a). A reflectivity
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 29
λ, µm spectrum of single-crystal Fe0.93 O obtained previously (Bowen,
400 50 25 16.7 12.5 10 Adler & Auker 1975) resembles Fig. 10(a), but additional weak
0.8
bands are seen at 300 and 569 cm−1 , the high-frequency side of the
FeO a peaks is shifted to longer wavelength, the maximum peak reflectiv-
Fe Mg O ity is about 10 per cent higher, and the minimum occurs at 270 cm−1 .
0.6 0.56 0.44
Bowen et al. (1975) did not perform a quantitative analysis, so an
K-K parameters
that R is constant at frequencies above the data, leads to a value of
n in the IR (Fig. 10b) that is larger than the n = 2.32 determined
1 for the visible region from optical microscopy (Bowen & Schairer
0.2
calc. R 1935). The high-frequency reflectivity is not correct because strong
final fit electronic transitions contribute negligibly to k in the visible, as im-
30 -0.4 ples studied by the Henning group. Consequently, the dielectric and
H&M (1997)
optical functions obtained by Kramers–Kronig analysis (Henning
et al. 1995; Henning & Mutschke 1997) err at high frequency.
-0.6
20 The upturn in R at frequencies below 125 cm−1 is a spectral
artefact for similar reasons, leading to n and k being suspect at
-0.8 all frequencies. Work in progress suggests that single crystals with
10 compositions between MgO and Mg0.44 Fe0.56 O have constant re-
ε
2 flectivity from 50 to 200 cm−1 . Henning and co-workers extrapo-
final fit -1
late their trend to R = 1 at ν = 0 cm−1 , assuming that FeO is a
0 free carrier. This assumption is contradicted by previous work. It is
true that the electrical conductivity of single-crystal wüstite is high,
6 c 10 ohm−1 cm−1 at room temperature, but neither a Hall voltage nor
intrinsic conduction has been observed (Bowen et al. 1975). FeO
5 is thought to be a partially compensated semiconductor where the
compensating defects are oxygen vacancies or trivalent impurities
Optical functions
C 2003 RAS, MNRAS 345, 16–38
30 A. M. Hofmeister, E. Keppel and A. K. Speck
rather suggest that the behaviour at low ν could depend on the im-
purity content and on the presence of exsolved magnetite or iron a b c
metal (typically of 0.03 < x < 0.07). More likely, the rise towards
low frequency is an instrumental problem in that the pyroxene glass
examined by Henning & Mutschke (1997) has a similar rise in R
towards ν = 0. d<λ
One possibility is inaccuracies in ratioing the sample spectrum
to the reference spectrum when the throughput is low. It is not triv-
ial to obtain precise values of R below 100 cm−1 . If a SiC globar
was used as the source in this frequency region, its intensity is low.
R=16.7% R from Eq.1
Mercury-arc lamps produce more light at low frequency, but their
output is not completely stable, leading to baseline drift. The partic- Figure 11. Ray paths (arrows) in a DAC. The trapezoids are the diamonds,
ular beamsplitter used also has an effect. The broad-range, coated the light grey represents an air gap, and the sample is the stippled grey area.
mylar beamsplitters have little throughput below 50 cm−1 , whereas (a) Empty cell used as the reference. The reflections are caused by a small
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 31
λ, µm λ, µm
50 25 20 12.5 100 50 25 20 16.7 12.5
MgO a 0.4 1 µm a
1.6 0.5
calculated absorbance calc. FeO
SCM absrorbance
SCM Absorbance
0.3
1.2
t = 1 µm
0.2
0.8
LO
0.6
µm
LO 0.2 meas.
0.1 µm
0
0 0.7
b b
2.5 t = 0.5 µm
0.6 0.5 µm
calculated for one
SCM absrorbance
sample-diamond
2 0.5 reflection
calculated: diamond 0.2
SCM Absrobance
air to sample to sample
to air to diamond 0.4 0.1
1.5 0.01 µm
0.3
1
thin film
0.2
0.5 0.1
meas.
no refl.
0
0 100 200 300 400 500 600 700 800
200 300 400 500 600 700 800
-1 Wavenumbers, cm-1
Wavenumber, cm
Figure 13. Comparison of measured absorbance of FeO (solid lines) with
Figure 12. Comparison of measured absorbance of MgO (grey line) with
calculations based on classical dispersion analyses (broken lines) for various
calculations based on Kramers–Kronig analyses (thin lines). Vertical lines
sample thicknesses. Dot-dashed = 1 µm; long dashes = 0.5 µm; short
show peak positions in the measured absorption spectra. (a) Comparison of
dashes = 0.2 µm; dots = 0.1 µm; broken = 0.01 µm. (a) Comparison
measurements with perfect absorption without surface reflections for various
of measurements with perfect absorption without surface reflections; (b)
thicknesses. Long dashes = 1 µm; dot–dashed = 0.6 µm; short dashes =
assuming that one diamond–sample reflection exists.
0.5 µm; dots = 0.2 µm; thin solid line = 0.1 µm. (b) Calculated spectra
for thickness of 0.5 µm, assuming that two additional surface reflections
exist. Short dashes = no reflections; dotted = two reflections arising from (v) Non-normal incidence exaggerates the LO mode (Berreman
the air–sample interfaces; long dashes = two reflections arising from the
1963).
diamond–sample interfaces.
Discerning the microscopic behaviour is difficult because some
of these effects compete: for example, light leakage lowers the ab-
(ii) The incident angle is high inside the diamond, owing both to sorbance from intrinsic values whereas a reflection raises it, and
the use of converging light and to the fact that diamond has a high excitation of the LO mode raises a for the high-frequency side of
index of refraction. Whilst n = 1 for air or vacuum, the dielectric the peak. However, the imperfect correspondence of Fig. 12 is most
constant 0 is 5.7 and n = 2.4099 at 15 237 cm−1 for diamond likely connected with light leakage (compare Figs 6b and 12b) and
(Fontanella et al. 1977), implying that n is close to 2.387 across also with the surface roughness of the sample (which reduces the
the far-IR. Effectively, the diamonds act like lenses, making the reflectance) compared with the smoothness of the diamond and the
relevance of Fig. 1 questionable. The sample, being thin, is also focusing effect. Fig. 11(c) probably best describes the sample, which
analogous to a coating, which is used in the optical industry to touches both diamonds, with some air gaps and some light leakage.
prevent surface reflections. Thus, equation (10) provides excessive reflection. A smaller amount
(iii) The film is not perfectly smooth, so the loss of light may than the ideal, roughly 10 per cent, and a mixture of sample–air
involve scattering rather than the specular reflection effect assumed and sample–diamond reflections would provide a better match. The
in deriving Fig. 12. spectrum is also affected by the light leakage depicted in Fig. 7. The
(iv) The film does not have a uniform thickness, which enhances excess intensity on the high-frequency side is ascribed to the non-
the shoulder. normal incidence. MgO, with its broad peak and strong LO mode,
C 2003 RAS, MNRAS 345, 16–38
32 A. M. Hofmeister, E. Keppel and A. K. Speck
is a worst test case (Fig. 12) with its extreme changes in R across the λ, µm
peak (Fig. 2) and with the reflectivity being high near the LO mode, 50 33.3 25 20 16.7
0.7
1
but the absorbance being moderate, rather than strong. It is also the
hardest of the materials studied, which makes it difficult to obtain FeS
an even film without cracks. The comparison for FeO (Fig. 13) with 0.8
Reflectivity
lower R and less contrast (Fig. 10), and a little softer, is better, but R 0.6
SCM absorbance
the reflectivity data are uncertain. Much better agreement is seen for
CaO which has a narrower peak, and is soft: see below. Excellent 0.6
agreement was seen for garnets (Hofmeister 1995). These complex a
silicates have 13 bands with widths about 1/10th those of the ox- 0.5
ides, and thus the reflectivity is only high where the sample absorbs 0.4
strongly.
The thicknesses created by compression are tenths of a µm to 1 R
vis
0.4
µm, as ascertained by transmission electron microscopy measure- 0.2
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 33
of the six absorbance measurements on two different samples of peaks reported for FeS and pyrrhotite, although it could create a
troilite. In fact, a depression is seen over this frequency range in weak peak. Moreover, absorption strengths obtained from the re-
some of our spectra (Fig. 8c). This band is not evident in the raw flectance data were shown above to be too large, and thus additional
reflectivity data, and we conclude that this feature between 17 and problems exist in the FeS and FeO absorption spectra reported by
30 µm is a spectral artefact, probably induced by back reflections Keller et al. (2002). To probe the source of the discrepancies, we
and the use of constant R as an extrapolation of the data to high collected mid-IR data from the above minerals as well as from pyrite
frequency. (FeS2 ).
The n- and k-values extracted by Kramers–Kronig analysis by The pyrrhotite spectra (Fig. 15), like that of troilite, show no ev-
Henning & Mutschke (1997) clearly err. This inference is corrob- idence of an extremely broad band from about 400 to 600 cm−1 .
orated by the large values of about 6 so obtained for n, whereas Pyrite has a fairly narrow bands at 600 and 1644 cm−1 and over-
equation (1) sets an upper limit of n = 4.4 in the visible (for k = 0) lapping peaks at 1000 and 1200 cm−1 . The pyrite structure contains
from independent observations of R. Lower values of about 2.5 for S2 dimers which should be active in the mid-IR, and probably pro-
n are seen for the compounds FeO, ZnS and CdS, but these com- duce the higher frequency modes. Pyrite peak widths are similar
pounds are more transparent. If the lower n holds also for FeS, then to those of the iron oxides. Our thin-film data on sulphides do not
C 2003 RAS, MNRAS 345, 16–38
34 A. M. Hofmeister, E. Keppel and A. K. Speck
the aperture near the source. The phase correction spectra were λ, µm
examined in all cases to ensure that saturation had indeed been 100 50 25 20 12.5
8
avoided. Because the throughput of our relatively dim globar was
too much, even when a DAC was used which reduces throughput Calculated functions
7
by a factor of 1000, one might reasonably conclude that a bright CaO
synchrotron source requires special measures to avoid saturation. 6
Because no peak exists for iron sulphides in the vicinity of 25 µm,
Optical Functions
identification of this material in protoplanetary discs is predicated 5
on a match with the weak features of troilite near 250 cm−1 (Figs 8c n
and 14). Of the objects examined by Keller et al. (2002), only M2-43 4
has features in this region, but these are at or barely above the noise
level. The spectral match is unconvincing. 3 nvis
MgS
6.6 Estimated optical constants for CaO, CaS and MgS 2 CaS
Data for 0 of CaO are available (Table 4), along with reflectivity
Figure 17. Calculated optical functions for CaS (heavy black lines), MgS
measurements (Jacobson & Nixon 1968; Galtier, Montaner & Vidal
(grey lines) and CaO (thin black lines). Solid lines = real part of the complex
1972). Because we found these sources after classical dispersion
index of refraction. Broken lines = imaginary part. Measurements of n in
analysis had been carried out, and because the oscillator strengths the visible are shown by arrows.
were not reported, we fit the thin-film data and compare the resulting
parameters with the previous reports.
For CaO, the film thickness was ascertained by comparison with
inferred position for the LO component. Relevant parameters (given
thick-film data (Fig. 7b). For CaS and MgS, thicknesses are esti-
in Table 2) must be considered as estimates for MgS and CaS. How-
mated as about 1 and 0.75 µm, respectively, by comparison with
ever, the results for CaO should be fairly accurate because thickness
results for the oxides (Figs 7, 12 and 13). These are rough numbers.
could be constrained through comparison with spectra obtained from
The fit is made by scaling the spectrum to correspond to t = 1 µm
a t = 6 µm film. The parameters are not affected by changing the
and to provide the true absorbance. This allows direct comparison
thickness by 20 per cent, which is roughly the uncertainty in t. The
with the calculated absorption coefficient.
resulting frequencies for CaO (Table 2) are within a few cm−1 of
Good fits were obtained for all substances (Fig. 16). The results
those obtained from reflectivity of single crystals (Table 4). The
are consistent, in that fitting the peak resulted in structure near the
uncertainty given in Table 4 is derived from the range of reported
frequencies (Jacobson & Nixon 1968; Galtier et al. 1972): our re-
λ, µm sults for the LO mode are slightly high, and as a result 0 is high,
100 50 25 20 cf. Tables 2–4. If we had altered the damped harmonic oscillator
parameters to fit the low-frequency dielectric measurements of Ja-
2.5 cobson & Nixon (1968), better agreement would have been obtained
Absorbance
CaS for the LO mode. As is, our results are within 1.2 per cent.
Absorption Coefficient, 1/µm
232,22,9.5 Results for MgS and CaS are less accurate, as thickness is uncer-
2 CaO 2
tain and because the spectrum consists of two overlapping bands.
298,32,9 The spectrum for MgS appears to have a baseline problem at the
1.5 high frequency, probably due to high reflectance. Even with these
MgS problems, for frequencies above about 1200 cm−1 , the results for n
240,31,11 and k (Fig. 17) are independent of the peak parameters, but instead
1 1 are determined by n vis . The results suggest that if the thin-film thick-
ness for MgS could be determined from interference fringes or if
the reflectance at high frequency could be minimized by collecting
0.5 data from a thicker film, then fitting the absorption spectra would
set a limit on n in the visible, and better constrain k.
0 0 The peaks of CaO, CaS and MgS are narrower than those of MgO
100 200 300 400 500 600 and FeO, resulting in a better match of ideal absorption coefficients
-1
Wavenumbers, cm with the measurements. The better match is largely due to the reflec-
tivity and absorptivity being high over the same frequency interval.
Figure 16. Classical dispersion analysis of thin-film absorption data. The
raw data were scaled to represent a 1 µm thick film, and converted to true
An additional factor is that the CaO and the sulphides are soft, which
absorbance units (solid lines). The fits to the data are shown as broken lines. is conducive to preparing thin, even films. The existence of multiple,
Heavy black lines = CaS; grey = MgS; thin black = CaO. We fitted the overlapping peaks contributes some uncertainty to the fits, but is of
low-frequency slope, and assumed that some rounding at the peak occurred, less importance. We conclude that materials such as silicates with
and that the spectra near the LO mode are overly high because of partial relatively narrow peaks, and which are frequently softer than MgO,
reflectance (for CaO) or were disturbed by interference fringes. would provide absorption spectra closer to intrinsic values.
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 35
Emissivity
50
solved peaks are present (Tables 1 and 2). Neither were the previous
data adequately fitted by two oscillators. An additional problem is
that high resolution is needed to measure the steep profiles found at 0.4 0.2
cryogenic temperatures (Bowey et al. 2001). Undersampling rounds
the peaks. Therefore the temperature response of MgO deduced by
C 2003 RAS, MNRAS 345, 16–38
36 A. M. Hofmeister, E. Keppel and A. K. Speck
λ, µm from flat, parallel interfaces (from equations 1 and 10) do not fit
100 50 25 20 12.5 10 the data and thus the amount of reflection is lower, consistent with
8
1.2 10 the non-normal incidence and some surface roughness of the films.
MgO However, this effect is most pronounced in the transparent regions
and near the LO mode, where the sample reflects, but is weakly
1 10
8
t = 1 µm absorbant. The mismatch of ideal and measured absorption wors-
Emissions, erg/sec/cm3
C 2003 RAS, MNRAS 345, 16–38
IR spectra of MgO compounds 37
collected under similar conditions. We cannot control the conditions ing and providing the Mg18 O. Comments by anonymous review
in space (temperature, dust density, grain size, phase and chemistry), substantially improved the manuscript.
and thus it is imperative that we understand the effect such variables
have on laboratory measurements in order to gauge the relevance of
the laboratory spectra to the astronomical environment. This paper REFERENCES
not only presents new optical data on oxides and sulphides, but also
Andermann G., Duesler E., 1970, J. Opt. Soc. Am., 60, 53
provides methods to ascertain the accuracy of such measurements.
Andermann G., Caron A., Dows D. A., 1965, J. Opt. Soc. Am., 55, 1210
Analysing IR spectra in detail reveals the limitations inherent in Arguello C. A., Rousseau D. L., Porto S. P. S., 1969, Phys. Rev., 181,
laboratory data. In particular, the absolute values of n and k and the 1351
positions of absorption features can be compromised by sampling Arnold R. G., 1967, Canad. Mineral., 9, 31
techniques. Understanding these issues allows astronomers to be Bass J. D., 1995, in Ahrens T. J., ed., Mineral Physics and Crystallogra-
able to judge whether imperfectly fitted observed spectra are due phy: A Handbook of Physical Constants. American Geophysical Union,
to variance of the conditions in the laboratory from those remotely Washington, DC, p. 45
probed, or to the presence of different phases. Bates J. B., 1978, Fourier Transform IR Spectrosc., 1, 99
C 2003 RAS, MNRAS 345, 16–38
38 A. M. Hofmeister, E. Keppel and A. K. Speck
Hofmeister A. M., Keppel E., Bowey J. E., Speck A. K., 2000, in Kessler Peckham G., Saunderson D. H., Sharpe R. I., 1967, Br. J. Appl. Phys., 18,
M. F., Salama A., eds, ESA SP-456, ISO Beyond the Peaks: The 2nd 473
ISO Workshop on Analytical Spectroscopy. ESA, Noordwijk, p. 343 Piriou B., 1964, Comptes Rendes Acad. Sci. Paris, 259, 1052
Hony S., Waters L. B. F. M., Tielens A. G. M., 2001, A&A, 378, L41 Refner J. A., Hornlein R. W., Hellgeth J. W., 1997, Micros. Microanalysis,
Horak M., Vitek A., 1978, Interpretation and Processing of Vibrational Spec- 3, 867
tra. John Wiley and Sons, New York Sangster M. J. L., Peckham G., Saunderson D. H., 1970, J. Phys. C, Solid
Jacobson J. L., Nixon E. R., 1968, J. Phys. Chem. Solids, 29, 967 State Phys., 3, 1026
Jasperse J. R., Karan A., Plendl J. N., Mitra S. S., 1966, Phys. Rev., 146, Shankland T. J., Nitsen U., Duba A. G., 1979, J. Geophys. Res., 84,
526 1603
Kachare A., Andermann G., Brantley L. R., 1972, J. Phys. Chem. Solids, Siegel R., Howell J. R., 1972, Thermal Radiation Heat Transfer. McGraw-
33, 467 Hill, New York
Keller L. P. et al., 2002, Nat, 417, 148 Speck A. K., Hofmeister A. M., Barlow M. J., 1999, ApJ, 513, L87
Kessler M. F. et al., 1996, A&A, 315, L27 Spitzer W. G., Miller R. C., Kleinman D. A., Howarth L. W., 1962, Phys.
Knittle E., 1995, in Ahrens T. J., ed., Mineral Physics and Crystallogra- Rev., 126, 1710
phy: A Handbook of Physical Constants. American Geophysical Union, Subramanian M. A., Shannon R. D., Chai B. H. T., Abraham M. M.,
C 2003 RAS, MNRAS 345, 16–38