Sie sind auf Seite 1von 5

Icarus 233 (2014) 101–105

Contents lists available at ScienceDirect

Icarus
journal homepage: www.elsevier.com/locate/icarus

Temperature dependence of the sublimation rate of water ice: Influence


of impurities
Konrad J. Kossacki ⇑, Jacek Leliwa-Kopystynski
Institute of Geophysics of Warsaw University, Pasteura 7, 02-093 Warsaw, Poland

a r t i c l e i n f o a b s t r a c t

Article history: The sublimation rate of ice is commonly calculated using the simple Hertz–Knudsen formula. This for-
Received 10 September 2013 mula is derived from the kinetic theory of gases and ignores microphysical processes determining the
Revised 22 January 2014 actual sublimation rate. The microphysical processes can be accounted for by including the temperature
Accepted 22 January 2014
dependent sublimation coefficient (Kossacki, K.J., et al. [1999]. Planet. Space Sci. 47, 1521–1530;
Available online 3 February 2014
Gundlach, B., Skorov, Y.V., Blum, J. [2011]. Icarus 213, 710–719). Kossacki and Markiewicz (Kossacki,
K.J., Markiewicz, W.J. [2013]. Icarus 224, 172–177) discussed to what extent inaccuracy of the simple
Keywords:
Hertz–Knudsen equation affects the calculated temperature of Comet 67P/Churyumov-Gerasimenko.
Comets
Comets, nucleus
Numerical simulations presented in Kossacki and Markiewicz (Kossacki, K.J., Markiewicz, W.J. [2013]. Ica-
Comets, composition rus 224, 172–177) indicate, that derivation of the temperature below the dust mantle from the measured
Ices water production rate ignoring temperature dependence of the sublimation coefficient can lead to an
underestimate of the temperature by more than 10 K. Thus, it is important to know the dependence
on the sublimation coefficient of the composition of the real cometary ice, which can be far from purity.
We intended to check whether a small amount of dissolved minerals can affect the temperature depen-
dence of the sublimation coefficient of ice. According to our experiments the answer is positive.
Ó 2014 Elsevier Inc. All rights reserved.

1. Introduction (Skorov and Rickman, 1995). In its original form the HK


equation contains the dimensionless coefficient a ¼ 1. This
Sublimation processes are significant for many materials of sci- assumption is not sufficient to explain correctly the sublima-
entific and industrial importance. Typically, the flux of sublimating tion-condensation phenomena, even for the pure materials.
molecules and the recession rate of a surface due to sublimation Moreover, materials encountered in nature are never the pure
are calculated using different forms of the classical Hertz–Knudsen elements or the pure chemical compounds but they always con-
equation (also called Hertz–Knudsen–Langmuir equation). Farther tain different contaminants. Influence of the volatile impurities
on we call it the HK equation. For the recession rate it is written and/or the non-subliming (solid) contaminants in the material
  can be taken into account by replacing a ¼ 1 by the temperature
dz a l 0:5 dependent sublimation coefficient as being also function of the
ðTÞ ¼ ðpsat  pÞ; ð1Þ
dt . 2pRg T contents and the properties of the impurities. The influence of
where T is the temperature on the sublimation surface, . is the den- the collisions and backscatter of molecules can be taken into
sity of the sublimin material, p is the pressure over the subliming account by adding an additional multiplicative coefficient. In
surface, psat is the equilibrium saturation pressure at temperature this case the coefficient is called the return flux coefficient B.
T, l is the molar mass of ice, Rg = 8.314 J mol1 K1 is the universal Thus, Eq. (1) becomes
gas constant, and alpha is a coefficient which we investigate exper-  
imentally (for sublimation of the water ice) in this work. dz Bas ðTÞ l 0:5
ðTÞ ¼ ðpsat  pÞ; ð2Þ
The HK equation is derived from the kinetic theory of gases dt . 2pRg T
assuming an equilibrium distribution of molecular velocity
or for the flux of subliming molecules

⇑ Corresponding author.  0:5


l
E-mail addresses: kjkossac@igf.fuw.edu.pl (K.J. Kossacki), jkopyst@mimuw. F s ðTÞ ¼ Bas ðpsat  pÞ: ð3Þ
edu.pl (J. Leliwa-Kopystynski). 2pRg T

http://dx.doi.org/10.1016/j.icarus.2014.01.025
0019-1035/Ó 2014 Elsevier Inc. All rights reserved.
102 K.J. Kossacki, J. Leliwa-Kopystynski / Icarus 233 (2014) 101–105

The aim of this work is an experimental investigation of the Kossacki et al. (1999) and Gundlach et al. (2011) investigated
dependence of as , on temperature, taking into account presence temperature dependence of the sublimation coefficient in wide
of dissolved impurities. We intended to check whether a small ranges of the temperature. Kossacki et al. (1999) considered the
amount of dissolved minerals, less than 0.1% by mass, may have pure ice samples, while Gundlach et al. (2011) investigated the
noticeable influence on the temperature dependence of the sub- ice samples covered by different layers of dust.
limation coefficient of ice. The samples were produced of the According to Kossacki et al. (1999) the temperature dependence
water ice made of the demineralized water, and of a spring water of the sublimation coefficient can be described by
with dissolved minerals. We performed series of measurements     
1 1 T  a2 p
of the sublimation rate at temperature within the range 190– as ¼ 1  þ tanh a3 tan p  ; ð4Þ
260 K.
a1 a1 273  a2 2
where a1 ¼ 2:342, a2 ¼ 150:5, and a3 ¼ 4:353.
Gundlach et al. (2011) investigated sublimation of a dust cov-
2. Sublimation coefficient
ered ice samples. The main reason was investigation of the diffu-
sion of subliming molecules through the layers of dust of
Before describing the coefficient a for the water ice we present a
different properties. The authors have found, that the diffusivity
short chronological review of representative papers related to
of dust is insufficient to describe decrease of the vapor flux when
determination of as for different materials.
compare to that predicted by HK equation for clean ice. Thus, they
proposed the new formula for the sublimation coefficient of ice
2.1. Sublimation of different solids b1
as ðTÞ ¼ þ b4 ; ð5Þ
1:0 þ exp½b2 ðT 1  b3 Þ
Hansen et al. (1969) studied sublimation from oriented single-
crystals and from powdered lead selenide. Analysis of the data where b1 ¼ 0:854, b2 ¼ 11:58  103 , b3 ¼ 4:99  103 , and
showed that the sublimation coefficient is unity over the experi- b4 ¼ 0:146.
mental range 887–1052 K. Both formulas predict as ¼ 1 at temperatures lower than 195 K,
Ishiyama and Fujikawa (2004) studied the evaporation and con- and as ¼ 0:15, at temperatures higher than 230 K. However, at
densation at an interface of argon vapor and its condensed phase. moderate temperatures where the sublimation coefficient features
They investigated the validity of kinetic boundary condition for the strongest temperature dependence the formulas predict sub-
the Boltzmann equation, which prescribes the velocity distribution stantially different values of as . The discrepancy between formulae
function of molecules outgoing from the interface, by the numeri- (4) and (5) can be the consequence of limited number of data avail-
cal method of molecular dynamics. They found that the condensa- able when the formula (4) was developed. It should also be noted,
tion coefficient is close to unity below the triple-point temperature that the qualitative temperature dependence of the sublimation
and decreases gradually as the temperature rises. coefficient predicted by Kossacki et al. (1999) and Gundlach et al.
Butman et al. (2008) studied the sublimation of praseodymium (2011) is in qualitative agreement with the theoretical predictions
tribromide PrBr3 under Knudsen and Langmuir conditions. The for subliming argon (Ishiyama and Fujikawa, 2004).
high-temperature mass spectrometry over the temperature range
804–957 K were applied. During sublimation from the open sur-
3. Measurements
face of a single crystal, the sublimation coefficient of PrBr3 changed
abruptly at the polymorphic transition point (UCl3 – type low-
3.1. Procedure
temperature polymorph transforms to PrBr3 – type high-temperature
polymorph, T = 934 ± 3 K). The change was reproducible in the
Sublimation of ice from a sample can be measured in different
heating–cooling cycles.
ways. Gundlach et al. (2011) determined the flux of water mole-
We note, that comparison of numerical results presented by dif-
cules escaping the ice samples, using measurements performed
ferent authors is rather difficults since it is not clear whether they
by the mass spectrometer. This method allowed building of auto-
described the values of as , or Bas .
matic measurement system, but produced very large scatter of
readings when ice slowly sublimed at low temperatures.
2.2. Water ice We decided to use another approach, direct non-contact mea-
surements of the position of the sample that recedes due to subli-
This sublimation of water ice was investigated many times, mation. Knowledge of the surface position versus time, when the
either with respect to the sublimation, or condensation of ice surface temperature is known makes possible calculation of the
(Kramers and Stemerding, 1951; Knake and Stranskii, 1959; sublimation rate versus temperature.
Davy and Samarjai, 1971; Beckmann and Lacmann, 1982;
Kossacki et al., 1999; Gundlach et al., 2011) and other authors. 3.2. Experimental set-up
Unfortunately, theoretical calculation of the sublimation coeffi-
cient (Knake and Stranskii, 1959) is too computationally inten- The surface temperature of the investigated samples is recorded
sive to be included into a model describing long term in two or three different points using resistance thermometers. The
evolution of comet nuclei. Thus, experimental values need to temperatures are measured every minute. For each of the sensors
be used. we create the separate profile of the temperature versus time.
According to Gadsden (1998) and Winkler et al. (2012) ‘‘for the We use a cylindrical vacuum chamber of 50 cm diameter and
water ice the value a = 0.83 is commonly used’’. Unfortunately, it 50 cm height. A cooler in the form of an open cylinder, 20 cm high,
is not clear, whether the mentioned authors took into account the is located inside the chamber. Liquid nitrogen is pumped through
return flux of subliming molecules. Thus, the values can be either the cooler. Similar cooled chambers are used by Kossacki et al.
as , or Bas . According to the kinetic theory of the non-equilibrium (1999). On the flat bottom of the cooler we placed samples within
boundary layer, the relative return flux of the water molecules to- plastic, or metal cylindrical holders up to 10 cm in diameter. Ther-
ward such a sublimating surface is 0.25 (Cercignani, 1981; Skorov mal isolation of the samples at the sides is not needed because we
and Rickman, 1998). Hence, B = 0.75. do not warm up the samples.
K.J. Kossacki, J. Leliwa-Kopystynski / Icarus 233 (2014) 101–105 103

Above the cooler is installed a device to measure evolving posi- For each of the sections we calculate recession of the ice surface.
tions of the surfaces of the samples. The distance to the sample The dependence of the surface position z versus time is expressed
needs to be measured with accuracy of the order of 0.1 mm. Thus, as
we used laser triangulation method. The laser beam is projected Z
F s ðTðtÞÞ
onto the target surface, and the diffuse element of the reflection zðtÞ ¼ z0 þ dt; ð6Þ
of the light spot is imaged by a CCD receiver located at some dis-
.
tance to the optical axis of the laser beam. The incidence angle of where T is the measured surface temperature, and F s is the flux of
the received light depends on the distance to the scattering target. subliming molecules. We were able to reduce pressure p in the
From the output signal of the CCD element a digital signal proces- cooled chamber with the sample inside to about 5  105 mbar.
sor (DSP) calculates the distance between the light spot on the ob- The corresponding ratio p=psat ðTÞ was of the order 0.01. However,
ject being measured and the sensor. We use integrated device when the chamber was not cooled to keep the sample at a temper-
Micro Epsilon ILD1700-200. It contains semiconductor laser with ature about 260 K the pressure was about 1 mbar. The sublimation
power less than 1 mW, 670 nm (red), of the spot diameter coefficient as was considered, due to moderate changes of the tem-
1300 lm. Accuracy of the measurements is 220 lm, and resolution perature in a section, as constant with the value of as being the
12 lm. parameter. The calculations are repeated for different values of as .
The surface temperature of the samples is measured by preci- The section n the value of as corresponding to the best fit of the cal-
sion resistance thermometers, 100 X and 500 X connected to the culated zðtÞ to the data is considered as as ðT n Þ, where
datalogger Agilent 34972A. We use 2, or 3 thermometers, located T n ¼ 0:5ðT n1 þ T n2 Þ. The temperatures T n1 and T n2 correspond to
close to the wall of a sample container. The spot of laser light is the beginning and the end of the section. In Fig. 1 we shown the
roughly in the center of the samples. The distance between the example set of data together with calculated profiles. It is shown,
spot and the thermometers was 1–3 cm, depending on the size of how the quality of the fit can depend on the value of as . In most
the sample. cases we determined as with the accuracy 0.05–0.1.

4. Results
3.3. Samples
In Fig. 2 we show values of the sublimation coefficient derived
We intended to answer the question, whether a small mass from the measured recession rates of the ice surface. The results
fraction of the dissolved minerals can significantly affect the tem- are for two types of the ice samples: (a) frozen demineralized
perature profile of the sublimation coefficient. Thus, we investi- water, and (b) frozen natural spring water of high content of dis-
gated two types of samples: (a) frozen demineralized water, and solved minerals. We have added fits to the data: curves calculated
(b) frozen natural spring water Cisowianka. The latter was selected according to the equation (Eq. (4)), proposed in Kossacki et al.
due to relatively high mineralization, mass of dissolved compo- (1999), but with new coefficients. It can be seen, that the results
nents 756 mg/l, and lack of added CO2 (see Table 1). for the investigated ices are different. This is important, because
This selection is sufficient for our purpose, but not to investigate the amount of admixtures dissolved in the used water was very
the influence of particular components. low, less than 1 g/l. It can be also seen, that the curve for pure
ice is steeper than for impure one.
In Fig. 3 we shown results obtained for ice produced of demin-
3.4. Data processing
eralized water, and in Fig. 4 for ice made of mineral water. In Fig. 3,
in addition to the data points are plotted curves as ðTÞ: (i) described
From each of the experimental runs we get two types of data:
in the literature (Kossacki et al., 1999; Gundlach et al., 2011), and
the profiles of the surface temperature versus time, and the profile
(ii) calculated according to the equations proposed in (Kossacki
of the surface position versus time.
et al., 1999; Gundlach et al., 2011) (Eqs. (4) and (5)), but with
The thermal contact between the sensors and the ice is not al-
the new coefficients. The coefficients of Eqs. (4) and (5), corre-
ways perfect and the sensors sometimes record different tempera-
sponding to the best fit to the data are presented in Table 2. The
tures. Cooling of the ice samples from the bottom means, that a
sensor which has poor contact with the sample is warmer than
2 270
the other. Thus, after each experimental run we compare the tem- Position of the surface
perature profiles and take for interpretation the coldest one. Measured
Calculated, αs = 0.7, p = 0.01
Experimental runs take few hours and correspond to the Calculated, αs = 0.8, p = 0.01
Calculated, αs = 0.9, p = 0.01
Position of the surface [mm]

changes of the surface temperature up to few tens of Kelvins. Thus, 1.5 T102
the recorded profiles of the surface position versus time are split
Temperature [K]

240
for sections satisfying two conditions:

(a) duration 30 min, and 1


(b) changes of temperature smaller than 10 K.
210

0.5

Table 1
Composition of the mineral water used to produce samples.

Component Content (mg/l) Component Content (mg/l) 0 180


16.6 16.8 17 17.2 17.4
Ca 128.3 HCO3 557.2
Mg 26.1 Cl <5 Time [hours]
Na 9.6 SO4 <5
K <5 F <0.5 Fig. 1. The example section of the data: the surface temperature TðtÞ, and position
SiO2 (undissociated) 18.38 of the surface zðtÞ. Added are the theoretical profiles zðtÞ calculated for different
values of as .
104 K.J. Kossacki, J. Leliwa-Kopystynski / Icarus 233 (2014) 101–105

Table 3
This work: The coefficients of Eq. (4) for the sublimation coefficient of H2O with minerals.
1 H2O, demineralized
fit to the data
H2O, mineral a1 ¼ 2:17  0:16
fit to the data
Sublimation coefficient α(T)

a2 ¼ 145:60  2:33
0.8 a3 ¼ 4:40  1:34

0.6

1.2
0.4 This work: data
1.1 fit, eq. 1

Sublimation coefficient α(T)


0.2 0.9

0.8

0.7
0
190 200 210 220 230 240 250 260 270
0.6
Temperature [K]
0.5
Fig. 2. Values of the sublimation coefficient derived from the measured recession 0.4
rate of the ice surface: (a) frozen demineralized water, and (b) frozen natural spring
0.3
water with 756 mg/l of dissolved minerals.
0.2

0.1
1.2
Kossacki et al. 1999 0
1.1 Gundlach et al. 2010 180 190 200 210 220 230 240 250 260 270
1 This work: data Temperature [K]
fit 1
Sublimation coefficient α(T)

fit 2
0.9
Fig. 4. Experimental results for the samples made of mineral water. In addition to
0.8 the values of as are presented curves as ðTÞ fitted to the data (Eq. (4) with new
coefficients). The coefficients are in Table 3.
0.7

0.6

0.5 Table 3. The coefficients a1 , and a2 are similar to these for pure
0.4 ice, but the coefficient a3 is two times smaller.
0.3

0.2
5. Summary and discussion
0.1

0 The main icy component of the cometary nuclei is the H2O ice.
180 190 200 210 220 230 240 250 260 270
However, it can be far from purity. Sunshine (2006) argued that the
Temperature [K] ice present on the surface of Comet Tempel 1 is thermally and
physically decoupled from the dust. However, this does not mean,
Fig. 3. Experimental results for ice made of demineralized water. In addition to the
values of as are presented curves as ðTÞ fitted to the data: fit 1 (Eq. (4) with new that this ice is clean. Spectral observations of the cometary comae
coefficients), and fit 2 (Eq. (5) with new coefficients). The latter is at low indicate presence of various gas molecules. Unfortunately, directly
temperatures nonphysically high, above 1. For comparison are drawn curves: observed are ions. The most abundant parent molecules are: CO,
as ðTÞ according to Kossacki et al. (1999), as ðTÞ according to Gundlach et al. (2011). CO2, CH4 and H2S, H2CO, CH3OH and NH3 (Bockelee-Morvan,
2011). Some of the molecules can be trapped and form clathrates,
but some (e.g. NH3) do not form clathrates. We investigated the
both formulas for the temperature dependence of the sublimation
temperature dependence of the sublimation coefficient for ice pro-
coefficient can fit to the data. However, at very low temperatures
duced of water: clean, and with dissolved minerals. We performed
only Eq. (4) predicts expected as ¼ 1. This suggests, that Eq. (5)
measurements of the sublimation rate at different temperatures,
can be used only when the available data cover the whole range
using new experimental approach.
of temperature. In the case of limited data set it seems better to
We intended to check, whether a small admixture of dissolved
fit Eq. (4). Another effect is, that the curves calculated using new
minerals can significantly affect the sublimation coefficient. Thus,
coefficients are significantly more steep than described in the ori-
we investigated samples of two compositions: pure ice, and frozen
ginal papers (Kossacki et al., 1999; Gundlach et al., 2011). The coef-
spring water with small concentrations of admixtures, about 0.1%
ficients of Eq. (4), corresponding to the best fit to the data are in
by mass. However, the temperature dependencies of the sublima-
tion coefficient are clearly different. At temperatures below 190 K
the sublimation coefficient as is in all cases equal to unity. At about
Table 2
The coefficients of Eqs. (4) and (5) for the sublimation coefficient of pure H2O, original 200 K the sublimation coefficient decreases versus temperature.
and fitted to the new data. This decreasing is less steep for the ice with impurities than for
the pure one. At high temperatures the values of as are: about
Eq. (4) Eq. (5)
0.14 for pure ice, and about 0.08 for ice with admixtures. However
Original New Original New it should also be noted, that our experimental setup is most suit-
a1 ¼ 2:342 a1 ¼ 2:33  0:08 b1 ¼ 0:854 b1 ¼ 0:13  0:03 able for measurements at temperatures about 200 K and we have
a2 ¼ 150:5 a2 ¼ 138:72  0:94 b2 ¼ 11; 580 b2 ¼ 18; 757  5926 very few data points at high temperatures, above 230 K. This
a3 ¼ 4:353 a3 ¼ 9:30  2:05 b3 ¼ 0:00499 b3 ¼ 0:00486  0:00027
makes our measurements at these temperatures less accurate than
b4 ¼ 0:146 b4 ¼ 0:88  0:18
at lower temperatures.
K.J. Kossacki, J. Leliwa-Kopystynski / Icarus 233 (2014) 101–105 105

Comparison of the temperature profiles of the sublimation coef- References


ficient with these results presented in Gundlach et al. (2011) indi-
cates, that: (a) our experimental technique gives the same results, Beckmann, W., Lacmann, R., 1982. Interface kinetics of the growth and evaporation
of ice single crystals from the vapour phase II. Measurements in a pure water
and (b) the mineral contaminants have significant influence on the vapour environment. J. Crystal Growth 58, 433–442.
temperature dependence of the sublimation coefficient. So, the Bockelee-Morvan, D., 2011. An overview of comet composition. In: Cernicharo, J.,
more detailed searches concerning dependence of that coefficient Bachiller, R. (Eds.), IAU Symposium Proceedings Series, vol. 280, pp. 261–274.
Butman, M.F., Motalov, V.B., Kudin, L.S., Kryuchkov, A.S., Grishin, A.E., Krmer, K.W.,
versus temperature and versus concentration of the impurities 2008. A jump change in the sublimation coefficient of the PrBr3 single crystal at
(the mineral ones as well as the volatiles) are needed. the polymorphic transition point. Russ. J. Phys. Chem. A Focus Chem. 82, 1972–
The main components of comets are dust and H2O ice. The dust/ 1974.
Cercignani, C., 1981. Strong evaporation of a polyatomic gas. Progr. Astron. Aeron.
ice mass ratio is commonly assumed to be 1. However, this is a
74, 305–320.
simplification due to lack of knowledge. Numerous observations Davy, J.G., Samarjai, G.A., 1971. Studies of the vaporization mechanism of ice single
indicate presence of non-volatile dust particles on the surfaces of crystals. J. Chem. Phys. 55, 3624–3630.
comet nuclei, and in the comae (e.g. sample return mission Star- Gadsden, M., 1998. Noctilucent clouds seen at 60N: Origin and development. J.
Atmos. Solar-Terr. Phys. 60, 1763–1772.
dust). Observations dealing with volatiles indicate, that the pro- Gundlach, B., Skorov, Y.V., Blum, J., 2011. Outgassing of icy bodies in the Solar
duction rates of molecules CO, CO2, CH3OH, CH4, H2S, and NH3 System – I. The sublimation of hexagonal water ice through dust layers. Icarus
relative to H2O are grater than 1% (Bockelee-Morvan, 2011). These 213, 710–719.
Hansen, E.E., Munir, A.A., Mitchell, M.J., 1969. Sublimation pressure and sublimation
volatiles could be present as separate ices, or as impurities in H2O coefficient of single-crystal lead selenide. J. Am. Ceram. Soc. 52, 610–612.
ice. Unfortunately, it is not known, how impure H2O ice really is. Ishiyama, T.Y., Fujikawa, S., 2004. Molecular dynamics study of kinetic boundary
Numerical simulations indicate, that derivation of the temperature condition at an interface between argon vapor and its condensed phase. Phys.
Fluids 16, 2899–2906.
below the dust mantle from the measured water production rate Knake, O., Stranskii, I.N., 1959. Mechanism of evaporating. Adv. Phys. Sci. 261, 6–88.
ignoring temperature dependence of the sublimation coefficient Kossacki, K.J., Markiewicz, W.J., Skorov, Y., Koemle, N.I., 1999. Sublimation
can lead to a significant underestimate of the temperature coefficient of water ice under simulated cometary-like conditions. Planet.
Space Sci. 47, 1521–1530.
(Kossacki and Markiewicz, 2013). Thus, further experiments Kossacki, K.J., Markiewicz, W.J., 2013. Comet 67P/CG: Influence of the sublimation
dealing with different concentrations of different minerals are very coefficient on the temperature and outgassing. Icarus 224, 172–177.
important. Kramers, H., Stemerding, S., 1951. The sublimation of ice in vacuum. Appl. Sci. Res.
A3, 73–82.
Skorov, Yu.V., Rickman, H., 1995. A kinetic model of gas flow in a porous cometary
mantle. Planet. Space Sci. 43, 1587–1594.
Acknowledgments Skorov, Y., Rickman, H., 1998. Gas flow and dust acceleration in a cometary Knudsen
layer. Planet. Space Sci. 47, 935–949.
Sunshine, J.M. et al., 2006. Exposed water ice deposits on the surface of Comet 9P/
This work was partially supported by the Polish Ministry of
Tempel. Science 311, 1453–1455.
Education and Science (Grant 4036/B/H03/2010/39). We thank also Winkler, R., Wang, L.Y., Lin, Y.H., Chu, C.S., 2012. Robust level coincidences in the
the Reviewers for their valuable suggestions. subband structure of quasi-2D systems. Solid State Commun. 152, 2096–2099.

Das könnte Ihnen auch gefallen