Subscriber access provided by IDAHO STATE UNIV
A: Spectroscopy, Molecular Structure, and Quantum Chemistry
Spectroscopic Measurements of Methane Solid-Gas Equilibrium Clapeyron Curve between 40 and 77 K Patrice Cacciani, Peter Cermak, Cedric Pardanaud, Gabriela Valentova, Jean Cosleou, Céline Martin, Stephane Coussan, Jennifer Anna Noble, Younes Addab, Corinne Boursier, Pascal Jeseck, Mathieu Bertin, Jean-Hugues Fillion, and Xavier Michaut J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.9b01278 • Publication Date (Web): 28 Mar 2019 Downloaded from http://pubs.acs.org on March 29, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
Spectroscopic Measurements of Methane Solid-Gas Equilibrium Clapeyron Curve between 40 and 77 K ˇ Patrice Cacciani,† Peter Cerm´ ak,∗,‡ C´edric Pardanaud,¶ Gabriela Valentov´a,‡ Jean Cosl´eou,† C´eline Martin,¶ St´ephane Coussan,¶ Jennifer A. Noble,†,¶ Younes Addab,¶ Corinne Boursier,§ Pascal Jeseck,§ Mathieu Bertin,§ Jean-Hugues Fillion,§ and Xavier Michaut§ †Laboratoire de Physique des Lasers, Atomes et Mol´ecules, Universit´e Lille, CNRS, UMR 8523, 59655 Villeneuve d’Ascq, France ‡Department of Experimental Physics, Faculty of Mathematics, Physics and Informatics, Comenius University, Bratislava, Slovakia ¶Aix Marseille Universit´e, CNRS, PIIM, Marseille, France §Sorbonne Universit´e, Observatoire de Paris, Universit´e PSL, CNRS, LERMA, F-75005, Paris, France E-mail:
[email protected] 1
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
1
March 28, 2019 Abstract
2
3
The infrared gas phase absorption spectrum of methane was used to determine its
4
Clapeyron solid-gas equilibrium curve in the 40-77 K temperature range. For com-
5
parative purposes, and to obtain more reliable results, two different optical experi-
6
mental setups were used. At higher temperatures (53-77 K), a single pass cryogeni-
7
cally cooled cell was coupled to a standard low-resolution FTIR spectrometer. The
8
second system was a state-of-the-art VECSEL tunable source operating at around
9
2.3 µm, combined with a 7 m path Herriott cell, to record methane absorption fea-
10
tures down to 40 K. From the measurements, the vapor pressure curve ln(p/Pa) =
11
−(1191.92 ± 8.92)/(T /K) + (22.49 ± 0.16) was derived in the range 40-77 K. This cor-
12
responds to a value of (9910 ± 75) J/mol for the sublimation enthalpy. The relation
13
was validated down to 40 K, increasing our knowledge of the saturation pressure by
14
two orders of magnitude. Data were compared with available pressure measurements
15
from the literature, obtained by manometric or mass spectrometry techniques, and the
16
sublimation enthalpy was compared with a thermodynamic approach based on heat
17
capacity measurements in the solid and gas phases.
18
Introduction
19
Methane (CH4 ) is a particularly interesting molecule of significant influence in many scien-
20
tific fields, including atmospheric sciences, astrophysics, geoscience, combustion science, the
21
chemical technology of hydrogen generation, and biotechnology. It is the main constituent
22
of natural gas that power stations and industrial and residential users consume. 1 Addition-
23
ally, there is a trend towards the increased use of methane as a fuel for appliances, vehicles,
24
industrial applications, and power generation. 2 CH4 plays an important role in the chem-
25
istry of the upper atmosphere, as it is the third most abundant greenhouse gas (after water
26
and carbon dioxide), and thus responsible for significant global warming. 3 Methane is also 2
ACS Paragon Plus Environment
Page 2 of 51
Page 3 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
27
present in the atmospheres of giant planets like Jupiter, Saturn, Uranus, and Neptune, 4 as
28
well as their satellites, such as Titan. 5 More recently, its presence in small but significant
29
quantities has been reported in the Martian atmosphere. 6 In all of these cases, the basic
30
thermodynamic properties of methane are of crucial importance.
31
In this study, we present a determination of the Clapeyron solid-gas equilibrium curve
32
for methane in the 40-77 K range. Commonly, the vapor pressure is measured using contact
33
pressure probes. However, at low temperatures, where the density of molecules in the gas
34
phase is so low that the collisions with the walls become non-negligible in comparison with
35
the inter-sample collisions, the so-called “thermal transpiration effect” prohibits manometric
36
measurements. 7 The motivation for this work was to investigate the performance of optical
37
measurements compared to these “conventional” contact methods, as applied to the case
38
of methane. The main advantage is that the use of light as a probe allows for vacuum
39
isolation of the studied sample in thermal and phase equilibrium at a given temperature.
40
Secondly, both the temperature and pressure of the sample can be retrieved from a single
41
optical spectrum, significantly simplifying the experimental scheme. A similar approach has
42
previously been applied to water, where the saturation pressure was measured below its triple
43
point using highly accurate measurements from Cavity Ring Down Spectroscopy. 8 Although,
44
in the CRDS experiment, the low pressure sample was not measured at low temperature,
45
but rather flushed by nitrogen from a humidity generation system at room temperature,
46
the commonality with our setup is the measurement of a low molecular density with an
47
appropriate spectroscopic method.
48
Despite the interest in methane in terms of its nuclear spin dynamics (e.g. Ref. 9 ), and
49
the fact that several extensive reviews on methane vapor pressure have been published, few
50
experimental vapor pressure measurements have been made below its triple point, TT P =
51
(90.6941 ± 0.0025) K and pT P = (116.96 ± 0.02) × 10−2 Pa. 10 In these studies, pressure was
52
determined using manometers in the case of Eucken et al. 11 and Armstrong et al., 12 or using
53
mass spectrometers by Tickner et al. 13 The temperature range of the sublimation was limited
3
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
54
to [76.89-90.6 K], [53-90.6 K] and [47.5-90.6 K], respectively. These references can be found
55
in the database of the National Institute of Standards and Technology Chemistry WebBook
56
(NIST), 14 which provides the methane vapor pressure temperature curves for sublimation
57
down to 50 K and for vaporization up to the critical point, TCP = (190.564 ± 0.012) K
58
and pCP = (4.5922 ± 0.002) MPa. 10 Indirectly, the sublimation vapor pressure can also be
59
retrieved from the calorimetric capacities, which can be measured with good precision and
60
accuracy even at very low temperatures. This was done by Colwell et al. 15 for methane and its
61
deuterated isotopologue CD4 in the 4-90 K range. Comparison of pressure and sublimation
62
enthalpy with our data will be discussed.
63
Our experimental work was aimed at establishing reliable vapor pressure data for methane
64
below 77 K and extending the lower temperature boundary to 40 K (from the previous 47.5 K
65
obtained by Tickner et al. 13 ). To perform this task, we have combined data from two separate
66
experiments. The “High-Resolution Setup” (HRS) laser source was a prototype frequency
67
tunable Vertical-External-Cavity Surface-Emitting Laser (VECSEL) operating in the 2.3 µm
68
region coupled to a 7 m path length cryogenically cooled Herriott cell, 16,17 and probed the
69
lowest rotational levels of the ν1 + ν2 , ν1 + ν4 , ν3 + ν4 , and 2ν2 + ν4 combination bands of
70
the ν1 (A), ν2 (E), ν3 (F), and ν4 (F) vibrational bands. 18 The second experiment, referred to
71
hereafter as the “Low-Resolution Setup” (LRS) was based upon a standard “low resolution”
72
Bruker IFS66 Fourier Transform Infra-Red (FTIR) spectrometer with a simple path cell
73
approximately 2 cm in length to cover the CH4 absorption from the strongest ν3 band, around
74
3 µm, to the components of the ν1 + ν4 and ν2 + ν3 combination bands, around 2.3 µm and
75
2.2 µm, respectively. 18 There were a number of reasons to combine these two experiments.
76
In terms of sensitivity, due to the longer path length, the HRS offered a better signal-to-
77
noise ratio down to low temperature but the accuracy was limited by the interferences in
78
the signal, characteristic of Herriott cells. 17 On the other hand, although the LRS allowed
79
the registration of the absorption signal only down to 55 K, its readings in the 60-77 K
80
range were more accurate and served as absolute calibration for the HRS data. Secondly,
4
ACS Paragon Plus Environment
Page 4 of 51
Page 5 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
81
as the “low resolution” FTIR spectrometer (∼0.1-10 cm−1 ) is a common tool found in many
82
research laboratories studying molecules in the solid or liquid phases, its ability to measure
83
gas pressure (limits and error sources) could be considered as an interesting alternative to
84
manometric measurements.
85
This paper is organized as follows: the first part is an overview of the methodology used
86
to determine pressure and temperature from the optical absorption measurements. Part two
87
details the two cryogenic experimental setups used to record the methane infrared absorption
88
spectra. Part three gives the experimental results and their analysis in terms of pressure and
89
sublimation enthalpy. This latter value is then compared to sublimation enthalpy derived
90
from a thermodynamic approach using heat capacities measurements of the solid and gas
91
phases, and finally, the vapor pressure curve including previous references 11–13 is addressed.
92
Some technical details regarding the LRS experiment are presented in the appendix (i.e. the
93
treatment of the interferogram and the determination of optical path length).
94
Methodology
95
The principle of the experiment is to retrieve both the pressure p and the temperature T
96
from the absorption pattern imprinted onto the light beam passing through a gas sample
97
at thermal and phase equilibrium. Therefore, by definition, the pressure retrieved is equal
98
to the sample vapor pressure at the corresponding temperature. The following text briefly
99
reviews the process of the extraction of p and T from the absorption spectra following
100
the methodology used by the High-resolution transmission molecular absorption database
101
(HITRAN). 18
102
To quantitatively characterize the absorption of a given substance we can use the di-
103
mensionless Napierian absorbance Aηη0 (˜ ν ). It describes the rate of light attenuation on the
104
interaction path of a length L at a given wavenumber (˜ ν ) due to the absorption by the
105
transition between the lower state η and upper state η 0 , separated by the energy difference
5
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 51
106
ν˜ηη0 . The total absorbance for a substance of concentration n is then a composition of
107
contributions Aηη0 , active in the targeted spectral range:
A(˜ ν) =
X
Aηη0 (˜ ν) =
∀(η 0 ←η)
X
n · L · Sηη0 · φηη0 (˜ ν − ν˜ηη0 ),
(1)
∀(η 0 ←η)
108
where the product Sηη0 · φ(˜ ν − ν˜ηη0 ) represents the rate of absorption by the molecule, with
109
Sηη0 the strength of the transition and φ(˜ ν − ν˜ηη0 ) its normalized line shape.
110
In the equation 1, the simplest variable to calculate is the concentration, from which the
111
pressure dependence is obtained from the ideal gas law: n = p/kB T , with kB the Boltzmann
112
constant. The line shape φηη0 (˜ ν − ν˜ηη0 ) is, to a first order approximation, described by a Voigt
113
function, in which the Lorentz component represents the pressure broadening and the Gauss
114
component accounts for the Doppler broadening of the transition. For a sample composed
115
of a single substance of molar mass M , the HITRAN database uses the following formulae
116
to calculate the corresponding Half-Widths at Half Maximum (given in cm−1 ): Tref nT · p · γself (Tref , pref ) T s ν˜0 RT = 2ln2 c M
wLor = wGauss
(2) (3)
117
where γself is the self-broadening coefficient for reference values of Tref = 296 K and pref =
118
1 atm (as defined by the HITRAN database), nT the coefficient of temperature dependence,
119
ν˜0 the transition wavenumber, c the speed of light, and R the molar gas constant. Finally,
120
for a given molecule and transition, Sηη0 depends only on the temperature, as:
Sηη0 (T ) = Sηη0 (Tref ) ·
Q(Tref ) exp[−Eη /kT ] 1 − exp[−c2 × ν˜ηη0 /T ] · · , Q(T ) exp[−Eη /kTref ] 1 − exp[−c2 × ν˜ηη0 /Tref ]
(4)
121
where Sηη0 (Tref ) is the tabulated line strength at the reference temperature (296 K), and Eη
122
is the energy of the initial state (given in cm−1 ). The total internal partition function Q
123
runs through all allowed rovibrational states i, including the vibrationally fundamental and 6
ACS Paragon Plus Environment
Page 7 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
124
The Journal of Physical Chemistry
excited states: Q(T ) =
X
gi exp[−Ei /kT ]
(5)
i 125
with gi the degeneracy of the level i. Due to its internal structure – having four half-integer
126
spin hydrogen atoms at equivalent positions – the main isotopologue of CH4 exists in three
127
possible nuclear spin (NS) configurations differing in the value of the total nuclear spin
128
(0/1/2 for para/ortho/meta isomers with spin symmetries E/F2 /A1 respectively). 9
129
Assuming there is no NS conversion, the para, ortho, and meta isomers constitute three
130
independent species and have their own partition functions Qp (T ), Qo (T ), and Qm (T ). Their
131
relative populations remain constant and are defined by the initial conditions in which the
132
gas is stored before being processed. In our experiments, gas was stored at 300 K, and
133
the statistical weight between the nuclear spin isomers at this temperature is 2/9/5, for
134
para/ortho/meta, respectively. This is an important phenomenon to consider, especially
135
at low temperatures. As some of the lowest rotational levels are forbidden for some NS
136
isomers (e.g. J = 0 of the vibrational ground state is allowed only for the meta isomer) 9
137
this will impact the dependence of Sηη0 (T ) differently for each NS isomer, especially at low
138
temperatures where the lowest J levels dominate the rotational distribution.
139
In the opposite case, assuming NS conversion, all three NS species can be considered as
140
a single ensemble with a single Q(T ) = Qp (T ) + Qo (T ) + Qm (T ). The principle and details
141
of the calculation on the relative line absorbance of the different nuclear spin species in the
142
two cases (NS Conversion or NS Conservation) for CH4 can be found in Ref. 17 Considering
143
the temperature range addressed by this study (down to 40 K), we find a maximum relative
144
difference of the partition functions between the two cases of around 2 %. This value is
145
beyond the accuracy of our model based on the HITRAN2016 line intensities S (with best
146
precision Serr in the 2-5 % range, see Tab. 1), therefore throughout the text, the effect of NS
147
dynamics is omitted. For more details on the values of different terms in the equations or
148
their relations to the internal molecular structure (e.g. dipole moments), see the appendix
149
of Ref. 19 or the HITRAN online database. 20 7
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 51
150
Finally, by applying the aforementioned relations, the p and T values can be retrieved,
151
by optimizing the Eq. 1 to obtain the best possible match with Aexp (˜ ν ). In the experiments
152
described here, however, we do not measure A(˜ ν ) directly. The transmitted signal is detected
153
as a power per unit area – intensity I(˜ ν ). The Napierian absorbance is then retrieved by
154
applying the Beer-Lambert law: A(˜ ν ) = ln(I0 /I), where I0 (˜ ν ) is the light intensity before
155
the interaction.
156
A standard approach to find the optimal model, composed of individual components
157
corresponding to observed transitions ν˜ηη0 (Eq. 1), is to minimize the χ2 function on the
158
digitized data points ν˜i :
χ2 =
X i
X
∀(η 0 ←η)
w(˜ νi ) A(˜ νi ) −
2
[Φηη0 · φV oigt (˜ νi − ν˜ηη0 )]
,
(6)
159
where w(˜ νi ) is the weight for given spectral point i, usually equal to the inverse of the square
160
of its standard deviation. Also for practical reasons, the integrated transition intensity Φηη0 =
161
n · Sηη0 is fitted rather than the pressure of the substance. Applying the relations Eq. 2 and
162
Eq. 3 to define the surface normalized Voigt line-shape φV oigt (˜ νi − ν˜ηη0 ) only three quantities
163
remain indetermined and are varied in the fit: the baseline function I0 , the temperature
164
T and the aforementioned transition intensity Φηη0 . After the convergence is achieved, the
165
pressure corresponding to the intensity of transition (η 0 ← η) is calculated using the gas law
166
and Eq. 4: "
pηη0
Φηη0 Q(T ) 1 1 kT · · exp −Eη − = L Sηη0 (Tref ) Q(Tref ) kTref kT
!#
.
(7)
167
In an ideal case, the shape of the detected intensity would be driven only by the absorption
168
of the substance under study. However, in many cases the detection scheme itself could
169
also influence the result. When this happens, the recorded Iexp is a convolution of the
170
apparatus function Fapp and the original transmitted signal I (Iexp = I ∗ Fapp ). This poses
171
an additional problem because, in a general case, Fapp does not commute with the Beer-
172
Lambert transformation. Thus, additional processing of the signal must be applied before 8
ACS Paragon Plus Environment
Page 9 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
173
treating the spectra. This was also the case in our work where each of the experiments
174
suffered from a different type of instrumental function distortion. As they are both different
175
and closely related to the experiment, the details of the treatments are provided in the
176
corresponding sections where the retrieval of p and T from the spectra is discussed.
177
Experiment
178
The spectral regions targeted by the two experimental setups are shown in Fig. 1, while
179
the list of selected transitions for analysis is presented in Tab.1. In both setups, whose
180
schematic outlines are given in Fig. 2, CH4 with a natural abundance of isotopologues and
181
99.9995 % purity was used (N55, Air Liquide). The target for the LRS was to focus on
182
the optimum accuracy and absolute precision in the higher temperature range, in order to
183
provide reference measurements. To this end, a relatively simple optical layout was used to
184
minimize the measurement uncertainties which can arise when using a more complex layout,
185
like the Herriott cell in the HRS case. Here, the aim was to exploit a long path length
186
system, and the high spectral resolution and fast mode-hop-free scanning properties of the
187
VECSEL to push the sensitivity limits to the lowest temperature limit. Combining the two
188
datasets, we were able to provide a coherent and accurate pressure analysis of methane at
189
its solid and gas phase equilibrium across the whole temperature range studied, 40-77 K.
190
Low resolution setup description
191
In the LRS, absorption spectra were recorded in the 1600-4500 cm−1 spectral range using a
192
Bruker IFS66 FTIR spectrometer. We used two wide range spectral sources: a halogen lamp
193
for near IR or a SiC globar lamp for mid IR, coupled to an infrared Mercury-Cadmium-
194
Telluride (MCT) detector cooled with liquid nitrogen. To avoid residual signal from remain-
195
ing atmospheric methane, the spectrometer and the volumes located on the line of sight were
196
purged with pure N2 (Alphagaz 2, Air Liquide).
9
ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 E -1 8
H IT R A N 2 0 1 6 - C H 4
V E C S E L - H R S
ν3
1 E -2 0
F T IR
- L R S
ν1+ν2
1 E -2 1
ν1+ν4
'
( c m /m o le c u le )
1 E -1 9
L in e s tr e n g th S
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 51
ν2+ν3
1 E -2 2
2ν2+ν4
1 E -2 3
1 E -2 4
1 E -2 5 2 5 0 0
3 0 0 0
3 5 0 0
4 0 0 0
W a v e n u m b e r (c m
-1
4 5 0 0
5 0 0 0
)
Figure 1: Methane line transitions strengths Sηη0 according to the HITRAN2016 database. 18 The arrows point to the lowest J transitions of the bands used in this work.
10
ACS Paragon Plus Environment
Page 11 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
Table 1: Selection of the targeted CH4 lines with their parameters according to the HITRAN2016 database. 18 Transition Band Br.(J, Sym.)1
ν3
ν1 + ν4
ν3 + ν4 2ν2 + ν4 ν1 + ν2 ν2 + ν3
R(0, A1 ) R(1, F1 ) R(2, F2 ) R(2, E) R(3, A2 ) R(3, F2 ) R(3, F1 ) Ra (0, A1 ) Ra (1, F1 ) Ra (2, F2 ) Ra (2, E) Ra (3, A2 ) Ra (3, F2 ) Ra (3, F1 ) Qb (1, F1 ) Rb (1, F1 ) Qb (3, F2 ) Rb (0, A1 ) Rb (3, F1 ) Q(2, F2 ) Q(4, F2 ) Q(4, F1 ) R(1, F1 ) R(0, A1 ) R(1, F1 ) R(2, F2 ) R(2, E) R(3, A2 ) R(3, F2 ) R(3, F1 )
ν˜0 2 (cm−1 )
Sηη0 3 (cm/mol)
3028.7523 3038.4985 3048.1533 3048.1691 3057.6874 3057.7265 3057.7607 4223.9053 4229.1940 4234.3498 4234.4173 4239.2506 4239.4278 4239.5765 4324.9331∗ 4328.4319∗ 4330.6300∗ 4333.6690∗ 4334.4357∗ 4330.1079∗ 4332.2360∗ 4333.3108∗ 4337.5489∗ 4555.5645 4566.9963 4578.6120 4578.6808 4590.2357 4590.4613 4590.6085
9.210 × 10−20 8.853 × 10−20 1.132 × 10−19 7.543 × 10−20 2.106 × 10−19 1.260 × 10−19 1.260 × 10−19 2.353 × 10−21 2.236 × 10−21 2.935 × 10−21 1.891 × 10−21 5.160 × 10−21 3.115 × 10−21 3.179 × 10−21 1.307 × 10−21 7.519 × 10−23 1.024 × 10−22 4.087 × 10−21 6.777 × 10−23 1.067 × 10−21 6.978 × 10−23 1.000 × 10−22 1.082 × 10−21 5.226 × 10−22 4.362 × 10−22 4.723 × 10−22 3.137 × 10−22 7.456 × 10−22 4.362 × 10−22 4.338 × 10−22
Serr
1
6 6 6 6 6 6 6 6 6 6 6 6 6 5 6 3 6 6 3 6 3 6 6 4 4 4 4 4 4 4
4
Ei 5 (cm−1 ) 0 10.4816 31.4424 31.4421 62.8781 62.8768 62.8757 0 10.4816 31.4424 31.4421 62.8781 62.8768 62.8757 10.4816 10.4816 62.8768 0 62.8757 31.4424 104.7799 104.7746 10.4816 0 10.4816 31.4424 31.4421 62.8781 62.8768 62.8757
γself 6 (cm−1 / atm) 0.081 0.081 0.083 0.083 0.079 0.079 0.079 0.076 0.087 0.087 0.076 0.080 0.093 0.082 0.078 0.081 0.083 0.091 0.079 0.081 0.078 0.079 0.77 0.080 0.081 0.083 0.075 0.076 0.083 0.081
γerr
4
3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4
Rotational Branch (Br.), J and symmetry (E, A1,2 or F1,2 ) of the lower state, the index letters ”a” and ”b” allow for distinguishing between lines with the same indicated quantum numbers. 2 Line center. 3 Line strength (Tref = 296 K). 4 HITRAN code for relative uncertainty: 3 (≥ 20 %); 4 (≥ 10 % and ≤ 20 %); 5 (≥ 5 % and ≤ 10 %); 6 (≥ 2 % and ≤ 5 %). 5 Energy of the lower state. 6 Self broadening coefficient. ∗ Transitions used in HRS.
11
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
(a)
Page 12 of 51
PD
Laser: VECSEL
He
T= 12-300 K
PD PD
Cooled Herriott cell
Reference cell
Fabry-Perot étalon
(b)
T = 4-300 K He
InSb detector
FTS Glowbar
10-7 mbar Cooled direct pass cell
N2 purged volume
Figure 2: Experimental setups: (a) the High Resolution Setup (HRS) and (b) the Low Resolution Setup (LRS). PD=Photodiode.
12
ACS Paragon Plus Environment
Page 13 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
197
The direct path optical cell was initially designed for adsorption isotherm measure-
198
ments. 21 The copper cell was equipped with two sapphire windows forming an optical path
199
of length LLRS = 1.746 ± 0.035 cm through the gas sample at thermal equilibrium (the de-
200
scription of the precise path length determination is presented in the Appendix ). The cell
201
was placed inside a cryostat equipped with KBr windows, continuously maintained under
202
a vacuum of 10−7 mbar and cooled by a CTI CRYOGENICS model 22 closed-cycle helium
203
cryocooler. The cell temperature was regulated using a heater and a Lakeshore controller
204
connected to a silicon diode (Lake Shore Cryogenics Model DT-500) fixed on the cell. The
205
cell temperature was adjusted between 40 and 300 K with a precision of 0.1 K. The cell was
206
connected to a stainless steel vessel at ambient temperature where the equilibrium pres-
207
sures were measured in the range 10−4 –10 mbar using two MKS baratron pressure gauges
208
(0–1 mbar and 0–10 mbar capacitive gauges with an accuracy of 0.25% ± 0.0001 mbar and of
209
0.25% ± 0.001 mbar, respectively).
210
To prepare the solid-gas mixture for the Clapeyron curve measurements, roughly 3 mmol
211
of methane were introduced at 100 K (corresponding to 100 Torr in the cryogenic cell and
212
ambient vessel). The cell was then cooled down to 45 K. At each fixed temperature between
213
45-80 K, the (p, T ) and the corresponding IR spectra were collected at thermodynamic
214
equilibrium. For each spectrum, 250 to 500 scans were recorded and averaged. In between
215
measurements, the vessel maintained at ambient temperature was pumped to 10−7 mbar to
216
avoid the contribution of molecules outgassing from the stainless steel walls.
217
After the measurement, the spectra were extracted from the recorded interferograms. To
218
concentrate spectral information around the central lobe of the transition line, a common
219
and standard procedure, named apodization, was applied. It consists of multiplying the
220
interferogram by a selected mathematical function. The effect is to obtain a cleaner spectrum,
221
useful when the density of lines is high but at the expense of the resolution RS . In our
222
case, the apodization function was set to the standard three-term Blackman-Harris function
223
(BH3). 22 The resulting value of RS = 0.13 cm−1 was then determined experimentally by
13
ACS Paragon Plus Environment
The Journal of Physical Chemistry
T = 5 5 K ra w 3 1 2 8
3 1 2 4
3 1 2 0
I ( ν~ ) ( a . u . )
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 51
P Q
R
c o rre c te d
R (0 ) R (3 )
3 1 5 0
R (2 )
3 1 0 0
3 0 5 0
W a v e n u m b e r (c m
n
R (1 ) -1
3
3 0 0 0
)
Figure 3: Effect of the interferogram correction on spectra recorded with the low resolution setup, in the ν3 asymmetric stretching vibrational band. The interferences due to the use of non wedged windows create a spike in the interferogram and give rise to oscillations on the raw spectrum. Replacing this spike by a straight line suppresses oscillations and gives the corrected spectrum with smooth baseline after Fourier transform (shifted in intensity on the figure, for clarity). 224
measuring the width of isolated R(0, A1 ) and R(1, F1 ) transitions. Also, before extracting
225
spectra from the interferogram, the spike associated with the reflections on the non-wedged
226
windows of the sample cell was replaced by a flat line (see the blue curve in Fig. 3). A
227
detailed description of the numerical process applied to our FTIR spectra is described in
228
Ref. 23
229
The resulting transmittance spectra, together with the absorptance signal αP,exp =
230
1 − Iexp /I0 , for the lowest components of the rovibrational P , Q and R branches are shown
231
in Fig. 4. The absorptance αP (˜ ν ) was preferred before the absorbance A(˜ ν ) for two reasons.
232
First, it is a recommended quantity to account for the effects inherent to the Fourier trans-
233
form spectra, 24,25 and second, it allows a better graphical comparison of the spectra. Note,
234
that in case of small absorption I ≈ I0 , the αP converges towards A/L.
235
The main target for the FTIR analysis was to record the rovibrational lines in the ν3
236
asymmetric stretching vibrational CH4 band around 3048 cm−1 (Fig. 4b). Lines from the
237
ν1 +ν4 band were used to obtain better signal at higher temperatures (Fig. 4c), while the lines 14
ACS Paragon Plus Environment
Page 15 of 51
(a ) 5 0 K 5 2 K 5 4 K 5 6 K 6 0 K 6 5 K 7 0 K 7 7 K
0 .2 5
0 .1 5
Ie
x p
( ν~ )
0 .2 0
ν2 + ν3
0 .1 0
ν1 + ν4 0 .0 5
ν3
0 .0 0 2 7 5 0
(b )
3 0 0 0
3 0 2 8
3 0 2 9
3 0 3 8
4 2 0 0
4 4 0 0
4 6 0 0
(c )
3 0 3 9
0 .0 4 1
R (1 ,F )
1
)
0 .1 5
ν3
0 .0 2
,e x p
R
ν1 + ν4 R
= 1 - I
e x p
/I0
R (0 ,A
A b s o r p t a n c e αP
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
a
(0 ,A
0 .1 0
1
) R
a
(1 ,F 1
a
(2 )
R )
a
(3 )
0 .0 0
R (2 )
R (3 ) 0 .0 5
0 .0 2
0 .0 0
0 .0 0 3 0 4 8
3 0 4 9
3 0 5 8
3 0 5 9
4 2 2 5
W a v e n u m b e r (c m
-1
4 2 3 0
4 2 3 5
4 2 4 0
)
Figure 4: Absorption spectra of methane in the spectral region of the ν3 and ν1 +ν4 vibrational bands recorded at different temperatures using the LRS. (a) Evolution of recorded light intensity, I(˜ ν ), with temperature. Spectra have been shifted vertically from each other for clarity. The transitions used in the analysis are indicated by red and pink arrows for the lowest and higher temperature range, respectively. The evolution of the absorptance αP = 1 − I/I0 , for selected rovibrational transitions in the ν3 band (b), and in the ν1 + ν4 band (c), is shown in detail on the lower plots. 15
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
238
from the ν2 +ν3 band were used for the precise measurement of the LLRS (see Appendix). The
239
baseline noise, estimated from the RMS of the fully transmitted signal, was approximately
240
0.002, comparable to that of the HRS. However, due to the much smaller optical path, the
241
minimum detectable absorption coefficient (Aexp (˜ ν )/L)min was only around 1 × 10−3 cm−1 .
242
High resolution setup description
243
The VECSEL source, used in the HRS, was a custom developed device based on the unit
244
described in detail in Ref. 26 The optically-pumped semiconductor core containing the active
245
medium on top of a Bragg mirror provided by IES Montpellier. 27 The 0.33 mm long external
246
cavity was formed by a highly reflective spherical mirror of 2 mm radius placed on a tubular
247
piezo transducer. It produced a single mode continuous emission with a narrow linewidth
248
below 3 MHz on a 1 ms time-scale. Variation of substrate temperature allowed VECSEL to
249
cover the spectral range between 4270 and 4340 cm−1 . The continuous and mode-hop-free
250
tuning of up to 15 cm−1 was performed by the external cavity piezo-driven scans.
251
The light from the VECSEL was coupled into the LHRS = 692 ± 5 cm path length Herriot
252
Cell (HC) and then analyzed by a photodiode (type G8423 from Hamamatsu coupled with a
253
home-made transimpedance amplifier with a bandwidth of 0.6 MHz). The details of the HC
254
design can be found in Ref. 16 and the precise LHRS determination in Ref. 17 A part of the
255
main beam was also diverted to a Fabry-Perot (FP) ´etalon (with Free Spectral Range, FSR
256
∼ 0.0242 cm−1 ) and to a reference cell filled with methane at low pressure (40 cm long and
257
about 10 mbar of CH4 ) to provide signal for precise frequency calibration (with a precision
258
better than 0.001 cm−1 ).
259
The HC temperature was stabilized to absolute variations below 0.1 K using the Lake
260
Shore Cryogenics Model 331 temperature controller. Cryogenic cooling was supplied by the
261
CTI-Cryogenics Model 22C Cryodyne cooler and two 25 Ω resistors were used for the heating.
262
The temperature was measured with two calibrated silicon diodes, Lake Shore Cryogenics
263
Model DT-670 with calibration accuracy ±0.25 K between 30 and 60 K, and ±0.15 K between 16
ACS Paragon Plus Environment
Page 16 of 51
Page 17 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
264
60 and 300 K. 16 One of the probes, TA , was placed close to the ”cold finger” cooler and was
265
used as a reference for the temperature-locking circuit. The second, TB , was located on the
266
cell far from the heating resistors, to provide an independent measurement of the actual
267
temperature. The difference between TA and TB varied from 0.12 to 0.20 K for the series
268
of measurements performed, in which the temperature was decreased slowly from 70 K to
269
40 K. The importance of this temperature measurement will be discussed later. For sample
270
manipulation and the pressure measurements (performed by the 0-10 mbar Pfeiffer CMR
271
capacitive gauge with a precision of 0.001 mbar) the HC was connected to a vacuum system
272
(at room temperature) through a 6 mm inner diameter steel pipe. Detailed descriptions of
273
the VECSEL, the Herriott cell, and the calibration process can be also found in a previous
274
study performed with the HRS. 28
275
To probe the entire rotational population of the lowest levels (J = 0 to 4) of combina-
276
tion bands of different vibrational modes (Tab. 1), we performed scans about 12 cm−1 wide
277
centered around 4330 cm−1 . The repetition rate of the scans was limited by the rapidity of
278
the amplified photo-diode detector. On the other hand, high recording speed was beneficial
279
in reducing the effects of mechanical noise in the system (caused by both the cryogenic cool-
280
ing system and the vacuum pumps). As a compromise, we set the tuning speed to around
281
18 THz.s−1 , corresponding to a 50 Hz spectral acquisition repetition rate.
282
At the beginning of the Clapeyron curve measurements, the HC was pumped down to
283
10−5 mbar and was then filled with around 9 mbar of methane at 77 K. Filling was stopped
284
once the absorption signal stabilized, i.e. once the saturated vapor pressure was reached.
285
The gaseous sample was then cooled progressively (∼ 4 hours) down to 40 K (when the
286
absorption signal disappeared in the baseline noise) with intermediate temperature plateaus
287
of a few minutes duration to record corresponding IR spectra. For each temperature, 32
288
scans were recorded, calibrated, and averaged (Fig. 5a). Fig. 5b and Fig. 5c show part of
289
the absorptance corresponding to a couple of the transitions used in the analysis. The
290
baseline noise, estimated from the root mean square of the fully transmitted signal was
17
ACS Paragon Plus Environment
The Journal of Physical Chemistry
(a ) 4 0 K 8
4 3 K 4 5 K 5 0 K 6
( ν~ )
5 5 K Ie
x p
6 0 K 6 3 K 4
6 5 K 7 0 K 2
0 4 3 2 5 .2
(b )
4 3 3 3 .6 3
,e x p
= 1 - I
e x p
/I0
0 .9
A b s o r p t a n c e αP
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 18 of 51
R
b
4 3 2 7 .4
4 3 3 3 .6 8
4 3 2 4 .9 2
Q
(0 ,A 1)
4 3 2 9 .6
b
4 3 3 1 .8
4 3 3 4 .0
(c )
4 3 2 4 .9 7
4 3 2 8 .3 6 0 .9
(1 ,F 1)
0 .6
0 .6
0 .3
0 .3
0 .0
0 .0 0 .0 9
Y -z o o m x 2 0
0 .0 4
b
R
4 3 2 8 .4 6
4 3 3 8 .4
4 3 3 4 .3 8
R
(1 ,F 1)
b
4 3 3 4 .4 8
(3 ,F 1)
Y -z o o m x 1 0
0 .0 6 0 .0 2
4 3 3 6 .2
0 .0 3
0 .0 0
0 .0 0 4 3 3 3 .6 3
4 3 3 3 .6 8
4 3 2 4 .9 2
4 3 2 4 .9 7
4 3 2 8 .3 6
W a v e n u m b e r (c m
-1
4 3 2 8 .4 6
4 3 3 4 .3 8
4 3 3 4 .4 8
)
Figure 5: Absorption spectra of methane in the spectral range around 4330 cm−1 recorded with the HRS. The intensity of light, I(˜ ν ), recorded at different temperatures, is shifted vertically for clarity(a). Red and black arrows correspond to the rovibrational transitions used at lower and higher temperatures, respectively. The panels (b) and (c) show the absorptance signal for an example of two transitions from each group.
18
ACS Paragon Plus Environment
Page 19 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
291
approximately 0.001, leading to a minimum detectable absorption coefficient (Aexp (˜ ν )/L)min
292
of about 1.4 × 10−6 cm−1 .
293
Analysis of the vapor pressure curve
294
In the first part of this section we discuss the procedure used to retrieve the p and T
295
values from the LRS and HRS experimental spectra. In the second part, we deal with
296
the thermodynamics interpretations of the results and their comparison with previously
297
published data.
298
Retrieval of p&T from the spectra
299
An important point to discuss before dealing with the treatment of the absorption spectra at
300
temperatures below the freezing point is the influence of solid phase absorption. In the case
301
of methane, this is a well-known phenomenon. For example, a detailed spectroscopic study
302
covering both the temperature and spectral ranges of our work was performed by Grundy
303
et al. 29 Above 20 K, methane is in its crystalline phase I, in which the absorption peaks are
304
tens of nanometers broad. 30 In our case (Fig. 5a and Fig. 4a), this effect was observable only
305
by a small decrease of the baseline level, with no impact on the retrieval process.
306
Due to differences in the apparatus function, two different approaches were used to treat
307
the signals from the LRS and HRS experiments. The common point of the analysis was
308
the application of the aforementioned methodology. This included the approximation of the
309
absorption profile by the Voigt function. Although it is well known that it does not exactly
310
reproduce the molecular transition profiles, neither the Hartmann-Tran profile (HTP), which
311
is currently suggested by the IUPAC task group to model the molecular transition line
312
shape, 31 nor any other “beyond Voigt” line shape models were used here, due to the fact that
313
(i) our measurements were performed at low temperatures, where the Doppler broadening
314
dominates the collisional one, and (ii) the uncertainty of Sηη0 (Tab. 1) in the range 2 to 5 %
19
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 20 of 51
315
was greater than the possible error of using the Voigt line shape model within the p and T
316
values of this work. 32
317
LRS
318
The first fact we had to consider before fitting the spectra was the role of the apparatus
319
function FLRS . Note, that at 60 K, neglecting the collisional broadening, transitions in the
320
3000 cm−1 range have wGauss around 120 MHz or 0.004 cm−1 (Eq. 3). This is almost a hundred
321
times less than the width of the instrument apparatus function (around 0.1 cm−1 – see the
322
section Low resolution setup description). It means that the retrieved signal ILRS (˜ ν ) is the
323
convolution of the absorption I0 exp[−A(˜ ν )] with the apparatus function FLRS (˜ ν ). Trying
324
to extract A(˜ ν ) using first Beer Lambert transformation ln(I0 /I(˜ ν )) and then deconvoluting
325
this by FLRS (˜ ν ) is not strictly correct as the two operations do not commute.
326
In case of Fourier transform spectra a boxcar function was used to represent the FLRS :
FLRS (˜ ν ) = 2/R · sinc(2˜ ν /R),
(8)
327
with R the LRS resolution. As in FTIRs the entrance intensity I0 (˜ ν ) is generated by a
328
light source which is spectrally broad compared to the width of the apparatus function, the
329
convolution I0 (˜ ν ) ∗ FLRS (˜ ν ) leaves I0 intact. Contrary to A(˜ ν ) the resulting absorptance
330
αP,exp is then directly proportional to the sample absorptance αP (˜ ν ): "
#
(I0 ∗ FLRS ) − (I ∗ FLRS ) I0 − ILRS ≈ αP (˜ ν ) ∗ FLRS . = 1− αP,LRS (˜ ν) = 1 − I0 (I0 ∗ FLRS )
331
(9)
In this case, the modified convergence criteria would be:
χ2αP =
X
w(˜ νi ) [αP,exp (˜ νi ) − αP,model (˜ νi )]2 .
(10)
∀˜ νi
332
There were two main reasons why we did not apply this procedure, but instead, a modified
20
ACS Paragon Plus Environment
Page 21 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
333
The Journal of Physical Chemistry
criterion based solely on the amplitudes of the observed transitions αP,exp (˜ νηη0 ) was used: ( 2
κ (p, T ) = log10
[αP,exp (˜ νηη0 ) − αP,model (˜ νηη0 )]2 . αP,model (˜ νηη0 )2 )
(11)
334
First, we had to take into account the low resolution of our LRS system hindering the
335
physical line shape of the transitions. As a consequence, the only reliable information we
336
can get from the spectra is the intensity of absorption peaks. Even more, this only stands
337
for individual transitions like R(0, A1 ). In cases like the R(2) line at 3048.2 cm−1 , the two
338
ortho and para transitions are separated only by 0.016 cm−1 (Tab. 1), and further information
339
about their relative intensities is necessary to retrieve their individual intensities from the
340
LRS spectra. Second point to be considered was the intensity noise. Although the filtering
341
of the interferogram removed the largest oscillation, the remaining interference patterns had
342
a greater influence on the surface than the height of the absorption lines. Considering both,
343
the low signal to noise ratio and the low resolution of the system, the choice of the amplitude
344
based convergence criteria (Eq. 11) over the integrated absorptance one (Eq. 10) was justified.
345
To further simplify the situation, the transitions for analysis were selected in such a way
346
that they could be treated individually in a narrow spectral region surrounding the line.
347
For T 65 K no transpiration effect occurs, and the value is equal to one. For
391
T < 50 K, the p/ph ratio roughly evolves as (T /Th )1/2 (Knudsen domain). For T between
392
50 and 65 K, the ratio evolves in an intermediate way, following an empirical law proposed
393
by Ref. 33 Trying to determine new values for coefficients of a more recent refined empirical
394
law proposed in Ref. 7 was not possible in this work because the outgassing from the walls
395
limits the methane pressure measurements.
396
HRS
397
Unlike in the FTS case, the combination of narrow VECSEL linewidth with its fast frequency
398
tuning rate ensures that the spectra recorded by the HSR contain negligible broadening by
399
the apparatus function. In that case the absorbance AHRS (˜ ν ) = A(˜ ν ) and the corresponding
400
p and T values can be found by minimizing Eq. 6 and then using Eq. 7.
401
However, as seen in Fig. 5, in our case a small asymmetry was observed in the line profiles,
402
which could affect the retrieval of the integrated area. Indeed, for strong lines this effect is
403
clearly visible, as can be observed in the example shown in Fig. 9. The reason was determined
404
to be that, compared with previous experiments, the recorded spectral range was enlarged (to
405
12 cm−1 to record in the same scan the lines Qb (1, F1 ) and R(1, F1 ) located at 4324.9331 and
406
4337.5489 cm−1 , respectively) while maintaining the same repetition rate. Consequently, the
407
scanning speed was increased and the response time of the detector became visible. In this
408
case, the convolution of the absorption spectrum with the apparatus function, which could
409
be represented by an exponential decay (defined by the time constant τ ), does not commute
26
ACS Paragon Plus Environment
Page 26 of 51
Page 27 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
410
with the logarithm in the Beer-Lambert law. This leads to a systematic underestimation of
411
the retrieved line intensities, in the case of strong peaks, when they are obtained by simply
412
fitting the absorbance (Eq. 10). To correctly account for this effect, the recorded signal must
413
be fitted by the numerical representation of this convolution:
χ2HRS =
X
w(˜ νi ) {IHRS (˜ ν (ti )) − I0 · exp [−A(˜ ν (ti ))] ∗ exp [−∆t/τ ]}2 ,
(12)
∀˜ νi
414
note, that in this case, the convolution is in the temporal space, rather than in the spectral
415
space like for the LRS, due to the nature of the instrument effect. To this end, we have
416
developed a dedicated fitting program which was used to treat all HRS signals with the
417
time-decay constant deduced from the HRS detector bandwidth (1.7 µs). An example is
418
given in Fig. 9 where the fitting curve is drawn for the Rb (0, A1 ) transition recorded at 40
419
and 55 K. The residual curves confirm the correctness of the fit, for both “small” symmetric
420
and “strong” asymmetric lines. A small excess of noise is visible at 55 K due to the fact
421
that, in this region where the signal nearly drops to zero level (complete absorption), a lesser
422
weight is used in the fit.
423
Considering the aforementioned issues, the selection of the transitions used for analysis
424
was limited to those with a reasonable level of transmittance signal (approximately A < 3) in
425
order to have the best possible estimation of surface. Due to the large temperature range ad-
426
dressed in the study, there was no transition which would fit this criterion across all measured
427
spectra. Therefore we had to create different selections for temperatures. Q(1, F1 ), Q(2, F2 ),
428
Rb (0, A1 ), and R(1, F1 ) lines (see Table 1 for attribution) with Sηη0 > 10−21 cm/mol were
429
used between 40 and 50 K (Fig. 5b), while less intense Rb (1, F1 ), Q(3, F2 ), Q(4, F2 ), Q(4, F1 ),
430
and Rb (3, F1 ) lines were preferred at higher temperatures (Fig. 5c) up to 70 K. Unfortunately,
431
for three of them (Rb (1, F1 ), Q(4, F2 ), and Rb (3, F1 ) ), the accuracy of S is claimed to be
432
above 20 % (Tab. 1). In the treatment, their weights were reduced accordingly. For each
433
temperature, up to four values of pressure were obtained and had to be averaged correctly.
27
ACS Paragon Plus Environment
The Journal of Physical Chemistry
2 .2 5
Z o o m
2 5 x
2 .2 0
2 .1 5
2 .0
ν1+ν4 b
R
(0 , A 1
)
R S
( ~ )
1 .5
IH
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 28 of 51
1 .0
0 .5
F it 5 5 K 4 0 K
0 .0
0 .0 0 5 0 .0 0 0 -0 .0 0 5
4 3 3 3 .6 5 6
4 3 3 3 .6 6 8
W a v e n u m b e r (c m
4 3 3 3 .6 8 0 -1
)
Figure 9: Example of HRS signal fit. The graph shows the fit and residuals for the Rb (0, A1 ) transition of ν1 + ν4 recorded at 40 and 55 K.
28
ACS Paragon Plus Environment
Page 29 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
434
In such measurements, error analysis is crucial, and in our case, was done considering the
435
factors present in Eq. 7. The relative error δp/p was obtained by summing the contribution
436
to L (0.8%), Aηη0 (given by the fit process from 1 to 10 %), S (given by the error code in HI-
437
TRAN Table 1), and temperature T . For the evaluation of the temperature contribution, an
438
absolute accuracy of 0.25 K (as mentioned in the experimental part) yields a relative error of
439
0.6 % for the lowest temperature of 40 K, this propagates in Eq. 7 to less than 1 %, including
440
the three T -dependent terms present. The value of 0.25 K is discussed below because of the
441
exponential dependence of the pressure with the temperature.
442
Fig. 10a presents the data points corresponding to the nine transitions in a (ln(Π), 1/θ−1)
443
plot as defined above compared to Armstrong’s formula 12 converted to dimensionless Π and
444
θ. In this case, the dynamical range of the variation of the saturated pressure does not allow
445
us to clearly visualize the differences between the different data. Thus, two residuals are
446
plotted on the lower panels: the difference with the pressure given by Armstrong’s formula
447
(Fig. 10b) and the difference with a linear fit ln(Π) function of (1/θ − 1) (Fig. 10c).
448
Although the error bars are wide compared to the dispersion of the data, it is clear that
449
one data set is shifted. This corresponds to the Q(4, F2 ) transition located at 4333.2360 cm−1 .
450
The value of SHIT RAN 2016 = 6.978×10−23 cm/mol appears to be too low, leading to an overes-
451
timation of the pressure. For this reason, different methane databases (experimental or the-
452
oretical) have been considered and compared to HITRAN2016: GEISA2016, 34 Theo296K, 35
453
EXOMOL, 36 Mecasda, 37 and ESA-Birk. 38 The more experimental ones (GEISA, ESA-Birk)
454
give a value around 8.1 × 10−23 cm/mol. The relative difference of 16 % remains within the
455
accuracy of >20 % assumed in HITRAN2016 but could be corrected.
456
The ESA-Birk database was developed for the atmospheric instrument TROPOMI and
457
gives positions, intensities and also air-broadening of molecular lines including HTP pro-
458
file and line-mixing effects. The given uncertainty on the linestrength is less than 0.5 %.
459
Fortunately, their range studied (4200 to 4340 cm−1 ) covers that of our HRS data, so it is
460
straightforward to use this database rather than HITRAN2016. Confirmation was given
29
ACS Paragon Plus Environment
The Journal of Physical Chemistry
T (K ) 6 9 .7 5
5 6 .6 7
2ν +ν
(a )
2
T P
ln ( p ) - ln ( p
ν +ν
-8
3
Q -1 2
) m s tr.
Q
4 2
(4 , F (4 , F
ν +ν 1
4
1
2
b
Q )
b
R )
b
)
Q R R
b b
2
(1 (1 (3 (0 (3
4
, F 1 ) , F 1 ) , F 2 ) , A 1 ) , F 1 )
R (1 , F 1)
-1 6
(b )
0 .0 -0 .3
l n ( ΠE
x p
) - l n ( ΠA
Q
(2 , F
ν +ν 1
0 .3
4 1 .2 1
4 7 .7 2
)
-4
-0 .6
H IT R A N
2 0 1 6
0 .3
l n ( ΠE
x p
) - F it
(c )
0 .0
-0 .3
E S A - B ir k
0 .3
x p
) - F it
(d )
l n ( ΠE
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 30 of 51
0 .0
-0 .3 0 .3
0 .6
1/θ−1
0 .9
1 .2
Figure 10: Upper panel (a): saturated pressure data points. The relative differences of saturated pressure, measured by the spectroscopy using the HITRAN2016 database, between Armstrong formula from Ref. , 12 and fitted linear curve are plotted in panel (b) and (c), respectively. The last panel (d) shows the relative differences between p retrieved using the ESA-Birk database and a fitted linear curve. 30
ACS Paragon Plus Environment
Page 31 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
461
within our experiment, where a reference spectrum of methane at room temperature was
462
simultaneously recorded to calibrate each cold spectrum. It was possible to compare the two
463
simulated spectra from the HITRAN2016 and ESA-Birk databases to our reference spectrum,
464
and although minor discrepancies were observed with the first, the second spectrum almost
465
perfectly matched the experimental data. Tab. 3 presents the comparison between the two
466
databases for the nine transitions used in this study. The corrected data sets are shown as
467
residuals in the lower panel of Fig. 10d, where it can be observed that the major discrepancy
468
has disappeared and the error bars have been reduced by one order of magnitude. Tab. 4
469
compares the pressures derived from the two databases including their uncertainties. Table 3: Comparison between HITRAN 2016 and ESA-Birk databases. Energy (cm−1 ) 4324.9331 4328.4319 4330.1079 4330.6300 4332.2360 4333.3108 4333.6690 4334.4357 4337.5489
Label Qb (1, F1 ) Rb (1, F1 ) Q(2, F2 ) Qb (3, F2 ) Q(4, F2 ) Q(4, F1 ) Rb (0, A1 ) Rb (3, F1 ) R(1, F1 )
Eη (cm−1 ) 10.4816 10.4816 31.4424 62.8768 104.7800 104.7747 0 62.8768 10.4816
SHIT RAN 2016 (cm.mol−1 ) 1.307E-21 7.519E-23 1.097E-21 1.024E-22 6.978E-23 1.000E-22 4.087E-21 6.777E-23 1.041E-21
SESA−Birk (cm.mol−1 ) 1.375E-21 7.803E-23 1.113E-21 1.040E-22 8.155E-23 1.058E-22 4.295E-21 6.727E-23 1.137E-21
∆ (%) 4.9 3.6 1.4 1.5 14.4 5.5 4.8 -0.7 8.4
Eη is the energy of the lower state of the transition and the last column ∆ is the relative difference in % for the two databases.
470
The Clapeyron curve
471
The measurement of the saturated pressure at temperatures below the triple point allowed
472
us to consider the thermodynamics of sublimation and to determine its enthalpy for the
473
methane molecule. The following paragraph recalls the basic aspects and tries to compare
474
the results obtained by spectroscopy to previous determinations.
475
The boundary between two phases of a single (pure) substance is described by the Clapey-
31
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 32 of 51
Table 4: Analysis of pressure values in HRS from the two databases.
Database T δT 40.189 0.25 41.109 0.25 42.123 0.25 43.115 0.25 45.138 0.25 50.107 0.25 55.046 0.25 60.098 0.25 65.036 0.25 70.079 0.25
476
HRS ESA-Birk p δp δp/p 0.000727 1.3E-05 0.019 0.001496 4.0E-05 0.027 0.00300 7E-05 0.023 0.00598 6E-05 0.009 0.0202 0.0002 0.010 0.278 0.004 0.014 2.375 0.041 0.017 14.60 0.17 0.012 65.2 0.8 0.012 238.5 4.4 0.019
HITRAN2016 p δp 0.000764 3.4E-05 0.001575 7.8E-05 0.00320 0.00017 0.00630 0.00031 0.0212 0.0010 0.294 0.017 2.44 0.15 15.10 0.89 67.6 4.0 249.0 14.7
δp/p 0.044 0.049 0.052 0.050 0.045 0.058 0.062 0.059 0.060 0.059
ron equation: dp ∆s = , dT ∆v
(13)
477
with ∆s the difference of molar entropies and ∆v the difference of molar volumes of the solid
478
and gas phase.
479
Considering that our case is obviously far below the critical temperature, ∆v can be ap-
480
proximated by the molar volume of the gas phase vg for which the ideal gas law applies :
481
vg = RT /p, with R = 8.3144598 J.mol−1 .K−1 the molar gas constant. Accordingly, ∆s can
482
be characterized by the molar enthalpy of sublimation ∆sub H as follows: ∆s = ∆sub H/T as,
483
during phase change, temperature and pressure remain constant. Therefore: ∆sub H dT dp = p R T2
(14)
484
and, if the molar enthalpy of sublimation is considered as constant in the temperature range,
485
the integration gives:
ln(p) = − 486
∆sub H 1 + const., R T
(15)
which is the classical form of Clapeyron’s law. If ∆Hsub is constant up to the triple point, its 32
ACS Paragon Plus Environment
Page 33 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
487
value is directly related to the slope of the straigth line written with dimensionless parameters
488
θ and Π: ln(Π) = −
∆Hsub RTP T
1 −1 θ
(16)
Table 5: Experimental data used for the final fit, including both LRS and HRS data. T (K) 40.18 41.10 42.12 43.11 45.14 50.10 55.05 60.10 65.03 70.08 53.13 54.14 55.14 59.13 64.14 69.14 76.14 90.69
θ = T /TT P 0.443 0.453 0.464 0.475 0.498 0.552 0.607 0.663 0.717 0.773 0.586 0.597 0.608 0.652 0.707 0.762 0.839 1
Π = p/pT P 6.22E-08 1.28E-07 2.56E-07 5.11E-07 1.72E-06 2.37E-05 2.03E-04 1.24E-03 5.58E-03 2.04E-02 6.95E-05 1.11E-04 1.93E-04 8.54E-04 3.99E-03 1.71E-02 7.98E-02 1
Type HRS HRS HRS HRS HRS HRS HRS HRS HRS HRS LRS LRS LRS LRS LRS LRS LRS TP
x = (1/θ − 1) 1.256 1.205 1.152 1.103 1.009 0.809 0.647 0.509 0.394 0.294 0.706 0.675 0.644 0.533 0.414 0.311 0.191 0
δx 1.40E-02 1.34E-02 1.28E-02 1.22E-02 1.11E-02 9.02E-03 7.48E-03 6.27E-03 5.36E-03 4.61E-03 1.61E-02 1.55E-02 1.49E-02 1.30E-02 1.10E-02 9.48E-03 7.82E-03 3.30E-04
y = lnΠ -16.592 -15.871 -15.177 -14.487 -13.272 -10.648 -8.500 -6.686 -5.187 -3.892 -9.573 -9.099 -8.548 -7.064 -5.524 -4.068 -2.528 0.000
δy = δp/p 1.80E-02 2.60E-02 1.80E-02 5.00E-03 7.00E-03 1.00E-02 1.50E-02 1.00E-02 1.00E-02 1.70E-02 8.20E-02 1.02E-01 5.90E-02 6.70E-02 5.80E-02 7.00E-02 5.80E-02 5.00E-03
σ 0.19 0.18 0.17 0.16 0.15 0.12 0.10 0.08 0.07 0.06 0.23 0.23 0.20 0.18 0.16 0.14 0.12 0.01
The last column σ is the error including the linear fit uncertainties δy on y and δx on x, calculated by σ 2 = δy 2 +a2 δx2 , where a = −13.15 ± 0.1 is the slope of the straight line, y = ax + b. The values of T P : TT P = (90.6941 ± 0.0025) K and pT P = (116.96 ± 0.02) × 10−2 Pa, were taken from Ref. 10 489
Indeed, as shown in Fig. 7 for LRS and Fig. 10a for HRS, the data points are nearly aligned
490
on a straight line, and a preliminary value of the sublimation enthalpy can be estimated using
491
Eq. 16. In fact, our data sets are limited to the range 40-77 K and Eq. 15 is preferred for
492
a weighted fit on this limited temperature range y = ax + b with y = ln(Π), x = 1/θ and
493
a the slope ∆Hsub /RTT P . To carefully evaluate uncertainties, it is important to consider
494
the absolute accuracy of T in each set-up and to propagate it on the pressure. So, the
495
uncertainty σy to be regarded on includes δy = δΠ/Π, and that of the measurement of 33
ACS Paragon Plus Environment
The Journal of Physical Chemistry
496
x = 1/θ, aδx = aδθ/θ2 . The expression to be minimized is thus:
χ2 =
X (y − ax − b)2
(17)
δy 2 + (aδx)2
q
δy 2 + (aδx)2 . The data points are presented in Tab. 5 and
497
considering an error σy equal to
498
the residuals are plotted in Fig. 11. The result of the fit is a slope of a = −13.15 ± 0.10 and
499
an intercept of b = 13.13 ± 0.160. With the value of RTP T = 753.872 J, the molar enthalpy
500
of sublimation is (9910 ± 75) J.mol−1 .
T (K ) 9 0 .6 7 0 .3
6 9 .7 5
5 6 .6 7
4 7 .7 2
4 1 .2 1
0 .3
0 .6
0 .9
1 .2
H R S L R S T P
0 .2 0 .1
l n ( Π) - F i t
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 34 of 51
0 .0 -0 .1 -0 .2 -0 .3 0 .0
1/θ − 1
Figure 11: Residuals from the fit ln(Π) function of (1/θ − 1) on the whole range 40-90 K. The Triple point is included in the fit with its uncertainties.
501
Looking at Fig. 11 we can conclude that the data points from the two independent exper-
502
iments overlap well on the common temperature interval confirming correct calibrations of
503
the two experiments (mainly the optical path measurements). Secondly, considering the un34
ACS Paragon Plus Environment
Page 35 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
504
certainties of our measurements, it would be tempting to extrapolate the curve to the triple
505
point. Putting θ = 1 gives a pressure of p = 11518(710) Pa which is compatible with the
506
value pT P = 11696(2) Pa. This extrapolation could suggest to take ∆sub H constant on the
507
whole 40-90.6941 K range and to include the triple point in the data. However, this happens
508
because of too large experimental uncertainties on the pressure and we will see in the next
509
paragraph that a thermodynamical approach based on the measurement of heat capacity Cp
510
for both gas and solid phases yields a small dependence of ∆sub H with the temperature.
511
512
The following equation expresses the enthalpy of sublimation at a temperature T below the triple point by:
∆sub H(T ) = ∆sub H(TT P ) +
Z
T
TT P
g s (CpM (T ) − CpM (T ))dT
(18)
513
g s (T ) and CpM (T ) are the isobaric molar heat capacities in the gas and solid phases, where CpM
514
respectively, and ∆sub H(TT P ) is the molar enthalpy of sublimation at the triple point tem-
515
perature TT P . Its value is equal to the sum of the easily derived enthalpies of vaporization
516
and fusion: ∆sub H(TT P ) = ∆vap H(TT P ) + ∆f us H(TT P )
(19)
517
The former value was calculated from a polynomial equation ∆vap H(T ) = f (T /TT P )
518
containing four fitted parameters and is suitable for temperatures from the triple point to
519
the critical point. 39 This value ∆vap H(TT P ) = 8721 J.mol−1 is not far from ∆vap H(TT P ) =
520
8731.29 J.mol−1 obtained by Setzman and Wagner 10 using a more complete analysis of ex-
521
perimental data obtained between 60 and 3000 K. The enthalpy of fusion at the triple point
522
is ∆f us H(TT P ) = 928.4 J.mol−1 , determined by Colwell et al. 15 in the range 90.659-92.325 K.
523
Here, ∆sub H(TT P ) will be taken to be (9659 ± 12) J.mol−1 . Below the triple point, the value
524
g of CpM (T ) was calculated from the molecular properties. 40 A very small change is observed
525
from 60 K (4.0014×R) to 90.6941 K (4.0021×R) very near to the value of 4R deduced from
526
g the kinetic theory. CpM (T ) is taken at 4.0021 × 8.3144598 = 33.275 J.K−1 .mol−1 . Solid
35
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 36 of 51
527
heat capacities were measured by Colwell et al. between 5 and 90.37 K. Taking into account
528
that solid methane exists in two differents solid phases, we considered only data in the
529
temperature range of the present work, between 30 and 88.4 K, corresponding to the so-
530
called Phase I solid methane. To perform the integration of Eq. 18, it was useful to express
531
g s CpM (T ) − CpM (T ) as a polynomial formula. Hence:
g s CpM (T ) − CpM (T ) = 34.2 − 1.20 T + 1.40 10−2 T 2 − 6.9 10−5 T 3
(20)
532
The model reproduces the experimental data better than the experimental uncertainty
533
estimated at 1 %. 15 The measurement at 90.4 K was not regarded in the fit, since it is
534
s too close to the triple point, for which a divergence of CpM is produced because of the
535
s discontinuity of the entropy. In order to estimate how the 1 % errors on CpM are propagated
536
to the enthalpy in the integration, two third order polynomial formulae similar to Eq. 20
537
have been calculated with Cps increased or decreased by this amount. The uncertainty of
538
12 J.mol−1 on ∆sub H(TT P ) in Eq. 18 was also included. Finally, the result is represented
539
on Fig. 12 by an uncertainty zone around the calculated enthalpy of sublimation ∆sub H(T ).
540
The maximum of the curve is obtained at about 53 K (∆sub H(T ) = 9870 J.mol−1 ) and the
541
variation of ∆sub H(T ) is less than 2 % within the range of temperature 30-90.6941 K. Due
542
to integration from the triple point Eq. 18, the uncertainty increases as the temperature
543
decreases reaching a value of about 20 J.mol−1 at 40 K. For comparison, on Fig. 12 we plot
544
a rectangle corresponding to our experimental determination ∆sub H = (9910 ± 75) J.mol−1
545
within our experimental range 40-77 K.
546
It can be observed that the difference between the experimental and calculated 15 ∆sub H(T ),
547
both determined independently, is not significant. This becomes all the more evident if we
548
consider that the uncertainties presented in the figure are one standard deviation. The physi-
549
cal meaning of the experimental enthalpy of sublimation is thus confirmed, and consequently,
550
the quality of experimental pressures and temperatures. However, their accuracy, and in par-
551
ticular that of the absolute value of T , is not yet high enough to observe a T -dependence of 36
ACS Paragon Plus Environment
Page 37 of 51
552
∆sub H(T ) as proposed by the model given in (Eqs. 18 and 20), the dependence being very
553
small as mentioned above. The required accuracy should be better than 1 %. On the other
554
hand, from the calculated curve it is clear that the average of molar enthalpy in the 40-77 K
555
range differs from the value at the triple point (9659 ± 12) J.mol−1 by more than 1%. This
556
confirms our reservations to include the triple point TT P , pT P in the data and to analyse
557
them restricted to our studied temperature range.
m o d e l a c c o r d in g to R e f.[1 4 ] + - σ1%
1 0 2 0 0
∆H
-1
S u b lim a tio n e n th a lp y ( J m o l )
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
( L in e b y lin e ) { T m in . , T m a x . }
1 0 1 0 0
∆H
( L R S + H R S fit) { + - σ ;T m in . , T m a x . }
4 3 3 7 .5 5 1 0 0 0 0
4 3 3 3 .6 7
4 3 3 0 .1 1
4 3 3 4 .4 4
4 3 2 4 .9 3
9 9 0 0
4 3 3 3 .3 1 4 3 2 8 .4 3
9 8 0 0
4 3 3 0 .6 3
4 3 3 2 .2 4
T P
9 7 0 0 3 0
4 0
5 0
6 0
7 0
8 0
9 0
T e m p e ra tu re (K ) Figure 12: Variation of the sublimation enthalpy with temperature from 30 K to the triple point calculated from measured solid and gas heat capacities Eq. 18 (blue curve) with 1 % uncertainty (greyed zone). Our experimental values are represented by the red squares corresponding to the nine individual transitions, by the green rectangle derived from both LRS and HRS pressures. The vertical error bars are the 1-σ uncertainty and the horizontal extension concerned the temperature range.
558
Following this remark, considering the HRS set-up, the relative uncertainty of the tem37
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
559
perature measurement is clearly better than its absolute accuracy of 0.25 K. Then, Eq. 14
560
tells us that the enthalpy, slowly temperature dependent, can be considered as the derivative
561
of the logarithm of the saturated pressure curve as a function of the inverse of the tempera-
562
ture. Thus, our measurements can be analysed in a different way, focusing on the variation
563
of the saturated pressure for a definite elevation of temperature (1 or 5 K in our case). In
564
addition, each transition can be separately treated over a limited temperature range where
565
its measurement is regarded confident. For our sets of transitions, we consider that enthalpy
566
remains constant within each range, such a value being attributed to the middle of the range.
567
In this process, the relative uncertainty on T is reduced to 0.03 K and the value of S does
568
not interfere as it only goes in the intercept value of the linear fit of ln(p) as a function of
569
1/T (see Eq. 7). The results are reported in Fig. 12 and show a small but not significant in-
570
crease of enthalpy with decreasing temperature. Considering a constant value over the range
571
40-70 K, an average value including the nine transitions leads to the value 9885 ± 75 J.mol−1 ,
572
practically identical value to the one obtained from the linear fit of the average pressures
573
(9910 ± 75 J.mol−1 ).
574
Finally, in Fig. 13 we compare the saturated pressures derived from the thermodynamics
575
(Colwell et al. 15 ) with the experimental measurements of Tickner et al. , 13 the derived curve
576
from Armstrong et al. 12 , and the present work. Let us note here that the pressure varies
577
over six orders of magnitude within the temperature range (and even seven orders if the
578
triple point is included). For clarity, the relative difference from the thermodynamic model
579
is plotted. An uncertainty of 1 % on the heat capacity measurement has been propagated in
580
the pressure curve (grey zone) and gives an error of 3 % at 40 K. Armstrong’s curve shows a
581
discrepancy of 18 % at 54 K, Tickner’s lowest points are 100 % off around 50 K but this was
582
already mentioned in his paper, our data points lie from 18 to 5 % lower in the range from
583
40 to 77 K.
38
ACS Paragon Plus Environment
Page 38 of 51
Page 39 of 51
1 .2
m o d e l b a s e d o n R e f. [1 4 ] ( + - σ1 % ) R e f. [1 2 ] R e f. [1 3 ] T h is w o r k
0 .8 0 .6
s o u rc e
- p
R e f.[1 4 ]
)/p
R e f.[1 4 ]
1 .0
(p
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
0 .4 0 .2 0 .0 -0 .2 -0 .4 4 0
5 0
6 0
7 0
8 0
9 0
T e m p e ra tu re (K )
Figure 13: Comparison between the thermodynamic model derived from Colwell et al. 15 and experimental data for the saturated vapor pressure below the triple point. For clarity, the relative difference is plotted. Green curve: Armstrong’s formula, 12 Red squares: Tickner et al. data, 13 Blue dots: our work.
39
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
584
Discussion
585
In our experiment the uncertainty, dominated by the absolute accuracy of the temperature
586
(Fig. 11), did not allow for the detection of a significant variation of the molar enthalpy of
587
sublimation nor for a more refined computed curve, 41 as was done for water. 8 However, the
588
retrieved value of ∆sub H = (9910 ± 75) J.mol−1 for the temperature range of 40-77 K was in
589
agreement with the variation of ∆Hsub calculated for the 30-90 K range from Eq. 18 (Fig. 12).
590
To go further, it is important to improve the pressure measurement below 10−2 P a as well
591
as the absolute value of temperature. As demonstrated, spectroscopic measurement of the
592
absorption of the gas is a good candidate for the first challenge. In principle, this can be done
593
considering only one transition. As an example, the absorption of the Rb (0, A1 ) transition
594
can be considered. Measurements in the 41-50 K range has an uncertainty better than 1 %,
595
which yields saturated pressure values given in Tab. 6 with a focus on the temperature
596
TB and TA (temperature of the second sensor used for the servo-locking process). The
597
choice of TB was guided by the argument of an independent measurement with no coupling
598
with electronics. However, here, TA is also a temperature measurement at another point
599
of the cell, and less than TB . Considering that saturated pressure is determined by the
600
lowest temperature in the cell, it seems to be more appropriate to choose TA instead of TB .
601
Do TA or TB really represents the real temperature of the gas? The difference TB − TA
602
varies from 0.081 K at 70 K to 0.19 K at 40 K and influences by less than 1 % the pressure
603
derived from Eq. 7. However, for enthalpy determination, these pressure values have to be
604
connected to a precise temperature. Using TA instead of TB in the treatment above including
605
the nine transitions has been tested, and led to a decrease in the enthalpy value of 1 % ,
606
∆sub H = 9807(57) J.mol−1 instead of 9885(75) J.mol−1 . Let us recall that around 40 K a
607
variation of 7 % on the pressure is observed for a temperature difference of 0.1 K, and 90 %
608
for a gap of 1 K.
609
Therefore, the question of the temperature measurement must be posed. Instead of the
610
contact sensors, there are two possible ways on our HRS set-up to derive the temperature 40
ACS Paragon Plus Environment
Page 40 of 51
Page 41 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
Table 6: Pressure derived from Eq. 7 for the transition Rb (0, A1 ) considering the two available temperatures of the cell in HRS, TB and TA . TB (K) 40.189 41.109 42.123 43.115 45.138 50.107
TA (K) 39.996 40.947 41.963 42.969 44.992 49.984
p(TB ) (Pa) 7.32E-04 1.47E-03 2.98E-03 5.99E-03 2.02E-02 2.75E-01
p(TA ) (Pa) 7.24E-04 1.46E-03 2.96E-03 5.95E-03 2.00E-02 2.73E-01
δp (Pa) 1.33E-05 1.23E-05 1.61E-05 1.79E-05 7.86E-05 9.07E-04
δp/p
(p(TB ) − p(TA ))/p
0.018 0.008 0.005 0.003 0.004 0.003
-0.011 -0.009 -0.009 -0.008 -0.007 -0.006
611
from the spectra. The first one is to consider the shape of each transition and to measure
612
their Gaussian width driven by the Doppler effect. The result is quite good but not enough to
613
reach even a 1 K precision, especially at low temperature where the signal to noise decreases.
614
The second method is well known to spectroscopists looking at the shape of R or P branches
615
which results from the Maxwell distribution of the fundamental state. This method can be
616
simplified in our case, by measuring the ratio of the integrated absorbances σ1 and σ2 of two
617
transitions with different lower state energies. An example is given with the R(0, A1 ) and
618
Qb (1, F1 ) transitions at 4333.669 and 4324.9331 cm−1 , respectively. T can be derived from
619
the following equation: σ1 ln σ2
S1 (T ) = ln S2 (T )
!
!
S1 (T0 ) 1 1 = ln − (E1 − E2 ) × − S2 (T0 ) kT kT0
(21)
620
where E1 and E2 are the energies of the lower level (fundamental state) of transitions 1 and
621
2 (accurately measured by rotational spectroscopy), and S1 (T0 ) and S2 (T0 ) the oscillator
622
strength at T0 = 296K. Unfortunately, the uncertainties on σ1 and σ2 and those on S1 (T0 )
623
and S2 (T0 ) were not low enough to reach uncertainties in the range of 1 K for the temperature.
624
Accurate values of the transition strengths Sηη0 are required to derive temperature but
625
also to obtain reliable pressure. Only nine transitions were used in the study, more lines
626
could be used in the relevant temperature range if their transition strengths are confident.
627
On the other hand, one can imagine reversing the problem and using the low-temperature
41
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
628
saturated vapor pressure spectra recorded at high resolution to increase the accuracy of
629
some transitions strengths. In particular, transitions corresponding to low J values are often
630
better measured at low temperature, taking advantage of larger absorption, better resolution
631
(Doppler effect), and less dense spectra (less blended transitions). Indeed, the work is
632
continuing in this direction as such an improvement would be useful in the perspective of
633
new space missions . 42
634
Also, as mentioned in the text, the temperature of 40 K is a limit where the population
635
ratio of the three nuclear spin isomers of methane starts to differ from its value at room
636
temperature. Data analysis at lower temperatures would allow consideration of the nuclear
637
spin conversion dynamics. 9
638
Conclusion
639
In this paper, we have applied absorption spectroscopy techniques to probe gas phase
640
methane at temperatures below 91 K to deduce its Clapeyron solid-gas equilibrium curve.
641
To this end, we used two distinct experimental setups.
642
The first one is a standard commercial “low resolution” spectrometer which allowed us
643
to measure the Clapeyron curve in the range 53-77 K and demonstrate that it is possible to
644
derive this type of data at low pressures (down to ∼ 10−1 Pa for methane) and temperatures
645
using a spectrometer that is not generally used to characterize the gas phase. This limit cor-
646
responds to the limit of detection raised by standard manometric measurements according
647
to Fray et al. 43 In that study, it is claimed that several molecules have no or poor measure-
648
ments of the p(T ) law in that pressure domain (C2 H2 , CH2 =C=CH2 , C6 H6 , HCN,. . .) and
649
that the empirical relations deduced at higher pressures cannot be extrapolated to lower
650
temperatures. With a low resolution set-up it becomes possible to extend our knowledge of
651
the saturation pressure curve for many molecules of astrophysical interest.
652
A second setup used a home-made high resolution laser source at 2.3 µm performing
42
ACS Paragon Plus Environment
Page 42 of 51
Page 43 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
653
high sensitivity absorption measurements which yielded novel data in the thermal range 40-
654
50 K (with p down to ∼ 10−3 Pa) increasing by two orders of magnitude the validity of the
655
Clapeyron law. On the other side, it is possible to measure pressure up to 70 K, by choosing
656
transitions with low line strength S as soon as this value is known with enough precision.
657
With the two techniques, we were able to provide reliable and precise information about
658
the methane vapor pressure in the 40-91 K range corresponding to a pressure variation of
659
seven orders of magnitude. We have shown that this in situ method is more convenient than
660
standard manometric measurements. Reasons include the absence of the transpiration effect,
661
which typically occurs when using gauges, and also that the measurement of the targeted
662
molecular density is not sensitive to any impurity.
663
Acknowledgments
664
This work was supported by the Programme National “Physique et Chimie du Milieu Inter-
665
stellaire” (PCMI) of CNRS/INSU with INC/INP co-funded by CEA and CNES. This work
666
was also supported by the French National Research Agency (Project ANR GASOSPIN
667
number 09-BLAN-0066-01). This project has received funding from the European Union’s
668
Horizon 2020 research and innovation programme under grant agreement No 692335. G.V.
669
thanks NSP of SR for financial support of her internship in Laboratoire PhLAM Lille.
670
References
671
672
673
674
(1) Chandra, V. Fundamentals of Natural Gas: An International Perspective; PennWell Corp., 2006. (2) Chynoweth, D. P.; Owens, J. M.; Legrand, R. Renewable Methane from Anaerobic Digestion of Biomass. Renewable Energy 2000, 22, 1–8.
43
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
675
676
(3) Wuebbles, D. J.; Hayhoe, K. Atmospheric Methane and Global Change. Earth Sci. Rev. 2002, 57, 177–210.
677
(4) Lutz, B. L.; Owen, T.; Cess, R. D. Laboratory Band Strengths of Methane and their
678
Application to the Atmospheres of Jupiter, Saturn, Uranus, Neptune, and Titan. As-
679
trophys. J. 1976, 203, 541–551.
680
681
682
683
(5) Cordier, D.; Mousis, O.; Lunine, J. I.; Lavvas, P.; Vuitton, V. An Estimate of the Chemical Composition of Titan’s Lakes. Astrophys. J. 2009, 707, L128–L131. (6) Formisano, V. Detection of Methane in the Atmosphere of Mars. Science 2004, 306, 1758–1761.
684
(7) Daud´e, B.; Elandaloussi, H.; Janssen, C. On the Gas Dependence of Thermal Tran-
685
spiration and a Critical Appraisal of Correction Methods for Capacitive Diaphragm
686
Gauges. Vacuum 2014, 104, 77–87.
687
(8) Bielska, K.; Havey, D. K.; Scace, G. E.; Lisak, D.; Harvey, A. H.; Hodges, J. T. High-
688
Accuracy Measurements of the Vapor Pressure of Ice Referenced to the Triple Point.
689
Geophys. Res. Lett. 2013, 40, 6303–6307.
690
ˇ (9) Cacciani, P.; Cosl´eou, J.; Khelkhal, M.; Cerm´ ak, P.; Puzzarini, C. Nuclear Spin Con-
691
version in CH4 : A Multichannel Relaxation Mechanism. J. Phys. Chem. A 2016, 326,
692
9–16.
693
(10) Setzmann, U.; Wagner, W. A New Equation of State and Tables of Thermodynamic
694
Properties for Methane Covering the Range from the Melting Line to 625 K at Pressures
695
up to 100 MPa. J. Phys. Chem. Ref. Data 1991, 20, 1061–1155.
696
697
(11) Eucken, A.; Karwat, E.; Fried, F. Die Konstantei der Thermodynamischen Dampfdruckgleichung bei Mehratomigen Molekeln. Z. Phys. 1924, 29, 1–35.
44
ACS Paragon Plus Environment
Page 44 of 51
Page 45 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
698
699
700
701
702
703
The Journal of Physical Chemistry
(12) Armstrong, G. T.; Brickwedde, F.; Scott, R. Vapor Pressures of Methanes. J. Res. Nat. Bur. Stand. 1955, 55, 39–52. (13) Tickner, A. W.; Lossing, F. P. The Measurement of Low Vapor Pressures by Means of a Mass Spectrometer. J. Phys. Chem. 1951, 55, 733–740. (14) “Methane - Phase Change data” in NIST Chemistry WebBook, NIST Standard Reference Database Number 69. http://webbook.nist.gov, (accessed June, 2018).
704
(15) Colwell, J. H.; Gill, E. K.; Morrison, J. A. Thermodynamic Properties of CH4 and CD4 .
705
Interpretation of the Properties of the Solids. J. Chem. Phys. 1963, 39, 635–653.
706
(16) Mondelain, D.; Camy-Peyret, C.; Mantz, A. W.; Tang, E.; Valentin, A. Performance
707
of a Herriott Cell, Designed for Variable Temperatures between 296 and 20 K. J. Mol.
708
Spectrosc. 2007, 241, 18–25.
709
ˇ (17) Cerm´ ak, P.; Cacciani, P.; Cosl´eou, J.; Khelkhal, M.; Hovorka, J.; Jeseck, P.;
710
Michaut, X.; Moudens, A.; Fillion, J.-H.; Bertin, M. et al. Observation of Methane
711
Nuclear Spin Isomers in Gas Phase at Low Temperature. J. Mol. Spectrosc. 2012, 279,
712
37–43.
713
(18) Gordon, I.; Rothman, L.; Hill, C.; Kochanov, R.; Tan, Y.; Bernath, P.; Birk, M.;
714
Boudon, V.; Campargue, A.; Chance, K. et al. The HITRAN2016 Molecular Spectro-
715
scopic Database. J. Quant. Spectrosc. Radiat. Transfer 2017, 203, 3–69.
716
(19) Rothman, L.; Rinsland, C.; Goldman, A.; Massie, S.; Edwards, D.; Flaud, J.-M.; Per-
717
rin, A.; Camy-Peyret, C.; Dana, V.; Mandin, J.-Y. et al. The HITRAN Molecular
718
Spectroscopic Database and HAWKS (HITRAN Atmospheric Workstation): 1996 Edi-
719
tion. J. Quant. Spectrosc. Radiat. Transfer 1998, 60, 665–710.
720
(20) HITRAN online. http://www.hitran.org, (accessed June, 2018).
45
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
721
(21) Manca, C.; Martin, C.; Allouche, A.; Roubin, P. Experimental and Theoretical Rein-
722
vestigation of CO Adsorption on Amorphous Ice. J. Phys. Chem. B 2001, 105, 12861–
723
12869.
724
725
(22) Harris, F. On the Use of Windows for Harmonic Analysis with the Discrete Fourier Transform. Proc. IEEE 1978, 66, 51–83.
726
(23) Georges, R.; Michaut, X.; Moudens, A.; Goubet, M.; Pirali, O.; Soulard, P.; Asselin, P.;
727
Huet, T.; Roy, P.; Fournier, M. et al. Nuclear Spin Symmetry Conservation in 1 H2 16 O
728
Investigated by Direct Absorption FTIR Spectroscopy of Water Vapor Cooled Down
729
in Supersonic Expansion. J. Phys. Chem. A 2017, 121, 7455–7468.
730
731
732
733
(24) Stohner, J.; Quack, M. In Handbook of High-resolution Spectroscopy; John Wiley & Sons, Ltd, 2011; pp 263–324. (25) Verhoeven, J. W. Glossary of Terms Used in Photochemistry (IUPAC Recommendations 1996). Pure Appl. Chem. 1996, 68, 2223–2286.
734
ˇ (26) Cerm´ ak, P.; Triki, M.; Garnache, A.; Cerutti, L.; Romanini, D. Optical-Feedback
735
Cavity-Enhanced Absorption Spectroscopy Using a Short-Cavity Vertical-External-
736
Cavity Surface-Emitting Laser. IEEE Photon. Technol. Lett. 2010, 22, 1607–1609.
737
(27) Cerutti, L.; Ducanchez, A.; Narcy, G.; Grech, P.; Boissier, G.; Garnache, A.;
738
Tourni´e, E.; Genty, F. GaSb-based VCSELs Emitting in the Mid-Infrared Wavelength
739
Range (2–3 µm) Grown by MBE. J. Cryst. Growth 2009, 311, 1912–1916.
740
ˇ (28) Cerm´ ak, P.; Hovorka, J.; Veis, P.; Cacciani, P.; Cosl´eou, J.; Romh, J. E.; Khelkhal, M. 14
Spectroscopy of
742
Laser Source. J. Quant. Spectrosc. Radiat. Transfer 2014, 137, 13–22.
743
NH3 and
15
741
NH3 in the 2.3 µm Spectral Range with a New VECSEL
(29) Grundy, W. The Temperature-Dependent Spectrum of Methane Ice I between 0.7 and
46
ACS Paragon Plus Environment
Page 46 of 51
Page 47 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
744
5 µm and Opportunities for Near-Infrared Remote Thermometry. Icarus 2002, 155,
745
486–496.
746
(30) Kobashi, K.; Okada, K.; Yamamoto, T. Theory of Phase Transitions in Solid Methanes.
747
XI. Infrared and Raman Spectra of the ν3 and ν4 Modes in Phase II of Solid CH4 . J.
748
Chem. Phys. 1977, 66, 5568–5577.
749
(31) Tennyson, J.; Bernath, P. F.; Campargue, A.; Cs´asz´ar, A. G.; Daumont, L.;
750
Gamache, R. R.; Hodges, J. T.; Lisak, D.; Naumenko, O. V.; Rothman, L. S. et al. Rec-
751
ommended Isolated-Line Profile for Representing High-Resolution Spectroscopic Tran-
752
sitions (IUPAC Technical Report). Pure Appl. Chem. 2014, 86 .
753
754
755
756
(32) Kochanov, V. On Systematic Errors in Spectral Line Parameters Retrieved with the Voigt Line Profile. J. Quant. Spectrosc. Radiat. Transfer 2012, 113, 1635–1641. (33) Takaishi, T.; Sensui, Y. Thermal Transpiration Effect of Hydrogen, Rare Gases and Methane. Trans. Faraday Soc. 1963, 59, 2503.
757
(34) Jacquinet-Husson, N.; Armante, R.; Scott, N.; Ch´edin, A.; Cr´epeau, L.; Boutam-
758
mine, C.; Bouhdaoui, A.; Crevoisier, C.; Capelle, V.; Boonne, C. et al. The 2015 Edition
759
of the GEISA Spectroscopic Database. J. Mol. Spectrosc. 2016, 327, 31–72.
760
(35) Rey, M.; Nikitin, A. V.; B´ezard, B.; Rannou, P.; Coustenis, A.; Tyuterev, V. G. New 12
CH4 in the 0–13400 cm−1 Range: Ap-
Accurate Theoretical Line Lists of
762
plication to the Modeling of Methane Absorption in Titan’s Atmosphere. Icarus 2018,
763
303, 114–130.
764
765
CH4 and
13
761
(36) Yurchenko, S. N.; Tennyson, J. ExoMol Line Lists – IV. The Rotation–Vibration Spectrum of Methane up to 1500 K. Mon. Not. R. Astron. Soc. 2014, 440, 1649–1661.
766
(37) Ba, Y. A.; Wenger, C.; Surleau, R.; Boudon, V.; Rotger, M.; Daumont, L.; Bonhom-
767
meau, D. A.; Tyuterev, V. G.; Dubernet, M.-L. MeCaSDa and ECaSDa: Methane and 47
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
768
Ethene Calculated Spectroscopic Databases for the Virtual Atomic and Molecular Data
769
Centre. J. Quant. Spectrosc. Radiat. Transfer 2013, 130, 62–68.
770
(38) Birk, M.; Wagner, G.; Loos, J.; Mondelain, D.; Campargue, A. ESA SEOM-IAS Mea-
771
surement Database 2.3 µm Region. https://doi.org/10.5281/zenodo.1009122, (ac-
772
cessed July, 2018).
773
774
775
776
(39) Somayajulu, G. R. New Equations for Enthalpy of Vaporization from the Triple Point to the Critical Point. Int. J. Thermophys. 1988, 9, 567–575. (40) McDowell, R. S.; Kruse, F. H. Thermodynamic Functions of Methane. J. Chem. Eng. Data 1963, 8, 547–548.
777
(41) Wagner, W.; Riethmann, T.; Feistel, R.; Harvey, A. H. New Equations for the Subli-
778
mation Pressure and Melting Pressure of H2 O Ice Ih. J. Phys. Chem. Ref. Data 2011,
779
40, 043103.
780
(42) Harrison, J. J.; Bernath, P. F.; Kirchengast, G. Spectroscopic Requirements for AC-
781
CURATE, a Microwave and Infrared-Laser Occultation Satellite Mission. J. Quant.
782
Spectrosc. Radiat. Transfer 2011, 112, 2347–2354.
783
784
(43) Fray, N.; Schmitt, B. Sublimation of Ices of Astrophysical Interest: A Bibliographic Review. Planet. Space Sci. 2009, 57, 2053–2080.
785
Appendix
786
Determination of the path length in the LRS
787
The precise optical path length, LLRS , was needed to model the spectra recorded with the
788
low resolution setup. For the high resolution setup, this length has been determined in a
789
previous work: LHRS = 692 ± 5 cm. 17 For the LRS, we introduced methane progressively 48
ACS Paragon Plus Environment
Page 48 of 51
Page 49 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
790
into the cell from 0.755 up to 6.938 Torr at a temperature of 76.19 K, recording spectra
791
corresponding to each pressure. The spectra of the ν2 + ν3 band are displayed in Fig. 14a,
792
where only the R(0) and R(1) lines correspond to an isolated transition. Fig. 14c displays for
793
each of the six pressures the difference between experimental and modeled spectra, defined
794
above, and varying the path length LLRS . Each curve has a minimum situated between
795
1.711 and 1.781 cm. From that, we have determined that LLRS = 1.746 ± 0.035 cm. Fig. 14b
796
displays the simulated spectra corresponding to the minima of Fig. 14c. We see that the
797
agreement is satisfactory.
49
ACS Paragon Plus Environment
The Journal of Physical Chemistry
-1
W a v e n u m b e r (c m
A
o b s
= 1 - I ( ~ ) / I 0 ( ~ )
4 5 5 5
4 5 5 6
4 5 6 7
4 5 7 8
0 .0 4
C H 0 .0 3
)
4 5 8 0
4 5 9 0
4 5 9 1
( ν2 + ν3 ) 4
(a )
0 .0 2 0 .0 1 0 .0 0
R (1 )
R (0 )
R (3 ) (b )
R (2 )
A
c a lc
0 .0 3 0 .0 2 0 .0 1 -0 .5
(c ) -1 .0
L
L R S
= 1 .7 4 6 + /-0 .0 3 5 c m
)
-1 .5
χ2 ( L
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 50 of 51
-2 .0 -2 .5 -3 .0
p (to 0 .7 1 .3 1 .9 2 .4 4 .1 6 .9
rr) 5 5 9 0 1 6 2 7 6 8 3 8
T = 7 6 .1 9 K
-3 .5 1 .5
1 .6
L
1 .7 c e ll
1 .8
1 .9
(c m )
Figure 14: Determination of the optical path length of the absorption cell used with the LRS. Plot (a) shows experimental absorption while plot (b) shows the best simulated absorption obtained for six pressures between 0.755 and 6.938 Torr at 76.19 K. Plot (c) shows the χ2 difference between experimental and modeled spectra.
50
ACS Paragon Plus Environment
Page 51 of 51
TOC Graphic:
2.3 mm R(0)
p (Pa)
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
T=40-77K
T (K)
798
51
ACS Paragon Plus Environment