This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:

The 12CO2 and 13CO2 Absorption Bands as Tracers of the Thermal History of Interstellar Icy Grain Mantles

, , , and

Published 2018 December 10 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Jiao He et al 2018 ApJ 869 41 DOI 10.3847/1538-4357/aae9dc

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/869/1/41

Abstract

Analyses of infrared signatures of CO2 in water-dominated ices in the ISM can give information on the physical state of CO2 in icy grains and on the thermal history of the ices themselves. In many sources, CO2 was found in the "pure" crystalline form, as signatured by the splitting in the bending mode absorption profile. To a large extent, pure CO2 is likely to have formed from segregation of CO2 from a CO2:H2O mixture during thermal processing. Previous laboratory studies quantified the temperature dependence of segregation, but no systematic measurement of the concentration dependence of segregation is available. In this study, we measured both the temperature dependence and concentration dependence of CO2 segregation in CO2:H2O mixtures, and found that no pure crystalline CO2 forms if the CO2:H2O ratio is less than 23%. Therefore, the segregation of CO2 is not always a good thermal tracer of the ice mantle. We found that the position and width of the broad component of the asymmetric stretching vibrational mode of 13CO2 change linearly with the temperature of CO2:H2O mixtures, but are insensitive to the concentration of CO2. We recommend using this mode, which will be observable toward low-mass protostellar envelopes and dense clouds with the James Webb Space Telescope, to trace the thermal history of the ice mantle, especially when segregated CO2 is unavailable. We used the laboratory measured 13CO2 profile to analyze the ISO-SWS observations of ice mantles toward Young Stellar Objects, and the astrophysical implications are discussed.

Export citation and abstract BibTeX RIS

1. Introduction

CO2 is abundant in quiescent and star-forming molecular clouds where it is found in ices with abundance in the 10–50% range with respect to water. Solid state CO2 is mostly detected in the mid-infrared absorption through the asymmetric stretching mode ν3 at ∼2350 cm−1 (Gerakines et al. 1999; Nummelin et al. 2001; Noble et al. 2013) and through the bending mode at ∼665 cm−1 (Gerakines et al. 1999; Pontoppidan et al. 2008; Ioppolo et al. 2013; Noble et al. 2013). Additional modes are detected as well, such as the combination modes ν1 + ν3 at 3708 cm−1 and 2ν2 + ν3 at 3600 cm−1 (Gerakines et al. 1999; Keane et al. 2001). Unlike CO, which has a high abundance only in highly shielded regions, CO2 has the same threshold of formation as water (Bergin et al. 2005; Whittet et al. 2009), which means that CO2 is mixed with water in pristine polar ices coating dust grains (Whittet et al. 2009). The column density ratio of CO2:H2O varies between 10% and 50%, depending on the specific cloud. Laboratory measurements of CO2-containing ice mixtures have found that the infrared absorption profile of CO2 strongly depends on the physical and chemical environment in the ice, such as the temperature and the ice compositions (Gerakines et al. 1999; Hodyss et al. 2008; Öberg et al. 2009; He & Vidali 2018). This makes CO2 a very good candidate to trace the composition and physical condition of the ice mantle.

Since the ${\nu }_{3}$ asymmetric stretch is particularly strong and often saturated, the ν2 bending mode at ∼650 cm−1 and the 13CO2 asymmetric stretching mode at ∼2280 cm−1 are often used instead to study CO2 in ice mantles. So far, most of the observations of solid state CO2 are through the 650 cm−1 feature. Pontoppidan et al. (2008) used Spitzer to systematically study the 650 cm−1 feature in Young Stellar Objects (YSOs). By comparing the observed absorption profile with laboratory measurements of different CO2-containing ice mixtures, they found that the observed spectra can be fit well using five different components, each representative of CO2 in different ice mixtures measured in the laboratory. However, there are redundancies in the derived ice compositions that are further amplified by the effect of the poorly constrained grain shapes on the observed spectral profiles. The latter is not the case for 13CO2 because it is diluted by almost two orders of magnitude in the ice, resulting in a low polarizability and thus negligible grain shape effects (Boogert et al. 2000). Therefore, the observed 13CO2 spectra can be readily compared with laboratory measured spectra. Its asymmetric stretching feature cannot be observed with ground-based telescopes because of strong telluric absorption. The only comprehensive study of this feature is by Boogert et al. (2000) who used ISO-SWS to observe 13CO2 in 13 sightlines. However, due to limited sensitivity, only a small sample of sightlines could be observed, lacking in particular the envelopes of low-mass YSOs and quiescent dense molecular clouds. The forthcoming James Webb Space Telescope (JWST) is expected to cover this spectral region at orders of magnitude better sensitivity and somewhat higher spectral resolution (R = λλ ∼ 3000 versus 2000). A comprehensive set of laboratory measurements of solid state 13CO2 would facilitate the interpretation of JWST observations of solid CO2. One of the motivations of this work is to measure the absorption profile of 13CO2 ν3 mode in CO2:H2O mixtures at different mixing ratios and different temperatures, providing improved insights into the composition as well the thermal history of interstellar and circumstellar icy mantles.

In some of the sightlines observed by Pontoppidan et al. (2008) and Gerakines et al. (1999), the CO2 bending profile shows double splitting features. This is interpreted as the Davydov splitting—which occurs in crystals with more than one identical molecular species or unit per unit cell. It is commonly interpreted as an indication of "pure" crystalline CO2 (Pontoppidan et al. 2008; Isokoski et al. 2013; Cooke et al. 2016; Baratta & Palumbo 2017). Since ices in the ISM are water-dominated, the appearance of the splitting indicates the formation of segregated CO2 solids due to thermal processing. Öberg et al. (2009) measured the temperature dependence of the segregation of CO2 from CO2:H2O mixtures. The majority of their experiments were carried out with a CO2:H2O mixing ratio of 1:2, with only one measurement for 1:4, a more representative ratio for the ices coating grains (Boogert et al. 2015). From an Arrhenius fitting to the experimental data, they found an energy barrier of 1090 ± 15 K for segregation of CO2. This translates in a segregation temperature of 30 ± 5 K, assuming a segregation timescale of 4000 yr. He et al. (2017) obtained an onset of segregation of CO2 on the surface of non-porous Amorphous Solid Water (np-ASW) at 65 K, corresponding to a temperature in space of 43 ± 3 K, assuming a diffusion pre-exponential factor of 1012 s−1 and a similar segregation timescale as in Öberg et al. (2009). This temperature range is somewhat higher than the result of Öberg et al. (2009).

Since CO2 is present in a wide range of concentrations with respect to water in ices (Boogert et al. 2015; Yamagishi et al. 2015), in order to correctly interpret IR spectra for studying the thermal evolution of ices, it is necessary to know how the level of CO2 concentration in mixed CO2:H2O ices impacts segregation. In this work, we comprehensively study the concentration dependence as well as temperature dependence of CO2 segregation from CO2:H2O mixtures. In a previous work (He & Vidali 2018), we found that the ν1 + ν3 mode at ∼3708 cm−1 and the ν1 + 2ν2 mode at ∼3600 cm−1 provide useful tools to quantify CO2 segregation in laboratory measurements. In this work, we study the profiles of the combination modes and show that they can be used to assess the degree of order of CO2 in CO2:H2O ices. To obtain the thermal history of a CO2:H2O ice and to compare it with spectra obtained with the ISO-SWS, we use the ν3 absorption profile of 13CO2 naturally occurring in laboratory CO2:H2O ice mixtures.

The remainder of this paper is organized as follows. Section 2 describes the experimental setup, followed by Section 3, which presents results and analysis. Section 4 compares our laboratory measured spectra of 13CO2 with ISO-SWS data and discusses the astrophysical implications.

2. Experimental Setup

Experiments were carried out in an ultra-high vacuum (UHV) apparatus at Syracuse University. The UHV chamber is pumped by a combination of turbopumps and a cryopump. After bake-out, the base pressure reaches 4 × 10−10 Torr, routinely. At the center of the UHV chamber, a gold-coated copper disk was used as the sample. The sample can be cooled down to 5 K by an Advanced Research Systems DE-204S cryocooler, or heated up to room temperature using a cartridge heater located right behind the sample. The sample temperature was measured by a calibrated silicone diode to an accuracy better than 50 mK. A Lakeshore 336 temperature controller was used to read and control the temperature.

The IR spectra of ices deposited on the sample were recorded using a Nicolet 6700 Fourier Transform InfraRed (FTIR) in the Reflection Absorption InfraRed Spectroscopy (RAIRS) configuration with an incident angle of ∼78°. The FTIR collects and averages nine spectra every 20 s at a resolution of 0.5 cm−1 in the range of 600–4000 cm−1. Because of the strong signals, we took averages of fewer scans than is typically done in order to obtain a good time resolution during the warming of the ice sample. The heating ramp rate during temperature-programmed desorption (TPD) was 0.1 K s−1 (except for dedicated flash heatings), which amounts to one infrared spectrum every 2 K. The modalities of deposition of CO2:H2O mixtures onto the gold-plated copper sample are discussed in the Appendix.

3. Results and Analysis

We carried out three sets of experiments. In the first set, we study the temperature dependence of IR absorption bands of CO2:H2O mixtures of different mixing ratios as they were heated linearly from 10 to 200 K. In the second set, we fix the CO2:H2O ratio and carried out isothermal experiments at different temperatures to find out the temperature at which CO2 segregation maximizes. A higher temperature facilitates segregation, but at too high a temperature, CO2 desorption begins to compete with segregation. There should exist an optimum temperature that maximizes segregation. In the third set of experiments, we fix the temperature for the isothermal experiments at the temperature of maximum segregation we found from the second set of experiments, and check how segregation depends on CO2 concentration.

3.1. Temperature Dependence of IR Bands

In this set of experiments, 50 ML of water and various amount of CO2 were co-deposited on the sample at 10 K, to make the following CO2:H2O mixtures: 5:100, 10:100, 15:100, 20:100, 25:100, 30:100, 40:100, and 50:100. After deposition, the ice mixtures were heated linearly from 10 to 200 K at 0.1 K s−1. Figures 13 show the absorption bands ν3, 13CO2 ν3, and the combination modes of a 50:100 CO2:H2O mixture at selected temperatures during heating. Figure 4 shows the integrated band area of the ν3 peak at around 2350 cm−1 for CO2:H2O mixtures of different mixing ratios during the heating up.

Figure 1.

Figure 1. RAIRS of ν3 band of 50:100 CO2:H2O mixture deposited at 10 K and heated at 0.1 K s−1. The temperature of each curve is marked. The small oscillations superimposed on the curves are due to gas-phase CO2 in the spectrometer.

Standard image High-resolution image
Figure 2.

Figure 2. Same as Figure 1, but for the ν3 band of 13CO2 that is present in natural abundance in CO2:H2O ice mixtures.

Standard image High-resolution image
Figure 3.

Figure 3. Same as Figure 1, but for the 3570 to 3750 cm−1 range.

Standard image High-resolution image
Figure 4.

Figure 4. Band area of the ν3 absorption peak of CO2 during warming up of the CO2:H2O mixtures with different mixing ratios (see inset).

Standard image High-resolution image

The CO2 ν3 mode shows an asymmetric peak centered at about 2365 cm−1 at low temperatures. He & Vidali (2018) has shown that the ν3 peak of CO2 on water ice surface is dependent on the CO2 coverage. As the coverage increases from zero to more than two layers (L), the peak shifts from ∼2347 to ∼2376 cm−1. The shape and position of the spectra in Figure 1 at low temperatures are qualitatively similar to the submonolayer spectra shown in Figure 1 of He et al. (2017). As the ice is heated to the 70–80 K range, a peak at ∼2378 cm−1 emerges, which is a signature of "pure" crystalline form of CO2 after segregation has taken place. Between 80 and 100 K, the ν3 band area of CO2 decreases. This is due to the desorption of weakly bound CO2 on the surface of water ice (including the surface of pores). The remaining CO2 molecules that are trapped inside the water ice matrix have an absorption peak at around 2345 cm−1 that redshifts with temperature. Between 100 and 150 K, the CO2 amount decreases linearly with temperature. This slow desorption is induced by the compaction of ASW, and CO2 molecules are pushed out of the water matrix during the pore collapse of ASW. Between 150 and 155 K, water crystallizes and all of the remaining CO2 desorbs from the ice. This is referred to as the "molecular volcano desorption" (Smith et al. 1997).

A similar trend is also seen in the ν3 13CO2 band. When CO2 segregates, a peak at 2282 cm−1 emerges. This peak is characteristic of 13CO2 with natural isotopic abundance (He et al. 2017) and is more sensitive to segregation than the ν3 peak of CO2 at around 2376 cm−1. During the heating from 10 to 140 K, the peak redshifts from about 2280 cm−1 to about 2276 cm−1, and the width becomes narrower with temperature. To see more clearly how the peak position and width change with temperature, we used one broad Gaussian lineshape and one narrow Lorentzian lineshape to fit 13CO2 in disordered and ordered (crystalline) CO2, respectively. Although the disordered component has an asymmetric shape, for simplicity we still use a Gaussian function. This fitting scheme is sufficient to reliably obtain the peak position and the width. Figures 5 and 6 show the center position (μ) and the full width at half maximum (FWHM) of the Gaussian fit, respectively. Both the center position μ and FWHM decrease with temperature roughly linearly. The best-fitting parameters with a 95% confidence interval are

Equation (1)

Equation (2)

where the unit is cm−1 for μ and FWHM, and Kelvin for T. These simple functions will be useful for the analysis of the observed 13CO2 profile, to be discussed below.

Figure 5.

Figure 5. Center position (μ) of the Gaussian disordered component of the 13CO2 ${\nu }_{3}$ absorption peak at ∼2280 cm−1 for different ratios of CO2:H2O mixtures during warm-up. The CO2:H2O ratio is in the inset.

Standard image High-resolution image
Figure 6.

Figure 6. Same as in Figure 5, but for FWHM.

Standard image High-resolution image

Figure 3 shows the region of the ν1 + ν3 12CO2 mode at around 3700 cm−1, ν1 + 2ν2 12CO2 mode at 3600 cm−1, as well as a peak at ∼3650 cm−1. The 3650 cm−1 peak can be attributed to 12CO2 on the surface of water. This feature is common in CO2:H2O mixtures or CO2 on the surface of water at low temperatures. The ν1 + ν3 and ν1 + 2ν2 combination modes are broad at low temperatures. As the CO2 segregates, a sharp feature emerges for both combination modes. After weakly bound CO2 has desorbed from the surface, the 3600 cm−1 peak is barely seen, but the peak at 3700 cm−1 is still clearly visible. Its position and width are no different from the 3-coordinated dangling bond of amorphous water annealed at similar temperatures. We attribute this peak to the dangling bond of ASW, although we do not exclude the possibility that the ν1 + ν3 mode of CO2 may also have a small contribution to this peak.

3.2. Segregation of CO2 in CO2:H2O Mixtures

Prior laboratory measurements (Hodyss et al. 2008; He & Vidali 2018) have shown that the segregation and crystallization of CO2 is accompanied by changes in the bending, asymmetric stretching (for both 12CO2 and 13CO2), and combination modes. The bending mode absorption at ∼650 cm−1 is an important feature that is used to characterize CO2 ice (Pontoppidan et al. 2008). Toward heavily embedded YSOs, the bending mode is easier to observe than the combination modes or the ν3 of 13CO2 because of the brighter continuum emission at 15 μm. However, the bending mode band is close to the lower limit of our infrared detector, and the signal is weak. Following our previous work on CO2 ice (He & Vidali 2018), we use the ν1 + ν3 combination mode at around 3700 cm−1 to quantify the segregation. The analysis of segregation based on the combination mode at 3700 cm−1 should not differ from that using the bending mode at ∼650 cm−1.

We decompose the profile of the combination mode at ∼3700 cm−1 into two components, one broad Gaussian component centered at ∼3703 cm−1 attributed to disordered CO2, and one sharp Lorentzian component centered at 3708 cm−1 attributed to (poly)crystalline CO2. An example of the fitting is shown in Figure 7. We defined the "degree of crystallinity" (DOC) as the fraction of CO2 in the (poly)crystalline form (the Lorentzian component),

Equation (3)

where Acrystalline and Aamorphous are the band area of the Lorentzian component and Gaussian component, respectively. We calculate DOC of the experiments presented in Section 3.1. The results are shown in Figure 8.

Figure 7.

Figure 7. Example of fitting the CO2 ν1 + ν3 mode absorption profile using a Lorentzian component and a Gaussian component.

Standard image High-resolution image
Figure 8.

Figure 8. Degree of crystallinity (DOC) of CO2:H2O mixtures when warming up the ice at a ramp rate of 0.1 K s−1. The mixing ratios are shown in the figure.

Standard image High-resolution image

From Figure 8, it can be seen that the segregation strongly depends on both concentration and temperature. Here we performed the experiments at a fixed heating ramp rate. It is possible that different ramp rates would also yield different segregation ratios. To use CO2 segregation to trace the temperature history of the ice mantle would require a systematic study of the segregation over the entire parameter space of time, temperature, and concentration, which is a very time consuming set of measurements. Here we take a simpler approach and focus on one parameter at a time. A set of isothermal experiments were devoted to determine the temperature that maximizes the segregation. We deposited 50 ML of water and 30 ML of CO2 simultaneously onto the sample at 10 K, then flash-heated the sample at a rate of 0.5 K s−1 to a target temperature and kept it at that temperature for 2 hr while monitoring the ν1 + ν3 mode. The band area of the Lorentzian component at different target temperatures as a function of time is shown in Figure 9. It can be seen that 72 K is the most favorable temperature to form crystalline CO2 in CO2:H2O mixtures. Below 72 K the mobility of CO2 is not enough for segregation of CO2 to occur to the fullest extent, while above 72 K the desorption of CO2 starts to play a significant role. At 70 and 68 K, the segregated CO2 does not reach maximum after 2 hr. To verify that the highest degree of segregation is indeed achieved at 72 K instead of 68 K or 70 K, we use a function to fit the curves and try to find the saturation level. Öberg et al. (2009) found that the segregation during isothermal experiments cannot be fit by a single exponential function. Two exponential functions are required to fit it. They attributed the two parts of the segregation to two distinct mechanisms of segregation—surface processes and bulk processes. Here we focus on the second part of the segregation only and use the function $a(1-\exp (-{bt}))+c$ to fit the 68–72 K curves after the first 10 minutes. The fittings are extrapolated to 4 hr to show the saturation level, from which it is clear that 72 K is the favorable temperature that maximizes the segregation. Based on Figure 8, the temperature at which the DOC maximizes is similar for all concentrations. Therefore, it is fair to assume that this favorable temperature, 72 K, works for all concentrations.

Figure 9.

Figure 9. Band area of the Lorentzian component of ν1 + ν3 at 3708 cm−1 during isothermal experiment of a 30:100 CO2:H2O mixture at different temperatures. The isothermal temperature is marked in the figure. The dashed lines are the fitting using a function $a(1-\exp (-{bt}))+c$.

Standard image High-resolution image

After finding this most favorable isothermal experimental temperature, we fix the temperature and try different concentrations to obtain the lowest concentration required for the formation of "pure" crystalline CO2. We fixed the amount of water deposited at 50 ML and selected the CO2 dose to be: 2.5, 5.0, 7.5, 10, 11.5, 12.5, 15, 20, and 25 ML. After the co-deposition at 10 K, the ice mixtures were heated to 72 K at a ramp rate of 0.5 K s−1 and then kept at 72 K for 2 hr. Figure 10 shows the DOC as a function of isothermal experimental time at different target temperatures. It can be seen that below the CO2:H2O = 23:100 concentration, the DOC is almost zero. We thus conclude that 23% is the threshold concentration to obtain "pure" crystalline CO2 in CO2:H2O mixtures.

Figure 10.

Figure 10. Degree of crystallinity (DOC) of CO2:H2O mixture for different concentrations (see inset) during isothermal experiments at 72 K.

Standard image High-resolution image

4. Astrophysical Implications

4.1. CO2:H2O Ices as Temperature Tracers

CO2 is one of the main components of ISM ice mantles. In some of the observed sightlines, CO2 is in the "pure" crystalline form, as seen from the double splitting of the bending absorption profile. This splitting feature has been proposed to be a candidate tracer of the thermal history of the ice mantle. Prior laboratory studies (Ehrenfreund et al. 1999; Hodyss et al. 2008; Öberg et al. 2009) have found that the segregation of CO2 from ice mixtures is a function of temperature, and the segregation is irreversible with temperature. Experiments in this work show that the segregation of CO2 from a CO2:H2O mixture is not only temperature dependent but also strongly affected by the concentration of CO2. In fact, if the concentration of CO2 is too low, pure crystalline CO2 never forms, regardless of the thermal history. According to our measurements, the concentration threshold for CO2 segregation is 23:100. This is larger than the average CO2/H2O column density ratio observed toward massive YSOs (17 ± 3%; Gerakines et al. 1999) and comparable to the ∼25% ratio of the low-mass YSOs (Pontoppidan et al. 2008). The segregated CO2 detected in these sightlines (e.g., S140 IRS1) might thus probe CO2/H2O concentrations that are enhanced at certain locations along the sightline or in certain ice layers.

Our experiments show that the 13CO2 absorption profile at around 2280 cm−1 is a good tracer of the thermal history of the ice mantle, even if there is no sign of the very narrow feature of segregated crystalline CO2. Figures 5 and 6 show that both the peak position (in cm−1) and width decrease with increasing temperature, but they are not sensitive to CO2 concentration. This isolates the effect of temperature from the effect of concentration, and thus provides an easy tool for the determination of the thermal history. Next we try to fit the 13CO2 spectra in Boogert et al. (2000) based on our laboratory measurements. We visually examine the observed spectra and separate them into two groups: group 1 without a significant narrow blue peak at 2283 cm−1, and group 2 with it. Spectra in group 1 are fit with a single Gaussian function, while spectra in group 2 are fit with one Gaussian function for disordered CO2 and one Lorentzian function for crystalline CO2, as shown in Figures 11 and 12. The best-fit peak position, FWHM as well as the calculated temperature based on the Gaussian component using Equation (2) are also shown for each spectrum. For group 2, the magnitude of both components are also shown.

Figure 11.

Figure 11. Fitting of selected ISO-SWS spectra of 13CO2 using one Gaussian function. The fitting parameters are shown in the figure. The temperature T is calculated using Equation (1) and corresponds to the temperature in the laboratory timescale. To convert the timescale to that of the warming up stage of a interstellar clouds, the temperature should be multiplied by a factor of 0.3–0.5.

Standard image High-resolution image
Figure 12.

Figure 12. Fitting of selected ISO-SWS spectra of 13CO2 using one Gaussian function and one Lorentzian function. The fitting parameters are shown in the figure. The temperature T refers to the laboratory timescale.

Standard image High-resolution image

The temperatures quoted in Figures 11 and 12 are based on the laboratory timescale. In a typical astrophysically relevant timescale, the warming up of the ice is over a much longer time, and a lower temperature is expected to yield the same structural changes in the ice mixture. To translate the ice's temperature in the laboratory timescale Tlab to the temperature in astronomical conditions Tast, an Arrhenius-type expression can be used to describe the rate of segregation:

Equation (4)

where ν and Eseg are the prefactor and the energy barrier (in unit of degree Kelvin) for CO2 segregation, respectively. The timescale for segregation to happen can be approximated by t ∼ k−1. The temperature at which segregation happens most efficiently is

Equation (5)

In laboratory experiments, the timescale tlab is on the order of seconds. Under astronomical conditions, the freefall time between distances that have temperatures of 70–90 K for a low-mass star is 75 yr (Pontoppidan et al. 2008). Rotation will slow down this process, but 102–103 yr for the temperature range where segregation takes place is a reasonable estimate. The lifetime of hot cores around massive YSOs is 30,000 yr (Charnley et al. 1992). These hot cores have temperatures exceeding 100 K, but there is a gradient outside of that where ices have not yet evaporated and are heated during that time. We adopt the range of 102–105 yr for the segregation timescale under astronomical conditions. The conversion factor from the laboratory timescale to an astronomical timescale is

Equation (6)

The prefactor ν is not well characterized. Öberg et al. (2009) performed isothermal experiments of 1:2 CO2:H2O mixtures and reported a prefactor of $2\times {10}^{5\pm 1}$ s−1 for segregation. This prefactor likely reflects a combined effects of CO2 diffusion on ASW and the collapse of pores in ASW. Using this prefactor, the conversion factor can be calculated to be in the 0.3–0.4 range. If we assume that the segregation is dominated by the diffusion of CO2 on the pore surface of ASW and take the laboratory determined prefactors for volatile molecules such as CO, N2, CH4, which are mostly in the 108 ± 1 range (He et al. 2018), then the conversion factor is in the 0.4–0.5 range. In summary, to convert the temperature in the laboratory timescale to that of the warming up of a interstellar cloud, a factor 0.4 ± 0.1 should be considered. With a correction for timescale taken into account, the dependence of segregation on both concentration and temperature shown in Figure 8 should be useful in models of YSO envelopes.

4.2. Comparison with Astronomical Spectra

In the sightline of W33A, the 13CO2 absorption peak is centered at 2277 cm−1, which corresponds to a temperature of 105 K in the laboratory timescale. But if we plug T = 105 K into Equation (2), the calculated FWHM = 6.3 cm−1 is much smaller than that for the observed one, 9.0 cm−1. Note that this is likely not due to contamination by CO mixed in the ices, because W33A has a smaller contribution from CO (10%) compared to other YSOs, e.g., NGC 7538 IRS9 (20%; Pontoppidan et al. 2008). This inconsistency is likely due to the nonuniform temperature along the sightline. For this reason, we use the peak position in Equation (1) instead of the FWHM in Equation (2) to determine the temperature. In general, the best-fit temperature in Figure 11 should be understood as the average temperature along the sightline.

So far, we used CO2 in water ice to obtain the temperature of the ice. Now we put the temperature traced by 13CO2 in the context of the observed sightlines. Of the targets with a single "disordered" 13CO2 component (Figure 11), the peak of R CrA IRS2 (TS13) has the largest wavenumber, and thus the lowest temperature. The space temperature of (0.4 ± 0.1) ×(23 ± 12) K is well below that for CO sublimation, and indeed this sightline harbors an exceptionally large apolar CO component (Vandenbussche et al. 1999). The two other high-quality spectra of the low-mass YSO HH 100 IR and the massive YSO W33A show 13CO2 bands peaking at smaller wavenumbers, corresponding to space temperatures of (0.4 ± 0.1) × (55 ± 9) K and (0.4 ± 0.1) × (105 ± 7) K, respectively. Indeed, the CO profile of W33A is dominated by polar CO ices (Pontoppidan et al. 2003), indicating that the volatile apolar CO ices have sublimated, and a significant abundance of warm CO gas is detected in this sightline (Mitchell et al. 1990). The relatively high 13CO2 temperature for HH 100 seems puzzling considering the large abundance of apolar CO (Pontoppidan et al. 2003). This likely reflects a large temperature gradient in the HH 100 YSO envelope, with colder apolar CO dominating the ices at larger radii. One should also consider the possibility that the CO2:H2O ices are formed earlier in the cloud history in less shielded conditions on warmer grains than the volatile CO. Then, the 13CO2 profile reflects the formation temperature. This can be tested by observations toward background stars tracing quiescent clouds. Unfortunately, the ISO/SWS spectrum of Elias 16, a background star of the Taurus Molecular Cloud, is of low quality, but will be vastly improved when observed with JWST in the near future. For a discussion regarding the targets with segregated crystalline 13CO2 (Figure 12), we refer the reader to Boogert et al. (2000) and van der Tak et al. (2000), who show that the degree of segregation correlates with the dust temperature as the YSO envelope becomes less massive and hotter over timescales of a few times 10000 yr.

In the second group (Figure 12), we used one broad Gaussian and one narrow Lorentzian to fit the spectra. The temperatures are derived from the broad component using Equation (1). Boogert et al. (2000) used a narrow Gaussian instead of a Lorentzian to fit the 2283 cm−1 feature. They found that the narrow feature is present only toward high-mass protostars, even though not all high-mass protostars have the narrow feature (see Boogert et al. 2000 for a detailed discussion). In these four sightlines, the temperatures are all relatively high, in agreement with the previous proposition that segregation indicates thermal processing. Because of more free parameters being used in the fitting, the uncertainty in temperature is much larger than in group 1. Furthermore, the signal-to-noise ratios of the spectra from S140:IRS1 and W3:IRS5 are not good enough for an accurate determination of the peak position of the disordered component. To better constrain the temperature would require a better signal-to-noise ratio of the spectra. The JWST is expected to cover these regions at orders of magnitude better sensitivity and somewhat larger spectral resolution (R = 3000 versus 2000), and more accurate temperature determination and sorting of features between different classes of objects will be possible. The bending mode profile at 15 μm can also be used as a supplementary tool to further constrain the temperature.

Previously, Boogert et al. (2000) compared the observed spectra with the laboratory measurement of Ehrenfreund et al. (1999). They found a similar redshift and a narrowing of the 13CO2 peak as the temperature of the CO2:H2O mixture was increased. They also found that during heating a 1:0.92:1 H2O:CH3OH:CO2 mixture, the peak position and width changed, but with a different temperature dependence than that of a CO2:H2O mixture. Therefore, Boogert et al. (2000) concluded that while 13CO2 traces segregation, it is not suitable for temperature determinations, because the temperature effect could not be separated from composition effects. In their experiment, the concentration of CH3OH is much higher than what is typically observed. CH3OH is formed mostly on dust grains by the consecutive hydrogenation of CO after the heavy CO freeze-out. This formation mechanism is corroborated by both laboratory measurements (Hama & Watanabe 2013) and observations (Boogert et al. 2011), which show that CH3OH is mostly found in high-extinction regions. Conversely, CO2 is mostly found in a water-rich environment, consistent with the scenario that CO2 is formed together with water. Although most recent laboratory experiments found that CH3OH can also be formed before the heavy freeze-out of CO by the reaction between CH3 and OH (Qasim et al. 2018), the question still remains of how much CH3OH can be formed this way. Based on the current state of knowledge, it is safe to assume that CO2 primarily interacts with water instead of with CH3OH in the ice, and therefore it is justifiable to ignore the effect of CH3OH on the 13CO2 absorption profile. This assumption seems less applicable to CO mixed in the ices, as discussed below.

4.3. CO:CO2 Ices

Previous analysis of the CO2 bending mode absorption profile at 15 μm and the blue component of CO absorption profile at 4.7 μm reveals that 10–30% of the CO2 molecules are mixed with CO (Pontoppidan et al. 2003, 2008). This group proposed that "pure" crystalline CO2 can be formed either by thermally induced CO2 segregation from a CO2:H2O mixture, or by CO desorption from a CO:CO2 mixture. The mechanism of segregation is closely related to the mechanism of CO2 formation. So far there are mainly two categories of mechanisms proposed to explain the formation of CO2 molecules on grains. The first category involves pure thermal reactions among CO, O, H, and OH. Although several questions still remains, such as the relative contribution of CO+O (Roser et al. 2001) and CO+OH (Zins et al. 2011), and whether the reaction involves HOCO (Ioppolo et al. 2011), it is clear that water is formed on grains at the same time as CO2 via reactions with hydrogen: O + H→OH, OH+H→H2O. The experimental results and the temperature tracing method proposed in this study mostly apply to the CO2 that is formed together with water by thermal processes. The second category of CO2 formation mechanism involves energetic processing of the pure CO in the top layers of the ice mantle. Laboratory experiments have shown that CO2 can be formed by the bombardment of analogs of cosmic rays with CO ice (Gerakines & Moore 2001; Loeffler et al. 2005; Jamieson et al. 2006). However, other molecules that should have also been produced in the energetic processing of CO, such as C3O2 (Jamieson et al. 2006), were not observed in the same sightline as CO2 (Pontoppidan et al. 2008). The reason why a fraction of CO2 is in a CO-rich environment is still puzzling. If cosmic ray bombardment is important for CO2 formation, the compaction of the ASW and the segregation of CO2 from CO2:H2O mixtures may also be affected by cosmic rays. It would be less relevant to characterize the ice mantle by temperature than to use the fluence of cosmic rays. In any case, the use of CO2 segregation or 13CO2 to trace the temperature history of the ice mantle is only valid under the assumption that cosmic ray irradiation is not the dominant mechanism for CO2formation.

5. Summary

We made measurements in the laboratory of infrared absorption features of CO2:H2O ices with different mixing ratios, at different temperatures, and subjected to thermal cycles in order to elucidate the thermal history of ices observed toward YSOs with ISO-SWS (Boogert et al. 2000). We found that at a CO2:H2O concentration below 23%, there is no formation of pure crystalline CO2. Thus, looking at pure crystalline CO2 alone is not a good proxy for establishing the thermal history of ices. We found that the ν3 feature of 13CO2 does not suffer from this threshold concentration dependence, and its peak position and linewidth, together with the pure crystalline feature of 12CO2, can be used to infer the temperature history of ices near YSOs. Data such as those presented here will help to characterize the segregation status and thermal history of ices in upcoming JWST observations with an extended range in the IR spectrum and improved sensitivity and spectral resolution.

We thank Francis Toriello for technical assistance. This research was supported by NSF Astronomy & Astrophysics Research grant No. 1615897.

Appendix: Deposition of Gases

CO2 gas and water vapor were deposited onto the sample through background deposition using two UHV precision leak valves activated by two stepper motors controlled by a LabVIEW program. For the deposition of a single molecular species, the program first measures the base pressure of the chamber, and then calculates the target partial pressure based on a user set deposition rate. The pressure readings from the hot cathode ion gauge are corrected for the gas species in the LabVIEW program. A PID control loop is used to maintain the pressure at the target value. In the deposition of CO2, it takes about 20 s for the pressure to stabilize at the set value. The ice thickness during deposition is calculated by the program in real time. When the thickness reaches the setpoint, the leak valve is closed quickly. Even after the valve is closed, the residual gas in the chamber continues being deposited on the sample, until it is pumped out. We correct for the additional amount deposited from the residual gas by closing the valve slightly before the target thickness is reached. The exact offset thickness is calculated from the deposition pressure and the pumping speed. After this correction, in the deposition of CO2, the relative uncertainty of thickness measured by the integration of pressure over time is usually less than 0.1%. For water deposition, the uncertainty is larger (1%) because of the difficulty in maintaining a stable water inlet pressure in the gas manifold.

In CO2:H2O co-depositions, because the ion gauge can only measure the total pressure but not the partial pressure of each gas, we start with the deposition of one gas. CO2 is deposited first because it is easier to control its deposition. Within 20 s of introducing CO2, the deposition rate is already stable. We tested the stability of pressure by fixing the valve position after 20 s, and found that the pressure does not change over time. The same is not true for water because of the instability of inlet pressure. After determining the stable valve position for CO2, we stop the PID loop for the CO2 valve and fix the valve position. We then use a PID loop for the water leak valve to obtain a stable pressure during co-deposition. When the set time is reached, both leak valves are closed immediately. For the co-depositions in this study, the deposition is over 25 minutes, and therefore the uncertainty in CO2 amount is about 20 s over 25 minutes, which is about 1%.

The impingement rate (IPR), which is the number of molecules deposited per unit surface area per unit time, is calculated as follows:

Equation (7)

where P is the chamber pressure after correction for the ion gauge gas specific ionization cross-section, m is the mass of gas molecule, T is the gas temperature (assumed to be room temperature), and kB is the Boltzmann constant. It is assumed that the sticking of both CO2 and H2O are unity at 10 K (He et al. 2016), and the pressure in the vacuum chamber is uniform. This is a fair assumption, because the leak valves opening does not face the sample or cold head directly. The IPR can be converted to units of monolayer per second (ML s−1) by assuming 1 ML = 1015 cm−1. The absolute uncertainty of deposition is mostly due to the uncertainty in pressure measurement, and can be as high as 30%, as this is the accuracy of a typical hot cathode ion gauge. In the experiments, the uncertainty in mixing ratio of the CO2:H2O mixtures is governed by the relative uncertainty, while that of the total thickness of the mixture is governed by the absolute uncertainty.

Please wait… references are loading.
10.3847/1538-4357/aae9dc