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.

The following article is Open access

Astronomical Detection of the Interstellar Anion C10H toward TMC-1 from the GOTHAM Large Program on the Green Bank Telescope

, , , , , , , , , , , , , , , and

Published 2023 February 21 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Anthony Remijan et al 2023 ApJL 944 L45 DOI 10.3847/2041-8213/acb648

Download Article PDF
DownloadArticle ePub

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

2041-8205/944/2/L45

Abstract

Using data from the Green Bank Telescope (GBT) Observations of TMC-1: Hunting for Aromatic Molecules (GOTHAM) survey, we report the first astronomical detection of the C10H anion. The astronomical observations also provided the necessary data to refine the spectroscopic parameters of C10H. From the velocity stacked data and the matched filter response, C10H is detected at >9σ confidence level at a column density of ${4.04}_{-2.23}^{+10.67}\times {10}^{11}$ cm−2. A dedicated search for the C10H radical was also conducted toward TMC-1. In this case, the stacked molecular emission of C10H was detected at a ∼3.2σ confidence interval at a column density of ${2.02}_{-0.82}^{+2.68}\times {10}^{11}$ cm−2. However, as the determined confidence level is currently <5σ, we consider the identification of C10H as tentative. The full GOTHAM data set was also used to better characterize the physical parameters including column density, excitation temperature, line width, and source size for the C4H, C6H, and C8H radicals and their respective anions, and the measured column densities were compared to the predictions from a gas/grain chemical formation model and from a machine learning analysis. Given the measured values, the C10H/C10H column density ratio is ∼${2.0}_{-1.6}^{+5.9}$—the highest value measured between an anion and neutral species to date. Such a high ratio is at odds with current theories for interstellar anion chemistry. For the radical species, both models can reproduce the measured abundances found from the survey; however, the machine learning analysis matches the detected anion abundances much better than the gas/grain chemical model, suggesting that the current understanding of the formation chemistry of molecular anions is still highly uncertain.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

It has now been more than 15 yr since the first reported molecular anion was detected in astronomical environments with the discovery of C6H (McCarthy et al. 2006). This detection was the culmination of the early predictions that anions should be present under interstellar conditions (Herbst 1981). As summarized by Millar et al. (2017) molecular anions are critically important to the physics and evolution of astronomical objects—including building up structures in the early universe, dominating the visible opacity for stars like the Sun, being the possible carriers to the diffuse interstellar bands, and determining the physical and chemical environments of astrophysical regions including measuring the impact of interstellar radiation fields and other molecular cloud properties given their reactivity. As such, the detection of interstellar molecular anions are far beyond just an astrochemical curiosity.

The initial detection then sparked the search for new molecular anions and the identification of C4H (Cernicharo et al. 2007; Agúndez et al. 2008), C8H (Brünken et al. 2007; Gupta et al. 2007; Kentarou et al. 2007; Remijan et al. 2007), C3N (Thaddeus et al. 2008), C5N (Cernicharo et al. 2008, 2020), and ${\mathrm{CN}}^{-}$ (Agundez et al. 2010), observed primarily toward the dark cloud source TMC-1 and the evolved star IRC+10216. Since these initial detections, searches for C6H have shown that this anion is abundant in a variety of sources, from quiescent dark clouds to active star-forming regions (Cordiner et al. 2013). These searches revealed a wide discrepancy in the anion-to-neutral column density ratios. The anion-to-neutral column density ratio for carbon chains varies markedly with the chain length and the astrophysical environment. In TMC-1, the C6H/C6H ratio is ∼2.5%, whereas the C4H/C4H ratio is only ∼0.0012% (Cordiner et al. 2013), increasing twentyfold to ∼0.024% toward IRC+10216 (Cernicharo et al. 2007). The enhanced anion fraction of longer chains is mirrored by the N-terminated species: C5N/C5N ≈12.5% and C3N/C3N ≈0.7% in TMC-1. (Cernicharo et al. 2020). These observations highlight that anion chemistry and the limit to which carbon-chain anions can grow in astronomical environments are still not well understood.

Recently, Siebert et al. (2022) reported the identification of the largest CH3-terminated carbon-chain molecule CH3C7N toward TMC-1 and in that work, determined the column densities of several large carbon-chain families and made predictions for the column densities for those species yet to be detected. These predictions were made using the three-phase gas/grain chemical network model nautilus (v1.1; Ruaud et al. 2016). In addition, a linear extrapolation as a function of the chain length was made from the measured column densities of the smaller carbon chains. While the predicted and measured column density ratios can differ up to an order of magnitude, it served as motivation to conduct an astronomical search for the elusive, larger carbon-chain species that may already be contained within our existing TMC-1 data set.

As such, to follow on from the recent detections and surveys of molecular anions, an extensive search for the decapentaynyl radical (C10H) and the decapentaynyl anion (C10H) was conducted toward the dark cloud TMC-1—the site of the first detection of interstellar anions—with the Green Bank Telescope (GBT) Observations of TMC-1: Hunting for Aromatic Molecules (GOTHAM) survey. The family of H-terminated carbon-chain radicals—C2n H (where n >1) and their associated anions have been studied in various astronomical environments, and these species are believed to share common chemical formation pathways (Bettens & Herbst 1997; Walsh et al. 2009). To best constrain the column density ratios between the radical and anion species, we performed a complete reanalysis of the C4H, C6H, C8H radicals and their associated anions taking advantage of the significantly improved signal-to-noise ratio (S/N) and spectral resolution of our data compared to their original detections. These observations provide the most rigorous measurement of the column densities of these species with which to compare to chemical formation and machine learning models. They have also set a new limit to the largest carbon-chain species detectable in astronomical environments as the first detection of C10H and a tentative detection of C10H are reported toward TMC-1. The observing parameters describing the search for these carbon-chain species are presented in Section 2. The new spectroscopic analyses for C4H, C6H, and C10H are given in Section 3, which includes a discussion of how the astronomical detection of C10H enabled the more accurate determination of its molecular constants. The observational analyses on how the physical parameters of TMC-1 are determined from these data are presented in Section 4. The results of the observational searches for C10H and C10H and the previously detected carbon-chain molecules are presented in Section 5. A comparison of the predicted-to-measured column density ratios for this family of carbon-chain molecules from both state-of-the-art gas/grain chemical models and from machine learning analyses are given in Section 6. Finally, our conclusions are highlighted in Section 7.

2. Observations

Observations for this study were obtained as part of the GOTHAM survey. GOTHAM is a large project on the 100 m Robert C. Byrd GBT currently managed by the Green Bank Observatory (GBO). The GOTHAM program is a dedicated spectral line observing program of TMC-1 covering almost 30 GHz of bandwidth at high sensitivity and spectral resolution. All data were taken with a uniform frequency resolution of 1.4 kHz (0.05–0.01 km s−1 in velocity) and an rms noise of ∼2–20 mK across most of the observed frequency range with the rms gradually increasing toward higher frequency because of the shorter integration times. This work uses the fourth data reduction (DR4) of GOTHAM targeting the cyanopolyyne peak (CP) of TMC-1, centered at αJ2000 = 04h41m42.5s., δJ2000 = +25°41'26farcs8. Briefly, the spectra in these data cover the entirety of the X-, K-, and Ka-receiver bands with nearly continuous coverage from 7.9 to 11.6 GHz, 12.7 to 15.6 GHz, and 18.0 to 36.4 GHz (24.9 GHz of total bandwidth). Data reduction involved removal of radio frequency interference and artifacts, baseline continuum fitting, and flux calibration using complementary Very Large Array observations of the source J0530+1331. Uncertainty from this flux calibration is estimated at ∼20%, and is factored into our statistical analysis described below (McGuire et al. 2020). A full description of the fourth data reduction can be found in Sita et al. (2022), and the observing strategy and reduction pipeline is fully described in McGuire et al. (2020).

3. Spectroscopic Analyses

Upon comparison of the astronomical data and the fitted laboratory spectra found in the publicly available Cologne Database for Molecular Spectroscopy (CDMS; Endres et al. 2016), it was determined that a reanalysis of the measured and predicted frequencies was needed for C4H and C6H. The spectroscopic data available for the other radicals and associated anions very closely matched the observational spectrum and no reanalysis was required. The subsections below describe the process for how the catalogs for C4H and C6H were revised and how the catalog for C10H was generated using the astronomically measured spectroscopic data.

3.1. C4H

Upon inspection of the signal from C4H in the GOTHAM data, it was immediately obvious that the available line frequencies in the CDMS (Endres et al. 2016) were insufficiently accurate to reproduce the observations. In particular, the hyperfine splitting, which is well resolved in the GOTHAM data, was poorly matched by the catalog data. In some cases, the predicted line frequencies varied by more than a FWHM line width compared to the astronomically detected features, thus necessitating new high-resolution measurements.

C4H was prepared in the laboratory using the same methods as previous experiments on its 13C isotopologues (Chen et al. 1995) and vibrationally excited states (Cooksy et al. 2015). A mixture of 0.1% acetylene was seeded in neon at a pressure of 2.5 kTorr and supersonically expanded along the axis of a cavity Fourier transform microwave (FTMW) spectrometer (Grabow et al. 2005) in 400 μs gas pulses at a rate of 6 Hz. During each gas pulse, a 1.0 kV discharge was struck between two copper electrodes placed immediately after the valve aperture, creating reactive products which combined to make, among other species, C4H. Three perpendicular pairs of Helmholtz coils are positioned around the spectrometer and tuned to null Earth's magnetic field to less than 50 mG throughout the cavity volume.

A total of 22 hyperfine-resolved transitions between 9 and 38 GHz were measured and assigned. These were combined with previous submillimeter measurements of a further 10 spin-rotation transitions between 143 and 200 GHz reported by Gottlieb et al. (1983), for which hyperfine structure was not resolved. The dipole moment, which has been a matter of some debate over the years due to the complicated electronic structure of C4H, was taken to be 2.1 Debye (D), as recently determined using high-level quantum chemical calculations (Oyama et al. 2020). Fitting was performed using the SPFIT/SPCAT suite of programs (Pickett et al. 1998). The full measured line list, including which lines are used from which data sources, along with the corresponding input and output files from SPFIT/SPCAT, are provided as Supplemental Information.

3.2. C6H

Although not quite as striking as for C4H, it was evident from our high-resolution GOTHAM observations that existing catalogs for C6H were insufficiently accurate to reproduce the observational data. To address this, we have refit all of the high-resolution experimental lines from the work of Gottlieb et al. (2010), which had not been previously included. The updated catalog was sufficiently accurate for our purposes. The full measured line list, along with the corresponding input and output files from SPFIT/SPCAT, are provided as Supplemental Information.

3.3. C10H

To our knowledge, no laboratory spectra exist for C10H, and our own efforts to produce detectable quantities in our instruments have not yet been successful. Unlike C4H and C6H, however, C10H is a closed-shell linear molecule, and as such, it is straightforward to predict its rotational spectrum, which presents as a series of lines spaced by ∼2B. Given a reliable prediction of the rotational constant, B, prior work has shown that it is possible to identify such species in interstellar spectra preceding their laboratory confirmation. Examples include C3H+ (Pety et al. 2012) in the Horsehead photodissociation region and C5H+ (Cernicharo et al. 2022), HC5NH+ (Marcelino et al. 2020), and HC7NH+ (Cabezas et al. 2022) in TMC-1, among others.

We began our search for C10H by using a value of B = 299.882 MHz and D = 1 Hz, obtained by extrapolating and scaling from the shorter members of the family of anions. The B value specifically is in excellent agreement with that obtained from a quantum chemical calculation carried out at the M06-2X/6–31+G(d) level of theory and basis set using the Psi4 suite of programs (Smith et al. 2020) of B = 303.761 MHz. This level of theory and basis set has been previously shown to reliably produce rotational constants in excellent agreement with experiments (Lee & McCarthy 2020).

Using our estimated values of B and D, we produced a catalog of lines and performed a simulation, using a set of fiducial values for vlsr, the column density ratios between velocity components, the excitation temperature (Tex), and the line width (ΔV). These values are based on our prior detection and treatment of benzonitrile within the source, and have been shown to be excellent starting points for the analysis of other molecules in the GOTHAM data (see Supplementary Information of McGuire et al. 2021). No individual transitions were seen above the noise level of the observations. A spectral stack of the data using this first-pass estimated catalog, however, revealed a strong (>5σ) signal in the stacked spectra shifted by just over 10 km s−1 from the expected central velocity (Figure 1).

Figure 1.

Figure 1. Velocity stacked spectra of C10H using our first-pass estimated catalog generated with B = 299.882 MHz. The stacked spectra from the GOTHAM DR4 data are displayed in black, overlaid with the expected line profile in red from our first-pass catalog and using our fiducial molecular parameters. The intensity of the simulation is arbitrarily scaled (no fit has been performed). The S/N is on a per-channel basis.

Standard image High-resolution image

We then performed a least-squares fit to determine what value of the rotational constant would be required to reproduce the observed signal at the expected central velocity. We derive a value of B = 299.87133 MHz, differing from the scaled prediction by 11 kHz or 0.004%. This derived value was then used to generate a final spectral line catalog for C10H that was used for the remainder of the analysis described in this paper; a summary of the values of the B rotational constant determined by our methods are summarized in Table 1. As described in detail in McGuire et al. (2021), a fractional accuracy of better than 10−5 (and ideally better than 10−6) in the rotational constants of a molecule is required to recover any significant signal from our spectral stacking techniques. Thus, we can infer that our derived value of B is likely accurate to at least 3 kHz.

Table 1. Rotational Constant of C10H from Quantum Chemistry, Scaled from Experimental Values of smaller Anions, and Determined in this Work from our Astronomical Observations

ParameterM06-2X/6–31+G(d)ScaledThis Work
B (MHz)301.353299.882299.87133
D (Hz)[1.0] a [1.0] a

Note.

a D was fixed to a value near that of C8H (D = 4.3 Hz; Gupta et al. 2007), but no attempt was made to refine it further, given the limitations of the analysis.

Download table as:  ASCIITypeset image

4. Observational Analysis

In order to derive physical parameters (column density [NT ], excitation temperature [Tex], line width [ΔV], and source size ['']) for the target molecules in our observations, we used the same Markov Chain Monte Carlo (MCMC) model employed in prior GOTHAM analyses (see, e.g., Lee et al. 2021a; Siebert et al. 2022; Sita et al. 2022) and discussed in detail in Loomis et al. (2021). In short, the MCMC model calculates the probability distributions and covariances for these parameters that are used to describe the emission of molecules observed in our data. The resulting corner plots for the molecules analyzed here are shown in Figures A2, A3, B1, C1, D1, E1, F1, and G1.

We adopt the 50th percentile value of the posterior probability distributions as the representative value of each parameter for the molecule and use the 16th and 84th percentile values for the uncertainties. For probabilities that show a Gaussian distribution, these correspond to the 1σ uncertainty level. Many of our resulting probability distributions are indeed either Gaussian or nearly Gaussian, and thus these values are usually quite representative of the 1σ uncertainties. One of the advantages of the MCMC technique over a traditional least-squares fit approach is that far more of parameter space is explored. Correspondingly, a much larger exploration of the uncertainty space is performed as well, including highlighting parameters that may be highly covariant with one another. This manifests as nonseparable distributions in the corner plots.

To explore this parameter space with our MCMC approach, a model of the molecular emission is generated for each set of parameters using the molsim software package (Lee et al. 2021b) and following the conventions of Turner (1991) for a single excitation temperature and accounting for the effect of optical depth. Prior observations from GOTHAM (Xue et al. 2020) and others (Dobashi et al. 2018, 2019) have found that most emission seen at centimeter wavelengths in TMC-1 can be separated into contributions from four distinct velocity components within the larger structure, at approximately 5.4, 5.6, 5.8, and 6.0 km s−1 (Loomis et al. 2021). In some cases, especially for less abundant species where there is not a clear detection in one of the velocity components, we find that a three-component model has performed better (McGuire et al. 2020).

To determine the statistical evidence that our model of the emission of these molecules is consistent with the data, we followed the procedures described in detail in Loomis et al. (2021) and performed a spectral stack and matched filtering analysis for C10H and C10H. Briefly, a weighted average of the observational spectra in the velocity space and centered on each spectral line of a target molecule was performed. The weights were determined by the relative intensity of the expected emission (based on the MCMC-derived parameters) and the local rms noise of the observations. Considering the weak expected intensities for both C10H and the C10H isomers, any observational windows containing emission at >5σ were ignored in the analysis of those molecules.

Simulated spectra of the molecular emission using the same MCMC-derived parameters were then also generated and stacked using identical weights. This simulation was then used as a matched filter, which is passed through the observational signal. The resulting impulse response function represents the statistical evidence that our model of the emission from the molecule—and thus our derived parameters for the molecule—is consistent with the observations. In addition to the details of the methodology provided in Loomis et al. (2021), the appendices of McGuire et al. (2021) include an extensive analysis of the robustness of the methodology, including the improbability of spurious signals and the minimal impact of red noise on the procedure.

We also performed MCMC fits for the more abundant C4H, C4H, C6H, C6H, C8H, and C8H molecules. Given the low line density in our spectra, it is extremely improbable that interfering signals from other species would be present. Still, the spectral regions containing these transitions were manually inspected to ensure there were no interloping signals or other concerns. All strongly detected lines in our data are shown in the Appendix for C4H (Figures B2B4), C4H (Figure C2), C6H (Figures D2 and D3), C6H (Figure E2), C8H (Figure F2), and C8H (Figure G2).

5. Observational Results

Figure 2 shows the stacked data, the stacked MCMC model, as well as the matched filter response and the first detection of interstellar C10H toward TMC-1. The stacked emission in the right panel of Figure 2 exhibits evidence at >9σ for the presence of this molecule. Figure A1 shows the individual spectral lines present in the survey. While no individual lines of C10H are present above the current noise level of the survey at >3σ, there are spectral features seen in the corresponding passbands that show some emission above the noise (see, e.g., transitions at 10195.41, 13194.04, 14993.2, and 19191.26 MHz). Table 2 lists the measured physical values determined from the C10H fits. In this case, three independent velocity components were detected, and a total C10H column density of ${4.04}_{-2.23}^{+10.67}\times {10}^{11}$ cm−2 was determined.

Figure 2.

Figure 2. Velocity stacked and matched filter spectra of C10H. The intensity scales are the S/Ns of the response functions when centered at a given velocity. The "zero" velocity corresponds to the channel with the highest intensity to account for blended spectroscopic transitions and variations in velocity component source sizes. (Left) The stacked spectra from the GOTHAM DR4 data are displayed in black, overlaid with the expected line profile in red from our MCMC fit to the data. The S/N is on a per-channel basis. (Right) Matched filter response obtained from cross correlating the simulated and observed velocity stacks in the left panel; the value annotated corresponds to the peak impulse response of the matched filter.

Standard image High-resolution image

Table 2. Summary Statistics of the Marginalized C10H Posterior

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1011 cm−2)(K)(km s−1)
${5.624}_{-0.014}^{+0.012}$ ${16}_{-7}^{+9}$ ${3.68}_{-2.23}^{+10.67}$   
${5.759}_{-0.020}^{+0.020}$ ${27}_{-9}^{+10}$ ${0.27}_{-0.12}^{+0.27}$   
${4.00}_{-0.21}^{+0.21}$ ${0.360}_{-0.025}^{+0.027}$
${6.040}_{-0.019}^{+0.019}$ ${659}_{-274}^{+234}$ ${0.09}_{-0.02}^{+0.02}$   
NT (Total): ${4.04}_{-2.23}^{+10.67}\times {10}^{11}$ cm−2

Download table as:  ASCIITypeset image

Figure 3 show the stacked observational data, the stacked MCMC model, as well as the matched filter response and a tentative detection of the C10H radical toward TMC-1. The stacked molecular emission of C10H in the right panel of Figure 3 exhibits evidence at ∼3.2σ for the presence of this molecule; no individual spectral lines were present. Table 3 lists the measured physical values determined from the C10H fits. In this case, four independent velocity components were fit and a total C10H column density of ${2.02}_{-0.82}^{+2.68}\times {10}^{11}$ cm−2 was reported.

Figure 3.

Figure 3. Similar to Figure 2. Velocity stacked and matched filter spectra of C10H.

Standard image High-resolution image

Table 3. Summary Statistics of the Marginalized C10H Posterior

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1011cm−2)(K)(km s−1)
${5.590}_{-0.009}^{+0.010}$ ${24}_{-4}^{+4}$ ${0.53}_{-0.16}^{+0.26}$   
${5.726}_{-0.010}^{+0.015}$ ${29}_{-4}^{+4}$ ${0.29}_{-0.09}^{+0.13}$ ${5.45}_{-0.29}^{+0.31}$ ${0.102}_{-0.011}^{+0.019}$
${5.859}_{-0.010}^{+0.010}$ ${13}_{-6}^{+12}$ ${1.15}_{-0.80}^{+2.67}$   
${6.036}_{-0.013}^{+0.015}$ ${633}_{-256}^{+246}$ ${0.05}_{-0.01}^{+0.01}$   
NT (Total): ${2.02}_{-0.82}^{+2.68}\times {10}^{11}$ cm−2

Download table as:  ASCIITypeset image

6. Discussion

6.1. Chemical Modeling Predictions

Building on the modeling efforts of carbon-chain chemistry in Siebert et al. (2022), we utilized the nautilus v1.1 code (Ruaud et al. 2016) which has been used previously to successfully study the formation of carbon-chain molecules detected with GOTHAM data (Xue et al. 2020; McGuire et al. 2021; Shingledecker et al. 2021). The model's physical conditions are identical to those studies (Tgas = Tgrain = 10 K, ${n}_{{{\rm{H}}}_{2}}$ = 2 × 104 cm−3, AV = 10, and ζCR = 1.3 × 10−17 s−1; Hincelin et al. 2011) as are the elemental abundances (Loomis et al. 2021). Based originally off of the Kinetic Database for Astrochemistry network, our network already contained some formation routes to the Cn H family from n = 2 to n = 10 and the Cn H family from n = 4 to n = 10.

The simulated abundances are compared with those observed in TMC-1 and the machine learning predictions discussed in Section 6.2 assuming a TMC-1 hydrogen column density of ${N}_{{{\rm{H}}}_{2}}$ = 1022 cm−2. For utility of comparison, we adopt the same source age as discussed in Siebert et al. (2022). However, it should be noted that the simulated time of peak abundance can vary between species, with heavier species and longer carbon chains typically requiring a longer time to form. This time dependence is shown in greater detail in Appendix I.

As Figure 4 shows for the Cn H family, the chemical model agrees within within a factor of 5 and the log-linear trend is generally reproduced. One of the primary production pathways of this family comes from atomic reactions of the form

Equation (1)

The rate coefficients of these reactions were estimated by Loison et al. (2014) through extrapolation of the calculations by Chastaing et al. (2001) and Chastaing et al. (2000) on C + alkenes, alkynes, dienes, and diynes. The other major formation routes involve atomic reactions with related anions:

Equation (2)

Equation (3)

For C10H, these reactions were studied theoretically in Harada & Herbst (2008) and experimentally in Eichelberger et al. (2007) and Barckholtz et al. (2001). The majority of the family is primarily destroyed through electron attachment, producing their carbon-chain length analogs among the Cn H family. These rates were also estimated by Harada & Herbst (2008).

Figure 4.

Figure 4. The abundance and column density of the Cn H (left) and Cn H (right) families as observed by the GBT (black, solid, circles), simulated by nautilus (blue, dashed, stars) at a model time of t = 2.5 × 105 yr, and predicted by machine learning (orange, dotted, ×s). The observed column density of C2H is taken from Pratap et al. (1997) with the Five College Radio Astronomical Observatory 14 m antenna.

Standard image High-resolution image

For the carbon-chain anions, Cn H, the abundances agree within an order of magnitude for C6H and longer. While the model predictions show the monotonically decreasing abundances found for the radical species, akin to the log-linear trend seen for the majority of long carbon chains, it is not able to reproduce the enhanced abundances found in C6H and C10H nor the relative positive correlation between Cn H abundance and carbon-chain length. Given the measured values, the C10H/C10H column density ratio is $\sim {2.0}_{-1.6}^{+5.9}$—the highest value measured between an anion and neutral species to date. Such a high ratio is impossible for our model to reproduce, even if every electron collision with C10H results in C10H formation. As such, the predictions from the gas/grain chemical model illustrate that there is much still uncertain about the chemistry needed to form molecular anions in astronomical environments. The chemical network for these species originates from estimations done by Harada & Herbst (2008). The primary production route of the anions (80%–90%) is through radiative electron attachment

Equation (4)

Another minor production route is also present involving reactions between longer carbon-chain anions and atomic oxygen,

Equation (5)

There are two product channels suggested for the destruction reactions between carbon-chain anions and atomic hydrogen (Harada & Herbst 2008). We considered the pathway involving associative electron detachment,

Equation (6)

for Cn H (n = 4, 6, 8, and 10). In contrast, the other pathway involves fragmentation and has only been considered for C10H,

Equation (7)

Harada & Herbst (2008) originally estimated all rates for route 6 to be 1.0 × 10−9 cm3 s−1. This was also the rate estimated for route 7. In order to keep the total C10H + H rate consistent with the other analogous anion reactions in route 6, we modified the rates of both C10H+H reactions from routes 6 and 7 from 1.0 × 10−9 cm3 s−1 to 5.0 × 10−10 cm3 s−1. This resulted in a factor of ∼2 increase in the abundance of C10H relative to simulations performed with the original rates.

There are limited experimental constraints on these pathways and the corresponding rate coefficients. In particular, major route 4 is very challenging to measure experimentally because of two reasons. First, anions are difficult to produce in a stable quantity. Second, it is difficult to observe the photoemission process due to competing collisional stabilization. In addition, anion–neutral reactions, for example, might also contribute to the Cn H's formation, but these need to be investigated further. Furthermore, while the current model mainly focuses on the gas chemistry of the Cn H species, electron attachment processes might occur on grains, allowing the superexcited anion intermediate to be stabilized efficiently rather than being dependent on radiative processes in the gas phase. Our understanding of molecular anions and carbon chains can be improved by constraining their grain chemistry.

6.2. Machine Learning Predictions

The column densities of both the observed and unobserved parent and anion pairs were predicted using the trained supervised learning regressors presented in Lee et al. (2021c). For brevity, we only explain the relevant methodological details here.

Starting from the Simplified Molecular-Input Line-Entry System (SMILES; Weininger 1988; O'Boyle 2012) representations of molecules, we rely on a pretrained MOL2VEC embedding model (Jaeger et al. 2018) to transform each species into corresponding molecular vectors. These vectors form a high-dimensional representation that captures chemical (e.g., charge state, bonding patterns) and structural (e.g., aromaticity) properties into a compact form usable by even simple regressors for property prediction; in this case, and that of Lee et al. (2021c), to predict column densities of unseen molecules. In contrast to conventional chemical modeling, our machine learning approach only requires molecular structure as input, and prior knowledge of the chemical and physical properties of the source is captured implicitly based on what molecules have been observed. Thus, when conditioned on an astrophysical environment, the trained model predictions can be interpreted as a baseline for column densities based solely on manifold distances from new molecules to those already observed.

The column density predictions from the machine learning approach are given in Table 4, combining two types of models as described in Lee et al. (2021c) using regression modules from scikit-learn (Pedregosa et al. 2011). The magnitude of the column densities were predicted using the gradient boosting regressor (GBR), given its high accuracy in many regression tasks. We also use a Gaussian process regressor (GPR) to estimate the uncertainty in the column density predictions. Aside from C4H, our machine learning predictions match the mean of the observed values well within an order of magnitude, and are captured within our observational and model uncertainties.

Table 4. Column Densities used to Train the Machine Learning Algorithm and the Predicted Column Density Outputs

Molecule NT NT + Std Dev. NT − Std Dev.
 (1011cm−2)(1011cm−2)(1011cm−2)
C2H223.06203138.721615.8525
C4H4.4814131.22130.1530
C6H46.8980512.21104.2940
C8H4.9852161.80650.1536
C10H6.2294255.19430.1521
C12H7.1877364.45810.1417
C2H 22.9619777.71980.6779
C4H 1.172413.52630.1016
C6H 1.298314.01850.1203
C8H 3.4520118.59920.1005
C10H 4.6145196.59180.1083
C12H 7.0534395.21620.1259

Note. The column densities predictions using gradient boosting regression. The standard deviations for the predictions were calculated using Gaussian process regression.

Download table as:  ASCIITypeset image

In case of C4H, the nearly degenerate low-lying electronic states (Senent & Hochlaf 2010) cause a discrepancy in the dipole moment trend that is observed across the other carbon chains investigated in this work. The average dipole resulting from a mixed state is not well understood, as discussed in Gratier et al. (2016). We attribute the low column density prediction for C4H to limitations in the ML predictions, as they are based purely on chemical similarity without dynamical/electronic effects being considered, and can be considered as a baseline; their main strength is generalizing across families of molecules to unseen species in a fast and data-driven manner.

The inherent value of the machine learning predictions is twofold: the capability to provide baseline expectation for the C10H/C10H ratio and a straightforward means to predict the abundance for the next in series carbon chains, C12H/C12H. The predictions from the machine learning approach are able to better match both the radical and anion species detected toward TMC-1 compared to the chemical formation model, except for the predictions of C4H due to electronic effects (as discussed above) and expected model uncertainty for species where representative inventory and knowledge is sparse or lacking. The machine learning approach also comes closer to predicting the observed C10H/C10H column density ratio compared to the chemical formation model. As such, it is now possible to make a prediction for the other nondetected carbon chains species, namely, C12H and C12H and the smaller molecular anion CCH. For comparison, the machine learning model prediction for CCH of 2.2 × 1013 cm−2 is within a factor of 3 of the previously measured value of 7.2 × 1013 cm−2 found by Pratap et al. (1997). For CCH, chemical model predictions are not as clear: earlier chemical models suggests CCH should have a lower abundance than the larger polyyne anions due to its greatly reduced radiative electron attachment rate (Cordiner et al. 2008; Herbst & Osamura 2008). Yet, the machine learning model prediction of 2.3 × 1012 cm−2 suggests that CCH may be abundant enough to be detected in sources such as TMC-1. However, the results from Agúndez et al. (2008) reported an upper limit for CCH of <2.2 × 1011 cm−2 toward TMC-1. As such, these disparate set of predictions compared to the reported upper limit again shows our limited understanding of the formation of the smaller CCH anion. For the larger species C12H and C12H, the predictions of 7.1 × 1011 cm−2 and 7.0 × 1011 cm−2, respectively, also suggest they may be detected in TMC-1 once the spectroscopy of these species (including calculated or measured dipole moments and partition functions), specifically C12H are fully characterized. The spectroscopy of C12H is reported in Gottlieb et al. (1998) and can be used to guide an astronomical search.

7. Conclusions

Using the GOTHAM data, we report the first astronomical detection of the C10H anion toward the dark cloud TMC-1 with the GBT. In fact, the astronomical observations provided the necessary data to the refinement of the spectroscopic parameters of C10H to enable the first astronomical detection. These new parameters will be essential for future studies of C10H in the laboratory. From the velocity stacked data and the matched filter response, C10H is detected at >9σ confidence interval. In addition, there is evidence for several individual lines of C10H in the GOTHAM data though none above the current noise level of the survey beyond the >3σ limit. In this case, three independent velocity components were detected and a total C10H column density of ${4.04}_{-2.23}^{+10.67}\times {10}^{11}$ cm−2 was determined.

A dedicated search for the C10H radical was also conducted toward TMC-1. The stacked molecular emission of C10H was detected at a ∼3.2σ confidence interval. As such, the presence of this molecule is currently considered tentative. In addition, no individual spectral lines from this species were detected. The measured physical values determined from the C10H fits include four independent velocity components, and a total C10H column density of ${2.02}_{-0.82}^{+2.68}\times {10}^{11}$ cm−2 was reported. Given the measured values, the C10H/C10H column density ratio is ∼${2.0}_{-1.6}^{+5.9}$ - the highest value measured between an anion and neutral species to date. Such a high ratio is at odds with current theories for interstellar anion chemistry.

The full GOTHAM data set was also used to better characterize the physical parameters including the column density [NT ], excitation temperature [Tex], line width [ΔV], and source size [''] for the more abundant C4H, C4H, C6H, C6H, C8H, and C8H molecules. These data were compared to predicted abundances from both a gas/grain chemical formation model and from a machine learning analysis. For the radical species, both models reproduce the measured abundances found from the survey better than an order of magnitude (except for the C4H predicted abundance from the machine learning analysis). However, the machine learning analysis matches the detected anion abundances much better than the gas/grain chemical model suggesting that the understanding of the formation chemistry of molecular anions is still highly questionable. Finally, using the machine learning analysis, it is possible to make a prediction for the larger species, namely, C12H and C12H and the smaller molecular anion CCH. For CCH, a model prediction of 2.3 × 1012 cm−2 shows that predicted column density of CCH is in stark contrast to the reported upper limit of <2.2 × 1011 cm−2. However, for the larger species C12H and C12H, the predictions of 7.1 × 1011 cm−2 and 7.0 × 1011 cm−2, respectively, suggest they may be detected in TMC-1 once the spectroscopy of these species has been fully characterized for an astronomical search.

The authors would like to thank the anonymous referees who provided valuable insight and suggestions that greatly improved the manuscript. We gratefully acknowledge support from NSF grants AST-1908576 and AST-2205126. I.R.C. acknowledges support from the University of British Columbia and the Natural Sciences and Engineering Research Council of Canada (NSERC). E.H. thanks the National Science Foundation (US) for support of his research program in astrochemistry through grant AST 19-06489. M.A.C. and S.B.C. were supported by the Goddard Center for Astrobiology. H.G. acknowledges support from the National Science Foundation for participation in this work as part of his independent research and development plan. Any opinions, findings, and conclusions expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. The Green Bank Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.

Appendix A: C10H/C10H Analysis

Figure A1 shows individual lines of C10H covered in the GOTHAM data through 25 GHz. The data and simulations are displayed at a smoothed resolution of 5.6 kHz (versus the 1.4 kHz native resolution of the GOTHAM observations) to show the (very weak) potential signal from these individual lines particularly between 13.1 and 19.2 GHz. Lines up to 30 GHz were included in the analysis (and are included in the catalog provided in the Supplementary Information), but are not shown here as the noise level of that data is much higher than the predicted line intensities.

Figure A1.

Figure A1. Individual lines of C10H covered in the GOTHAM observations through 25 GHz (black). Both the spectra and the simulations have been smoothed to a resolution of 5.6 kHz (from the native 1.4 kHz full resolution) to better show potential features. The spectra for lines falling at higher frequencies are substantially noisier and have been omitted from the plot to reduce the number of panels. Simulations of C10H emission using the parameters given in Table 2 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 6.0 km s−1 in total width.

Standard image High-resolution image

Figures A2 and A3 show the corner plots from the MCMC analysis of C10H and C10H, respectively.

Figure A2.

Figure A2. Corner plot for C10H showing parameter covariances and marginalized posterior distributions for the C10H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1 sigma for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure A3.

Figure A3. Corner plot for C10H showing parameter covariances and marginalized posterior distributions for the C10H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image

A.1. Jacknife Analysis

To further ensure that the signal attributed to C10H is molecular in origin, we performed a jacknife analysis similar to that described in detail in McGuire et al. (2021). The catalog for C10H was divided in half, with every other line assigned to one of two different catalogs. A spectral stack and matched filter was then performed on each catalog separately. Assuming the signal is indeed molecular and coming from C10H, the resulting impulse response functions should, when added in quadrature, closely reproduce the signal from the full catalog. The results of the analysis are shown in Figure A4, and we find impulse response functions of 5.3σ and 7.8σ. When added in quadrature, this results in a total of 9.2σ, in very good agreement with the value determined from the entire catalog and shown in Figure 2 of 9.3σ. As discussed in McGuire et al. (2021), the small mismatch is almost certainly due to minor contributions of red noise in the data at the level of, in this case, 0.1σ.

Figure A4.

Figure A4. Velocity stacked and matched filter spectra of C10H from a jacknife analysis where the full catalog was split into two. The intensity scales are the S/Ns of the response functions when centered at a given velocity. The "zero" velocity corresponds to the channel with the highest intensity to account for blended spectroscopic transitions and variations in velocity component source sizes. (Left) The stacked spectra from the GOTHAM DR4 data are displayed in black, overlaid with the expected line profile in red from our MCMC fit to the data. The S/N is on a per-channel basis. (Right) Matched filter response obtained from cross correlating the simulated and observed velocity stacks in the left panel; the value annotated corresponds to the peak impulse response of the matched filter.

Standard image High-resolution image

Appendix B: C4H Analysis

The best-fit parameters from the MCMC fit for C4H are shown in Table B1. The corner plot from the analysis is shown in Figure B1. Figures B2, B3, and B4 show the individual lines of C4H detected in the GOTHAM observations.

Figure B1.

Figure B1. Corner plot for C4H showing parameter covariances and marginalized posterior distributions for the C4H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure B2.

Figure B2. Individual lines of C4H detected in the GOTHAM observations (black). Simulations of C4H emission using the parameters given in Table B1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image
Figure B3.

Figure B3. Individual lines of C4H detected in the GOTHAM observations (black). Simulations of C4H emission using the parameters given in Table B1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image
Figure B4.

Figure B4. Individual lines of C4H detected in the GOTHAM observations (black). Simulations of C4H emission using the parameters given in Table B1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image

Table B1. C4H Values

vlsr Size NT a Tex ΔV
(km s−1)(arcsec)(1013cm−2)(K)(km s−1)
${5.657}_{-0.009}^{+0.008}$ ${570}_{-300}^{+289}$ ${6.95}_{-0.94}^{+0.85}$   
${5.754}_{-0.030}^{+0.029}$ ${29}_{-5}^{+5}$ ${6.52}_{-1.83}^{+2.24}$ ${5.10}_{-0.42}^{+0.46}$ ${0.178}_{-0.010}^{+0.013}$
${5.864}_{-0.032}^{+0.045}$ ${543}_{-332}^{+311}$ ${1.42}_{-0.68}^{+0.75}$   
${6.047}_{-0.025}^{+0.019}$ ${453}_{-292}^{+369}$ ${1.33}_{-0.26}^{+0.30}$   
NT (Total): b ${1.62}_{-0.22}^{+0.25}\times {10}^{14}$ cm−2

Notes. The quoted uncertainties represent the 16th and 84th percentile (1σ for a Gaussian distribution) uncertainties.

a Column density values are highly covariant with the derived source sizes. b Uncertainties derived by adding the uncertainties of the individual components in quadrature.

Download table as:  ASCIITypeset image

Appendix C: C4H Analysis

The best-fit parameters from the MCMC fit for C4H are shown in Table C1. The corner plot from the analysis is shown in Figure C1. Figure C2 shows the individual lines of C4H detected in the GOTHAM observations.

Figure C1.

Figure C1. Corner plot for C4H showing parameter covariances and marginalized posterior distributions for the C4H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure C2.

Figure C2. Individual lines of C4H detected in the GOTHAM observations (black). Simulations of C4H emission using the parameters given in Table C1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image

Table C1. C4H Values

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1010cm−2)(K)(km s−1)
${5.652}_{-0.032}^{+0.013}$ ${40}_{-4}^{+4}$ ${2.51}_{-1.07}^{+0.62}$   
${5.740}_{-0.034}^{+0.047}$ ${31}_{-4}^{+4}$ ${0.95}_{-0.68}^{+1.32}$ ${5.13}_{-0.47}^{+0.48}$ ${0.181}_{-0.018}^{+0.014}$
${5.916}_{-0.047}^{+0.040}$ ${29}_{-5}^{+5}$ ${0.44}_{-0.39}^{+0.62}$   
${6.018}_{-0.035}^{+0.058}$ ${10}_{-3}^{+4}$ ${2.89}_{-2.22}^{+4.41}$   
NT (Total): ${6.79}_{-2.59}^{+4.68}\times {10}^{10}$ cm−2

Download table as:  ASCIITypeset image

Appendix D: C6H Analysis

The best-fit parameters from the MCMC fit for C6H are shown in Table D1. The corner plot from the analysis is shown in Figure D1. Figures D2 and D3 show the individual lines of C6H detected in the GOTHAM observations.

Figure D1.

Figure D1. Corner plot for C6H showing parameter covariances and marginalized posterior distributions for the C6H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure D2.

Figure D2. Individual lines of C6H detected in the GOTHAM observations (black). Simulations of C6H emission using the parameters given in Table D1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 6.0 km s−1 in total width.

Standard image High-resolution image
Figure D3.

Figure D3. Individual lines of C6H detected in the GOTHAM observations (black). Simulations of C6H emission using the parameters given in Table D1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 6.0 km s−1 in total width.

Standard image High-resolution image

Table D1. C6H Values

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1012 cm−2)(K)(km s−1)
${5.600}_{-0.005}^{+0.012}$ ${356}_{-110}^{+99}$ ${1.15}_{-0.07}^{+0.15}$   
${5.744}_{-0.008}^{+0.020}$ ${35}_{-5}^{+55}$ ${3.06}_{-1.78}^{+0.58}$ ${5.12}_{-0.06}^{+0.13}$ ${0.154}_{-0.008}^{+0.025}$
${5.882}_{-0.019}^{+0.021}$ ${339}_{-136}^{+111}$ ${0.50}_{-0.42}^{+0.08}$   
${6.016}_{-0.023}^{+0.014}$ ${284}_{-121}^{+144}$ ${0.46}_{-0.07}^{+0.14}$   
NT (Total): ${5.17}_{-1.83}^{+0.62}\times {10}^{12}$ cm−2

Download table as:  ASCIITypeset image

Appendix E: C6H Analysis

The best-fit parameters from the MCMC fit for C6H are shown in Table E1. The corner plot from the analysis is shown in Figure E1. Figure E2 shows the individual lines of C6H detected in the GOTHAM observations.

Figure E1.

Figure E1. Corner plot for C6H showing parameter covariances and marginalized posterior distributions for the C6H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure E2.

Figure E2. Individual lines of C6H detected in the GOTHAM observations (black). Simulations of C6H emission using the parameters given in Table E1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image

Table E1. C6H Values

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1010cm−2)(K)(km s−1)
${5.646}_{-0.009}^{+0.009}$ ${91}_{-13}^{+19}$ ${8.76}_{-1.27}^{+1.49}$   
${5.784}_{-0.018}^{+0.017}$ ${32}_{-4}^{+4}$ ${9.94}_{-2.24}^{+2.52}$ ${4.32}_{-0.18}^{+0.22}$ ${0.177}_{-0.012}^{+0.013}$
${5.905}_{-0.031}^{+0.024}$ ${18}_{-4}^{+4}$ ${6.66}_{-3.53}^{+4.95}$   
${5.991}_{-0.010}^{+0.011}$ ${636}_{-274}^{+249}$ ${3.08}_{-0.43}^{+0.41}$   
NT (Total): ${2.84}_{-0.44}^{+0.58}\times {10}^{11}$ cm−2

Download table as:  ASCIITypeset image

Appendix F: C8H Analysis

The best-fit parameters from the MCMC fit for C8H are shown in Table F1. The corner plot from the analysis is shown in Figure F1. Figure F2 shows the individual lines of C8H detected in the GOTHAM observations.

Figure F1.

Figure F1. Corner plot for C8H showing parameter covariances and marginalized posterior distributions for the C8H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure F2.

Figure F2. Individual lines of C8H detected in the GOTHAM observations (black). Simulations of C8H emission using the parameters given in Table F1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Note that for C8H, the two separated features seen in each window represent the e (lower frequency) and f (higher frequency) states; the two listed hyperfine components are blended into the features. Each window is 6.0 km s−1 in total width.

Standard image High-resolution image

Table F1. C8H Values

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1011cm−2)(K)(km s−1)
${5.632}_{-0.004}^{+0.004}$ ${696}_{-253}^{+209}$ ${1.35}_{-0.07}^{+0.06}$   
${5.775}_{-0.008}^{+0.007}$ ${20}_{-3}^{+4}$ ${5.00}_{-1.22}^{+2.05}$ ${7.15}_{-0.18}^{+0.18}$ ${0.159}_{-0.008}^{+0.010}$
${5.897}_{-0.043}^{+0.031}$ ${654}_{-265}^{+237}$ ${0.47}_{-0.14}^{+0.10}$   
${6.010}_{-0.021}^{+0.027}$ ${672}_{-267}^{+227}$ ${0.44}_{-0.15}^{+0.15}$   
NT (Total): ${7.25}_{-1.23}^{+2.06}\times {10}^{11}$ cm−2

Download table as:  ASCIITypeset image

Appendix G: C8H Analysis

The best-fit parameters from the MCMC fit for C8H are shown in Table G1. The corner plot from the analysis is shown in Figure G1. Figure G2 shows the individual lines of C8H detected in the GOTHAM observations.

Figure G1.

Figure G1. Corner plot for C8H showing parameter covariances and marginalized posterior distributions for the C8H MCMC fit. The 16th, 50th, and 84th confidence intervals (corresponding to ±1σ for a Gaussian posterior distribution) are shown as vertical lines.

Standard image High-resolution image
Figure G2.

Figure G2. Individual lines of C8H detected in the GOTHAM observations (black). Simulations of C8H emission using the parameters given in Table G1 are shown in colors, with the total simulation in red. The quantum numbers for each transition are given in the upper left of each panel, and the central frequency of the window (in the sky frame) is given in the top right. Each window is 4.0 km s−1 in total width.

Standard image High-resolution image

Table G1. C8H Values

vlsr Size NT Tex ΔV
(km s−1)(arcsec)(1010cm−2)(K)(km s−1)
${5.646}_{-0.004}^{+0.004}$ ${87}_{-27}^{+51}$ ${0.70}_{-0.10}^{+0.17}$   
${5.812}_{-0.009}^{+0.005}$ ${29}_{-4}^{+4}$ ${2.20}_{-0.55}^{+0.74}$ ${6.17}_{-0.29}^{+0.30}$ ${0.115}_{-0.005}^{+0.005}$
${5.896}_{-0.030}^{+0.037}$ ${18}_{-4}^{+5}$ ${0.51}_{-0.38}^{+1.05}$   
${6.021}_{-0.005}^{+0.005}$ ${12}_{-5}^{+7}$ ${4.59}_{-2.50}^{+7.68}$   
NT (Total): ${8.00}_{-2.59}^{+7.79}\times {10}^{10}$ cm−2

Download table as:  ASCIITypeset image

Appendix H: Spectroscopy

The sources of the spectroscopic catalogs used for the analysis, as well as the literature references from which those catalogs were created, are provided in Table H1. The rotational partition function values for each of the molecules analyzed here are provided in Table H2 at the standard set of temperatures from SPFIT/SPCAT.

Table H1. Sources for the Spectroscopic Catalogs Used in the Analysis for Each Molecule

MoleculeCatalog SourceLab Ref.
C4HThis workGottlieb et al. (1983), this work
C4H CDMSGupta et al. (2007), McCarthy & Thaddeus (2008), Amano (2008)
C6HThis workGottlieb et al. (2010)
C6H CDMSMcCarthy et al. (2006)
C8HCDMSMcCarthy et al. (1996), McCarthy et al. (1999)
C8H CDMSGupta et al. (2007)
C10HCDMSGottlieb et al. (1998)
C10H This work

Download table as:  ASCIITypeset image

Table H2. Values of the Rotational Partition Function Used in the Analysis for Each of the Molecules at the Standard Set of Temperatures from SPFIT/SPCAT

T (K)C4HC4H C6HC6H C8HC8H C10HC10H
9.375165.539242.3000623.9222142.21011407.8848335.20481336.6177651.7587
18.75329.744684.26511478.7845284.08803275.2896670.07923029.91811303.1895
37.5660.0312168.19923503.2976567.84997869.63661339.83827308.60952605.6555
75.01420.6195336.07887850.95141135.396518005.02832679.395717004.51645146.2785
150.04095.2962671.880916755.01772270.580738992.12925358.668937330.05839254.1018
225.08848.39361007.739925714.48623405.886260175.21408038.153357920.628311998.1575
300.015542.21301343.656034689.69334541.312881412.899210717.848678585.629413864.1185

Download table as:  ASCIITypeset image

Appendix I: Chemical Modeling Time Dependence

Figure I1 shows the time dependence of the simulated abundances within the nautilus chemical models in relation to the observed values, shown as hashed horizontal boxes. As it can be seen, the relative trend lines of these species can be quite time dependent, the time of peak abundance is strongly dependent on the carbon-chain length. For ease of comparison to previous studies of carbon chains in TMC-1, we adopt the same source age as Siebert et al. (2022) of 2.5 × 105 yr, as also shown in Figure I1.

Figure I1.

Figure I1. Simulated gas-phase abundance and column densities of the Cn H (left) and Cn H (right) families from nautilus chemical models in comparison to the observed values with uncertainties shown as horizontal bars with hashed patterns. The time used in Figure 4 is shown as a vertical dashed gray line. The same hash and color scheme are used for molecules containing the same number of carbon atoms.

Standard image High-resolution image
Please wait… references are loading.
10.3847/2041-8213/acb648