Probing the Chemical Complexity of Amines in the ISM: Detection of Vinylamine (C2H3NH2) and Tentative Detection of Ethylamine (C2H5NH2)

, , , , , , , , , , and

Published 2021 October 14 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Shaoshan Zeng et al 2021 ApJL 920 L27 DOI 10.3847/2041-8213/ac2c7e

Download Article PDF
DownloadArticle ePub

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

2041-8205/920/2/L27

Abstract

Amines, particularly primary amines (R-NH2), are closely related to the primordial synthesis of amino acids since they share the same structural backbone. However, only a limited number of amines has been identified in the interstellar medium, which prevents us from studying their chemistry as well as their relation to prebiotic species that could lead to the emergence of life. In this Letter, we report the first interstellar detection of vinylamine (C2H3NH2) and tentative detection of ethylamine (C2H5NH2) toward the Galactic center cloud G+0.693-0.027. The derived abundance with respect to H2 is (3.3 ± 0.4) × 10−10 and (1.9 ± 0.5) × 10−10, respectively. The inferred abundance ratios of C2H3NH2 and C2H5NH2 with respect to methylamine (CH3NH2) are ∼0.02 and ∼0.008, respectively. The derived abundance of C2H3NH2, C2H5NH2, and several other NH2-bearing species are compared to those obtained toward high-mass and low-mass star-forming regions. Based on recent chemical and laboratory studies, possible chemical routes for the interstellar synthesis of C2H3NH2 and C2H5NH2 are discussed.

Export citation and abstract BibTeX RIS

1. Introduction

The term prebiotic molecules refers to species that are considered to be involved in the processes leading to the origin of life. From the prospective of understanding the prebiotic chemistry in the interstellar medium (ISM), primary amines have attracted much attention because they contain the same NH2 group as those considered to be the fundamental building blocks of life (e.g., amino acids, nucleobases, nucleotides, and other biochemical compounds). Although the attempts to observe glycine (NH2CH2COOH), the simplest amino acid, in the ISM have not succeeded (e.g., Ceccarelli et al. 2000; Belloche et al. 2013), an increasing number of NH2-bearing species have been reported in the last few years, which increases the chance to discover prebiotic complex organics in space. This is in fact well-demonstrated toward G+0.693-0.027 (hereafter G+0.693), a molecular cloud located within the Sgr B2 complex in the Galactic center. As revealed by the census of N-bearing species toward G+0.693 (Zeng et al. 2018), several species containing NH2 such as cyanamide (NH2CN), formamide (NH2CHO), and methylamine (CH3NH2) have been detected with abundance ≥10−9. Following the search, two key precursors in the synthesis of prebiotic nucleotides, hydroxylamine (NH2OH) and urea (NH2CONH2), have also been identified toward G+0.693 (Jiménez-Serra et al. 2020; Rivilla et al. 2020). The very recent discovery of by far the most complex amine, ethanolamine (NH2CH2CH2OH), the simplest head group of phospholipids in cell membranes, toward G+0.693 attests the potential of this source for finding more complex molecules of prebiotic relevance (Rivilla et al. 2021). This has thus prompted us to keep hunting for more amines in order to understand the chemical processes yielding related ingredients for life in space.

Structurally analogous to amino acids, CH3NH2 is the only primary amine that has been unambiguously observed toward different astronomical objects (e.g., Kaifu et al. 1974; Zeng et al. 2018; Bøgelund et al. 2019; Ohishi et al. 2019). Vinylamine (also known as ethenamine, C2H3NH2) and ethylamine (C2H5NH2) have not yet been reported in the ISM although the latter has been identified in multiple meteorites (Aponte et al. 2020, and references therein) and materials returned by the Stardust mission from comet 81P/Wild 2 (e.g., Glavin et al. 2008). Due to some concerns regarding the possible contamination of the stardust sample, their cometary origin could not be confirmed before the robust detection in the coma of comet 67P/Churyumov-Gerasimenko (Altwegg et al. 2016). The detection of CH3NH2 and C2H5NH2 together with glycine in these solar system objects enhanced the probability of amino acids being formed in space. Indeed, the retrosynthesis of amino acids revealed that molecules containing the -NH2 functional group are likely precursors of amino acids (Förstel et al. 2017). For example, CH3NH2 and C2H5NH2 may be the major constituents of glycine and alanine (NH2CH3CHCOOH), respectively. According to the detailed quantum chemical calculation, C2H3NH2 is the next energetically stable isomer in C2H5N group after E- and Z-conformer of ethanimine (CH3CHNH; Sil et al. 2018), both of which have been detected in Sgr B2 (Loomis et al. 2013) as well as in G+0.693 (Rivilla et al. 2021, in preparation). C2H5NH2 exists in two forms: anti- and gauche-conformer. The former is known to be more stable and has a higher expected intensity ratio than the latter conformer (see Sil et al. 2018, for details). Therefore anti-C2H5NH2 should be the most viable candidate for the astronomical detection in the C2H7N group. In this Letter, we present the first detection of C2H3NH2 and tentative detection of C2H5NH2 in the ISM toward G+0.693 through the identification of several rotational transitions of its millimeter spectrum.

2. Observations

We have carried out high-sensitivity spectral surveys at 7, 3, and 2 mm toward G+0.693 molecular cloud using the IRAM 30 m 9 and Yebes 40 m 10 telescopes. The observations were centered at α(J2000) = 17h47m22s, δ(J2000) = −28°21'27''. The position switching mode was used in all observations with the reference position located at Δα, Δδ = −885'', 290'' with respect to the source position. The half-power beamwidth of the IRAM 30 m and Yebes 40 m telescopes are in a range of 14''–36'' at observed frequencies between 30 GHz and 175 GHz. The intensity of the spectra was measured in units of antenna temperature, ${T}_{{\rm{A}}}^{* }$ as the molecular emission toward G+0.693 is extended over the beam (Zeng et al. 2020). The IRAM 30 m observations were performed in three observing runs during 2019: April 10–16, August 13–19, and December 11–15, from project numbers 172-18 (PI Martín-Pintado), 018-19 (PI Rivilla), and 133-19 (PI Rivilla). It covered spectral ranges of 71.76–116.72 GHz and 124.77–175.5 GHz. We refer to Rivilla et al. (2020) for a full description of the IRAM 30 m observations. The Yebes 40 m observations were carried out during six observing sessions in 2020 February, as part of the project 20A008 (PI Jiménez-Serra). The new Q-band (7 mm) HEMT receiver was used to allow broadband observations in two linear polarizations. The spectral coverage ranges from 31.075 GHz to 50.424 GHz. We refer to Zeng et al. (2020) and Rivilla et al. (2020) for more detailed information on the Yebes 40 m observations.

3. Analysis and Results

The line identification and analysis were carried out using the Spectral Line Identification and Modeling (SLIM) tool implemented within the madcuba package 12 (version 21/12/2020, Martín et al. 2019). The spectroscopic information of C2H3NH2, within 0+ & 0 (CDMS entry 43504) 11 was obtained from Brown et al. (1990) and Mcnaughton & Robertson (1994). And the spectroscopic information of C2H5NH2, anticonformer (CDMS entry 45515) was obtained from Fischer & Botskor (1982) and Apponi et al. (2008). Table 1 summarizes the unblended or only partially blended transitions of C2H3NH2 and C2H5NH2 detected toward G+0.693. Note that all the C2H5NH2 lines in Table 1 are a-type transitions with selection rules of ΔKa = 0 and ΔKc = ±1. The rotational spectrum of C2H3NH2 and C2H5NH2 are characterized by inversion doubling due to the large amplitude inversion motion of the NH2 group. Each rotational energy level, specified by the rotational quantum numbers J and K, is thus split into an inversion doublet. The 0+ or 0 in Table 1 indicate the inversion doublet from which the transition arises. For C2H5NH2, the variation of upper state degeneracy and integrated intensity of the same transition is due to the spin statistical weight of 3:1 between the symmetric and antisymmetric sublevels.

Table 1. List of Observed Transitions of C2H3NH2 and C2H5NH2

FrequencyTransitionlog Aul gu Eu rms $\int {T}_{A}^{* }\,{dv}$ S/N a Blending
(GHz)(${J}_{{K}_{a}}$, Kc )(s−1) (K)(mK)(mK km s−1)  
C2H3NH2
92.3122950,5–40,4, 0+ −5.316111113.31.232546aGg-(CH2OH)2
92.3153950,5–40,4, 0 −5.413091178.31.28  
*92.9208552,4–42,3, 0+ −5.382641122.41.317022clean
*96.5136951,4–41,3, 0+ −5.275331116.12.528219clean
112.6247962,4–52,3, 0+ −5.100951327.85.71986g-C2H5SH
*128.3203070,7–60,6, 0+ −4.876141524.78.73507clean
*129.9244572,6–62,5, 0+ −4.895881533.96.82005clean
*146.0393480,8–70,7, 0+ −4.704201731.72.931620clean
149.1429083,6–73,5, 0+ −4.741221752.42.9936CH3C13CH
159.8870191,9-81,8, 0+ −4.587631940.75.32419C2H5CN
C2H5NH2
*32.1460421,2–11,1, 0 −6.88737153.41.184clean
* 32.1460521,2–11,1, 0+ −6.88739453.41.124  
*34.0465921,1–11,0, 0 −6.81259153.51.284clean
*34.0466221,1–11,0, 0+ −6.81261453.51.225  
*49.5287530,3–20,2, 0+ −6.16968634.72.8676clean
*49.5287530,3–20,2, 0 −6.16966214.72.822  
82.1687850,5–40,4, 0+ −5.485289911.82.81098 13CH3CH2OH
82.1687850,5–40,4, 0 −5.485263311.82.836  
* 83.2428752,3–42,2, 0+ −5.543129916.43.1644clean
*83.2428752,3–42,2, 0 −5.543103316.43.121  
84.9807851,4–41,3, 0+ −5.458289913.33.3966H15NCO
84.9807851,4–41,3, 0 −5.458263313.33.332  
145.8717490,9-80,8, 0+ −4.7210517135.32.3566S18O
145.8717490,9-80,8, 0 −4.721035735.32.319  

Note. The following parameters are obtained from the CDMS catalog entries 43504 and 45515: frequencies, quantum numbers, upper state degeneracy (gu ), the logarithm of the Einstein coefficients (log Aul ), and the energy of the upper levels (Eu ). The derived root mean square (rms) of the analyzed spectra region, integrated intensity ($\int {T}_{A}^{* }\,{dv}$), signal-to-noise ratio (S/N), and the information about the species with transitions slightly blended with C2H3NH2 or C2H5NH2 lines are provided in the last column.

a S/N is calculated as $(\int {T}_{A}^{* }{dv})/[\mathrm{rms}{\left(\tfrac{{\rm{\Delta }}\nu }{\mathrm{FWHM}}\right)}^{0.5}\mathrm{FWHM}]$, where Δν is the spectral resolution of the data, ranging between 1.5 and 2.2 km s−1. The clean transitions are denoted by an asterisk. A single common S/N is given for the overlapping transitions.

Download table as:  ASCIITypeset image

The analysis was performed under the assumption of local thermodynamic equilibrium (LTE) conditions due to the lack of collisional coefficients of C2H3NH2 and C2H5NH2. Due to the low density of G+0.693 (∼104–105 cm−3; Zeng et al. 2020), molecules are subthermally excited in the source and hence their excitation temperatures (in a range of 5–20 K; e.g., Requena-Torres et al. 2008; Rivilla et al. 2018; Zeng et al. 2018) are significantly lower than the kinetic temperature of the source (∼150 K; e.g., Zeng et al. 2018). However, note that the transitions of C2H3NH2 and C2H5NH2 are well fitted using one excitation temperature for each species. Considering the effect of line opacity, MADCUBA-SLIM generated synthetic spectra that can be compared to the observed spectra. The MADCUBA-AUTOFIT tool was then used to provide the best nonlinear least-squares LTE fit to the data using the Levenberg–Marquardt algorithm. It is important to note that not a single transition of C2H3NH2 and C2H5NH2 predicted by the LTE spectrum is missing in the data. To properly evaluate the line contamination by other molecules, over 300 species have been searched for in our data set. This included not only all the molecules detected toward G+0.693 in previous studies (Requena-Torres et al. 2008; Rivilla et al. 2018, 2019, 2020, 2021; Zeng et al. 2018; Jiménez-Serra et al. 2020; Rodríguez-Almeida et al. 2021), but also those reported in the ISM. 13 The molecule blended with the transitions of C2H3NH2 and C2H5NH2 is listed in the last column of Table 1. We note that the line identification and fitting file used in this analysis is the same as the one used in previous works (e.g., Zeng et al. 2018; Rodríguez-Almeida et al. 2021) as well as ongoing works. Therefore, the excitation temperatures and column densities of blending species are consistent with the ones reported in all these studies.

For C2H3NH2, we detect five clean transitions and four slightly blended transitions with a blending contribution of <10% (see Figure 1). For C2H5NH2, only four clean transitions are reported and three are slightly blended (Figure 2). Since the clean transitions of C2H5NH2 are weak (S/N = 4 in integrated intensity; Table 1), we conclude that this species is tentatively detected. The free parameters that can be fitted are molecular column density (Ntot), excitation temperature (Tex), radial velocity (VLSR), full width at half-maximum (FWHM, Δν), and source size (θ). For G+0.693, we assumed that the source size is extended in madcuba. As the algorithm did not converge when fitting C2H3NH2 with all parameters left free, we fixed the FWHM and VLSR to 18 km s−1 and 67 km s−1, respectively, by visual inspection of the most unblended lines. These values are consistent with those from many other molecules previously analyzed in G+0.693 (FWHM ∼ 20 km s−1, VLSR ∼ 68 km s−1; Requena-Torres et al. 2008; Zeng et al. 2018; Rivilla et al. 2018, 2019, 2020; Jiménez-Serra et al. 2020; Rodríguez-Almeida et al. 2021). The resulting LTE fit gives Tex = (18 ± 3) K and Ntot = (4.5 ± 0.6) × 1013 cm−2. In the case of C2H5NH2, the VLSR was fixed to 67 km s−1 and the LTE fit gives Tex = (12 ± 5) K, FWHM = (18 ± 5) km s−1, and Ntot = (2.5 ± 0.7) × 1013 cm−2.

Figure 1.

Figure 1. Unblended or only slightly blended transitions of C2H3NH2 detected toward G+0.693. The red line shows the best LTE fit to the C2H3NH2 lines while the blue line shows the total contribution, including the emission from other molecular species (labeled) identified in G+0.693. The cleanest detected transitions are denoted by a red ▾.

Standard image High-resolution image
Figure 2.

Figure 2. Unblended or only slightly blended transitions tentatively identified for C2H5NH2 toward G+0.693. The cleanest transitions are denoted by a red ▾.

Standard image High-resolution image

The resulting best LTE fit to the C2H3NH2 and C2H5NH2 lines is shown by the red line in Figure 1 and Figure 2 while the blue line indicates the best fit considering also the total contribution of the LTE emission from all the other identified molecules. Adapting the H2 column density inferred from observations of C18O (${N}_{{{\rm{H}}}_{2}}$ = 1.35 × 1023 cm−2; Martín et al. 2008), the resulting abundance is (3.3 ± 0.4) × 10−10 and (1.9 ± 0.5) × 10−10 for C2H3NH2 and C2H5NH2, respectively.

4. Discussion

Figure 3 presents the abundance with respect to H2, in decreasing order, of NH2-bearing species detected toward G+0.693. The results are compared to those derived toward three high-mass and low-mass star-forming regions that are chemically rich, i.e., Sgr B2(N), Orion KL, and IRAS 16293-2422 B. The low abundance or nondetection of -NH2 species may indicate that their formation is less efficient toward Orion KL and IRAS 16293-2422 B. On the other hand, the detection with abundance >10−11 suggests G+0.693 is a prominent -NH2 molecule repository, which allows us to study their origin as well as their chemical relation to other prebiotic molecules.

Figure 3.

Figure 3. Derived abundance of NH2-bearing species with respect to H2 toward G+0.693-0.027, Sgr B2(N), Orion KL, and IRAS 16293-2422 B. For G+0.693-0.027 (N(H2) =1.35 × 1023 cm−2; Martín et al. 2008), molecular column densities are obtained from this work, Zeng et al. (2018), Jiménez-Serra et al. (2020), Rivilla et al. (2020), and Rivilla et al. (2021). For IRAS 16293-2422 B (N(H2) = 2.8 × 1025 cm−2; Martín-Doménech et al. 2017), molecular column densities are from Martín-Doménech et al. (2017), Ligterink et al. (2018), and Jiménez-Serra et al. (2020). For Sgr B2(N), CH3NH2, NH2CHO, and NH2CN are from Belloche et al. (2013) with N(H2) = 1.3 × 1025 cm−2 (Belloche et al. 2008); the abundance of C2H5NH2 and the upper limit of NH2OH are taken directly from Apponi et al. (2008) and Pulliam et al. (2012), respectively; NH2CONH2 is derived from Belloche et al. (2019) by assuming Sgr B2(N1S) has the same N(H2) = 3.5 × 1024 cm−2 (Li et al. 2021) as Sgr B2 (N1E). For Orion KL, the column density of CH3NH2 and N(H2) = 3.1 × 1023 cm−2 are from Pagani et al. (2017); NH2CHO is from Motiyenko et al. (2012) with N(H2) = 4.2 × 1023 cm−2 (Tercero et al. 2010); upper limits of NH2OH and NH2CH2CH2OH are from Pulliam et al. (2012) and Wirström et al. (2007), respectively, with N(H2) = 7.0 × 1023 cm−2 (Womack et al. 1992).

Standard image High-resolution image

In contrast to the three compared sources, G+0.693 lacks an internal heating source responsible for the rich chemistry. The high level of molecular complexity is attributed to dust grain sputtering by low-velocity shocks (≤20 km s−1), which is driven by the possible cloud–cloud collision occurring in the Sgr B2 complex (Zeng et al. 2020). This is indicated by the high abundances of shock tracers such as HNCO and SiO (Martín et al. 2008; Rivilla et al. 2018) and the presence of molecules that are known to be formed on grain surfaces (Requena-Torres et al. 2008; Zeng et al. 2018). In addition, the abundance ratio of HC3N/HC5N and C2H3CN/C2H5CN derived in Zeng et al. (2018) suggested that an enhanced cosmic-ray ionization rate may also play a role in the chemistry of N-bearing species toward G+0.693. In the following section, we evaluate the possible formation routes for the amines detected toward G+0.693 and the discussed formation routes are summarized in Figure 4.

Figure 4.

Figure 4. Summary of the chemical routes proposed for the formation of C2H3NH2 and C2H5NH2 in the ISM. The molecular species in red are those that have been detected toward G+0.693. The solid arrows denote surface chemistry reactions and dashed arrows denote gas-phase chemistry.

Standard image High-resolution image

4.1. Formation of Primary Amines

Despite the low number of detections in the ISM, the CH3NH2 chemistry under astrophysical conditions has been studied in theoretical, experimental, and chemical modeling work. In the gas-phase, CH3NH2 is proposed to form via the radiative association between ammonia (NH3) and the methyl radical cation (CH3 +) followed by recombination dissociation (Herbst 1985). On the grain surfaces, experimental work demonstrated that CH3NH2 can form via sequential hydrogenation of hydrogen cyanide (HCN): HCN → CH2NH → CH3NH2 (Theule et al. 2011). Alternatively, the gas-grain chemical model by Garrod et al. (2008) has suggested that CH3NH2 is formed by simple addition of CH3 from CH4 to azanyl radical (NH2) from NH3 during warm-up phases. This radical–radical recombination has recently been studied in laboratory ice simulations revealing that CH3NH2 can be synthesized in irradiated ices composed of CH4 and NH3 (Kim & Kaiser 2011; Förstel et al. 2017) but also in cold and quiescent molecular clouds (Ioppolo et al. 2021). On the contrary, little is known about the chemistry of C2H3NH2 and C2H5NH2.

C2H5NH2 —Although there are no chemical routes included in current astrochemical databases, different formation mechanisms of C2H5NH2 have been discussed in the literature. For instance, the addition of the CH3 radical to CH2NH followed by hydrogenation of the resulting product would produce C2H5NH2 on grain surfaces (Bernstein et al. 1995); the photochemistry of a mixture of ethylene (C2H4) and NH3 in which C2H5NH2 is formed from radical–radical reaction of the ethyl radical (C2H5) with NH2 (Danger et al. 2011, and references therein). Reactions of CH3NH2 with carbene (CH2) or ethane (C2H6) with nitrene (NH) are also expected to form C2H5NH2, but with lower probability (Förstel et al. 2017). More recently, experimental simulation of interstellar ice analogs containing CH3NH2 revealed that C2H5NH2 can be formed efficiently from CH3 and methanamine radical (CH2NH2; Carrascosa et al. 2021).

In addition to the aforementioned formation pathways, one might also expect C2H5NH2 to be formed through the same mechanism as suggested for CH3NH2: successive hydrogenation starting from acetonitrile (CH3CN): CH3CN → CH3CHNH → C2H5NH2. However, CH3CN has been proposed to be a product of the surface chemistry of CH3NH2 (Carrascosa et al. 2021). Furthermore, laboratory work showed that CH3CN does not react with H atoms between 10 and 60 K (Nguyen et al. 2019). Considering the temperature of dust grains in G+0.693 is Tdust ≤ 30 K (Zeng et al. 2018, and reference therein), this proposed hydrogenation leading to C2H5NH2 is unlikely to occur on dust grains in G+0.693. Based on the reaction rate coefficients provided in Sil et al. (2018), C2H5NH2 is less likely produced in the same hydrogenation reaction in the gas phase due to its low efficiency at typical kinetic temperature in Galactic center clouds, Tkin = 50−120 K (Zeng et al. 2018, and reference therein). To our best knowledge, with no other possible gas-phase reaction to form C2H5NH2, this species likely forms on the surface of dust grains in G+0.693. Regardless of the chemical formation route on grains, the sputtering of grain icy mantles by large-scale low-velocity (≤20 km s−1) shocks in G+0.693 would release C2H5NH2 into gas phase from grains (see, e.g., Martín et al. 2008; Requena-Torres et al. 2008; Zeng et al. 2020).

C2 H3 NH2 —With the available reaction rate coefficients in the Kinetic Database for Astrochemistry (Wakelam et al. 2012), C2H3NH2 is proposed to form from the reaction between the CH radical and CH3NH2 in the gas-phase, most efficient at temperatures T = 50−200 K. With Tkin = 50−120 K, C2H3NH2 is thus expected to be formed efficiently via this pathway in G+0.693. In particular, CH3NH2 is found to be abundant (∼10−8; Zeng et al. 2018) in G+0.693, which would be readily available for this chemical reaction to proceed. Another possible gas-phase formation route is through the reaction involving C2H5NH2 + H+ or H3 + followed by recombination dissociation, analogous to the formation of C2H3CN from C2H5CN proposed by Caselli et al. (1993). As discussed in Zeng et al. (2018) for -CN group species, this ion-molecule gas-phase reaction can be efficient thanks to (i) the presence of high cosmic-ray ionization rate in the Galactic center and (ii) the relatively low densities of ∼104 cm−3 of G+0.693. This, and the fact that similar abundance ratios are found for C2H3CN/C2H5CN = 2.2 ± 0.3 and C2H3NH2/C2H5NH2 = 1.7 ± 0.5 toward G+0.693, makes this formation route plausible.

On the grain surface, C2H3NH2 might be a photoproduct of C2H5NH2 (Hamada et al. 1984). But recent experimental investigation showed that its isomer, ethanimine (CH3CHNH), appears to be the primary product of photolysis of C2H5NH2 (Danger et al. 2011). This has also been recently found by Carrascosa et al. (2021), who even synthesized large N-heterocycles in interstellar ice analogs under UV radiation. We thus propose that the formation of C2H3NH2 likely occurs in the gas phase toward G+0.693 although further theoretical or laboratory work is needed to determine the rate constant of the reaction C2H5NH2 + H+/H3 +.

In summary, we report the discovery of two new amines in the ISM: C2H3NH2 and tentatively C2H5NH2. The abundance ratios with respect to CH3NH2 are ∼0.02 and ∼0.008, i.e., about a factor >10. This trend has been found for other species such as ethanol (with respect to methanol) or ethyl cyanide (with respect to CH3CN; Zeng et al. 2018; Rodríguez-Almeida et al. 2021). Primary amines are known to be involved in the synthesis of proteinogenic alpha-amino acids (Förstel et al. 2017). Therefore, their discovery provides crucial information about the connection between interstellar chemistry and the prebiotic material found in meteorites and comets.

The authors wish to thank the referees for constructive comments that significantly improved the paper. We are grateful to the IRAM 30 m and Yebes 40 m telescope staff for help during the different observing runs. IRAM is supported by the National Institute for Universe Sciences and Astronomy/National Center for Scientific Research (France), Max Planck Society for the Advancement of Science (Germany), and the National Geographic Institute (IGN) (Spain). The 40 m radio telescope at Yebes Observatory is operated by the IGN, Ministerio de Transportes, Movilidad y Agenda Urbana. This study is supported by a Grant-in-Aid from the Ministry of Education, Culture, Sports, Science, and Technology of Japan (20H05845) and by a RIKEN pioneering Project (Evolution of Matter in the Universe). We also acknowledge partial support from the Spanish National Research Council (CSIC) through the i-Link project number LINKA20353. I.J.-S. and J.M.-P. have received partial support from the Spanish State Research Agency through project number PID2019-105552RB-C41. V.M.R., L.R.-A., and L.C. have received funding from the Comunidad de Madrid through the Atracción de Talento Investigador (Doctores con experiencia) Grant (COOL: Cosmic Origins Of Life; 2019-T1/TIC-15379).

Footnotes

Please wait… references are loading.
10.3847/2041-8213/ac2c7e