Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-27T18:16:59.801Z Has data issue: false hasContentIssue false

Radiation-induced reactions in comet analogues

Published online by Cambridge University Press:  01 December 2022

A. López-Islas*
Affiliation:
Instituto de Ciencias Nucleares, Cto. Exterior S/N, C.U., Coyoacán, 04510 Ciudad de México, México
A. Negrón-Mendoza
Affiliation:
Instituto de Ciencias Nucleares, Cto. Exterior S/N, C.U., Coyoacán, 04510 Ciudad de México, México
*
Author for correspondence: A. López-Islas, E-mail: anayelly.lopez@correo.nucleares.unam.mx
Rights & Permissions [Opens in a new window]

Abstract

Comets are a source of prebiotic molecules that likely enriched the early Earth during the Late Heavy Bombardment period. Laboratory experiments that replicate cometary conditions may facilitate understanding of the chemical reactions and supplement observational studies of these icy bodies. Prebiotic compounds, such as formic acid and formaldehyde, have been observed in comets. Furthermore, these compounds can easily be formed in experimental models using a variety of gas combinations and energy sources. We conducted experimental cometary simulations using radiation chemistry tools to obtain insight into the possible fate of formic acid and formaldehyde. The main results suggest a redundant system, signifying that the irradiation of formic acid forms formaldehyde molecules and vice versa. This phenomenon ensures the permanence of prebiotic molecules in high-radiation environments. Additionally, the potential role of forsterite and graphite was explored in cometary simulations. Our experimental results show the differential formation of aldehydes and other carbonyl-containing compounds dependent on the mineral phase present.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

Introduction

In the 1980s, explorations of the solar system provided new insight into the distribution of organic compounds in extraterrestrial environments. Specifically, organic compounds have been found in meteorites and comets (Kobayashi, Reference Kobayashi2019). The organic molecules detected in comets are highly diverse. Amino acids, the building blocks of proteins, have been detected in the Murchison meteorite, a CM2 carbonaceous chondrite (Martins et al., Reference Martins, Watson, Sephton, Botta, Ehrenfreund and Gilmour2006). Additionally, HCN-, NH3 and CN-containing compounds have been detected in comets (Despois and Cottin, Reference Despois, Cottin, Gargaud, Barbier, Martin and Reisse2005; Goesmann et al., Reference Goesmann, Rosenbauer, Bredehöft, Cabane, Ehrenfreund, Gautier, Giri, Krüger, Le Roy, MacDermott, McKenna-Lawlor, Meierhenrich, Muñoz Caro, Raulin, Roll, Steele, Steininger, Sternberg, Szopa, Thiemann and Ulamec2015), and simple molecules, such as H2O, H2CO and HCOOH, are abundant in cometary ices (Delsemme, Reference Delsemme1988; Bockelée-Morvan et al., Reference Bockelée-Morvan, Crovisier, Mumma and Weaver2004; Llorca, Reference Llorca2005). In this respect, comets are considered a source of prebiotic molecules that may have enriched the early Earth during the Late Heavy Bombardment (Chyba et al., Reference Chyba, Thomas, Brookshaw and Sagan1990).

Interaction of ionizing radiation with comets

Different types of radiation induce chemical transformations in cometary ices. Irradiation with cosmic rays (an ionizing radiation source) and ultraviolet light can induce the formation of organic compounds via radical reactions. To a lesser extent, embedded radionuclides in the cometary core (40K, 235U, 238U, 232Th, 244Pu, 129I, 247Cm, 10Be and 237Np) could function as a minor energy source (Draganić and Draganić, Reference Draganić and Draganić1984), which has been estimated to deposit up to 14 MGy. This corresponds to the expected total dose delivered by the natural decay of the radionuclides in the comet over 4.6 × 109 years (Cataldo et al., Reference Cataldo, Ursini, Angelini, Iglesias-Groth and Manchado2011; Iglesias-Groth et al., Reference Iglesias-Groth, Cataldo, Ursini and Manchado2011).

Protons from cosmic rays are the most abundant ionizing particles in the solar system: they have an energy of approximately 2 GeV (Draganić and Draganić, Reference Draganić and Draganić1988) and a linear energy transfer (LET) comparable to gamma rays. As a result, gamma-ray irradiation is an excellent candidate for replicating GeV proton irradiation of cometary nuclei.

Minerals detected in comets

Minerals have also been detected in comets, and they may significantly impact the chemical behaviour of organics. The primary mineral phases in these bodies are olivine and pyroxene. Forsterite (olivine) and enstatite (pyroxene) were detected in cometary dust samples recovered by the Stardust mission from 81P/Wild 2 (Brownlee et al., Reference Brownlee, Tsou, Aléon, O'D Alexander, Araki, Bajt and Baratta2006). Additionally, inorganic carbon has been identified as a component of cometary dust (Kissel et al., Reference Kissel, Brownlee, Büchler, Clark, Fechtig, Grün, Hornung, Igenbergs, Jessberger, Krueger, McDonnell, Morfill, Rahe, Schwehm, Sekanina, Utterback, Volk and Zook1986a, Reference Kissel, Sagdeev, Bertaux, Angarov, Audouze, Blamont, Büchler, Evlanov, Fechtig, Fomenkova, von Hoerner, Inogamov, Khromov, Knabe, Krueger, Langevin, Leonas, Levasseurregourd, Managadze, Podkolzin, Shapiro, Tabaldyev and Zubkov1986b). According to studies of particles from 67P/Churyumov–Gerasimenko, comets contain carbon-rich and non-hydrated minerals (Bardyn et al., Reference Bardyn, Baklouti, Cottin, Fray, Briois, Paquette, Stenzel, Engrand, Fischer, Hornung, Isnard, Langevin, Lehto, Le Roy, Ligier, Merouane, Modica, Orthous-Daunay, Rynö, Schulz, Silén, Thirkell, Bastien, Bland, Bleuet, Borg and Zolensky2006; Fray et al., Reference Fray, Bardyn, Cottin, Altwegg, Baklouti, Briois, Colangeli, Engrand, Fischer, Glasmachers, Grün, Haerendel, Henkel, Höfner, Hornung, Jessberger, Koch, Krüger, Langevin, Lehto, Lehto, Le Roy, Merouane, Modica, Orthous-Daunay, Paquette, Raulin, Rynö, Schulz, Silén, Siljeström, Steiger, Stenzel, Stephan, Thirkell, Thomas, Torkar, Varmuza, Wanczek, Zaprudin, Kissel and Hilchenbach2016). Limited data exist on the role of minerals in prebiotic reactions in comets. This work aimed to evaluate the reactivity of formic acid and formaldehyde in an experimental model of a simple cometary core.

Methodology

The experimental model used in the present work was designed according to the previous one proposed by Draganić et al. (Reference Draganić, Vujosevic, Negrón-Mendoza, Azamar and Draganić1985). The model consisted of a frozen aqueous solution in a glass container. First, the aqueous solution is evacuated to eliminate the atmospheric oxygen and then is frozen with liquid nitrogen. Finally, the glass container with the frozen solution is irradiated with gamma radiation at different doses.

In the present study, the experimental model included a frozen aqueous solution of formic acid or formaldehyde and a mineral phase. Next, this simple cometary core was exposed to an ionizing radiation energy source.

The experimental sets are presented in Scheme 1 below: (1) frozen formic acid solutions with and without minerals, and (2) frozen formaldehyde solutions with and without forsterite.

Scheme 1. Experimental sets: (1) frozen (77 K) solutions of formic acid (0.02 M) with and without minerals and (2) frozen (77 K) solutions of formaldehyde (0.3 M) with and without minerals.

Irradiation source

The samples were irradiated in a cobalt-60 source facility (Gammabeam 651 PT) at the Instituto de Ciencias Nucleares, UNAM, Mexico. The dose rates of the gamma source (266 Gy min−1 at room temperature and 44 Gy min−1 at 77 K, [Gray, 1 Gy = J kg−1]) were measured using the ferrous sulphate–cupric sulphate dosimeter (O'Donnell and Sangster, Reference O'Donnell and Sangster1970; Draganić and Draganić, Reference Draganić and Draganić1971). In the dosimeter, the transformation of Fe2+ to Fe3+ induced by radiation was measured via UV spectroscopy at 304 nm. These data can evaluate the energy deposited in the system over time. If the time of exposure is known, the dose can be evaluated.

The samples were placed 10 cm from the source inside a Dewar flask filled with liquid nitrogen. The dosimetry was conducted under the same conditions as the samples to account for the attenuation of the flask and nitrogen. Using these data, the absorbed dose in our model was established as previously described (Cruz-Castañeda et al., Reference Cruz-Castañeda, Camargo, López-Islas and Negrón Mendoza2018; Leal-Acevedo and Gamboa-de Buen, Reference Leal-Acevedo and Gamboa-de Buen2020). The irradiation was performed at a fixed position, and the irradiation time was varied.

Preparation and irradiation of a simple cometary core of formic acid

A 95% formic acid solution (Sigma-Aldrich, Saint Louis, Missouri, USA) and triple-distilled water were used to prepare formic acid at 0.02 M. Oxygen-free aliquots (5 ml) of formic acid were frozen in glass tubes using liquid nitrogen and maintained in a Dewar flask (77 K). The samples were exposed to various doses of gamma radiation (0–90 kGy).

Measurement of formic acid decomposition by titration

After exposure to gamma radiation, the decomposition of the formic acid was measured by titration. Sodium hydroxide was used as a titrant solution, 0.5% phenolphthalein (Sigma-Aldrich) was used as an indicator, and formic acid (1 × 10−3 M) was used as the standard solution.

Aldehyde and ketone detection via derivatization using 2,4-dinitrophenylhydrazine

The detection of aldehydes and ketones in the irradiated formic acid samples (42 kGy) was performed through derivatization using 2,4-dinitrophenylhydrazine (DNPH) (Sigma-Aldrich). The derivatives were recovered by centrifugation (14 000 rpm × 10 min) and dried at room temperature. Finally, the dried derivatives were dissolved in acetonitrile (Merck, Germany) and analysed using high-performance liquid chromatography (HPLC; Knauer Azura P 4.1S, Germany) coupled with a UV detector. A C-18 ODS 2.7 μm column, (Beckman Coulter, USA), 4.6 mm × 150 mm was employed. A mixture of acetonitrile and water (70:30) was used as the eluent; the flow rate was 1 ml min−1.

Additionally, DNPH-derivatives (the irradiated samples with minerals) were analysed by HPLC-MS. A 515 HPLC pump was used coupled with a SQ-2 Single Quadrupole Mass Detector system, with electrospray ionization in negative (ESI–) mode (Waters®, Massachusetts, USA). The cone energy was 24 V, the capillary energy was 1.46 kV and the desolvation temperature was 350°C. The DNPH-derivatives were separated in a Symmetry C18 3.5 μm column (Waters®), 4.6 mm × 75 mm. A mixture of acetonitrile and water (70:30) was used as the eluent; the flow rate was 0.8 ml min−1.

Formic acid cometary core with mineral phases

Two sets of experiments were performed for this part of the study: one using graphite and the other using forsterite. Graphite powder (0.1 g) or forsterite grains (1 g) in formic acid solution (5 ml) were placed into glass tubes. The oxygen-free mixtures were frozen using liquid nitrogen (77 K). Later, the glass tubes were placed in a Dewar flask containing liquid nitrogen and irradiated (0–70 kGy). The formic acid decomposition was evaluated by titration, and the radiolytic products (aldehydes) were identified using DNPH derivatization, as described above.

Preparation and irradiation of a simple cometary core of formaldehyde

A 37% formaldehyde solution (Sigma-Aldrich) and triple-distilled water were used to prepare formaldehyde at 0.3 M. Oxygen-free aliquots (5 ml) of formaldehyde were frozen in glass tubes using liquid nitrogen (77 K). The samples were kept in a Dewar flask to maintain the temperature throughout the irradiation. The applied doses ranged from 0 to 600 kGy.

Measurement of formaldehyde decomposition by gas chromatography

The decomposition of the formaldehyde was measured by gas chromatography using a Varian 2400 chromatograph coupled with a flame ionization detector (California, USA) and a Chromosorb 102 column (Sigma-Aldrich), i. d. 1/8-inch × 4 m. Nitrogen was used as the carrier gas at a flow rate of 30 ml min−1. The temperature ramp started at 60°C and ended at 230°C. The rate of heating was 6°C min−1.

Identification of radiolytic products in the irradiated formaldehyde samples

Titration assays were conducted to quantify the formic acid formation in the irradiated formaldehyde samples. The experimental procedure was performed as previously described in section ‘Measurement of formic acid decomposition by titration’.

Formaldehyde cometary core with forsterite

Forsterite (1 g) and formaldehyde solutions (5 ml) were placed in glass tubes. The oxygen-free mixtures were frozen using liquid nitrogen. Later, the glass tubes were placed in a Dewar flask containing liquid nitrogen (77 K) and irradiated (0–300 kGy). Gas chromatography was performed to measure the decomposition of the irradiated formaldehyde under these conditions. Additionally, titration was performed to quantify the radiolytic products.

Results

The analysis of the results presented in this work was based on measuring the decomposition of formic acid and formaldehyde upon exposure to a high radiation environment in the presence of minerals. Additionally, their radiolytic products were identified.

Decomposition of formic acid and the identification of its radiolytic products

Figure 1 shows the percentage of the formic acid samples recovered after irradiation. With gamma radiation exposure from 0 to 90 kGy, there were minimal changes in the decomposition of the formic acid. The recovery rate of this molecule remained greater than 90% after irradiation.

Fig. 1. Recovery of formic acid after irradiation, quantified by titration. Frozen formic acid solutions (0.02 M) were exposed to gamma radiation (0–90 kGy) at a fixed dose rate.

Identification of aldehydes in frozen solutions of formic acid

Frozen solutions of formic acid were exposed to 42 kGy of irradiation. These samples were treated with DNPH and analysed by HPLC. A representative chromatogram of the identified aldehydes is shown in Fig. 2. Formaldehyde was detected in the irradiated samples of formic acid.

Fig. 2. HPLC chromatograms of (a) unirradiated formic acid treated with DNPH (control) and (b) DNPH-formaldehyde detected in frozen solutions of irradiated formic acid.

Decomposition of formic acid and identification of its radiolytic products in frozen solutions containing minerals

In both mineral systems (forsterite and graphite), the recovery of formic acid remained greater than 80% (Fig. 3). A similar observation was made regarding the frozen solutions that did not contain minerals (dotted line, Fig. 3), in which the recovery was 90%. No significant change in the decomposition of the molecule was observed in the systems, regardless of the presence of minerals. However, the distribution of the radiolytic products changed in the systems studied (Fig. 4).

Fig. 3. Recovery of formic acid after irradiation, quantified by titration. Frozen formic acid solutions (0.02 M) with and without minerals were exposed to gamma radiation (0–90 kGy) at a fixed dose rate.

Fig. 4. DNPH derivatives detected by HPLC in frozen formic acid solutions without and with minerals. The systems were irradiated at 42 kGy. (a) Glyoxylic acid, (b) unidentified carbonyl-containing molecule, (c) formaldehyde, (d) unidentified carbonyl-containing molecule and (e) acetaldehyde.

Figure 4 shows the radiolytic products identified in the frozen solutions of formic acid with and without minerals. The system containing graphite favours the formation of glyoxylic acid, formaldehyde and an unidentified carbonyl-containing molecule (D). In the system containing forsterite, glyoxylic acid, formaldehyde, acetaldehyde and two carbonyl-containing molecules (B and D) were identified. In the absence of minerals, only formaldehyde is detected.

Although B and D molecules have not been fully characterized, these compounds contain a carbonyl group in their structure because DNPH reacts specifically with the carbonyl groups of aldehydes or ketones (Shriner et al., Reference Shriner, Herman, Morril, Curtin and Fuson2004). Additionally, the irradiated samples with minerals were analysed by HPLC-MS. DNPH-derivatives with molecular masses of 363 and 325 g mol−1 were detected (data not shown), that could correspond to the unidentified molecules (B and D). The differences observed in the distribution of the radiolytic products are discussed in the subsequent sections.

Decomposition of formaldehyde in the control and frozen solutions

The recovery of formaldehyde after irradiation was quantified by gas chromatography. The measurements are presented in Fig. 5. In an interval of exposure of 0–300 kGy, a recovery of greater than 90% was observed, whereas between 400 and 600 kGy, the recovery remained above 80%.

Fig. 5. Recovery of formaldehyde from frozen solutions after irradiation (0–600 kGy). Gas chromatography was performed to quantify the amount of formaldehyde recovered.

Formic acid, a radiolytic product identified in the irradiated formaldehyde samples

In previous studies conducted by our research group (López-Islas et al., Reference López-Islas, Colín-García and Negrón-Mendoza2018), formic acid in irradiated formaldehyde samples was detected by mass spectrometry. In the present study, formic acid in the formaldehyde samples that were exposed to gamma radiation was measured by titration. Molarity calculations show low concentrations (0.007–0.002 M) of formic acid (Fig. 6). These data are consistent with the recovery of formaldehyde. The data presented in Fig. 5 show that the formaldehyde molecules remained stable at high doses of gamma radiation; therefore, the formation of radiolytic products was low.

Fig. 6. Quantification of formic acid in the irradiated formaldehyde samples by titration. The frozen solutions of formaldehyde were irradiated from 0 to 300 kGy.

Decomposition of formaldehyde in systems containing forsterite and the quantification of its radiolytic product: formic acid

According to the findings of the present study, the recovery of formaldehyde after irradiation was similar in the systems with and without forsterite. More than 90% of the formaldehyde was preserved in the samples exposed to 300 kGy of irradiation. Under these conditions, the mineral phase did not influence the formaldehyde decomposition, and these results agreed with the quantification of the formic acid (radiolytic product of formaldehyde irradiation). The highest concentration of formic acid observed in the formaldehyde–forsterite system was 0.002 M (Figs. 7 and 8).

Fig. 7. The recovery of formaldehyde in the frozen systems with and without forsterite after irradiation (0–300 kGy) quantified by gas chromatography.

Fig. 8. Formic acid quantification by titration of the frozen solutions of formaldehyde with and without forsterite.

Discussion

It has been proposed that comets delivered organic compounds to the early Earth (Chyba et al., Reference Chyba, Thomas, Brookshaw and Sagan1990). The organic content of comets is crucially important in prebiotic chemistry and the origin of life. The presence of important molecules, such as H2O, C2H6, C2H6, HCN, CO, CH3OH, H2CO, C2H2 and CH4, was confirmed on Comet 9P/Tempel 1 by the Deep Impact mission (Mumma et al., Reference Mumma, DiSanti, Magee-Sauer, Bonev, Villanueva, Kawakita, Dello Russo, Gibb, Blake, Lyke, Campbell, Aycock, Conrad and Hill2005). When the Giotto, Vega 1 and Vega 2 missions flew by the comae of the comets Halley, Hyakutake and Hale-Bopp in 1986 (Henbest, Reference Henbest1986; Hornung et al., Reference Hornung, Jessberger, Koch, Krüger, Langevin, Lehto, Lehto, Le Roy, Merouane, Modica, Orthous-Daunay, Paquette, Raulin, Rynö, Igenbergs, Jessberger, Krueger, Kuczera, Mcdonnell, Morfill, Rahe, Schwehm, Sekanina, Utterback, Völk and Zook1986; Fray et al., Reference Fray, Bardyn, Cottin, Altwegg, Baklouti, Briois, Colangeli, Engrand, Fischer, Glasmachers, Grün, Haerendel, Henkel, Höfner, Hornung, Jessberger, Koch, Krüger, Langevin, Lehto, Lehto, Le Roy, Merouane, Modica, Orthous-Daunay, Paquette, Raulin, Rynö, Schulz, Silén, Siljeström, Steiger, Stenzel, Stephan, Thirkell, Thomas, Torkar, Varmuza, Wanczek, Zaprudin, Kissel and Hilchenbach2016), they confirmed 30 organic and inorganic compounds, including formic acid and formaldehyde. As comets are maintained in the outer solar system, they have conserved the solar nebula composition. However, researchers have proposed that comets have been exposed to radiation for over 4.6 billion years from an internal source due to radionuclides embedded in the comet and from external sources of cosmic radiation (Donn, Reference Donn1977; Whipple, Reference Whipple and Delsemme1977; Draganić and Draganić, Reference Draganić and Draganić1984).

Comet irradiation

Significant radiation processing of comets is expected when they enter the solar system. Therefore, the molecules observed during the passage of comets near the Sun result from the irradiation of icy materials at low temperatures (Hudson and Moore, Reference Hudson and Moore1999), among other factors. Many of the products observed in the comae of comets are evidence of continuous transformation rather than direct sublimation of ice in the nuclei.

Considering their exposure to high-energy radiation, a radiation chemistry approach can be advantageous to study the behaviour of the molecules detected in comets. Cobalt-60 generates two high-energy emission lines (1.17 and 1.33 MeV) that are extremely penetrating. Numerous experiments have shown that radiation using the same or similar LET causes the same or similar radiation-induced chemical changes in a particular chemical system (Draganić and Draganić, Reference Draganić and Draganić1971). The cobalt-60 photons, which have a low LET of 0.2 keV μ−1, are a suitable tool for examining potential alterations brought on by high-energy cosmic ray protons, which have LET values comparable to the gammas of radioactive cobalt-60 at energies between 1 and 5 GeV. Thus, the average energy deposits along the radiation paths are comparable. The interaction with charged particles at GeV relativistic energies is undoubtedly far more complex than the interaction with gamma rays from cobalt-60, yet the average energy deposition per unit path length is comparable in both cases (Draganić and Draganić, Reference Draganić and Draganić1984).

Experimental simulations aid in studying the chemistry of the formation and destruction of molecules in a comet and predict the formation of more complex molecules based on those already detected. Due to the high penetration of gamma rays, the irradiation reaches the bulk of the solution (in the experimental model). Our focus was to study the behaviour of formic acid and formaldehyde under high-energy irradiation and in the presence of an inorganic compound (graphite or forsterite).

Water ice is the most abundant icy component in astrophysical environments and comets (Ehrenfreund et al., Reference Ehrenfreund, Irvine, Becker, Blank, Brucato, Colangeli, Derenne, Despois, Dutrey, Fraaije, Lazcano, Owen and Robert2002; Zheng et al., Reference Zheng, Jewitt and Kaiser2006). Energy deposition from the radiation occurs primarily in water ice, forming radicals that react with one or more compounds present in low concentrations (~0.1 M). In comet simulation experiments, ionizing radiation is an efficient promoter of chemical processes, potential synthesis and the subsequent chemical evolution of important prebiotic molecules (Spinks and Woods, Reference Spinks and Woods1990).

Frozen systems

As previously mentioned, ionizing radiation induces the radical formation, for example in the water molecule, as it is summarized in equation (A) (O'Donnell and Sangster, Reference O'Donnell and Sangster1970; Spinks and Woods, Reference Spinks and Woods1990).

(A)$${\rm H}_2{\rm O}\to \cdot {\rm H} + {{\bar{\rm e}}}_{{\rm eq}} + \cdot {\rm OH} + {\rm H}_2{\rm O} + {\rm H}_2$$

The radicals formed react with themselves and with a solute present in the medium. However, at low temperatures the radical production and diffusion rate is lower than in liquid systems (Sanner, Reference Sanner1965; Burton and Magee, Reference Burton and Magee1969; Spinks and Woods, Reference Spinks and Woods1990), this will yield less decomposition in a frozen system than in a liquid phase. To compare the rate of decomposition of formic acid and formaldehyde under irradiation in frozen solutions, the kinetic model proposed by Mincher and Curry (Reference Mincher and Curry2000), Criquet and Leitner (Reference Criquet and Leitner2011, Reference Criquet and Leitner2012) and Criquet et al. (Reference Criquet, Nebout and Leitner2010) was applied (Fig. 9).

Fig. 9. Reactivity of formic acid and formaldehyde in frozen solutions (77 K).

Although the model is purely descriptive and does not attempt to explain the entire physical and chemical processes of radiolysis, it can provide an understanding of the reactivity of the target compounds towards irradiation (Mincher and Curry, Reference Mincher and Curry2000; Criquet and Leitner, Reference Criquet and Leitner2012). A pseudo-first-order approximation is made at a constant dose rate:

$${\bf ln} C/C_{0} = -{\rm k}_ 0{\rm D},$$

where k 0 is the dose constant; C 0 is the initial concentration; and C is the concentration at the absorbed dose, D (kGy). When the trend is linear, the slope is the dose constant. For the systems examined in the present study (Fig. 9), the evaluated dose constants were 0.0002 kGy−1 for formaldehyde and 0.0004 kGy−1 for formic acid (at 77 K).

These data show that formic acid is more sensitive to irradiation (Fig. 1) than formaldehyde (Fig. 5) because of the different rate constants (Fig. 9). At 77 K, the dose constant for formaldehyde is ten times lower than at 298 K (0.0038 kGy−1), and the probability of survival in those conditions increases.

The role of minerals in frozen solutions

Mineral phases, such as olivine (forsterite) and solid carbon (as non-crystalline amorphous matter), are important components of comets (Messenger et al., Reference Messenger, Keller and Dante2005). The catalysis capacity of some minerals and their role in forming organic compounds have been explored in meteorites (Anders et al., Reference Anders, Hayatsu and Studier1973). However, limited data exist for comets. In the present study, two mineral phases, forsterite and graphite, were used to simulate the olivine and amorphous carbon detected on comets.

Frozen solutions of formic acid: Figures 2 and 4 show the chromatograms of the aldehydes and carbonyl-containing compounds detected in the irradiated frozen formic acid solutions under various experimental conditions. Our results suggest a differential distribution of aldehydes and carbonyl-containing compounds (Fig. 4, Table 1) that depends on the mineral phase used in the experiments.

Table 1. Carbonyl-containing molecules identified in frozen solutions of formic acid with and without minerals

The irradiated formic acid samples that did not contain minerals yielded an extremely limited number of compounds, and formaldehyde was the only aldehyde detected (Fig. 2). In equation (3) of Scheme 2, a potential mechanism for the formation of formaldehyde through abstraction reactions by radical COH (Scheme 2) is proposed.

Scheme 2. Proposed mechanisms for forming formaldehyde in frozen solutions of formic acid.

In contrast, the variety and number of radiolytic products increased in the presence of minerals. Graphite and forsterite promoted the synthesis of carbonyl-containing compounds in the formic acid solutions irradiated at low temperatures (77 K). In particular, the major number of carbonyl-containing compounds was detected in the systems that contained forsterite (Fig. 4, Table 1).

Although further studies are required to establish a mechanism for synthesizing these molecules, Li et al. (Reference Li, Dai, Liu, Sarre, Xie and Cheung2018) have proposed adsorption processes as a first step in the catalysis reactions of organic compounds. The authors used forsterite and methanol in their experiments as starting materials to synthesize PAHs. They observed the oxygen atom of methanol binds to the silicon atom of forsterite, which leads to the elongation of the O-H bond of methanol and its subsequent rupture, resulting in the formation of radicals (Li et al., Reference Li, Dai, Liu, Sarre, Xie and Cheung2018).

Although limited data exist regarding the role of carbonaceous material in the synthesis of prebiotic compounds, several aspects of its reactivity and structure are known. The properties of amorphous carbon suggest that the randomness of its structure and its lack of long-range order results in a higher proportion of chemically active sites (Duley and Williams, Reference Duley and Williams1981). Therefore, amorphous carbon is reactive concerning the sorption of atoms and molecules at these sites. This high reactivity promotes interaction with a wide variety of chemical species and potential catalytic reactions.

Frozen solutions of formaldehyde: As illustrated in Figs. 7 and 8, in the formaldehyde systems containing forsterite that were irradiated, no significant changes were observed in the decomposition of formaldehyde or the formation of its radiolytic product (formic acid). This was most likely driven by the low reactivity of formaldehyde at 77 K (Fig. 9) and not by the catalytic activity of forsterite.

Relevance of the radiolytic products of formic acid and formaldehyde in prebiotic chemistry

Aldehydes are of prebiotic interest because they are precursors to pre-biological molecules such as sugars and amino acids (Cleaves, Reference Cleaves2008). In the present study, formaldehyde, acetaldehyde and glyoxylic acid were detected in the formic acid samples after irradiation. These compounds are useful precursors to more complex molecules.

Under basic conditions, formaldehyde is the raw material for synthesizing sugars through the formose reaction (Cleaves, Reference Cleaves2008). Conversely, acetaldehyde has been reported as a precursor of acrolein, which is involved in synthesizing amino acids and alternative genetic materials (Cleaves, Reference Cleaves2002, Reference Cleaves2003). Similarly, glyoxylic acid participates in the synthesis of nucleobases (Menor-Salván and Marín-Yaseli, Reference Menor-Salván and Marín-Yaseli2013).

Conclusion

Comets are the prime candidates to link cosmic phenomena to the origin of life. They are a source of raw materials and water to initiate chemical evolution on early Earth. However, comets are surrounded by a radiative environment that may affect prebiotic molecule stability. The ionizing radiation is primarily generated by highly penetrating energetic particles, and the energy delivery is mainly produced on the surfaces of icy bodies up to several metres deep. Embedded radionuclides are the other source of ionizing radiation. Our investigation suggests that prebiotic molecules, such as formaldehyde, in comet simulations can withstand high radiation via two mechanisms: (a) a frozen system is a protective mechanism in degradative environments. Solid material, such as water ice, tends to modulate the energy of radiation. Thus, the radical production and diffusion rate decreases, as well as the decomposition of formaldehyde. (b) The separate irradiation of frozen solutions of formic acid and formaldehyde constitutes a redundant system, i.e. the irradiation of formic acid forms formaldehyde and vice versa. This system is capable of regenerating prebiotic molecules through interconversion.

Finally, the response of the minerals will depend on the type of organic molecule with which they interact. In the frozen solutions of formic acid that contained forsterite, aldehydes and other carbonyl-containing compounds were detected. In contrast, in the frozen solutions of formaldehyde, the presence of forsterite did not contribute to the formation of products.

Acknowledgments

This work was supported by PAPIIT grant IN114122 and CONACYT GRANT 319818. Technical support from C. Camargo, B. Leal-Acevedo, J. Rangel Gutiérrez, J. Gutierrez-Romero, M. Cruz, E. Palacios A. Meléndez, C. Fuentes and L. González is also acknowledged.

Conflict of interest

None.

References

Anders, E, Hayatsu, R and Studier, MH (1973) Organic compounds in meteorites. Science 182, 781790.CrossRefGoogle ScholarPubMed
Bardyn, A, Baklouti, D, Cottin, H, Fray, N, Briois, C, Paquette, J, Stenzel, O, Engrand, C, Fischer, H, Hornung, K, Isnard, R, Langevin, Y, Lehto, H, Le Roy, L, Ligier, N, Merouane, S, Modica, P, Orthous-Daunay, FR, Rynö, J, Schulz, R, Silén, J, Thirkell, L, Bastien, R, Bland, P, Bleuet, P, Borg, J and Zolensky, M (2006) Comet 81P/Wild 2 under a microscope. Science 314, 17111716.Google Scholar
Bockelée-Morvan, D, Crovisier, J, Mumma, MJ and Weaver, HA (2004) The Composition of Cometary Volatiles in Comets II. Tucson: University of Arizona Press, pp. 391423.Google Scholar
Brownlee, D, Tsou, P, Aléon, J, O'D Alexander, CM, Araki, T, Bajt, S and Baratta, GA (2006) Comet 81P/Wild 2 under a microscope. Science 314, 17111716.CrossRefGoogle ScholarPubMed
Burton, M and Magee, JL (1969) Advances in Radiation Chemistry. New York: Wiley-Interscience.Google Scholar
Cataldo, F, Ursini, O, Angelini, G, Iglesias-Groth, S and Manchado, A (2011) Radiolysis and radioracemization of 20 amino acids from the beginning of the solar system. Rendiconti Lincei. Scienze Fisiche e Naturali 22, 8194.CrossRefGoogle Scholar
Chyba, CF, Thomas, PJ, Brookshaw, L and Sagan, C (1990) Cometary delivery of organic molecules to the early earth. Science 249, 366373.CrossRefGoogle Scholar
Cleaves, HJ (2002) The reactions of nitrogen heterocycles with acrolein: scope and prebiotic significance. Astrobiology 2, 403415.Google ScholarPubMed
Cleaves, HJ (2003) The prebiotic synthesis of acrolein. Monatshefte für Chemie 134, 585593.CrossRefGoogle Scholar
Cleaves, HJ (2008) The prebiotic geochemistry of formaldehyde. Precambrian Research 164, 111118.CrossRefGoogle Scholar
Criquet, J and Leitner, NKV (2011) Radiolysis of acetic acid aqueous solutions-effect of pH and persulfate addition. Chemical Engineering Journal 174, 504509.CrossRefGoogle Scholar
Criquet, J and Leitner, NKV (2012) Electron beam irradiation of citric acid aqueous solutions containing persulfate. Separation and Purification Technology 88, 168173.CrossRefGoogle Scholar
Criquet, J, Nebout, P and Leitner, NKV (2010) Enhancement of carboxylic acid degradation with sulfate radical generated by persulfate activation. Water Science and Technology 61, 12211226.CrossRefGoogle ScholarPubMed
Cruz-Castañeda, J, Camargo, C, López-Islas, A and Negrón Mendoza, A (2018) Dosimetría de la Fuente de Irradiación Gamma ‘Gammabeam-651 Pt’. Informe Técnico Q-01-2018 Departamento de Química de Radiaciones y Radioquímica-UNAM, August 2018.Google Scholar
Delsemme, AH (1988) The chemistry of comets. Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences 325, 509523.Google Scholar
Despois, D and Cottin, H (2005) Comets: potential sources of prebiotic molecules for the early earth. In Gargaud, M, Barbier, B, Martin, H and Reisse, J (eds), Lectures in Astrobiology. Berlin, Heidelberg: Springer, pp. 289352.CrossRefGoogle Scholar
Donn, B (1977) Comparison of the composition of new and evolved comets. Proceedings of the International Astronomical Union 39, 1523.Google Scholar
Draganić, IG and Draganić, ZD (1971) Primary Products of Water Radiolysis: Oxidizing Species–the Hydroxyl Radical and Hydrogen Peroxide in the Radiation Chemistry of Water. Cambridge, Massachusetts: Academic Press.Google Scholar
Draganić, IG and Draganić, ZD (1984) Radiation chemical experiments relevant to studies of cometary nuclei: remarks on working conditions. Advances in Space Research 4, 115119.CrossRefGoogle ScholarPubMed
Draganić, IG and Draganić, ZD (1988) Radiation-chemical approaches to comets and interstellar dust. Journal de Chimie Physique 85, 5561.Google Scholar
Draganić, ZD, Vujosevic, SI, Negrón-Mendoza, A, Azamar, JA and Draganić, IG (1985) Radiation chemistry of a multicomponent aqueous system relevant to chemistry of cometary nuclei. Journal of Molecular Evolution 22, 175187.CrossRefGoogle ScholarPubMed
Duley, WW and Williams, DA (1981) The infrared spectrum of interstellar dust: surface functional groups on carbon. Monthly Notices of the Royal Astronomical Society 196, 269274.CrossRefGoogle Scholar
Ehrenfreund, P, Irvine, W, Becker, L, Blank, J, Brucato, JR, Colangeli, L, Derenne, S, Despois, D, Dutrey, A, Fraaije, H, Lazcano, A, Owen, T and Robert, F (2002) Astrophysical and astrochemical insights into the origin of life. Reports on Progress in Physics 65, 14271487.CrossRefGoogle Scholar
Fray, N, Bardyn, A, Cottin, H, Altwegg, K, Baklouti, D, Briois, C, Colangeli, L, Engrand, C, Fischer, H, Glasmachers, A, Grün, E, Haerendel, G, Henkel, H, Höfner, H, Hornung, K, Jessberger, EK, Koch, A, Krüger, H, Langevin, Y, Lehto, H, Lehto, K, Le Roy, L, Merouane, S, Modica, P, Orthous-Daunay, FR, Paquette, J, Raulin, F, Rynö, J, Schulz, R, Silén, J, Siljeström, S, Steiger, W, Stenzel, O, Stephan, T, Thirkell, L, Thomas, R, Torkar, K, Varmuza, K, Wanczek, KP, Zaprudin, B, Kissel, J and Hilchenbach, M (2016) High-molecular-weight organic matter in the particles of comet 67P/Churyumov-Gerasimenko. Nature 538, 7274.CrossRefGoogle ScholarPubMed
Goesmann, F, Rosenbauer, H, Bredehöft, JH, Cabane, M, Ehrenfreund, P, Gautier, , Giri, C, Krüger, H, Le Roy, L, MacDermott, AJ, McKenna-Lawlor, S, Meierhenrich, UJ, Muñoz Caro, G, Raulin, F, Roll, R, Steele, A, Steininger, H, Sternberg, R, Szopa, C, Thiemann, W and Ulamec, S (2015) Organic compounds on comet 67P/Churyumov-Gerasimenko revealed by COSAC mass spectrometry. Science 349, 13.CrossRefGoogle ScholarPubMed
Henbest, N (1986) Giotto and VEGA results on Halley's comet. Journal of the British Astronomical Association 96, 125127.Google Scholar
Hornung, K, Jessberger, EK, Koch, A, Krüger, H, Langevin, Y, Lehto, H, Lehto, K, Le Roy, L, Merouane, S, Modica, P, Orthous-Daunay, FR, Paquette, J, Raulin, F, Rynö, J, Igenbergs, EB, Jessberger, EK, Krueger, FR, Kuczera, H, Mcdonnell, JAM, Morfill, GM, Rahe, J, Schwehm, GH, Sekanina, Z, Utterback, NG, Völk, HJ and Zook, HA (1986) Composition of comet Halley dust particles from Giotto observations. Nature 321, 336337.Google Scholar
Hudson, RL and Moore, MH (1999) Laboratory studies of the formation of methanol and other organic molecules by water + carbon monoxide radiolysis: relevance to comets, icy satellites, and interstellar ices. Icarus 140, 451461.CrossRefGoogle Scholar
Iglesias-Groth, S, Cataldo, F, Ursini, O and Manchado, A (2011) Amino acids in comets and meteorites: stability under gamma radiation and preservation of the enantiomeric excess. MNRAS 410, 14471453.Google Scholar
Kissel, J, Brownlee, DE, Büchler, K, Clark, BC, Fechtig, H, Grün, E, Hornung, K, Igenbergs, EB, Jessberger, EKH, Krueger, K, McDonnell, JAM, Morfill, GM, Rahe, J, Schwehm, GH, Sekanina, Z, Utterback, NG, Volk, HJ and Zook, HA (1986a) Composition of comet Halley dust particles from Giotto observations. Nature 321, 336337.CrossRefGoogle Scholar
Kissel, J, Sagdeev, RZ, Bertaux, JL, Angarov, VN, Audouze, J, Blamont, JE, Büchler, K, Evlanov, EN, Fechtig, H, Fomenkova, MN, von Hoerner, H, Inogamov, NA, Khromov, VN, Knabe, W, Krueger, FR, Langevin, Y, Leonas, VB, Levasseurregourd, AC, Managadze, GG, Podkolzin, SN, Shapiro, VD, Tabaldyev, SR and Zubkov, BV (1986b) Composition of comet Halley dust particles from Vega observations. Nature 321, 280282.CrossRefGoogle Scholar
Kobayashi, K (2019) Prebiotic Synthesis of Bioorganic Compounds by Simulation Experiments in Astrobiology. Singapore: Springer Nature Singapore.Google Scholar
Leal-Acevedo, B and Gamboa-de Buen, I (2020) Dose distribution calculation with MCNP code in a research irradiator. Radiation Physics and Chemistry 167, 15.Google Scholar
Li, Q, Dai, W, Liu, BS, Sarre, PJ, Xie, MH and Cheung, AS (2018) Catalytic conversion of methanol to larger organic molecules over crystalline forsterite: laboratory study and astrophysical implications. Molecular Astrophysics 13, 2229.CrossRefGoogle Scholar
Llorca, J (2005) Organic matter in comets and cometary dust. International Journal of Microbiology 8, 512.Google ScholarPubMed
López-Islas, A, Colín-García, M and Negrón-Mendoza, A (2018) Stability of aqueous formaldehyde under γ irradiation: prebiotic relevance. International Journal of Astrobiology 18, 420425.CrossRefGoogle Scholar
Martins, Z, Watson, JS, Sephton, MA, Botta, O, Ehrenfreund, P and Gilmour, I (2006) Free dicarboxylic and aromatic acids in the carbonaceous chondrites Murchison and Orgueil. Meteoritics & Planetary Science 41, 10731080.CrossRefGoogle Scholar
Menor-Salván, C and Marín-Yaseli, MR (2013) A new route for the prebiotic synthesis of nucleobases and hydantoins in water/ice solutions involving the photochemistry of acetylene. Chemistry – A European Journal 19, 64886497.CrossRefGoogle ScholarPubMed
Messenger, S, Keller, LP and Dante, LS (2005) Supernova olivine from cometary dust. Science 309, 737741.CrossRefGoogle ScholarPubMed
Mincher, BJ and Curry, RD (2000) Considerations for choice of a kinetic fig. of merit in process radiation chemistry for waste treatment. Applied Radiation and Isotopes 52, 189193.CrossRefGoogle ScholarPubMed
Mumma, MJ, DiSanti, MA, Magee-Sauer, K, Bonev, BP, Villanueva, GL, Kawakita, H, Dello Russo, N, Gibb, EL, Blake, GB, Lyke, JE, Campbell, RD, Aycock, J, Conrad, A and Hill, GM (2005) Parent volatiles in Comet 9P/Tempel 1: before and after impact. Science 310, 270274.CrossRefGoogle ScholarPubMed
O'Donnell, JH and Sangster, DF (1970) Principles of Radiation Chemistry. Cambridge, Massachusetts: Elsevier Publishing Company.Google Scholar
Sanner, T (1965) Transfer of radiation energy to solute molecules in irradiated frozen aqueous solution. Radiation Research 25, 586600.CrossRefGoogle Scholar
Shriner, RL, Herman, CKF, Morril, TC, Curtin, DY and Fuson, RC (2004) The Systematic Identification of Organic Compounds. New York: John-Wiley and Sons.Google Scholar
Spinks, JWT and Woods, RJ (1990) An Introduction to Radiation Chemistry. New York: John-Wiley and Sons, Inc.Google Scholar
Whipple, FL (1977) The constitution of cometary nuclei. In Delsemme, AH (ed). Comets, Asteroids, Meteorites: Interrelations, Evolution and Origins. Toledo, Ohio: University of Toledo Press, pp. 2532.Google Scholar
Zheng, W, Jewitt, D and Kaiser, RI (2006) Formation of hydrogen, oxygen, and hydrogen peroxide in electron-irradiated crystalline water ice. The Astrophysical Journal 639, 534548.CrossRefGoogle Scholar
Figure 0

Scheme 1. Experimental sets: (1) frozen (77 K) solutions of formic acid (0.02 M) with and without minerals and (2) frozen (77 K) solutions of formaldehyde (0.3 M) with and without minerals.

Figure 1

Fig. 1. Recovery of formic acid after irradiation, quantified by titration. Frozen formic acid solutions (0.02 M) were exposed to gamma radiation (0–90 kGy) at a fixed dose rate.

Figure 2

Fig. 2. HPLC chromatograms of (a) unirradiated formic acid treated with DNPH (control) and (b) DNPH-formaldehyde detected in frozen solutions of irradiated formic acid.

Figure 3

Fig. 3. Recovery of formic acid after irradiation, quantified by titration. Frozen formic acid solutions (0.02 M) with and without minerals were exposed to gamma radiation (0–90 kGy) at a fixed dose rate.

Figure 4

Fig. 4. DNPH derivatives detected by HPLC in frozen formic acid solutions without and with minerals. The systems were irradiated at 42 kGy. (a) Glyoxylic acid, (b) unidentified carbonyl-containing molecule, (c) formaldehyde, (d) unidentified carbonyl-containing molecule and (e) acetaldehyde.

Figure 5

Fig. 5. Recovery of formaldehyde from frozen solutions after irradiation (0–600 kGy). Gas chromatography was performed to quantify the amount of formaldehyde recovered.

Figure 6

Fig. 6. Quantification of formic acid in the irradiated formaldehyde samples by titration. The frozen solutions of formaldehyde were irradiated from 0 to 300 kGy.

Figure 7

Fig. 7. The recovery of formaldehyde in the frozen systems with and without forsterite after irradiation (0–300 kGy) quantified by gas chromatography.

Figure 8

Fig. 8. Formic acid quantification by titration of the frozen solutions of formaldehyde with and without forsterite.

Figure 9

Fig. 9. Reactivity of formic acid and formaldehyde in frozen solutions (77 K).

Figure 10

Table 1. Carbonyl-containing molecules identified in frozen solutions of formic acid with and without minerals

Figure 11

Scheme 2. Proposed mechanisms for forming formaldehyde in frozen solutions of formic acid.