Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2021 Nov 3.
Published in final edited form as: Nat Chem. 2021 Oct 11;13(11):1126–1132. doi: 10.1038/s41557-021-00789-w

Prebiotic photoredox synthesis from carbon dioxide and sulfite

Ziwei Liu 1, Long-Fei Wu 1, Corinna L Kufner 2, Dimitar D Sasselov 2, Woodward W Fischer 3, John D Sutherland 1,*
PMCID: PMC7611910  EMSID: EMS133087  PMID: 34635812

Abstract

Carbon dioxide (CO2) is the major carbonaceous component of many planetary atmospheres, including the Earth throughout its history. Carbon fixation chemistry— that reduces CO2 to organics—utilizing hydrogen as stoichiometric reductant usually requires high pressures and temperatures, and yields of products of potential use to nascent biology are low. Here we demonstrate efficient ultraviolet photoredox chemistry between CO2 and sulfite that generates organics and sulfate. The chemistry is initiated by electron photodetachment from sulfite giving sulfite radicals and hydrated electrons, which reduce CO2 to its radical anion. A network of reactions—generating citrate, malate, succinate and tartrate by irradiation of glycolate in the presence of sulfite—was revealed. The simplicity of this carboxysulfitic chemistry and the widespread occurrence and abundance of its feedstocks suggest that it could have readily taken place on the surfaces of rocky planets. The availability of the carboxylate products on early Earth could have driven the development of central carbon metabolism before the advent of biological CO2 fixation.


Many CO2 reduction reactions have been discussed in the context of prebiotic chemistry, but all are problematic in that they require very special conditions and/or materials that are simply rare on planetary surfaces. For example, reduction by hydrogenation of bicarbonate (HCO3 ) over a Ni-Fe alloy under hydrothermal conditions1 requires high temperatures and pressures, and predominantly generates the C1 product methane, a poor feedstock for elaboration into (proto)biomolecules. By separating H2 and CO2 with a thin Fe(Ni)S precipitate barrier across which there is a large pH difference, milder conditions enable reduction, but the product formate (HCO2 ) is only produced in trace amounts2. Reduction of CO2 using metallic Fe powder in water generates acetate, methanol, formate and pyruvate – the latter only transiently – but the widespread occurrence of Fe powder on rocky planets such as early Earth or Mars is unlikely3. Finally, UV photoreduction of CO2 on colloidal ZnS semiconductor particles using hydrogen sulfide/hydrosulfide (H2S/HS) as a hole scavenger gives formate, acetate and propionate in low yield4, but these conditions are not likely to be common in a planetary context.

We previously demonstrated that hydrogen cyanide (HCN) can be reductively homologated using hydrated electrons (and/or hydrogen atoms derived therefrom by protonation) generated by UV irradiation of sulfidic anions in a process we termed cyanosulfidic chemistry56. For this chemistry, we originally used H2S/HS as stoichiometric reductant, but switched to using bisulfite7 (HSO3 , pKa ~7.2)/SO3 2– because sulfur dioxide (SO2) and H2S are outgassed in a ~10:1 or greater ratio on Earth89, and there is substantial evidence from the geological records of both Earth and Mars via the anomalous mass fractionation of sulfur isotopes that these sulfur species were important constituents of the early sulfur cycle1011. The Henry’s law constant for SO2 is greater than that of H2S and the first pKa of hydrated SO2 (~1.9) is far lower than that of H2S (~7.1)12, so dissolution and hydration of SO2 in surficial water followed by dissociation would therefore have been greater than dissolution and dissociation of H2S on early Earth and Mars. Based on reports that hydrated electrons generated by UV illuminating diamond surfaces reduce CO2 to carbon monoxide (CO) in acidic aqueous solution13, and the aforementioned semiconductor UV photoreduction of CO2, we now wondered if HSO3 /SO3 2– could serve as the source of hydrated electrons for CO2 reduction by UV photodetachment14 (Supplementary Table 1). Given that alkaline lakes can simultaneously absorb atmospheric CO2 and SO2 to give HCO3 and SO3 2– and a growing body of evidence that suggests that such lakes could have concentrated other prebiotically important species on early Earth and maybe Mars1516, we started to explore reduction chemistry at mildly alkaline pH.

Results

Photoredox CO2 fixation reaction

We subjected an aqueous solution of the sodium salts of HCO3 1 (50 mM) and SO3 2– (100 mM) at pH = 9 to UV irradiation from Hg-lamps with principal emission at 254 nm in a standard laboratory UV photoreactor and analyzed the resultant mixture by 1H-NMR spectroscopy, integrating signals relative to those of a subsequently added standard to quantitate products. After 4 hours irradiation, formate 2 (18 mM), hydroxymethanesulfonate 3 (200 μM), methanol 4 (200 μM), glycolate 5 (200 μM), acetate 6 (50 μM), tartronate 7 (600 μM), and malonate 8 (300 μM) had been produced alongside both rac- and meso-tartrate 9a (30 μM) and 9b (30 μM) (structures of products shown in Fig. 1, Supplementary Fig. 1). Sulfate was detected as a photoredox co-product14 by precipitation of barium sulfate upon addition of barium chloride under conditions where barium sulfite is soluble17. The bicarbonate-sulfite irradiation experiment was repeated using 13C-labelled HCO3 1 to confirm that all the products were generated from the photoreduction of CO2, and all product assignments were confirmed by spiking with authentic standards (Supplementary Fig. 1 and 2). Surprisingly, we were able to detect elemental hydrogen (H2) by 1H-NMR spectroscopy (δ = 4.5 ppm) if it was generated in situ by performing the irradiation experiment in a quartz NMR tube. This peak decreased/disappeared simply by shaking the NMR tube presumably because this accelerated degassing. The signal assignment for H2 was confirmed by running an NMR spectrum of the products of mixing zinc with hydrochloric acid solution in an NMR tube (Supplementary Fig. 3). Taken together, these results show that HCO3 1 is reductively converted to C2, C3 and (traces of) C4 compounds as well as being reduced to other C1 compounds in a process that also generates H2 and SO4 2–. If the initial concentration of HCO3 1 was reduced to 5 mM and the concentration of SO3 2– reduced to 10 mM, formate 2 (30 μM), glycolate 5 (20 μM), acetate 6 (10 μM), tartronate 7 (120 μM), and malonate 8 (30 μM) were observed by 1H-NMR spectroscopy after 4 hours irradiation. The combined yield of organics in these experiments exceeded 10% demonstrating the remarkably high efficiency of this chemistry compared to other potentially prebiotic CO2 fixation processes (Supplementary Fig. 4, Extended Data Fig. 1). The results were similar in experiments starting from carbon dioxide instead of sodium bicarbonate (Supplementary Fig. 5). In addition to the protiated products observed by 1H-NMR spectroscopy, oxalate 10 was observed by 13C-NMR spectroscopy in yields as high as 11% (Supplementary Fig. 6). At higher concentrations of reactants, the yield of C1 products, especially formate 2, went up relative to the yield of C>1 products and after prolonged irradiation, a new C3 product, β-hydroxypropionate 11 was identified (Supplementary Fig. 7).

Fig. 1. Carboxysulfitic photoredox reaction network starting from bicarbonate (HCO3 ) 1.

Fig. 1

Starting in the top left, the reaction network starts with the addition of hydrated electrons (produced by photodetachment from sulfite) to CO2 (blue) to give the carboxyl radical 12 after which point the network splits. Sequential reduction of the carboxyl radical 12 leads to the observed C1 products (green) whilst dimerization of 12 to oxalate 10 initiates a path to C>1 products (orange). Various reactions enable crossing between the C1 manifold and the C>1 manifold (the two manifolds are separated by a dashed line). The key oxidation of formate 2 back to the carboxyl radical 12 and the slow reduction of 2 that together divert flux from C1 to C>1 products are highlighted (fuchsia arrows). Photochemical reactions of oxalate 10, glyoxylate 18 and mesoxalate 23 (purple arrows) also contribute to the network.

Mechanistic study of the photoredox CO2 fixation reaction network

We next investigated the photoreaction of the various products and some putative intermediates in the presence of SO3 2– with a view to gaining information concerning the mechanism of the fixation chemistry. The results – summarized in Extended Data Fig. 2 (Supplementary Figs. 8 – 19) – can be rationalized by a reaction network based on photoredox radical chemistry (Fig. 1). Photodetachment of an electron from SO3 2– gives a hydrated electron and a sulfite radical (.SO3 )14. At pH 9, both loss of hydroxide from HCO3 1 and loss of water from its conjugate acid, H2CO3, furnish CO2. The latter process is efficiently catalyzed by nucleophiles1820, especially sulfite21, so it is unlikely that the otherwise slow kinetics of equilibration limit the photoredox chemistry. Although the equilibrium concentration of CO2 is very low in a solution containing HCO3 1 at pH = 9 relative to the concentration of 1 (Supplementary Fig. 20a)22, the rate constant for reaction of CO2 with hydrated electrons to give the carboxyl radical 12 is extremely high23 and the rate greatly exceeds the rate for protonation of hydrated electrons by 1 giving hydrogen atoms24. The carboxyl radical 12 can either be reduced by hydrogen atom transfer (HAT) from HSO3 , which has a ~1% abundance relative to SO3 2− at pH = 9, to give formate 2, or undergo dimerization to give oxalate 10, both directly and indirectly25. Focussing on the chemistry of formate 2 first, one electron reduction, though relatively slow26, gives the radical anion 13 and thence, through acid-base and hydration equilibria, the radicals 14 and 15 (although the latter is unfavoured relative to 13 and 14). The radicals 13 and 14 have two main fates; reduction by HAT from HSO3 , or recombination with the carboxyl radical 12. Coupled with acid-base and hydration equilibria, the first process (shown only for 13), generates formaldehyde 16 and its hydrate 17 and the second (shown only for 14) generates glyoxylate 18 via its hydrate 19. Formaldehyde 16, in equilibrium with the bisulfite adduct 3, can be reduced to the radical 20 which gives methanol 4 by HAT and glycolate 5 by recombination with the carboxyl radical 12 27. Another major reaction of formate 2 is oxidation back to carboxyl radicals 12 by reaction with sulfite radicals. This is inferred from the observation that irradiation of formate 2 and SO3 2– gives significant amounts of what appear to be products deriving from oxalate 10 in addition to C1 products (Extended Data Fig. 2).

The other initial product of the carboxyl radical 12 – its dimer oxalate 10 – can be reduced by addition of a hydrated electron to give the radical anion 21 28. This reduction is much faster than the corresponding reduction of formate 2 (Ultrafast pump-probe experiments and discussion are in Method). The radical anion 21 can undergo HAT leading to glyoxylate hydrate 19, or recombination with another carboxyl radical 12 to give mesoxalate hydrate 22, which equilibrates with mesoxalate 23 29. Reduction of mesoxalate 23 by addition of a hydrated electron, or electron transfer from a carboxyl radical 12, followed by HAT gives tartronate 7 and deoxygenation of the latter followed by HAT gives malonate 8. In the same multistep way that formate 2 can be reduced to methanol 4, reduction of one of the carboxylate groups of malonate 8 leads to β-hydroxypropionate 11. Reduction of glyoxylate 18 (in equilibrium with the hydrate 19 and a bisulfite adduct)30 and protonation of the initially formed radical anion31 leads to the key hydroxy-carboxymethyl radical 24 (pKa~8.8)32 which can recombine with the carboxyl radical 12 to give tartronate 7, dimerize to give the tartrates 9, or undergo HAT to give glycolate 5. Deoxygenation of glycolate 5 gives the carboxymethyl radical 25, which by recombination with the carboxyl radical 12 can give malonate 8 33 and by HAT, acetate 6.

Finally, we identified a number of photochemical steps other than the photodetachment of electrons from SO3 2−, which initiates the reaction network. Norrish type I reactions of glyoxylate 18 and mesoxalate 23 generate radicals 12, 15 and 26 (similar photocleavage of malonsemialdehyde 27, en route to β-hydroxypropionate 11, would generate radicals 15 and 25) and photodetachment of an electron from oxalate 10 3435 gives radical 28 which is thought to decarboxylate to the carboxyl radical 12. These additional photochemical steps set up futile cycles in the network, but also forge links from the C>1 parts of the network to the C1 part (Supplementary Figs. 21 – 26).

Based on the foregoing analysis, we thought that it might be possible to increase the amount of the C>1 products by adding sulfite portionwise – this would ensure that at any one time, the concentration of HSO3 would be low, so reaction flux through oxalate 10 would be favoured, but overall, there would be more reduction capacity. In accordance with expectation, at the end of this experiment, the combined yield of C>1 products (>25%) greatly exceeded C1 products (<1%) and the combined yield of malonate 8 (16.2%) and acetate 6 (1.0%) was greater than twice that of tartronate 7 (6.6%) and glycolate 5 (0.8%). Unexpectedly, a new product, sulfoacetate 29 (0.8%) was formed in low yield, presumably through recombination of carboxymethyl radicals 25 with sulfite radicals. (Supplementary Fig. 27, Extended Data Fig. 1, Entry 6). The general features of the time course of the CO2 reduction network were also revealed by this experiment. After 1 hour, formate 2 was the major product accompanied by traces of glycolate 5 and tartronate 7. After 2 hours, the amount of formate 2 had decreased and glycolate 5 and tartronate 7 were now the major products along with smaller amounts of acetate 6 and malonate 8. After further irradiation, the levels of formate 2, glycolate 5 and tartronate 7 stayed at about the same level and malonate 8 became the major product with minor amounts of acetate 6 and methanol 4. This time course behaviour can be understood from the reaction network (Fig. 1). At the outset of the experiment, the only carbonaceous species for the hydrated electrons to reduce is CO2. Carboxyl radicals 12 thereby produced apparently undergo HAT from HSO3 faster than they dimerise, and so formate 2 increases. However, the conversion of carboxyl radicals 12 to 2 is reversible, and so after some time a sufficient amount of oxalate 10 is produced for it to be reduced by the hydrated electrons as well. The reduction of oxalate 10 is much faster than the reduction of formate 2 (investigated by ultrafast pump-probe spectroscopy and further discussed in Methods), so 2 is consumed at the expense of making reduction products of 10. The appearance of the radical anion 21 opens up a new path for consumption of carboxyl radicals 12, including recombination to give mesoxalate hydrate 22 that is rapidly converted to tartronate 7 and a new path for the consumption of HSO3 , namely HAT to 21 giving glyoxylate 18 and thence, through rapid further reduction, glycolate 5. The opening of these new reaction paths reduces the level of formate 2 to a steady state where its consumption is balanced by continuous production from CO2 via carboxyl radicals 12. Eventually, the deoxygenation of tartronate 7 coupled to the slow reduction of malonate 8 means that the latter becomes the predominant product. At higher initial concentrations of sulfite, the early formate 2 pulse lasts longer and produces higher early amounts of 2, but eventually the paths to C>1 products start to operate and levels of formate 2 drop. Even if formate 2 is reduced, the reversibility of the downstream steps to C1 products and other paths from the C1 part of the network to the C>1 part means that products more complex than 2 eventually accumulate.

It has been reported that CO2 can be reduced to CO with hydrated electrons produced by UV irradiation of diamond in the presence of bisulfite as hole scavenger at pH = 3.213. The carboxyl radical 12 is formed as in our chemistry, but absent the chromophoric sulfite acting to absorb UV, it is photolyzed36 to CO and O.−, the conjugate base of the hydroxyl radical, which is then reduced by bisulfite. Hydrogen was not observed as a product because the reaction was carried out under a sufficiently high concentration of CO2 to out-compete H+ in reaction with electrons.

Glycolate photoredox reaction

As we investigated the photoreactions of the products and putative intermediates of the CO2 reduction network with SO3 2–, the photoredox chemistry of one product – glycolate 5 – stood out. Acetate 6 (16.2 mM), malonate 8 (0.1 mM), sulfoacetate 29 (7.8 mM), citrate 30 (0.2 mM), rac-tartrate 9a (1.5 mM), meso-tartrate 9b (1.0 mM), malate 31 (2.9 mM), succinate 32 (1.1 mM) and hydroxycitrate 33 (0.19 mM) along with C1 products were detected by 1H-NMR spectroscopy after 6 hours irradiation of glycolate 5 (50 mM) and SO3 2– (100 mM) (Supplementary Fig. 12, Extended Data Fig. 2, Entry 4). Particularly noteworthy is the fact that citrate 30, malate 31 and succinate 32 are key constituents of the Krebs cycle – a major cycle of central carbon metabolism, the consequences of which are discussed below. Remarkably, when the concentration of glycolate 5 was reduced to 5 mM and the concentration of SO3 2– reduced to 10 mM, after 2 hours irradiation, C1 products were no longer detected but the higher products were still formed in a comparable overall yield albeit with a different relative abundance distribution (Supplementary Fig. 13, Extended Data Fig. 2, Entry 5). The chemistry that generates acetate 6, malonate 8 and the tartrates 9 is the same as some of that of the CO2 fixation reaction network, but additional reactions now contribute to the detectable products (Fig. 2). Abstraction of a hydrogen atom from glycolate 5 by a sulfite radical generates the hydroxy-carboxymethyl radical 24 whilst redox compensatory reduction of 5 generates the carboxymethyl radical 25. Dimerization of 24 produces the tartrates 9 whereas dimerization of 25 produces succinate 32 as well as acetate 6 and glycolate 5 37. Recombination of radicals 24 and 25 provides one route to malate 31, a second would be from reduction of 9. Similar reduction of malate 31 would give a second path to succinate 32. Oxidation of the tartrates 9 and malate 31 to the corresponding hydroxyalkyl radicals 34 and 35 followed by recombination of these radicals with radicals 24 or 25 would give dihydroxycitrate 36, hydroxycitrate 33 and citrate 30. Reduction of 36 would constitute another reaction channel to hydroxycitrate 33 and further reduction of 33, another channel to citrate 30. In contrast to the reaction network starting from CO2 where all products are reduced relative to the starting material, the network starting from glycolate 5 is more subtle and contains both carbon oxidations and reductions. Thus malate 31 and citrate 30 are at the same oxidation level as glycolate 5, succinate 32 and acetate 6 are more reduced and the tartrates 9 and hydroxycitrate 33 are, on average, more oxidized.

Fig. 2. Carboxysulfitic photoredox reaction network starting from glycolate 5.

Fig. 2

Glycolate 5 (blue) can be both oxidized by sulfite radicals and reduced by hydrated electrons to give the radicals 24 and 25 which then react further to give the observed products (orange). Recombination of two C2 radicals (24 and/or 25) gives C4 products (tartrate 9, malate 31 and succinate 32). Tartrate 9 and malate 31 can be oxidized by sulfite radicals giving C4 radicals 34 and 35 and recombination of C2 and C4 radicals gives C6 compounds (including the products hydroxycitrate 33 and citrate 30). Reduction of dihydroxycitrate 36 and tartrate 9 provides additional reaction pathways to the C6 and other C4 products.

Photoredox reactions under lower photon flux

We also evaluated the bicarbonate reduction chemistry using a less intense broadband UV source, StarLab38 - an in-house constructed photoreactor designed to deliver UV radiation with a wavelength distribution representative of that from the Sun incident on the surface of early Earth, at a ~100 fold higher intensity than the Sun in a quiescent state and ~10 fold lower intensity than that during maximum flaring. After irradiation for 7 days in this apparatus, an aqueous solution of the sodium salts of HCO3 1 (5 mM) and SO3 2– (50 mM) at pH = 9 gave a mixture of protiated products similar to that obtained upon higher intensity irradiation in the standard laboratory photoreactor (at 254 nm for shorter time intervals) plus ethanol 37, confirming the utility of using 254 nm UV light to study this chemistry (Supplementary Fig. 28, Extended Data Fig. 1, Entry 7). Oxalate 10 was also detected in a similar experiment using 13C-labelled bicarbonate (Supplementary Fig. 29). Ethanol 37 could plausibly be obtained via dimerization of the hydroxymethyl radical 20 giving ethylene glycol 38, dehydration of 38 through radical 39 and the enoloxy radical 40 39 to acetaldehyde 41 and reduction (Fig. 1). Alternatively, acetate 6 could be reduced to acetaldehyde 41 and thence ethanol 37.

We then investigated the carboxysulfitic photoredox chemistry of glycolate 5 in the StarLab photoreactor. After 8 hours irradiation of glycolate 5 (50 mM) and SO3 2– (100 mM) with this less intense light source, acetate 6 (1.7 mM), sulfoacetate 29 (0.2 mM), rac-tartrate 9a (0.4 mM), meso-tartrate 9b (0.4 mM), malate 31 (0.2 mM), and succinate 32 (trace) along with C1 products were detected by 1H-NMR spectroscopy (Supplementary Fig. 14, Extended Data Fig. 2, Entry 6). Longer irradiation of more dilute samples of glycolate 5 (5 mM) and SO3 2– (50 mM) in the StarLab photoreactor resulted in higher yields of the same species and additionally produced malonate 8 and hydroxypropionate 11 (Supplementary Fig. 30, Extended Data Fig. 2, Entry 7).

Discussion

Planetary relevance

An important aspect of this chemistry is that the conditions and materials necessary to foster carboxysulfitic carbon fixation (short wave UV light, CO2 and SO2 derived from volcanism, and bodies of standing and flowing water on the crust) are mild, widespread, and expected to be common on rocky planets. Notably, there is geological evidence from the rock records of Earth and Mars that these conditions were met early in their history. Oxygen isotope ratios from Hadean zircons4041 and sedimentological observations from the earliest sedimentary record42 indicate abundant surface liquid water. Silicate weathering reactions occurred that sourced the alkalinity necessary to enable the dissociated hydrates, bicarbonate and sulfite, to partition from the atmosphere and accumulate in bodies of water in contact with the atmosphere43. Moreover, the anomalous fractionation of multiple sulfur isotopes in the early geological record10 provides a direct measure of SO2 photochemistry that establishes a valuable atmospheric correlate of the aqueous carbon fixation processes described herein. Finally, each of these observations for the early Earth that illustrates the plausibility of this chemistry occurring now has its complement in the Mars geological record11, 4447. Thus, the ingredients and basic conditions for carboxysulfitic chemistry to take place would have been present on both Earth and Mars. Depending on conditions, carboxylates such as formate 2, oxalate 10 or acetate 6 and malonate 8 are likely to have been the major initial products. Decarboxylation of malonate 8 to acetate 6 occurs on a short geological timescale in solution (~10 years at neutral pH and at 25°C)48 whereas oxalate 10 (in the absence of ferric ions and light)49, like acetate 6 is long-term stable and so it seems likely that these latter two products would have become the most abundant C>1 organics on early Earth had life not emerged – they might still be the most abundant organics on Mars if life did not emerge there.

The linkage between carboxysufitic chemistry and cyanosulfidic chemistry

The case for conditions conducive to cyanosulfidic chemistry being present on both young planets has also been made50. We note that for the full range of cyanosulfidic chemistry products to result, a scenario involving the mixing of bodies of water or flows (e.g. stream water) having subtly different reaction histories would probably be necessary. In locations where the basic conditions for cyanosulfidic chemistry were met, but the mixing of streams was absent or different, a limited set of products would have been generated and the first product of the restricted reaction network, glycolonitrile 42, would probably have been the most widespread. In addition, glycolonitrile 42 could have resulted from reaction of HCN with formaldehyde 16 rained in after production in the upper atmosphere by photoreduction of CO2 51. Hydrolysis of the nitrile group of glycolonitrile 42, however produced, generates glycolate 5, which could be converted by subsequent carboxysulfitic chemistry to the range of carboxylate products previously described. (Fig. 3) As the hydrolysis of glycolonitrile 42 generates ammonia in addition to glycolate 5, we also carried out the irradiation of 5 and sulfite in the presence of ammonia. Ammonia did not affect the outcome of the photoredox chemistry – the same set of products was formed with or without ammonia (Supplementary Fig. 31).

Fig. 3. Connections between environmental chemistry and the development of metabolism.

Fig. 3

Progression from heterotrophy fed by photochemical products of inorganic carbon reduction to autotrophy as the available products of environmental chemistry become less complex. The preformed building blocks of RNA, peptides and lipids produced by cyanosulfidic chemistry provision the origin and early evolution of life5,52, but gradually become depleted (fading of blue colour in timeline arrow) triggering the development of metabolism starting from simpler but more abundant products derived from glycolate 5 by cyanosulfidic chemistry (dark orange colour in timeline arrow). In turn, these materials become scarce (fading of orange colour in timeline arrow) and biology adapts to using carboxysulfitic products of CO2 and eventually, CO2 itself.

Biochemical relevance

Use of the products of cyanosulfidic chemistry as building blocks by nascent biology would eventually lead to their environmental depletion and biology would then be under evolutionary pressure to synthesize these building blocks from anything else that happened to be available and usable. Biology could either spread to encounter these materials in their place of synthesis, or fluvial advection could move them to the location of biology. It is fascinating that the majority of the carboxylate products deriving from the carboxysulfitic chemistry of glycolate 5 are key nodes of central carbon metabolism in extant biology and it seems likely that their synthesis by carboxysulfitic chemistry set the stage for the development of this metabolic network. At first glance, tartrate 9 seems to be somewhat an outlier, but its dehydration would lead through an enol to oxaloacetate53 and its oxidation, to dihydroxyfumarate which spontaneously decarboxylates to give glycolaldehyde54, a precursor of higher sugars.

With time, supply of most of the products of the carboxysulfitic chemistry of glycolate 5 would also dwindle and biology would have to evolve to make do with simpler, more abundant carbonaceous materials in the environment. The major long term stable products of the carboxysulfitic chemistry of CO2 – formate 2, acetate 6 and oxalate 10 – could then provision central carbon metabolism through development of a pyruvate-formate lyase activity and the glyoxylate shunt of the Krebs cycle via reduction of oxalate 10 to glyoxylate 18. Finally, even oxalate 10 and acetate 6 would become depleted and biology would be under evolutionary pressure to use the only remaining abundant carbon source, namely CO2.

Methods

General Methods

All reagents and deuterated solvents used for reactions and spiking experiments were purchased from Sigma-Aldrich and were used without further purification. All photochemical reactions were carried out in Norell Suprasil quartz NMR tubes purchased from Sigma-Aldrich using Hg lamps with principal emission at 254 nm in a Rayonet photochemical chamber reactor RPR-200, acquired from The Southern New England Ultraviolet Company. StarLab is an in-house constructed photoreactor that delivers broadband UV-Vis (from 220 nm to ~750 nm by using water as an optical filter) irradiation to a sample from a 75W xenon lamp manufactured by Horiba38. A Mettler Toledo SevenEasy pH Meter S20 was used to monitor the pH, and degassed H2O or D2O was achieved by four rounds of freeze-pump-thaw cycling. 1H-, and 13C-nuclear magnetic resonance (NMR) spectra were acquired using a Bruker Ultrashield 400 Plus or Bruker Ascend 400 operating at 400.1, and 100.6 MHz, respectively. Samples consisting of H2O/D2O mixtures were analyzed using HOD suppression to collect 1H-NMR data. Chemical shifts (δ) are shown in ppm. Coupling constants (J) are given in Hertz and the notations s, d, t represent the multiplicities singlet, doublet, and triplet. The conversion yields were determined by relative integrations of the signals using a known amount of acetamide as internal reference in the 1H-NMR spectrum.

General method of photoreaction of carboxylates with sulfite

Carboxylates and sodium sulfite (final concentrations were mentioned in Extended Data Fig. 1 and Extended Data Fig. 2) were dissolved in degassed H2O/D2O (9:1, 0.5 mL). After the pH was adjusted to the reported value with NaOH/HCl, the mixture was transferred to a quartz NMR tube which was sealed and irradiated for the reported time (Extended Data Fig. 1 and Extended Data Fig. 2). The resultant solution was analysed by 1H- and/or 13C-NMR spectroscopy. The yield was calculated by spiking with 4,5-dicyanoimidazole (final concentration of 0.5 mM, 1 mM or 5 mM) and relative integration.

Preparing hydrogen gas in an NMR tube

Metallic zinc (~6 mg) was added to 0.5 mL HCl (0.1 M) aqueous solution. This solution was transferred to an NMR tube after being vortexed for 5 seconds, and was then analysed by 1H-NMR spectroscopy.

Sulfate identification17

Sodium bicarbonate (21 mg, 0.25 mmol) and sodium sulfite (63 mg, 0.5 mmol) were dissolved in degassed water (10 mL) and the pH of the resultant solution was adjusted to 9 by adding NaOH/HCl solution. The mixture was then sealed in a quartz tube and irradiated with 254 nm light in the Rayonet photoreactor for 4 hours. 3 mL of the resulting solution was diluted to 20 mL with water and acidified to pH = 1 by the addition of concentrated HCl. The acidified solution was heated to nearly boiling for at least 30 min to remove all carbon dioxide and sulfur dioxide. Barium chloride solution was then added to the solution to give a precipitate which persisted upon boiling for another 30 min. The precipitate did not dissolve in dilute HCl solution.

Ultrafast pump-probe experiments

The general principles of pump-probe spectroscopy are described in the following references5557. The fundamental of the excitation pulses (800 nm) was generated by a Ti:Sa based laser-amplifier system (Solstice Ace by Spectra-Physics, Newport Co.) with a repetition rate of 1 kHz and a pulse duration of ~90 fs. The excitation pulses (251 nm) were generated in a nonlinear amplifier system (Topas Prime + NIRUVis, Light Conversion, Ltd.) and stretched by a 25 cm fused silica block (Corning) to ~ 1.7 ps to suppress two-photon ionization of the solvent. The excitation energy at the sample position was ~1 μJ and the spot diameter of ~250 μm (fwhm). For our microsecond ultrafast pump-probe spectroscopy (Supplementary Table 2) the broadband probe light (unpolarized) was generated, delayed, and detected in an EOS Fire system (Ultrafast Systems, LLC), with a nominal spectral range of 350 – 950 nm. For our picosecond ultrafast pump-probe spectroscopy (Supplementary Fig. 32) the broadband probe light was generated, delayed, and detected in a HELIOS Fire system (Ultrafast Systems, LLC), with a nominal spectral range of 400 – 750 nm. These spectral ranges are ideal for monitoring the broad absorption feature of the hydrated electron, which is centered near 700 nm. The experiments were carried out at a temperature of 23°C.

The transient pump-probe data were cropped to the spectral range 450 nm – 913 nm and 20 adjacent channels were averaged (Surface Xplorer, Ultrafast Systems, LLC). A global fitting analysis to determine transient lifetimes was performed5860.

Ultrafast pump-probe spectroscopic investigation of reduction chemistry

The rate constant reported in the literature for reaction of hydrated electrons with formate 2 (k = 2.4 x 104 M-1s-1)26 is considerably lower than that for reduction of oxalate (average of three values given in reference61 is 3.1 x 107 M-1s-1). However, in our experiments, oxalate 10 is also prone to photoionization3435 and so it was not clear if reaction of hydrated electrons with oxalate in our experiments is, or is not, faster than reaction of hydrated electrons with formate 2. The rate constant for reaction of hydrated electrons with glycolate 5 (k = 8.2 x 106 M-1s-1)32 is high, although the authors of this paper caution that this “unexpectedly high value may be due to trace impurity in the sample”. Furthermore, although the rate constant for reaction of solvated electrons with CO2 is known23, we do not know the concentration of CO2 in our experiments, or whether the catalysis of its interconversion with carbonic acid and bicarbonate by sulfite affects this rate. Accordingly, we used ultrafast pump-probe spectroscopy both to confirm the photoionization of oxalate 10 and compare it to that of sulfite (Supplementary Fig. 32A), and to measure hydrated electron decay kinetics in mixtures representative of the mixtures used in our continuous irradiation experiments (Supplementary Table 3). These pump-probe experiments confirmed the photoionization of oxalate 10 and further revealed that bicarbonate, formate 2 and glycolate 5 react at similar rates with hydrated electrons in our experiments, and that oxalate 10 reacts considerably faster (Supplementary Fig. 32 and Supplementary Tables 2 and 3).

Extended Data

Extended Data Fig. 1. Product concentrations and percentage yields after UV irradiation of solutions of NaHCO3 and Na2SO3.

Extended Data Fig. 1

Extended Data Fig. 2. Product concentrations and percentage yields after irradiation of individual bicarbonate reduction products (50 mM) and Na2SO3 (100 mM).

Extended Data Fig. 2

Supplementary Material

Supporting_information

Acknowledgements

The authors thank J.D.S., D.D.S. and W.W.F. group members for helpful discussions. This research was supported by the Medical Research Council (MC_UP_A024_1009 to J.D.S.), the Simons Foundation (290362 to J.D.S., 290360 to D.D.S. and 554187 to W.W.F.). C.L.K. and D.D.S. thank Wolfgang Zinth, Pablo Dominguez, Daniel Yahalomi, Gabriella Lozano for helpful discussions and experimental assistance, and acknowledge the Harvard Origins of Life Initiative.

Footnotes

Author contributions Z.L. discovered this carboxysulfitic chemistry and explored its scope under the supervision of J.D.S. and with the assistance of L.-F.W., C.L.K. performed the pump-probe experiments under the supervision of D.D.S., W.W.F. evaluated the geochemical relevance of the chemistry. All authors co-wrote the manuscript.

Competing Interests The authors declare no competing interests.

Data availability

All data generated or analysed during this study are included in the manuscript and the Supplementary Information.

References

  • 1.Horita J, Berndt ME. Abiogenic methane formation and isotopic fractionation under hydrothermal conditions. Science. 1999;285:1055–1057. doi: 10.1126/science.285.5430.1055. [DOI] [PubMed] [Google Scholar]
  • 2.Hudson R, et al. CO2reduction driven by a pH gradient. Proc Natl Acad Sci USA. 2020;117:22873–22879. doi: 10.1073/pnas.2002659117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Varma SJ, Muchowska KB, Chatelain P, Moran J. Native iron reduces CO2to intermediates and end-products of the acetyl-CoA pathway. Nature Ecol Evol. 2018;2:1019–1024. doi: 10.1038/s41559-018-0542-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Zhang XV, et al. Photodriven reduction and oxidation reactions on colloidal semiconductor particles: Implications for prebiotic synthesis. J Photochem Photobiol Chem. 2007;185:301–311. [Google Scholar]
  • 5.Patel BH, Percivalle C, Ritson DJ, Duffy CD, Sutherland JD. Common origins of RNA, protein and lipid precursors in a cyanosulfidic protometabolism. Nat Chem. 2015;7:301–307. doi: 10.1038/nchem.2202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Green NJ, Xu J, Sutherland JD. Illuminating life’s origins: UV photochemistry in abiotic synthesis of biomolecules. J Am Chem Soc. 2021;143:7219. doi: 10.1021/jacs.1c01839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Xu J, et al. Photochemical reductive homologation of hydrogen cyanide using sulfite and ferrocyanide. Chem Commun. 2018;54:5566–5569. doi: 10.1039/c8cc01499j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Zahnle K, Claire M, Catling D. The loss of mass-independent fractionation in sulfur due to a Palaeo-proterozoic collapse of atmospheric methane. Geobiology. 2006;4:271–283. [Google Scholar]
  • 9.Gerlach TM. Evaluation and restoration of the 1970 volcanic gas analyses from Mount Etna, SicilY J Volcanol. Geotherm Res. 1979;6:165–178. [Google Scholar]
  • 10.Farquhar J, Bao H, Thiemens M. Atmospheric influence of Earth’s earliest sulfur cycle. Science. 2000;289:756–758. doi: 10.1126/science.289.5480.756. [DOI] [PubMed] [Google Scholar]
  • 11.Farquhar J, Savarino J, Jackson T, Thiemens MH. Evidence of atmospheric sulphur in the martian regolith from sulphur isotopes in meteorites. Nature. 2000;404:50–52. doi: 10.1038/35003517. [DOI] [PubMed] [Google Scholar]
  • 12.Ranjan S, Todd ZR, Sutherland JD, Sasselov DD. Sulfidic anion concentrations on early earth for surficial origins-of-life chemistry. Astrobiology. 2018;18:1023–1040. doi: 10.1089/ast.2017.1770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Zhang L, Zhu D, Nathanson GM, Hamers RJ. Selective photoelectrochemical reduction of aqueous CO2 to CO by solvated electrons. Angew Chem Int Ed. 2014;126:9904–9908. doi: 10.1002/anie.201404328. [DOI] [PubMed] [Google Scholar]
  • 14.Fischer M, Warneck P. Photodecomposition and photooxidation of hydrogen sulfite in aqueous solution. J Phys Chem. 1996;100:15111–15117. [Google Scholar]
  • 15.Toner JD, Catling DC. A carbonate-rich lake solution to the phosphate problem of the origin of life. Proc Natl Acad Sci USA. 2020;117:883–888. doi: 10.1073/pnas.1916109117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Toner JD, Catling DC. Alkaline lake settings for concentrated prebiotic cyanide and the origin of life. Geochim Cosmochim Acta. 2019;260:124–132. [Google Scholar]
  • 17.Malati MA. Experimental inorganic/physical chemistry: an investigative, integrated approach to practical project work. Woodhead Publishing Limited; 1999. [Google Scholar]
  • 18.Murphy LJ, et al. A simple complex on the verge of breakdown: isolation of the elusive cyanoformate ion. Science. 2014;344(6179):75–78. doi: 10.1126/science.1250808. [DOI] [PubMed] [Google Scholar]
  • 19.Hering C, von Langermann J, Schulz A. The elusive cyanoformate: an unusual cyanide shuttle. Angew Chem Int Ed. 2014;53:8282–8284. doi: 10.1002/anie.201405339. [DOI] [PubMed] [Google Scholar]
  • 20.Juhl M, Petersen AR, Lee J-W. CO2-enabled cyanohydrin synthesis and facile iterative homologation reactions. Chem Eur J. 2021;27:228–232. doi: 10.1002/chem.202003623. [DOI] [PubMed] [Google Scholar]
  • 21.Roughton FJW, Booth VH. The catalytic effect of buffers on the reaction CO2 + H2O = H2CO3 . Biochem J. 1938;32:2049–2069. doi: 10.1042/bj0322049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Zeebe RE, Wolf-Gladrow D. CO2 in seawater: equilibrium, kinetics, isotopes. Vol. 65. Gulf Professional Publishing; 2001. [Google Scholar]
  • 23.Gordon S, Hart EJ, Matheson MS, Rabani J, Thomas JK. Reactions of the hydrated electron. Discussions Faraday Soc. 1963;36:193–205. [Google Scholar]
  • 24.Hentz RR, Farhataziz, Milner DJ, Burton M. γ-Radiolysis of liquids at high pressures. III. Aqueous solutions of sodium bicarbonate. J Chem Phys. 1967;47:374–377. [Google Scholar]
  • 25.Flyunt R, Schuchmann MN, von Sonntag C. A common carbanion intermediate in the recombination and proton-catalysed disproportionation of the carboxyl radical anion CO2 .− in aqueous solution. Chem Eur J. 2001;7:796–799. doi: 10.1002/1521-3765(20010216)7:4<796::aid-chem796>3.0.co;2-j. [DOI] [PubMed] [Google Scholar]
  • 26.Swallow AJ. Recent results from pulse radiolysis. Photochem Photobiol. 1968;7:683–694. [Google Scholar]
  • 27.Getoff N, Gütlbauer F, Schenck GO. Strahlenchemische carboxylierung von ameisensäure und methanol in wässriger lösung. Int J Appl Radiat Isot. 1966;17:341–349. doi: 10.1016/0020-708x(66)90126-8. [DOI] [PubMed] [Google Scholar]
  • 28.Getoff N, Schwörer F, Markovic VM, Sehested K, Nielsen SO. Pulse radiolysis of oxalic acid and oxalates. J Phys Chem. 1971;75:749–755. [Google Scholar]
  • 29.Doussin J-F, Monod A. Structure-activity relationship for the estimation of OH-oxidation rate constants of carbonyl compounds in the aqueous phase. Atmos Chem Phys. 2013;13:11625–11641. [Google Scholar]
  • 30.Olson TM, Hoffmann MR. Formation kinetics, mechanism and thermodynamics of glyoxylic acid–S(IV) adducts. J Phys Chem. 1988;92:4246–4253. [Google Scholar]
  • 31.Laroff GP, Fessenden RW. 13C Hyperfine interactions in radicals from some carboxylic acids. J Chem Phys. 1971;55:5000–5008. [Google Scholar]
  • 32.Bell JA, Grunwald E, Hayon E. Kinetics of deprotonation of organic free radicals in water Reaction of HOC.HCO2 −, HOC.HCONH2 and HOC.CH3CONH2 with various bases. J Am Chem Soc. 1975;97:2995–3000. [Google Scholar]
  • 33.Getoff N. CO2 and CO utilization: radiation-induced carboxylation of aqueous chloroacetic acid to malonic acid. Radiat Phys Chem. 2003;67:617–621. [Google Scholar]
  • 34.Arvis M, Lustig H, Hickel B. Étude par photolyse éclair de la photoionisation des anions formiate, acetate et oxalate dans l’eau. J Photochem. 1980;13:223–232. [Google Scholar]
  • 35.Huie RE, Clifton CL. Kinetics of the reaction of the sulfate radical with the oxalate anion. Int J Chem Kinetics. 1996;28:195–199. [Google Scholar]
  • 36.Habteyes T, Velarde L, Sanov A. Photodissociation of CO2 − in water clusters via Renner-Teller and conical interactions. J Chem Phys. 2007;126:154301. doi: 10.1063/1.2717932. [DOI] [PubMed] [Google Scholar]
  • 37.Wang W-F, Schuchmann MN, Schuchmann H-P, von Sonntag C. The importance of mesomerism in the termination of α-carboxymethyl radicals from aqueous malonic and acetic acids. Chem Eur J. 2001;7:791–795. doi: 10.1002/1521-3765(20010216)7:4<791::aid-chem791>3.0.co;2-2. [DOI] [PubMed] [Google Scholar]
  • 38.Rimmer P, et al. Timescales for prebiotic photochemistry under realistic surface UV conditions. Astrobiology Accepted. doi: 10.1089/ast.2020.2335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Gilbert BC, Larkin JP, Norman ROC. Electron spin resonance studies Part XXXIIi Evidence for heterolytic and homolytic transformations of radicals from 1,2-diols and related compounds. J Chem Soc, Perkin Trans. 1972;2:794–802. [Google Scholar]
  • 40.Wilde S, Valley J, Peck W, Graham CM. Evidence from detrital zircons for the existence of continental crust and oceans on the Earth 4.4 Gyr ago. Nature. 2001;409:175–178. doi: 10.1038/35051550. [DOI] [PubMed] [Google Scholar]
  • 41.Valley JW, et al. 4.4 billion years of crustal maturation: oxygen isotope ratios of magmatic zircon. Contrib Mineral Petrol. 2005;150:561–580. [Google Scholar]
  • 42.Rosing MT. 13C-Depleted carbon microparticles in >3700-Ma sea-floor sedimentary rocks from west Greenland. Science. 1999;283:674–676. doi: 10.1126/science.283.5402.674. [DOI] [PubMed] [Google Scholar]
  • 43.Kasting JF. The Goldilocks planet? How silicate weathering maintains Earth “just right”. Elements: An International Magazine of Mineralogy, Geochemistry, and Petrology. 2019;15:235–240. [Google Scholar]
  • 44.Grotzinger JP, et al. Deposition, exhumation, and paleoclimate of an ancient lake deposit, Gale crater, Mars. Science. 2015;350 doi: 10.1126/science.aac7575. [DOI] [PubMed] [Google Scholar]
  • 45.DiBiase RA, Limaye AB, Scheingross JS, Fischer WW, Lamb MP. Deltaic deposits at Aeolis Dorsa: sedimentary evidence for a standing body of water on the northern plains of Mars. J Geophys Res Planets. 2013;118:1285–1302. [Google Scholar]
  • 46.Hurowitz JA, et al. Redox stratification of an ancient lake in Gale Crater, Mars. Science. 2017;356 doi: 10.1126/science.aah6849. [DOI] [PubMed] [Google Scholar]
  • 47.Milliken RE, Fischer WW, Hurowitz JA. Missing salts on early Mars. Geophys Res Letts. 2009;36 doi: 10.1029/2009GL038558. [DOI] [Google Scholar]
  • 48.Wolfenden R, Lewis CA, Jr, Yuan Y. Kinetic challenges facing oxalate, malonate, acetoacetate and oxaloacetate decarboxylases. J Am Chem Soc. 2011;133:5683–5685. doi: 10.1021/ja111457h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Goldstein S, Rabani J. The ferrioxalate and iodide–iodate actinometers in the UV region. J Photochem Photobiol A. 2008;193:50–55. [Google Scholar]
  • 50.Sasselov DD, Grotzinger JP, Sutherland JD. The origin of life as a planetary phenomenon. Sci Adv. 2020;6 doi: 10.1126/sciadvaax3419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Cleaves HJ., II The prebiotic geochemistry of formaldehyde. Precambrian Res. 2008;164:111–118. [Google Scholar]
  • 52.Liu Z, Wu LF, Xu J, Bonfio C, Russell DA, Sutherland JD. Harnessing chemical energy for the activation and joining of prebiotic building blocks. Nat Chem. 2020;12:1023–1028. doi: 10.1038/s41557-020-00564-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Yew WS, et al. Evolution of enzymatic activities in the enolase superfamily: D-tartrate dehydratase from Bradyrhizobium japonicum . Biochemistry. 2006;45:14598–14608. doi: 10.1021/bi061688g. [DOI] [PubMed] [Google Scholar]
  • 54.Sagi VN, Punna V, Hu F, Meher G, Krishnamurthy R. Exploratory experiments on the chemistry of the “glyoxylate scenario”: formation of ketosugars from dihydroxyfumarate. J Am Chem Soc. 2012;134:3577–3589. doi: 10.1021/ja211383c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Schrader T, et al. Vibrational relaxation following ultrafast internal conversion: comparing IR and Raman probing. Chem Phys Lett. 2004;392:358–364. [Google Scholar]
  • 56.Ryseck G, et al. The excited‐state decay of 1‐methyl‐2(1H)‐pyrimidinone is an activated process. ChemPhysChem. 2011;12:1880–1888. doi: 10.1002/cphc.201001063. [DOI] [PubMed] [Google Scholar]
  • 57.Haiser K, et al. Mechanism of UV‐induced formation of Dewar lesions in DNA. Angew Chem Int Ed. 2012;51:408–411. doi: 10.1002/anie.201106231. [DOI] [PubMed] [Google Scholar]
  • 58.Satzger H, Zinth W. Visualization of transient absorption dynamics - towards a qualitative view of complex reaction kinetics. Chem Phys. 2003;295:287–295. [Google Scholar]
  • 59.Dominguez PN, Himmelstoss M, Michelmann J, Lehner FT, Gardiner AT, Cogdell RJ, Zinth W. Primary reactions in photosynthetic reaction centers of Rhodobacter sphaeroides-Time constants of the initial electron transfer. Chem Phy Lett. 2014;601:103–109. [Google Scholar]
  • 60.Gutierrez-Osuna R, Nagle HT, Schiffman SS. Transient response analysis of an electronic nose using multi-exponential models. Sens Actuators B Chem. 1999;61:170–182. [Google Scholar]
  • 61.Buxton GV, Greenstock CL, Helman P, Ross AB. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals in aqueous solution. J Phys Chem Reference Data. 1988;17:513–886. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting_information

Data Availability Statement

All data generated or analysed during this study are included in the manuscript and the Supplementary Information.

RESOURCES