Skip to main content
PLOS Biology logoLink to PLOS Biology
. 2022 Oct 4;20(10):e3001437. doi: 10.1371/journal.pbio.3001437

A prebiotic basis for ATP as the universal energy currency

Silvana Pinna 1, Cäcilia Kunz 1, Aaron Halpern 1, Stuart A Harrison 1, Sean F Jordan 1, John Ward 2, Finn Werner 3, Nick Lane 1,*
Editor: Matthew Pasek4
PMCID: PMC9531788  PMID: 36194581

Abstract

ATP is universally conserved as the principal energy currency in cells, driving metabolism through phosphorylation and condensation reactions. Such deep conservation suggests that ATP arose at an early stage of biochemical evolution. Yet purine synthesis requires 6 phosphorylation steps linked to ATP hydrolysis. This autocatalytic requirement for ATP to synthesize ATP implies the need for an earlier prebiotic ATP equivalent, which could drive protometabolism before purine synthesis. Why this early phosphorylating agent was replaced, and specifically with ATP rather than other nucleoside triphosphates, remains a mystery. Here, we show that the deep conservation of ATP might reflect its prebiotic chemistry in relation to another universally conserved intermediate, acetyl phosphate (AcP), which bridges between thioester and phosphate metabolism by linking acetyl CoA to the substrate-level phosphorylation of ADP. We confirm earlier results showing that AcP can phosphorylate ADP to ATP at nearly 20% yield in water in the presence of Fe3+ ions. We then show that Fe3+ and AcP are surprisingly favoured. A wide range of prebiotically relevant ions and minerals failed to catalyse ADP phosphorylation. From a panel of prebiotic phosphorylating agents, only AcP, and to a lesser extent carbamoyl phosphate, showed any significant phosphorylating potential. Critically, AcP did not phosphorylate any other nucleoside diphosphate. We use these data, reaction kinetics, and molecular dynamic simulations to infer a possible mechanism. Our findings might suggest that the reason ATP is universally conserved across life is that its formation is chemically favoured in aqueous solution under mild prebiotic conditions.


This study shows that acetyl phosphate is unique among a large panel of plausible prebiotic agents in phosphorylating ADP to ATP in water in the presence of ferric iron. Acetyl phosphate does not phosphorylate other nucleoside diphosphates, suggesting that ATP became established as the universal energy currency in a prebiotic, monomer world on the basis of its unusual chemistry in water.

Introduction

ATP is casually referred to as the “universal energy currency” of life. Why it gained this ascendency in metabolism, in place of many possible equivalents, is an abiding mystery in biology. There is nothing particularly special about the “high-energy” phosphoanhydride bonds in ATP. Rather, its ability to drive phosphorylation or condensation reactions reflects the extraordinary disequilibrium between ATP and ADP—about 10 orders of magnitude in modern cells, pushed by free energy derived from respiration [1]. ATP drives intermediary metabolism through the coupling of exergonic to endergonic reactions via phosphorylation and hydrolysis, but other phosphorylating agents could be pushed equally far from equilibrium and accomplish equivalent coupling.

A partial explanation is that ATP links energy metabolism with genetic information [2]. ATP-coupled monomer activation promotes the polymerisation of macromolecules, including RNA, DNA, and proteins. Unlike the simple phosphorylation of intermediary metabolites, the leaving group during nucleotide polymerization is pyrophosphate (PPi) [3]. Likewise, the activation of amino acids by adenylation liberates PPi as the leaving group [47]. The hydrolysis of PPi renders these steps strongly exothermic, if not practically irreversible—a ratchet towards polymerization [3,8,9]. Only nucleoside triphosphates can release PPi while still retaining a phosphate for the sugar-phosphate backbone of RNA and DNA, or for amino acid activation. But the fact that the canonical nucleosides can all form triphosphates, with equivalent free-energy profiles, only serves to reemphasise the prominence of ATP over GTP, TTP, UTP, or CTP in RNA, DNA, and protein synthesis. While GTP is not uncommon in metabolic processes, including ribosomal GTPases [10], it hardly displaces ATP from its preeminent position in biology.

In our view, the prominence of ATP is unlikely to reflect a frozen accident, as adenine nucleotide cofactors are also ubiquitous across intermediary metabolism, including the ancient cofactors NADH, FADH, and coenzyme A. The centrality of these cofactors to intermediary metabolism, combined with their ability to catalyse the same reactions in the absence of enzymes [1114], suggests that adenosine arose very early in biology, possibly even in a monomer world before the advent of RNA, DNA, and proteins [15,16]. This hypothesis is consistent with the idea that life acts as a guide to its own origin [17,18]. Phylogenetics indicates that the earliest cells grew autotrophically from H2 and CO2 [16,1923]. Recent experimental work shows that the core of autotrophic metabolism can occur spontaneously in the absence of genes and enzymes. This includes nonenzymatic equivalents of the acetyl CoA pathway and parts of the reverse Krebs cycle [2426], glycolysis and the pentose phosphate pathway [27,28], gluconeogenesis [29], and amino acid biosynthesis [3032]. Recent work demonstrates that some nucleobases can also be formed following the universally conserved biosynthetic pathways, using transition metal ions as catalysts [33]. The idea that ATP could have arisen as a product of protometabolism starting from H2 and CO2 is therefore not unreasonable (though we do not dispute that other mechanisms could potentially also give rise to purine nucleotides; [34,35]). Nonetheless, biological purine synthesis specifically involves 6 phosphorylation steps that are catalysed by ATP in modern cells—an autocatalytic feedback loop. If ATP was indeed formed in a monomer word via a biomimetic protometabolism, then an earlier ATP equivalent must have driven the phosphorylation steps in purine synthesis. Why this earlier phosphorylating agent was replaced, and specifically with ATP rather than other nucleoside triphosphates, might explain why ATP later rose to prominence in metabolism.

Arguably the most plausible ancestral mechanism of ATP synthesis is through the substrate-level phosphorylation of ADP to ATP by acetyl phosphate (AcP), which is still the fulcrum between thioester and phosphate metabolism in bacteria and archaea [36,37]. In modern bacteria, AcP is formed by the phosphorolysis of acetyl CoA; in archaea and eukaryotes, AcP remains bound to the active site of the enzyme but is still formed as a transient intermediate [37]. The notion that AcP played an important role at the origin of life goes back to Lipmann [38] and has been advocated by de Duve, Ferry and House, Martin and Russell, and others [23,36,37,3941]. While CoA itself is derived from ATP, simpler thioesters, with equivalent functional chemistry to acetyl CoA, have long been linked with prebiotic chemistry and the core metabolic networks in cells [36,38,4249]. Recent work suggests that thioesters such as methyl thioacetate can be synthesised under hydrothermal conditions [50]. AcP can also be made in water under ambient or mild hydrothermal conditions by phosphorolysis of thioacetate, which, as a thiocarboxylic acid, is even simpler than thioesters [15].

AcP will phosphorylate various nucleotide precursors in water, including ribose to ribose-5-phosphate, and adenosine to AMP [15]. Importantly, AcP can also phosphorylate ADP to ATP at approximately 20% yield in water in the presence of Fe3+ ions, suggesting that substrate-level phosphorylation could indeed take place in aqueous prebiotic conditions [51,52]. But whether this serendipitous discovery holds real relevance to protometabolism is uncertain, as other metal ions, phosphorylating agents and nucleoside diphosphates have not been tested under equivalent conditions. We have therefore explored the phosphorylation of ADP more systematically using a range of prebiotically plausible and biologically relevant phosphorylating agents, and a panel of metal ions as possible catalysts. We find that the combination of Fe3+ and AcP is unique: No other metal ions or phosphorylating agents are as effective at phosphorylating ADP. Equally striking, we find that ADP is also unique: The combination of AcP and Fe3+ will phosphorylate ADP but not GDP, CDP, UDP, or IDP. We use these data, reaction kinetics, and molecular dynamic (MD) simulations to propose a possible mechanism. Our results suggest that ATP became established as the universal energy currency in a prebiotic, monomeric world, on the basis of its unusual chemistry in water.

Results

Fe3+ is unique in promoting ADP phosphorylation by acetyl phosphate

We analysed a panel of metal ions commonly used as cofactors in metabolism, and likely available at the origin of life, to compare their effect on the phosphorylation of ADP by AcP. We first confirmed the results of Kitani and colleagues [51,52] in demonstrating that Fe3+ catalyses the formation of ATP by AcP at approximately 15% to 20% yield depending on the conditions (Fig 1A). We corroborated our HPLC results using MS/MS (Fig 1C) and established that Fe3+ was not catalysing the disproportionation reaction between 2 ADP molecules (Fig A in S1 Results). Surprisingly, we found that Fe3+ is uniquely effective at catalysing ADP phosphorylation, at least among the large panel of metal ions we tested. FeS clusters chelated by monomeric cysteine initially seemed to produce small yields of ATP, as shown in Fig 1B. However, Cys-FeS clusters are unstable and break down over hours except under strictly anoxic conditions [53]. We therefore suspected that the ATP yield actually reflected the release of Fe3+ into the medium. This was confirmed under more strictly anoxic conditions in an anaerobic glovebox, wherein FeS clusters failed to catalyse ATP formation (Fig B in S1 Results).

Fig 1. ATP synthesis with metal ion catalysts.

Fig 1

(a) HPLC trace of ATP control (0.7 mM) and ATP produced by the reaction ADP (1 mM) + AcP (4 mM) + Fe3+ (500 μM) at 30°C and pH approximately 5.5–6. (b) Test of reaction ADP (1 mM) + AcP (4 mM) at 30°C and pH approximately 5.5–6 with Fe3+ (Fe2(SO4)3), Mg2+ (MgCl2), Ca2+ (CaCl2), Mn2+ (Mn(NO3)2), Cr3+ (Cr(NO3)3), Mo3+ (MoCl3), Co3+ ([Co(NH₃)₆]Cl₃), Co2+ (CoCl2), CuSO4, Cu(NO3)2, FeS clusters (500 μM), and hematite (Fe₂O₃, 50 mg). The bars represent the ATP yield after 5 h. N = 3 ±SD. (c) Mass spectrometry analysis on a reaction sample at t = 0 h (upper panel) and 24 h (middle panel). The MS/MS spectrum and proposed structures of the products of the fragmentation of the ATP mass detected at 24 h (m/z = 506.19) is shown in the lower panel and was confronted to commercial standards and public data [137]. Conditions: ADP (1 mM) + AcP (4 mM) + Fe3+ (500 μM) at 30°C and pH approximately 5.5–6. The data underlying this figure can be found in Table A–C in S1 Data (sheet 1).

Metal ions that are commonly associated with ATP in metabolism, notably Mg2+ [54,55], failed to catalyse ATP formation either as free ions, or when coordinated by the monomeric amino acid aspartate, or in mineral form as brucite (Fig C in S1 Results). We had anticipated that chelated metal ions would show a stronger catalytic efficacy than free ions, as the coordination environment partially mimics the active site of enzymes, in this case acetate kinase or RNA polymerase (where glutamate or aspartate chelates Mg2+ at the active site). Brucite is a hydroxide mineral (Mg(OH)2) with a unit-cell structure that is also reminiscent of the Mg2+ coordination by the carboxylate of aspartate in the RNA polymerase. Surface catalysis may play an important role in prebiotic chemistry, but in this case failed to promote ATP synthesis. Mn2+, which has a similar activity to Mg2+ in acetate kinase [56] also failed to promote ATP synthesis.

ADP phosphorylation occurs in a range of aqueous prebiotic environments

We next explored the conditions under which Fe3+ catalyses the phosphorylation of ADP by AcP, specifically pH, temperature, water activity, and pressure. We found that the reaction is strongly sensitive to pH, and occurs most readily under mildly acidic conditions, with an optimum pH of approximately 5.5 to 6, the uncorrected default pH of the reaction (Fig 2A). Slightly more acidic conditions (pH 4) suppressed the yield a little, but more alkaline conditions had a much stronger suppressive effect. ATP yield fell by around three-quarters at pH 7 and collapsed to nearly zero at pH 9. This collapse of phosphorylation under alkaline conditions most likely reflected the precipitation of the catalyst as Fe(OH)3. While this sharp sensitivity to pH might seem at first sight limiting, in the Discussion, we show that, on the contrary, it could be valuable in generating disequilibria, enabling ATP hydrolysis to power work.

Fig 2. ATP synthesis by AcP and Fe3+ at different conditions.

Fig 2

(a) Effect of pH on reaction ADP (1 mM) + AcP (4 mM) + Fe3+ (500 μM) at 30°C. The optimal pH of the reaction is approximately 5.5–6. Rate of reaction: 0.0079 μM/s (optimal pH), 0.0074 μM/s (pH 4), and 0.0011 μM/s (pH 7). N = 3 ±SD. (b) Effect of temperature on reaction ADP (1 mM) + AcP (4 mM) + Fe3+ (500 μM), pH approximately 5.5–6. Rate of reaction: 0.0066 μM/s (20°C), 0.0079 μM/s (30°C), and 0.028 μM/s (50°C). N = 3 ±SD. (c) Comparison of ATP yield from the reaction ADP (1 mM) + AcP (4 mM) at 30°C, pH approximately 5.5–6 in water (reaction ionic strength = 3.75 mM), a modern ocean mix (600 mM NaCl, 50 mM MgCl2, and 10 mM CaCl2, reaction ionic strength = 783.75 mM), 300 mM NaCl (reaction ionic strength = 303.75 mM), modern ocean concentration of NaCl (600 mM, reaction ionic strength = 603.75 mM), 1 mM NaCl (reaction ionic strength = 1.004 M), dissolved silicate (10 mM SiO2, reaction ionic strength = 123.75 mM), and suspended silica in water (50 mg). N = 3 ±SD. (d) Comparison of ATP yield from the reaction ADP (1 mM) + AcP (4 mM) at 30°C and pH approximately 5.5–6 with and without Fe3+ (500 μM) at 80 bar (striped yellow) and at atmospheric pressure (1 bar, solid blue). N = 2 ±SD. The data underlying this figure can be found in Table D–G in S1 Data (sheet 2).

ATP yield was less acutely sensitive to temperature, at least between 20 and 50°C. Over 24 h, the overall ATP yield reflects both synthesis and hydrolysis. We found that 30°C optimised yield across 24 h, by promoting synthesis within the first 4 h while limiting hydrolysis over the subsequent 20 h (Fig 2B). The rate of synthesis was a little lower at 20°C, but this was offset by slightly less hydrolysis over 24 h. ATP synthesis was markedly faster at 50°C, but so too was hydrolysis, which already lowered yields within the first 2 h and cut them to less than a quarter of those at 30°C after 24 h. If ATP is to power work, as in modern cells, then hydrolysis in itself is not an issue, but rather needs to be coupled to other reactions such as the phosphorylation or condensation of substrates. Such processes also tend to take place over minutes to hours [15], meaning that temperature has a relatively trivial effect, with the yield after 2 to 3 h being similar at all 3 temperatures studied, at around 10% to 15% (Fig 2B). This implies that temperature would not be a strong limiting factor on many possible prebiotic environments.

More surprisingly, ATP yield was greatest at high water activity, either in HPLC-grade water or in suspended silica (Fig 2C). Adding NaCl lowered ATP yield, albeit not dramatically. Moderate NaCl concentration (300 mM, giving a total reaction ionic strength of 303.75 mM) lowered ATP yield by around a fifth. Modern ocean salinity (600 mM NaCl, reaction ionic strength 603.75 mM) and higher salinity (1 M NaCl, reaction ionic strength 1.004 M) both roughly halved the yield. This suggests that the effect of solutes does not only reflect ionic strength, which was confirmed by the addition of other solutes. Dissolved silicate (10 mM SiO2) also halved ATP yield, even though the ionic strength in this case was only 123.75 mM (Fig 2C). Likewise, higher Mg2+ and Ca2+ concentrations (50 mM and 10 mM, respectively) as part of a modern ocean mix collapsed ATP yields to nearly zero (Fig 2C), presumably because Ca2+ and Mg2+ promote ATP hydrolysis [57,58]. While this might suggest that ATP synthesis could not occur in modern oceans, Mg2+ and Ca2+ concentrations can in fact vary considerably in ocean environments (see Discussion). We show later that lower Mg2+ and Ca2+ concentrations (approximately 2 mM) actually promote ATP synthesis.

High pressure (80 bar) had very little effect on ATP synthesis (Fig 2D). This is consistent with the work of Leibrock and colleagues [59], who showed that high pressure promotes ATP hydrolysis, but only at pressures ≥300 bar. The slightly greater ATP yield at ambient pressure in our experiment may be attributable to greater evaporation in the open (nonpressurized) system. This was clearly the case in the absence of Fe3+, where most of the ATP detected was not produced by phosphorylation of ADP, but contamination of the ADP commercial standard via the manufacturing process, then concentrated by evaporation at ambient pressure (Fig D in S1 Results).

Acetyl phosphate is more effective than other prebiotic phosphorylating agents

We compared AcP with a panel of 6 other potentially prebiotic phosphorylating agents, including a number still used by cells today (Table 1).

Table 1. Phosphorylating agents tested.

Name ID Formula Prebiotic/biochemical prominence
Cyclic trimetaphosphate cTMP Na3P3O9 [55,6063]
Pyrophosphate PPi(V) K4P2O7 [64]
Pyrophosphite PPi(III) Na2H2P2O5 Has been detected in meteorites and can be generated from phosphite under hot acidic hydrothermal conditions; phosphate can be reduced to phosphite by serpentinization [6569]
Phosphoenolpyruvate PEP KC3H5O6P Has the highest phosphoryl-transfer potential found in living organisms (ΔGo´ = −62 kJ/mol) [70], and is an intermediate in gluconeogenesis and glycolysis, where its conversion to pyruvic acid by pyruvate kinase generates ATP via substrate-level phosphorylation
Carbamoyl phosphate CP Li2CH2NO5P·xH2O Can be made abiotically and has a role in extant biochemistry [71]
Trimethyl phosphate TMP (CH3)3PO4 Has been studied for its potential role in the nonenzymatic conversion of hypoxanthine to adenine [72]

Given the diverse reaction kinetics anticipated with these different phosphorylating agents, we carried out experiments at both at 30°C (the optimal temperature for AcP) and 50°C (as most phosphate donors are less labile than AcP and so might be more effective at higher temperatures), as well as pH 5.5 to 6, 7, and 9. As shown in Fig 3, no other phosphorylating agent was as effective as AcP at synthesising ATP in the presence of Fe3+. The only other phosphorylating agent to show any notable efficacy was carbamoyl phosphate (CP), which is similar in structure to AcP; it has a carbamate (-CO-NH2) rather than acetate (-CO-CH3) bound to phosphate. CP produced about half the ATP yield of AcP at 20°C and pH 5.5 to 6 (Fig 3A), but barely a quarter of the yield at pH 7 (Fig 3B). At pH 9, only cyclic trimetaphosphate (cTMP) produced any ATP at all, albeit after a delay of more than 20 h (Fig 3C).

Fig 3. ATP synthesis with different phosphorylating agents.

Fig 3

1 mM:4 mM ADP:phosphorylating agent reaction catalysed by Fe3+ with various phosphorylating agents at different pH and temperature. AcP, acetyl phosphate; CP, carbamoyl phosphate; cTMP, trimetaphosphate; PEP, phosphoenolpyruvate; PO4: potassium phosphate; PPi(III), pyrophosphite; PPi(V), pyrophosphate; TMP, trimethyl phosphate. N = 3 ±SD. (a) ATP yield (%) over 24 h at pH approximately 5.5–6 and 30°C; (b) ATP yield (%) over 24 h at pH 7 and 30°C; (c) ATP yield (%) over 24 h at pH 9 and 30°C; (d) ATP yield (%) over 24 h at pH approximately 5.5–6 and 50°C; (e) ATP yield (%) over 24 h at pH 7 and 50°C; (f) ATP yield (%) over 24 h at pH 9 and 50°C. The data underlying this figure can be found in Table H–M in S1 Data (sheet 3).

At 50°C, CP generated ATP continuously over 24 h at pH 5.5 to 6, despite producing only half the yield in the first 2 h (Fig 3D). The fact that ATP yield declined over time with AcP indicates that ATP was hydrolysed over hours at 50°C; it was not replenished because AcP also hydrolysed at that temperature [15]. While CP has a similarly low thermal stability, the primary decomposition product is cyanate [73], which is itself a proficient condensing agent [74,75]. This likely contributed to a balance between the synthesis and hydrolysis of ATP over 24 h. Only AcP formed any ATP at 50°C and pH 7 (Fig 3E), consistent with the pH sensitivity of CP seen at 30°C. The main conclusion here is that from a panel of 7 plausibly prebiotic phosphorylating agents, only AcP was capable of generating an ATP yield of >10% in water at both 30 and 50°C. The only other agent to show remotely comparable efficacy at mildly acidic pH was CP, but its maximal yield was half that of AcP.

Phosphorylation of ADP to ATP is unique among nucleoside diphosphates

We next explored the propensity of AcP to phosphorylate other canonical nucleoside diphosphates (NDPs), specifically cytidine diphosphate (CDP), guanosine diphosphate (GDP), uridine diphosphate (UDP), and inosine diphosphate (IDP). While not a canonical base, inosine is the precursor to both adenosine and guanosine in purine synthesis. Importantly, from a mechanistic point of view, inosine lacks the amino group incorporated at different positions onto the purine rings of adenosine and guanosine, but like GDP, IDP has an oxygen in place of the N6 amino group of adenosine. The results clearly show that AcP will phosphorylate ADP but not other NDPs (Fig 4A–4E), demonstrating a strong dependence on the structure of the nucleobase. For all NDPs, a peak for the corresponding triphosphate was present at the start of the reaction, but this did not change over 3 h for any NDP except ADP. As noted above for ADP, the presence of the NTP at 0 h can be ascribed to minor contamination of the commercial standard during the manufacturing process (more striking in the case of ATP). It was also striking that although AcP could also phosphorylate AMP to ADP (Fig 4F) the yields were much lower than the phosphorylation of ADP to ATP, suggesting that Fe3+ interacts most strongly with the beta phosphate of ADP, and much less strongly with the alpha phosphate.

Fig 4. Phosphorylation of NDPs by AcP.

Fig 4

HPLC chromatogram of the resulting NTP of the phosphorylation of (a) ADP, (b) IDP, (c) GDP, (d) CDP, and (e) UDP by AcP catalysed by Fe3+ at 30°C and pH approximately 5.5–6 at the beginning of the reaction (0 h, broken line, teal) and after 3 h (solid line, green). The molecular structure of each nucleobase forming the nucleotides is shown. (f) HPLC chromatogram showing the progressive ADP synthesis over 24 h via phosphorylation of AMP by AcP in the presence of Fe3+ at 30°C. The data underlying this figure can be found in Table N–S in S1 Data (sheet 4). AcP, acetyl phosphate; CDP, cytidine diphosphate; GDP, guanosine diphosphate; IDP, inosine diphosphate; NDP, nucleoside diphosphate; UDP, uridine diphosphate.

The fact that neither pyrimidine NDP could be phosphorylated suggests that the purine ring (or at least adenosine) is essential for positioning the interactions between Fe3+ and AcP. ADP has an amine group at N6, whereas GDP has a carbonyl at C6 and an amine group at N2; inosine has a carbonyl group at C6; and both GDP and IDP have a protonated N at N1. It is possible that the N6 amino group of adenosine interacts with the carboxylate oxygen on AcP (see below), but otherwise, it is difficult to identify a specific mechanism from these results, as the N7 nitrogen, which is known to interact with metal ions, is equivalent in all 3 purine rings [7679]. Nonetheless, the N6 amine group clearly influences electron delocalisation in purine rings and, therefore, the basicity of the other nitrogen moieties. We return to this in MD simulations below.

Catalysis of ADP phosphorylation does not involve nucleotide stacking

To understand how Fe3+ catalyses the phosphorylation of ADP to ATP, we tested the effect of varying the Fe3+ ion concentration. Holding the ADP and AcP concentrations constant at 1 mM and 4 mM, respectively, we varied the Fe3+ concentration from 0.05 to 2 mM. We found that the maximal ATP yield was produced by 1 mM Fe3+, indicating that the optimal ADP:Fe3+ stoichiometry of the reaction was 1:1 (Fig 5A). Following Kitani and colleagues [52], we confirmed that low concentrations of either Mg2+ or Ca2+ (up to 2 mM) slightly increased the ATP yield in the presence of 1 mM Fe3+. This suggests that either of these divalent cations can stabilise the newly formed ATP against hydrolysis and possibly liberate Fe3+ to catalyse the next phosphorylation of ADP (Fig 5A). We note that chelation of Fe3+ by ligands such as EDTA prevented the phosphorylation reaction altogether (Fig E in S1 Results), as did the use of clusters such as 4Fe4S clusters (Fig B in S1 Results) or hematite (Fig 1).

Fig 5. Mechanism studies.

Fig 5

(a) Effect of varying concentration of Fe3+ (teal circles) and adding increasing concentrations of Mg2+ (purple circles) and Ca2+ (green diamonds) on ATP yield at 2 h from the reaction ADP (1 mM) + AcP (4 mM) with 0.5 mM Fe3+ (solid line) and 1 mM Fe3+ (broken line) at 30°C and pH approximately 5.5–6. (N = 3 ±SD and 2 ±SD, respectively). (b) Michaelis–Menten kinetic analysis on the ADP + AcP reaction catalysed by Fe3+ (0.5 mM). N = 3 ±SD. (c) MALDI-ToF spectra of ADP control (top) and a reaction sample at 1 h (bottom). The data underlying this figure can be found in Table T and U and Fig A in S1 Data (sheet 5).

We next conducted a kinetic study of the phosphorylation reaction, specifically varying the ADP concentration and monitoring the reaction rate over the first 5 h of reaction (i.e., until the ATP yield starts to plateau, as in Fig 2), using the first-order rate equation r = Δ[ATP]/Δt. The resulting curve resembled a characteristic Michaelis–Menten mechanism for an enzyme, indicating that Fe3+ does indeed act as a catalyst (Fig 5B). The question remained whether a single Fe3+ was interacting directly with a single ADP and AcP, or whether larger units such as stacked ADP rings were involved. Stacking can alter the geometry of which group interacts with Fe3+ (Fig F in S1 Results) and has previously been suggested as a possible mechanism [80]. However, MALDI-ToF analysis, which can sensitively detect stacked nucleotides, showed no difference between the ADP control and the reaction sample; the main visible peaks appeared to be dimers of ADP/AMP present in the commercial ADP standard, possibly due to freeze-drying during production of ADP [81] (Fig 5C). While it is possible that this negative result reflects an issue with the MALDI-Tof analysis, we have previously detected stacks of AMP monomers using the same instrument and methodology [15]. It therefore seems likely that the difference in this case reflects the use of ADP rather than AMP.

The tentative conclusion that stacking of bases does not facilitate ADP phosphorylation was supported by MD simulations (Fig 6). These do not simulate chemical reactions but rather the forces acting between atoms and molecules under equivalent simulated experimental conditions. MD simulations can therefore predict the proportion of the simulation time in which specific atoms are oriented within plausible reaction distance [82]. For AcP to phosphorylate ADP to ATP, the phosphate of AcP needs to be positioned within a few Ångstroms of the beta phosphate oxygens of ADP. This is clearly seen in Fig 6A, where there is a strong peak representing a favourable conformation within 5 Å (and a secondary peak around 5 Å) in which the bond-forming groups are oriented within a feasible reaction space. This orientation was seen in the presence Fe3+ but was strongly subdued with Fe2+ or Mg2+ ions, indicating that only Fe3+ had strong enough interactions with phosphate to bring the 2 phosphate groups into close proximity for extended periods, corroborating our experimental findings. In this context, Fig 6B shows that a single ADP is clearly better than 2 (potentially stacked) ADPs in the same system. While the molar ratios have been kept consistent, the scope for interactions between 2 ADPs clearly suppressed the interaction time between the beta phosphate of any ADP and the AcP phosphate. This corroborates our MALDI-ToF results, as both sets of data indicate that the reaction between ADP and AcP does not occur between stacked dimers but rather between monomers. The bottom panels are consistent with these findings, showing that only Fe3+ ions spend a significant proportion of the simulation time positioned close to the ADP beta phosphate (Fig 6C) or the AcP phosphate (Fig 6D), thereby positioning them suitably close to react.

Fig 6. MD simulations.

Fig 6

(a). Proximity of AcP phosphate to ADP beta phosphate oxygens in the presence of either Fe3+, Fe2+, or Mg2+. System components: 2 ADP, 8 AcP, 8 Li+, 4 Na+, 12 K+, 18 Cl, and either 2 Fe3+, 2 Fe2+, or 2 Mg2+. ADP and AcP both have charges of 2−. (b). Proximity of AcP phosphate to ADP beta phosphate oxygens with either 1 or 2 ADPs with constant molecular and ionic ratios. The 2-ADP system is the same experiment as the ferric iron simulation from panel (a). The 1-ADP system uses the same molecule and ion ratios but with the total quantities halved. To maintain comparable concentrations, the periodic box size was reduced to 31.75 Å in the 1-ADP system. (c). Proximity of ADP beta phosphate to either Fe3+, Fe2+, or Mg2+ in the simulations from panel (a). (d). Proximity of AcP phosphate to either Fe3+, Fe2+, or Mg2+ in the simulations from panel (a). The data underlying this figure can be found in Table V–Y in S1 Data (sheet 6). AcP, acetyl phosphate; MD, molecular dynamic.

Our MD simulations also show striking differences between ADP and GDP in terms of the proportion of simulation time that Fe3+ interacts with N atoms in the purine ring (Fig G in S1 Results). With the exception of the N7 nitrogen in ADP (Fig Gc in S1 Results), which is known from experimental work to interact strongly with Fe3+ [7679], facilitating phosphorylation, the other N atoms in the guanine ring interact more strongly with Fe3+ than those in the adenine ring (Fig Ga, Gb, and Gd in S1 Results). This presumably has the effect of partially abstracting the catalyst, such that it is less available to interact with AcP or the beta phosphate on GDP than on ADP. Thus, MD simulations corroborate our experimental findings and show that Fe3+ ions can interact with the N7 on the ADP ring as well as the AcP phosphate and the ADP beta phosphate, forming a macro-chelate complex that facilitates ATP synthesis.

These findings suggest that the high charge density of Fe3+ allows it to interact directly with the N7 on the adenine ring, while anchoring AcP in position for its phosphate group to interact with the diphosphate tail of ADP, giving a taut conformation of ADP (Fig 7A). The interaction with the dianion has been proposed before [83,84] and is key because at the optimal pH of 5.5 to 6, the first 2 hydroxyl groups of ADP (pKa 0.9 and 2.8) are deprotonated, while the external OH group (pKa 6.8) remains protonated, and is therefore not available for nucleophilic attack [85]. The interaction of the 2 deprotonated OH groups with Fe3+ has the effect of lowering the pKa of the outermost OH group, thus deprotonating it and enhancing its nucleophilicity (Fig 6B). Possibly stabilised by interactions between the carboxylate oxygen of AcP and the N6 amino group on the adenosine ring (Fig 7B), the phosphate group of AcP is now readily positioned for nucleophilic attack by the newly deprotonated Oof ADP, forming ATP (Fig 7C). This mechanism might also help explain why Ca2+ and Mg2+ slightly increase the rate of reaction; these ions could displace Fe3+ from the ATP product (as they interact better with the triphosphate tail; Fig 7D) or alternatively protect newly formed ATP from hydrolysis after displacement of Fe3+. Given the approximation to Michaelis–Menten reaction kinetics, we assume that Fe3+ is displaced from the ATP, being freed to catalyse further rounds of ADP phosphorylation (Fig 7E).

Fig 7. Potential mechanism of ADP phosphorylation in water.

Fig 7

Fe3+, stabilised by the N7 group on adenine, interacts with the dianion of ADP, lowering the pKa of the outermost OH group and enhancing its nucleophilicity (a). Fe3+ interacts with the oxygens of a molecule of the surrounding AcP, bringing it close enough to facilitate the phosphate transfer (b). This interaction might be stabilised by further interactions between the N6 amino group and the carboxylate oxygen on AcP (b). Note that we only depict interactions between the Fe3+ and moieties on the ADP, adenosine ring, and AcP that have been established by our experiments and MD simulations, and this is not intended as a full depiction of the coordination sphere; for clarity, we do not include any interactions with water, which would certainly be part of the coordination sphere. Fe3+ is then likely to move from Pα to the Pβ and Pγ of ATP, although this is uncertain (c), and ultimately abandons the ATP chelated by acetate groups facilitated by the favourable association of Mg2+ as suggested by the reaction kinetics (d). Fe3+ is then available to catalyse another phosphorylation of ADP (e). AcP, acetyl phosphate; MD, molecular dynamic.

Discussion

Our results support the following conclusions: (i) AcP efficiently phosphorylates ADP to ATP, but only in the presence of Fe3+ ions as catalyst (Fig 1); (ii) the reaction takes place in water and can occur in a wide range of aqueous environments (Fig 2); (iii) no other phosphorylating agent tested was as effective as AcP (Fig 3); and (iv) adenine is unique among canonical nucleobases in facilitating the phosphorylation of its nucleoside diphosphate to the triphosphate (Fig 4). Taken together, these findings suggest that the preeminence of ATP in biology might have its roots in aqueous prebiotic chemistry. The substrate-level phosphorylation of ADP to ATP by AcP is uniquely facilitated in water under prebiotic conditions and remains the fulcrum between thioester and phosphate metabolism in bacteria and archaea today [2]. This implies that ATP could have become the universal energy currency of life not as the endpoint of genetic selection or as a frozen accident, but for fundamental chemical reasons, and arguably in a monomer world before the polymerization of RNA, DNA, and proteins.

The work presented here provides a compelling basis for each of these statements, but also raises a number of questions. Why ferric iron? Unlike AcP or ATP itself, there is no clear link with life in this case; we had expected other ions more commonly associated with nucleotides, notably Mg2+ or Ca2+ [54,55], to play a more clear-cut role. In fact, their catalytic effect was only noticeable in the presence of Fe3+, as previously reported [52], whereas higher concentrations, equivalent to modern ocean conditions, precluded ATP synthesis. We infer that the reason Fe3+ plays a unique role relates in part to its high charge density and small ionic radius. The fact that ADP is phosphorylated more readily than AMP (Fig 4) indicates that Fe3+ interacts with the diphosphate tail of ADP, which is also borne out by molecular dynamic simulations (Fig 6). The fact that the optimal stoichiometry of Fe3+ to ADP is 1:1, coupled with the absence of evidence for stacking of bases by MALDI-ToF (Fig 5), suggests that a single Fe3+ ion interacts with a single ADP, and necessarily also with a single AcP. Again, this interpretation is supported by MD simulations, which show that interactions between 2 ADP molecules suppress their interactions with AcP (Fig 6B). Interestingly, the MD simulations also showed that with the exception of N7, Fe3+ interacts more strongly with all the N atoms on the guanosine ring compared with those on the adenosine ring (Fig G in S1 Results). This finding is consistent with earlier work showing that the replacement of the N6 amine on the adenosine ring with the carbonyl group on the inosine and guanosine rings affects the properties of the other Ns on the purine ring, in particular their basicity [86]. Our results suggest that the stronger interactions between Fe3+ and the guanosine Ns likely abstract the catalyst, hindering phosphorylation reactions.

As shown in Fig 7, these stipulations require a taut molecular configuration of ADP, which is far from the loose conformation usually depicted, if only for ease of presentation. The orientation of the adenine ring in nucleotides has long been disputed, with some arguing that it should face the opposite way, with a more “rigid” (anti-gg) conformation [87]. Others have suggested an equivalent orientation to that proposed here [84,88], some specifically with Fe3+ [76,77]. Furthermore, a similar complex between ATP and Cu2+ has been reported to efficiently catalyse Diels–Alder reactions, thus confirming that such an orientation of the molecule is possible [89]. In any case, this taut conformation almost certainly requires the interacting ion to have a high charge density and small ionic radius, to draw each of these groups into close enough proximity to react. Among the cations tested here, Fe3+ has the highest charge density and the smallest ionic radius [90]. Nonetheless, some of the other ions studied, notably Cr3+ and Co3+, have a similar ionic radius and charge density, yet do not have a remotely comparable catalytic effect, so the size and charge density cannot be the only explanation for our results. The electronic configuration of Fe3+ may also play a role: Unlike Cr3+ and Co3+, Fe3+ has the electronic configuration [Ar]3d5, having all 5d orbitals half occupied. However, Mn2+, which can substitute Mg2+ in the catalytic centre of acetate kinase, has an equivalent 3d orbital, yet yielded negative results in our experiments. If so, then size, charge density, and electronic configuration might all play a role, potentially by stabilizing the phosphorylation transition state as a macro-chelate complex with ADP and AcP [86,91,92]. This hypothesis needs to be explored in future work. A related question, given our emphasis on life as a guide to protometabolism, is why Fe3+ is no longer used to catalyse ADP phosphorylation in biology. The most likely possibility is simply that Fe3+ always had limited availability (but see below) and undesirable reactivity, such as catalysing Fenton chemistry [93]. We note that chelated forms of Fe3+ do not catalyse ADP phosphorylation—neither FeS clusters (Fig B in S1 Results) nor EDTA-chelated Fe3+ (Fig E in S1 Results) generated any ATP at all. Nor did Fe3+ on surfaces such as hematite (Fig 1). For these reasons, enzymatic catalysis most likely displaced Fe3+ catalysis from later metabolism, especially given the high demand for ATP.

Why AcP? The idea that this small (2-carbon) molecule might have acted as a prebiotic phosphoryl donor has a long history, going back to Lipmann himself [15,36,38,4248], as indeed does its confounding potential as an acetyl donor [15,37,9497]. AcP still plays a global signalling and energy transduction role in bacteria [98], in part because its free energy of hydrolysis (and therefore its phosphorylating potential) is greater than that of ATP (ΔGo´ = −43 kJ mol−1 versus −31 kJ mol−1, respectively). When complexed in a 1:1 ratio with ADP, therefore, AcP has the potential to transfer its phosphate to form ATP, and so serves as a labile energy source in cells, linked to the excretion of acetate as waste. But the actual change in ΔG depends on how far from equilibrium the ratio of AcP/Ac + Pi or ATP/ADP + Pi has been pushed, and so varies depending on conditions. In our experiments, all phosphoryl donors were added at equivalent excess. The fact that the ΔGo´ for hydrolysis of PEP (−62 kJ mol−1) and CP (−51 kJ mol−1) are markedly greater than that for AcP means that free-energy change is only part of the explanation for the efficacy of AcP. As ATP was primarily formed by AcP in the presence of Fe3+ ions, the critical factors may instead have been (i) the position of the 2 phosphoester oxygen atoms in relation to the Fe3+; (ii) the phosphate group in relation to the diphosphate tail of ADP; and (iii) the carboxyl oxygen of AcP in relation to the N6 amine group of adenine (Fig 7). This latter point might also discriminate ADP from GDP. In other words, both AcP and ADP are favoured not for selective or thermodynamic reasons, but kinetic—because their chemistry is facilitated by molecular geometry in aqueous prebiotic environments.

The only other molecule with equivalent geometry in this regard is CP, which our model would therefore predict should have some phosphorylating efficacy. CP was indeed the only other species to show significant phosphorylating activity in our system (Fig 3). CP has long been considered a plausible prebiotic phosphorylating agent [71,99101], albeit possibly in the form of cyanate and Pi, which is in equilibrium with CP [102105]. CP can also promote the formation of ATP in the presence of Ca2+ or Ba2+ ions [71,106108]. Like AcP, CP retains a place in modern metabolism, for example, as a substrate for carbamate kinase, phosphorylating ADP to ATP in microbial fermentation of arginine, agmatine, and oxalurate/allantoin [109], as well as the de novo synthesis of pyrimidines (although not as a phosphorylator) [110]. Taken together with our own results, these findings suggest that both AcP and CP are molecular “living fossils” of prebiotic chemistry, retaining a role in modern metabolism due to their felicitous aqueous chemistry (Fig 3). Nonetheless, in general, CP does not perform phosphorylation reactions in extant life, preferring primary N-carbamoylation reactions [111]. The only other phosphorylating agent that showed any phosphorylating potential in our panel was cTMP, which is known to self-condense from diamidophosphate (DAP) in solution [112]. DAP has previously been shown to phosphorylate NDPs to the triphosphates [112]. However, we have not pursued this further as phosphoamidates have little relevance to extant biology. No N–P bonds exist in central metabolism common to archaea and bacteria, and creatine phosphate is restricted to eukaryotes, so is unlikely to reflect a biomimetic protometabolism.

Surprisingly, our results demonstrate that maximal ATP synthesis occurred at high water activity and low ion concentrations, indicating that prebiotic ATP synthesis would be most feasible in freshwater systems. Likewise, ferrous iron can be oxidized to ferric iron by photochemical reactions or oxidants such as NO derived from volcanic emissions, meteorite impacts, or lightning strikes, which also points to terrestrial geothermal systems as a plausible environment for aqueous ATP synthesis [113,114]. Conversely, high concentrations of Mg2+ (50 mM) and Ca2+ (10 mM) precluded ATP synthesis, implying that this chemistry would not be favoured in modern oceans. Nonetheless, our results do not exclude submarine hydrothermal systems as potential environments for this chemistry. Some shallow submarine systems such as Strytan in Iceland are sustained by meteoritic water and feature Na+ gradients as well as H+ gradients [115]. Such mixed systems could have been common in shallow Hadean oceans. Moreover, the concentration of divalent cations in the Hadean oceans may have been lower than modern oceans, with estimates varying widely [55,116]. Regardless of mean ocean concentrations, alkaline hydrothermal systems tend to precipitate Ca2+ and Mg2+ ions as aragonite and brucite, so their concentrations are typically much lower than mean ocean values. Modelling work in relation to Hadean systems indicates that hydrothermal concentrations of Ca2+ and Mg2+ would likely have been <1 mM [117,118], which is in the range that enhanced phosphorylation here. Other conditions considered here, including salinity and high pressure [59], would have only limited effects on ATP synthesis in submarine hydrothermal systems (which typically have pressures in the range of 100 to 300 Bars). Alkaline hydrothermal systems might also have generated Fe3+ in situ for ADP phosphorylation. Thermodynamic modelling shows that the mixing of alkaline hydrothermal fluids with seawater in submarine systems can promote continuous cycling between ferrous and ferric iron, potentially forming soluble hydrous ferric chlorides [118], which our experiments show have the same effect as ferric sulphate (Fig H in S1 Results). The availability of ferric iron is critical for other prebiotic catalysts including cysteine-FeS clusters [25,53,119,120] and has been discussed in more detail elsewhere [53].

A major question for prebiotic chemistry is how could an energy currency power work? As noted in the Introduction, there is nothing special about the bonds in ATP; rather, the ATP synthase powers a disequilibrium in the ratio of ADP to ATP, which amounts to 10 orders of magnitude from equilibrium in the cytosol of modern cells. Molecular engines such as the ATP synthase use ratchet-like mechanical mechanisms to convert environmental redox disequilibria into a highly skewed ratio of ADP to ATP [121]. But how could a simple prebiotic system composed mostly of monomers sustain a disequilibrium in ATP to ADP ratio that powers work? One possibility is that dynamic environments could sustain critical disequilibria across short distances such as protocell membranes. For example, alkaline hydrothermal systems sustain steep pH gradients across thin inorganic barriers, as mildly acidic Hadean ocean waters (pH 5 to 6) continually mix with strongly alkaline hydrothermal fluids (pH 9 to 11) in microporous labyrinths that operate as electrochemical flow reactors [20,47,122,123]. Protocells with mixed amphiphile membranes could bind to mineral barriers and potentially use these proton gradients to drive work, including ATP synthesis [124127] Mixed amphiphile membranes are highly permeable to protons, because fatty acid flip flop continuously transfers protons from the acid exterior to the alkaline interior of protocells [124,128,129]. A continuous influx of protons across protocell membranes could in principle promote ADP phosphorylation under locally acidic conditions in the immediate vicinity of the membrane within protocells. The ATP formed inside protocells would then be more vulnerable to hydrolysis linked to phosphorylation under more alkaline conditions in the bulk water of the cytosol. At face value, the ATP yield reported here at pH 5.5 to 6 after 10 h was 17.4% (corresponding to 156.5 μM) while the yield at pH 9 was 0.043%, corresponding to 0.4 μM, a difference of 400-fold. Thus, a geologically sustained difference in pH across membranes could drive a local disequilibrium in the ATP/ADP ratio of 2 to 3 orders of magnitude, enough to power work even in the absence of other possible factors such as temperature. Higher temperatures (50°C) promote both the rapid synthesis and hydrolysis of ATP (Fig 2B), which should amplify this driving force. We stress that these considerations require further elucidation, but in principle, steep pH gradients can drive a disequilibrium in the ATP/ADP ratio that could power work.

What would constitute “work” under these conditions? We are thinking in particular of the coupling of exergonic ATP hydrolysis to endergonic reactions in protometabolism, most pertinently the polymerization of nucleotides, amino acids, or both. ATP itself is relatively stable and hence is unlikely to adenylate or phosphorylate other molecules in the absence of suitable catalysts [130]. Which catalysts would best promote ATP hydrolysis coupled to phosphorylation in the aqueous prebiotic environments discussed here is a separate question that will be addressed elsewhere; metal ions, amino acids, short peptides, short ribozymes, or nucleotide cofactors all deserve consideration. We suspect that ATP only displaced earlier phosphorylating agents such as AcP or CP when coupled by catalysts to polymerization reactions linked to the hydrolysis of pyrophosphate, as discussed in the Introduction, or potentially as an autocatalytic feedback loop in purine synthesis, which could have amplified purine nucleotide availability in early evolution [18,129]. Nonetheless, the work presented here shows that ATP might have entered protometabolism at an earlier stage than generally supposed and could have contributed to the transition from a monomer to a polymer world.

In conclusion, AcP is unique among a panel of relevant phosphorylating agents in that it can phosphorylate ADP to ATP, in water, in the presence of Fe3+. AcP is formed readily through prebiotic chemistry and remains central to prokaryotic metabolism, making it the most plausible precursor to ATP as a biochemical phosphorylator [15]. Critically, AcP does not phosphorylate other nucleoside diphosphates, giving a compelling new insight into how ATP might have come to be so dominant in modern metabolism. Our findings indicate that the high charge density and electronic configuration of Fe3+ can position molecules in water to react in the absence of macromolecular catalysts such as RNA or proteins, or even mineral surfaces. Beyond that, our results suggest that steep pH gradients could in principle generate disequilibria in the ratio of ATP to ADP of several orders of magnitude, enabling ATP to drive work even in a prebiotic monomer world. Given suitable catalysts, ATP could eventually displace earlier phosphorylating agents and promote the polymerization of amino acids and nucleotides to form RNA, DNA, and proteins by liberation of pyrophosphate as a leaving group. If so, then ATP became established as the universal energy currency for reasons of prebiotic chemistry before the emergence of genetically encoded macromolecular engines.

Materials and methods

Materials

All salts were purchased from Sigma-Aldrich, except for copper nitrate hemipentahydrate (Cu(NO3)2·2.5H2O), copper sulphate pentahydrate (CuSO4·5H2O) and manganese nitrate hexahydrate (Mn(NO3)2·6H2O, Alfa Aesar), TEAA (triethylammonium acetate, Fluka), and CTP (Cytidine 5′-triphosphate sodium salt, Cambridge Bioscience). All solvents were HPLC-grade and purchased from Fischer. All reagents used were analytical grade (≥96%).

Reaction setup

Depending on the solubility of the analytes, reactions were carried out in either a stationary (SciQuip HP120-S) or a shaking (ThermoMix HM100-Pro) dry block heater.

For the reaction, stock solutions of di-nucleotides (sodium salts, ≥96%, Sigma-Aldrich), phosphorylating agents, and metal catalyst were freshly prepared as to avoid freeze-thawing (10 mM for reactions to be analysed via HPLC, 1 M for reactions to be analysed via NMR). Except where indicated, the ratios of analytes in a solution were 1(ADP):4(AcP) and 1(Fe3+):2(ADP). When needed, the pH was adjusted using aqueous HCL and NaOH (1 M or 3 M)

After checking the pH (Fisher Scientific accumet AE150 meter with VWR semi-micro pH electrode), samples were taken at time points (0, 10, and 30 min, 1 to 5, 10, and 24h) and, unless otherwise specified, immediately frozen at −80°C for next-day analysis.

Pressure reactor

Experiments under pressure were performed in a pressure vessel (Series 4600-1L-VGR with single inlet valve, Parr Instrument Company), pressurised with N2 gas and placed on a hotplate (Fisherbrand Isotemp Digital Stirring Hotplate) at 30°C. Samples for both the high pressure experiment and ambient pressure control experiment were prepared in 2 mL glass headspace vials (Agilent Technologies) whose caps were pierced with a needle.

FeS clusters

FeS clusters coordinated by 5 mM of L-cysteine were prepared under anaerobic conditions, and water sparged with N2 was used to prepare all solutions. Stock solutions of 10 mM Na2S, 10 mM FeCl3, 50 mM of L-cysteine, and 1 M of NaOH were prepared either in water or in 10 mM bicarbonate buffer (pH 9.1). A volume of 4 mL of Na2S and 4 mL of L-cys were added to 28 mL of water/buffer, and the pH adjusted to approximately 9.8 using NaOH. A volume of 4 mL of FeCl3 was then added and the volume adjusted to 40 mL to obtain a 1-mM FeS solution.

Oxygen levels in the anaerobic glovebox were maintained below 5 ppm when possible, and no work was conducted if this level was surpassed.

UV/Vis spectroscopy

UV/Vis spectroscopy was used to verify the formation of FeS clusters. A volume of 1 mL of FeS stock solution was placed in a crystal cuvette, which was sealed with parafilm under anaerobic conditions. Spectra were obtained using a Thermofisher NanoDrop 2000c, with a baseline correction of 800 nm.

Analysis

HPLC

Samples were prepared at collection by spinning at 4,000 rpm for 2 min and diluting 200 μL in 800 μL of EDTA solution (500 μL in 100 mM PO4 buffer at pH 7.1) prior to freezing, in order to chelate the Fe3+ ions in solution that would otherwise block the HPLC column.

Thawed samples were filtered using syringe filters (ANP1322, 0.22 μm PTFE Syringe filter, Gilson Scientific) attached to a 1-mL sterile syringe (BD Plastipak Syringes) in 2 mL headspace vials and analysed on an HPLC instrument (Agilent Technologies, 1260 Infinity II); peaks were identified using pure standards. The wavelengths for UV detection were usually set at 254 nm and 260 nm (most suitable for cyclic rings such as adenosine), while the column tray temperature was maintained at room temperature. Two different columns were used depending on the pH of the sample being analysed: Poroshell 120 EC-C18 for pH 2 to 8 and Poroshell HPH-C18 for pH 9 to 11.

Mobile phase A consisted of 80 mM phosphate buffer (made by mixing equal parts of potassium phosphate dibasic (40 mM) and potassium phosphate monobasic (40 mM) salts dissolved in water) adjusted to pH 5.8 using 3 M HCl and filtered with 0.2 μm nylon membrane filters (GNWP04700, 0.2 μm pore size, Merck Millipore), while mobile phase B consisted of 100% methanol. The injection volume was 1 μL, with a flow rate of 1 mL/min, and the run was an isocratic gradient that consisted of 95% B for 5 min.

For experiments using nucleoside diphosphates with different bases, analyses were carried out on a Polaris C18-A column, with mobile phase A consisting of 10 mM potassium phosphate monobasic buffer with 10 mM tetrabutylammonium hydroxide (TBAH) adjusted to pH 8 using 3 M HCl and filtered with 0.2 μm nylon membrane filters (GNWP04700, 0.2 μm pore size, Merck Millipore), while mobile phase B consisted of 100% methanol (method described in Table 2). The wavelengths for UV detection were set at 254, 260, and 271 nm for guanosine, uridine and inosine, and cytidine, respectively.

Table 2. HPLC method for G, C, I, and U nucleotides experiments.
Mobile phase A 10 mM KH2PO4 + 10 mM TBAH in HPLC-grade water
Mobile phase B 100% HPLC-grade methanol
Gradient 5% B → 50% B (up during 25 min) → 50% B (for 2 min) → 95% B (up during 6 s) → 95% B (for 3 min) → 5% B (down during 6 s) → 5% B (for 2 min)
Flow rate 1.5 mL/min
Injection volume 1 μL

Two flush methods (Table 3) were employed to preserve the column: Flush 1 was used every 12 to 15 samples, then 3 rounds of Flush 1 followed by 1 run of Flush 2 were run prior to switching off the machine.

Table 3. HPLC flush methods.

These are run at the end of a set number of sample analyses.

FLUSH 1 FLUSH 2
Mobile phase A HPLC-grade water HPLC-grade water
Mobile phase B 100% HPLC-grade methanol 100% HPLC-grade methanol
Gradient 5% → 95% B (up during 15 min) → 95% B (for 5 min) → 5% B (down during 10 min) Initial: 5% B (for 17 min) → 95% B (up during 18 min) → 95% B (for 17 min) → 60% B (down during 6 min) → 60% B (for 17 min)
Flow rate 1 mL/min 1 mL/min

Computational analysis was done using Agilent OpenLAB software (ChemStation Edition). Each peak was manually integrated using the calibration curves as reference and the raw file was exported for data manipulation. As residual ATP is present in the ADP commercial standard, the yield of the reaction is calculated by subtracting the reading for ATP at time point 0 from all subsequent time point readings. Except where noted, the rate of the reaction was calculated over the first hour of reaction (to enable direct comparison between 30°C and 50°C) using the first-order rate equation r = Δ[ATP]/Δt.

ESI MS

Electrospray ionisation mass spectrometry was used to confirm the identity of ATP through MS/MS. After purification through SPE (see previous section), the sample was loaded into a 0.5-mL glass syringe (Gastight Syringe Model 1750 RN, Hamilton) and directly infused into the mass spectrometer (Finnigan LTQ Linear Ion Trap mass spectrometer) at a flow rate of 10 μL/min. To avoid contaminations, the syringe and line were flushed with 100% methanol before and after sample infusion, and the spectra recorded.

The mass spectrometer was operated in negative ion mode, and the capillary voltage was set at −16 V. Data were collected from 100 to 2,000 m/z with an acquisition rate of 5 spectra per second. For the MS/MS, Ar was used as the collision gas and the collision energy was adjusted to 30 eV. The software Xcalibur (Thermo Scientific) was used for method setup and data processing.

MALDI-ToF MS

Samples were thawed and desalted using a protocol adapted from Burcar and colleagues [131]. Two solvents were prepared: an ACN solution consisting of 50% acetonitrile in water and a 0.1-M TEAA solution in water.

Using a Millipore C18 zip tip (Sigma), 10 μL of ACN solution were aspirated and discarded 3 times. The 3 rinses were repeated with 10 μL of the TEAA solution. To allow for the retention of the analyte by the zip tip matrix, 10 μL of sample were aspirated up and down 8 times and then discarded. A volume of 10 μL of water were aspirated and discarded, followed by 10 μL of the TEAA solution and once again 10 μL of water. A volume of 4 μL of ACN were slowly aspirated up and down 3 times and deposited into a small Eppendorf microcentrifuge tube.

The MALDI-ToF protocol used was designed by Whicher and colleagues [15]. The matrix consisted of 2,4,6-trihydroxyacetophenone monohydrate (THAP) and ammonium citrate dibasic and was freshly prepared before the analysis using equal volumes of stocks that were maintained at 4°C for a maximum of a week.

A volume of 2 μL of matrix solution was mixed with 2 μL of sample, deposited onto a clean steel MALDI-ToF plate and allowed to evaporate for 30 min before the introduction of the steel plate into the instrument (Waters micro MX mass spectrometer). The analytical conditions were as follows: reflectron and negative ion mode, 400 au of laser power, 2,000 V of pulse, 2,500 V of the detector, 12,000 V of flight tube, 5,200 V of reflector, 3,738 V of negative anode, and 500 to 5,000 amu of scan range. The mass spectrometer was calibrated using a low-molecular weight oligonucleotide standard (comprising of a DNA 4-mer, 5-mer, 7-mer, 9- mer, and 11-mer (Bruker Daltonics)). Each oligonucleotide standard was initially dissolved in 100 μL water, divided in aliquots and frozen at −80°C. A fresh aliquot was used at each analytical calibration.

Molecular dynamics simulations

Simulation inputs were prepared using CHARMM-GUI’s ligand modeller and multicomponent assembler [132,133]. The simulation parameters utilized a 40-Å periodic box, 25°C NVT ensemble, CHARM-GUI’s automatic PME FFT grid information generation, Monte Carlo ion placement, and the CHARMM36m forcefield [134]. The system components were 2 ADP, 8 AcP, 8 Li+, 4 Na+, 12 K+, 18 Cl, and either 2 Fe3+, 2 Fe2+, or 2 Mg2+. ADP and AcP both have protonation states of −2, but as noted previously, the proximity of Fe3+ lowers the pKa of the outermost OH group, resulting in its deprotonation and increasing its nucleophilicity. We did not explicitly model this in the MD simulations as this would require quantum simulations of the reactions themselves. For the 1-ADP experiment, the same molecule and ion ratios were used but with the total quantities halved. To maintain comparable concentrations in this simulation, the periodic box size was reduced to 31.75 Å. Simulations were run using NAMD 2 [135] on UCL’s myriad cluster. Simulations were minimized and equilibrated for CHARMM-GUI’s default number of timesteps, 10,000 and 125,000, respectively (500 timesteps = 1 ps), before being run for 48 h using 2 MPI cores, 200 times in parallel. Each parallel repeat had randomized initial velocities with constant starting positions. This produced simulations on approximately 1.5 microsecond timescales. The data from the simulation trajectories were extracted using the MDToolbox package for MATLAB [136] and analysed in MATLAB R2020b.

Supporting information

S1 Results. Document reporting all Supporting information figures.

(DOCX)

S1 Data. Excel document detailing raw data for all analyses.

(XLSX)

Abbreviations

AcP

acetyl phosphate

CDP

cytidine diphosphate

CP

carbamoyl phosphate

cTMP

cyclic trimetaphosphate

DAP

diamidophosphate

GDP

guanosine diphosphate

IDP

inosine diphosphate

MD

molecular dynamic

PEP

phosphoenolpyruvate

PPi(III)

pyrophosphite

PPi(V)

pyrophosphate

NDP

nucleoside diphosphate

TMP

trimethyl phosphate

UDP

uridine diphosphate

Data Availability

All relevant data are within the paper and its Supporting Information files.

Funding Statement

We are grateful to the Biotechnology and Biological Sciences Research Council to NL, FW and JW (BB/V003542/1; https://www.ukri.org/councils/bbsrc/) and HR (LIDo Doctoral Training Programme; https://www.lido-dtp.ac.uk/), to Gates Ventures (formerly bgc3) to NL, and to the Natural Environment Research Council to AH and NL (2236041; https://www.ukri.org/councils/nerc/). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Nicholls DG, Ferguson S. Bioenergetics. 4th edition. Academic Press; 2013. [Google Scholar]
  • 2.de Duve C. Singularities: Landmarks on the Pathways of Life. New York, USA: Cambridge University Press; 2005. [Google Scholar]
  • 3.Kottur J, Nair DT. Pyrophosphate hydrolysis is an intrinsic and critical step of the DNA synthesis reaction. Nucleic Acids Res. 2018;46:5875–5885. doi: 10.1093/nar/gky402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Berg P, Ofengand EJ. An Enzymatic Mechanism for Linking Amino Acids to RNA. Proc Natl Acad Sci. 1958;44:78–86. doi: 10.1073/pnas.44.2.78 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Ofengand EJ, Dieckmann M, Berg P. The enzymic synthesis of amino acyl derivatives of ribonucleic acid. J Biol Chem. 1961;236:1741–1747. doi: 10.1038/181303a0 [DOI] [PubMed] [Google Scholar]
  • 6.Ibba M, Söll D. Aminoacyl-tRNA Synthesis. Annu Rev Biochem. 2000;69:617–650. doi: 10.1146/annurev.biochem.69.1.617 [DOI] [PubMed] [Google Scholar]
  • 7.do Nascimento Vieira A, Kleinermanns K, Martin WF, Preiner M. The ambivalent role of water at the origins of life. FEBS Lett. 2020;594:2717–2733. doi: 10.1002/1873-3468.13815 [DOI] [PubMed] [Google Scholar]
  • 8.Herdewijn P, Marlière P. Redesigning the leaving group in nucleic acid polymerization. FEBS Letters. 2012;586 (5):2049–2056. doi: 10.1016/j.febslet.2012.02.033 [DOI] [PubMed] [Google Scholar]
  • 9.Wimmer JLE, Kleinermanns K, Martin WF. Pyrophosphate and Irreversibility in Evolution, or why PPi Is Not an Energy Currency and why Nature Chose Triphosphates. Front. Microbiol. 2021;12:759359. doi: 10.3389/fmicb.2021.759359 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Voorhees RM, Schmeing TM, Kelley AC, Ramakrishnan V. The mechanism for activation of GTP hydrolysis on the ribosome. Science. 2010;330 (6005):835–838. doi: 10.1126/science.1194460 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Raffaelli N. Nicotinamide Coenzyme Synthesis: A Case of Ribonucleotide Emergence or a Byproduct of the RNA World? In: Egel R., Lankenau DH., Mulkidjanian A. editors. Origins of Life: The Primal Self-Organization. Berlin, Heidelberg: Springer; 2011. doi: 10.1007/978-3-642-21625-1_9 [DOI] [Google Scholar]
  • 12.Wimmer JLE, Xavier JC, Vieira AdN, Pereira DPH, Leidner J, Sousa FL, et al. Energy at Origins: Favorable Thermodynamics of Biosynthetic Reactions in the Last Universal Common Ancestor (LUCA). Front Microbiol. 2021;12:793664. doi: 10.3389/fmicb.2021.793664 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Wimmer JLE, Vieira AN, Xavier JC, Kleinermanns K, Martin WF, Preiner M. The Autotrophic Core: An Ancient Network of 404 Reactions Converts H2, CO2, and NH3 into Amino Acids, Bases, and Cofactors. Microorganisms. 2021;9 (2):458. doi: 10.3390/microorganisms9020458 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Preiner M, Xavier JC, Vieira AN, Kleinermanns K, Allen JF, Martin WF. Catalysts, autocatalysis and the origin of metabolism. Interface Focus. 2019;9:20190072. doi: 10.1098/rsfs.2019.0072 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Whicher A, Camprubí E, Pinna S, Herschy B, Lane N. Acetyl Phosphate as a Primordial Energy Currency at the Origin of Life. Orig Life Evol Biosph. 2018;48:159–179. doi: 10.1007/s11084-018-9555-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Copley SD, Smith E, Morowitz HJ. The origin of the RNA world: Co-evolution of genes and metabolism. Bioorg Chem. 2007;35:430–443. doi: 10.1016/j.bioorg.2007.08.001 [DOI] [PubMed] [Google Scholar]
  • 17.Harrison SA, Lane N. Life as a guide to prebiotic nucleotide synthesis. Nat Commun. 2018;9:5176. doi: 10.1038/s41467-018-07220-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Harrison S, Nunes Palmeira R, Halpern A, Lane N. A biophysical basis for the emergence of the genetic code in protocells. BBA Bioenergetics. 2022;1863:148597. doi: 10.1016/j.bbabio.2022.148597 [DOI] [PubMed] [Google Scholar]
  • 19.Wächtershäuser G. Evolution of the first metabolic cycles. Proc Natl Acad Sci. 1990;87:200–204. doi: 10.1073/pnas.87.1.200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Russell MJ, Hall AJ. The emergence of life from iron monosulphide bubbles at a submarine hydrothermal redox and pH front. J Geol Soc London. 1997;154:377–402. doi: 10.1144/gsjgs.154.3.0377 [DOI] [PubMed] [Google Scholar]
  • 21.Smith E, Morowitz HJ. Universality in intermediary metabolism. Proc Natl Acad Sci. 2004;101:13168–13173. doi: 10.1073/pnas.0404922101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Russell MJ, Martin WF. The rocky roots of the acetyl-CoA pathway. Trends Biochem Sci. 2004;29:358–363. doi: 10.1016/j.tibs.2004.05.007 [DOI] [PubMed] [Google Scholar]
  • 23.Martin W, Russell MJ. On the origin of biochemistry at an alkaline hydrothermal vent. Philos Trans R Soc Lond B Biol Sci. 2007;362:1887–1925. doi: 10.1098/rstb.2006.1881 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Muchowska KB, Varma SJ, Chevallot-Beroux E, Lethuillier-Karl L, Li G, Moran J. Metals promote sequences of the reverse Krebs cycle. Nat Ecol Evol. 2017;1:1716–1721. doi: 10.1038/s41559-017-0311-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Varma SJ, Muchowska KB, Chatelain P, Moran J. Native iron reduces CO2 to intermediates and end-products of the acetyl-CoA pathway. Nat Ecol Evol. 2018;2:1019–1024. doi: 10.1038/s41559-018-0542-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Preiner M, Igarashi K, Muchowska KB, Yu M, Varma SJ, Kleinermanns K, et al. A hydrogen-dependent geochemical analogue of primordial carbon and energy metabolism. Nat Ecol Evol. 2020;4:534–542. doi: 10.1038/s41559-020-1125-6 [DOI] [PubMed] [Google Scholar]
  • 27.Keller MA, Turchyn AV, Ralser M. Non-enzymatic glycolysis and pentose phosphate pathway-like reactions in a plausible Archean ocean. Mol Syst Biol. 2014;10:725. doi: 10.1002/msb.20145228 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Piedrafita G, Varma SJ, Castro C, Messner CB, Szyrwiel L, Griffin JL, et al. Cysteine and iron accelerate the formation of ribose-5-phosphate, providing insights into the evolutionary origins of the metabolic network structure. PLoS Biol. 2021;19 (12):e3001468. doi: 10.1371/journal.pbio.3001468 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Messner CB, Driscoll PC, Piedrafita G, De Volder MFL, Ralser M. Nonenzymatic gluconeogenesis-like formation of fructose 1,6-bisphosphate in ice. Proc Natl Acad Sci. 2017;114:7403–7407. doi: 10.1073/pnas.1702274114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Huber C, Wächtershäuser G. Primordial reductive amination revisited. Tetrahedron Lett. 2003;44:1695–1697. doi: 10.1016/S0040-4039(02)02863-0 [DOI] [Google Scholar]
  • 31.Barge LM, Flores E, Baum MM, VanderVelde DG, Russell MJ. Redox and pH gradients drive amino acid synthesis in iron oxyhydroxide mineral systems. Proc Natl Acad Sci. 2019;116:4828–4833. doi: 10.1073/pnas.1812098116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Mayer RJ, Kaur H, Rauscher SA, Moran J. Mechanistic Insight into Metal Ion-Catalyzed Transamination. J Am Chem Soc. 2021;143:42. doi: 10.1021/jacs.1c08535 [DOI] [PubMed] [Google Scholar]
  • 33.Yi J, Kaur H, Kazöne W, Rauscher SA, Gravillier LA, Muchowska KB, et al. A Nonenzymatic Analog of Pyrimidine Nucleobase Biosynthesis. Angew Chemie Int Ed. 2022;e202117211. doi: 10.1002/anie.202117211 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Deiana C, Sakhno Y, Fabbiani M, Pazzi M, Vincenti M, Martra G. Direct synthesis of amides from carboxylic acids and amines by using heterogeneous catalysts: Evidence of surface carboxylates as activated electrophilic species. ChemCatChem. 2013;5:2832–2834. [Google Scholar]
  • 35.Akouche M, Jaber M, Maurel MC, Lambert JF, Georgelin T. Phosphoribosyl pyrophosphate: A molecular vestige of the origin of life on minerals. Angew Chem Int Ed Engl. 2017;56:7920–7923. doi: 10.1002/anie.201702633 [DOI] [PubMed] [Google Scholar]
  • 36.Ferry JG, House CH. The stepwise evolution of early life driven by energy conservation. Mol Biol Evol. 2006;23:1286–1292. doi: 10.1093/molbev/msk014 [DOI] [PubMed] [Google Scholar]
  • 37.Schönheit P, Buckel W, Martin W. On the Origin of Heterotrophy. Trends Microbiol. 2016;24:12–25. doi: 10.1016/j.tim.2015.10.003 [DOI] [PubMed] [Google Scholar]
  • 38.Lipmann F. Wanderings of a Biochemist. New York, USA: Wiley-Interscience; 1971. Available from: https://books.google.com/books/about/Wanderings_of_a_Biochemist.html?id=jHuBAAAAIAAJ. [Google Scholar]
  • 39.Lipmann F, Tuttle LC. Acetyl phosphate: chemistry, determiantion, and synthesis. J Biol Chem. 1944;153:571–582. Available: http://www.jbc.org/content/153/2/571.full.pdf. [Google Scholar]
  • 40.Decker K, Jungermann K, Thauer RK. Energy Production i n Anaerobic Organisms. Angew Chemie Int Ed. 1970;9:138–158. [DOI] [PubMed] [Google Scholar]
  • 41.Thauer RK, Jungermann K, Decker K. Energy conservation in chemotrophic anaerobic bacteria. Bacteriol Rev. 1977;41:100–180. doi: 10.1128/br.41.1.100-180.1977 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Martin WF, Sousa FL, Lane N. Energy at life’s origin. Science (80-). 2014;344:1092–1093. doi: 10.1111/soru.12033.11 [DOI] [PubMed] [Google Scholar]
  • 43.de Duve C. Did God make RNA? Nature. 1988;336:209–210. doi: 10.1038/336209b0 [DOI] [PubMed] [Google Scholar]
  • 44.de Duve C. Blueprint for a Cell—The Nature and Origin of Life. Burlington, North Carolina, USA: Neil Patterson Publishers; 1991. [Google Scholar]
  • 45.de Duve C. Clues from present-day biology: the thioester world. In: Brack A, editor. The Molecular Origins of Life. Cambridge University Press; 1998. p. 219–236. doi: 10.1017/cbo9780511626180.012 [DOI] [Google Scholar]
  • 46.Sousa FL, Thiergart T, Landan G, Nelson-Sathi S, Pereira IAC, Allen JF, et al. Early bioenergetic evolution. Philos Trans R Soc B Biol Sci. 2013;368:20130088–20130088. doi: 10.1098/rstb.2013.0088 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Sojo V, Herschy B, Whicher A, Camprubí E, Lane N. The origin of life in alkaline hydrothermal vents. Astrobiology. 2016;16:181–197. doi: 10.1089/ast.2015.1406 [DOI] [PubMed] [Google Scholar]
  • 48.Goldford JE, Hartman H, Smith TF, Segrè D. Remnants of an Ancient Metabolism without Phosphate. Cell. 2017;168:1126–1134.e9. doi: 10.1016/j.cell.2017.02.001 [DOI] [PubMed] [Google Scholar]
  • 49.Hartman H. Speculations on the Origin and Evolution of Metabolism. J Mol Evol. 1975;4:359–370. doi: 10.1007/BF01732537 [DOI] [PubMed] [Google Scholar]
  • 50.Kitadai N, Nakamura R, Yamamoto M, Okada S, Takahagi W, Nakano Y, et al. Thioester synthesis through geoelectrochemical CO2 fixation on Ni sulfides. Commun Chem. 2021;4:1–9. doi: 10.1038/s42004-021-00475-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Kitani A, Tsunetsugu S, Sasaki K. Fe(III)-ion-Catalysed Non-enzymatic Transformation of ADP into ATP. J Chem Soc Perkin Trans. 1991; 329–331. Available: http://pubs.rsc.org/en/content/articlepdf/1991/p2/p29910000329. [Google Scholar]
  • 52.Kitani A, Tsunetsugu S, Suzuki A, Ito S, Sasaki K. Fe(III)-ion-catalysed non-enzymatic transformation of adenosine diphosphate into adenosine triphosphate part II. Evidence of catalytic nature of Fe ions. Bioelectrochemistry Bioenerg. 1995;36:47–51. doi: 10.1016/0302-4598(94)01751-L [DOI] [Google Scholar]
  • 53.Jordan SF, Ioannou I, Rammu H, Halpern A, Bogart LK, Ahn M, et al. Spontaneous assembly of redox-active iron-sulfur clusters at low concentrations of cysteine. Nat Commun. 2021;12 (1):5925. doi: 10.1038/s41467-021-26158-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Storer AC, Cornish-Bowden A. Concentration of MgATP2- and other ions in solution. Calculation of the true concentrations of species present in mixtures of associating ions. Biochem J. 1976;159:1–5. doi: 10.1042/bj1590001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Holm NG. The significance of Mg in prebiotic geochemistry. Geobiology. 2012;10:269–279. doi: 10.1111/j.1472-4669.2012.00323.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Winzer K, Lorenz K, Dürre P. Acetate kinase from Clostridium acetobutylicum: A highly specific enzyme that is actively transcribed during acidogenesis and solventogenesis. Microbiology. 1997;143:3279–3286. doi: 10.1099/00221287-143-10-3279 [DOI] [PubMed] [Google Scholar]
  • 57.Ramirez F, Marecek JF, Szamosi J. Magnesium and Calcium Ion Effects on Hydrolysis Rates of Adenosine 5’-Triphosphate. Journal of Organic Chemistry. Academic Press. 1980:4748–4752. doi: 10.1039/c2dt90214a [DOI] [Google Scholar]
  • 58.Williams NH. Magnesium ion catalyzed ATP hydrolysis. J Am Chem Soc. 2000;122:12023–12024. doi: 10.1021/ja0013374 [DOI] [Google Scholar]
  • 59.Leibrock E, Bayer P, Lüdemann HD. Nonenzymatic hydrolysis of adenosinetriphosphate (ATP) at high temperatures and high pressures. Biophys Chem. 1995;54:175–180. doi: 10.1016/0301-4622(94)00134-6 [DOI] [PubMed] [Google Scholar]
  • 60.Etaix E, Orgel LE. Phosphorylation of Nucleosides in Aqueous-Solution Using Trimetaphosphate-Formation of Nucleoside Triphosphates. J Carbohydrates-Nucleosides-Nucleotides. 1978;5:91–110. [Google Scholar]
  • 61.Yamagata Y, Inoue H, Inomata K. Specific effect of magnesium on 2’,3’-cyclic AMP synthesis from adenosine and trimetaphosphate in aqueous solution. Orig Life Evol Biosph. 1995;25:47–52. Available: https://link.springer.com/content/pdf/10.1007/BF01581572.pdf. [DOI] [PubMed] [Google Scholar]
  • 62.Ozawa K, Nemoto A, Imai EI, Honda H, Hatori K, Matsuno K. Phosphorylation of nucleotide molecules in hydrothermal environments. Orig Life Evol Biosph. 2004;34:465–471. doi: 10.1023/b:orig.0000043121.65714.05 [DOI] [PubMed] [Google Scholar]
  • 63.Britvin SN, Murashko MN, Vapnik Y, Vlasenko NS, Krzhizhanovskaya MG, Vereshchagin OS, et al. Cyclophosphates, a new class of native phosphorus compounds, and some insights into prebiotic phosphorylation on early Earth. Geology 2020;49(4): 382–386. doi: 10.1130/G48203.1 [DOI] [Google Scholar]
  • 64.Holm NG, Baltscheffsky H. Links Between Hydrothermal Environments, Pyrophosphate, Na +, and Early Evolution. Orig Life Evol Biosph. 2011;41:483–493. doi: 10.1007/s11084-011-9235-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Bryant DE, Greenfield D, Walshaw RD, Johnson BRG, Herschy B, Smith C, et al. Hydrothermal modification of the Sikhote-Alin iron meteorite under low pH geothermal environments. A plausibly prebiotic route to activated phosphorus on the early Earth. Geochim Cosmochim Acta. 2013;109:90–112. doi: 10.1016/j.gca.2012.12.043 [DOI] [Google Scholar]
  • 66.Kee TP, Bryant DE, Herschy B, Marriott KER, Cosgrove NE, Pasek MA, et al. Phosphate activation via reduced oxidation state phosphorus (P). Mild routes to condensed-P energy currency molecules. Life. 2013;3:386–402. doi: 10.3390/life3030386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Pasek MA, Harnmeijer JP, Buick R, Gull M, Atlas Z. Evidence for reactive reduced phosphorus species in the early Archean ocean. Proc Natl Acad Sci U S A. 2013;110:10089–10094. doi: 10.1073/pnas.1303904110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kaye K, Bryant DE, Marriott KER, Ohara S, Fishwick CWG, Kee TP. Selective Phosphonylation of 5′-Adenosine Monophosphate (5′-AMP) via Pyrophosphite [PPi(III)]. Orig Life Evol Biosph. 2016;46:425–434. doi: 10.1007/s11084-016-9497-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Pasek M, Omran A, Lang C, Gull M, Abbatiello J, Feng T, et al. Serpentinization as a route to liberating phosphorus on habitable worlds. Res Sq [Preprint]. 2020. doi: 10.21203/rs.3.rs-37651/v1 [DOI] [Google Scholar]
  • 70.Coggins AJ, Powner MW. Prebiotic synthesis of phosphoenol pyruvate by α-phosphorylation-controlled triose glycolysis. Nat Chem. 2017;9:310–317. doi: 10.1038/nchem.2624 [DOI] [PubMed] [Google Scholar]
  • 71.Liu Z, Rossi JC, Pascal R. How prebiotic chemistry and early life chose phosphate. Life. 2019;9:1–16. doi: 10.3390/life9010026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lagoja IM, Herdewijn P. A potential prebiotic route to adenine from hypoxanthine. Chem Biodivers. 2005;2:923–927. doi: 10.1002/cbdv.200590069 [DOI] [PubMed] [Google Scholar]
  • 73.Allen CM, Ellen Jones ME. Decomposition of Carbamylphosphate in Aqueous Solutions*. Biochemistry. 1964;3:1238–1247. doi: 10.1021/bi00897a010 [DOI] [PubMed] [Google Scholar]
  • 74.Lohrmann R, Orgel LE. Prebiotic synthesis: phosphorylation in aqueous solution. Science. 1968;161:64–66. doi: 10.1126/science.161.3836.64 [DOI] [PubMed] [Google Scholar]
  • 75.Danger G, Boiteau L, Cottet H, Pascal R. The Peptide Formation Mediated by Cyanate Revisited. N-Carboxyanhydrides as Accessible Intermediates in the Decomposition of N-Carbamoylamino Acids. J Am Chem Soc. 2006;128 (23):7412–7413. doi: 10.1021/ja061339+ [DOI] [PubMed] [Google Scholar]
  • 76.Rabinowitz IN, Davis FF, Herber RH. Mossbauer-Effect Studies on Metal Binding in Purine Compounds1a. J Am Chem Soc. 1966;88:4346–4354. doi: 10.1021/ja00971a008 [DOI] [PubMed] [Google Scholar]
  • 77.Izatt RM, Christensen JJ, Rytting JH. Erratum: Sites and thermodynamic quantities associated with proton and metal ion interaction with ribonucleic acid, deoxyribonucleic acid, and their constituent bases, nucleosides, and nucleotides. Chemical Reviews. 1972. pp. 439–481. doi: 10.1021/cr60280a005 [DOI] [PubMed] [Google Scholar]
  • 78.Gao Y, gui Sriram M, Wang AHJ. Crystallographic studies of metal ion—DNA interactions: Different binding modes of cobalt(II), copper(II) and barium(II) to N7 of guanines in Z-DNA and a drug-DNA complex. Nucleic Acids Res. 1993;21:4093–4101. doi: 10.1093/nar/21.17.4093 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Šponer J, Sabat M, Gorb L, Leszczynski J, Lippert B, Hobzat P. The effect of metal binding to the N7 site of purine nucleotides on their structure, energy, and involvement in base pairing. J Phys Chem B. 2000;104:7535–7544. doi: 10.1021/jp001711m [DOI] [Google Scholar]
  • 80.Sigel H. Have adenosine 5′-triphosphate (ATP^4−) and related purine-nucleotides played a role in early evolution? ATP, its own ‘enzyme’in metal ion facilitated hydrolysis! Inorganica Chim Acta. 1992;198:1–11. Available from: http://www.sciencedirect.com/science/article/pii/S0020169300923429. [Google Scholar]
  • 81.Morasch M, Mast CB, Langer JK, Schilcher P, Braun D. Dry polymerization of 3′,5′-cyclic GMP to long strands of RNA. ChemBioChem. 2014;15:879–883. doi: 10.1002/cbic.201300773 [DOI] [PubMed] [Google Scholar]
  • 82.Oku H, Mutanda I, Fukuta M, Inafuku M. Non-enzymatic formation of isoprene and 2-methyl-3-buten-2-ol (2-MBO) by manganese. Scientific Reports. 2022;12(1):1–9. doi: 10.1038/s41598-022-06520-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Goucher CR, Taylor JF. Compounds of Ferric Iron with Adenosine Triphosphate and Other Nucleoside Phosphates. J Biol Chem. 1964;239:2251–2255. [PubMed] [Google Scholar]
  • 84.Ouameur AA, Arakawa H, Ahmad R, Naoui M, Tajmir-Riahi HA. A comparative study of Fe(II) and Fe(III) interactions with DNA duplex: Major and minor grooves bindings. DNA Cell Biol. 2005;24:394–401. doi: 10.1089/dna.2005.24.394 [DOI] [PubMed] [Google Scholar]
  • 85.Swain M. chemicalize.org. J Chem Inf Model. 2012;52:613–615. doi: 10.1021/ci300046g [DOI] [Google Scholar]
  • 86.Sigel H, Massoud SS, Corfu NA. Comparison of the Extent of Macrochelate Formation in Complexes of Divalent Metal Ions with Guanosine (GMP2-), Inosine (IMP2-), and Adenosine 5’-Monophosphate (AMP2-). The Crucial Role of N-7 Basicity in Metal Ion-Nucleic Base Recognition. J Am Chem Soc. 1994;116 (7):2958–2971. doi: 10.1021/ja00086a028 [DOI] [Google Scholar]
  • 87.Yathindra N, Sundaralingam M. Correlation between the backbone and side chain conformations in 5′-nucleotides. The concept of a ‘rigid’ nucleotide conformation. Biopolymers. 1973;12:297–314. doi: 10.1002/bip.1973.360120208 [DOI] [Google Scholar]
  • 88.Auffinger P, Grover N, Westhof E. Metal ion binding to RNA. Metal ions in life sciences. 2011. doi: 10.1515/9783110436648-006 [DOI] [PubMed] [Google Scholar]
  • 89.Wang C, Qi Q, Li W, Dang J, Hao M, Lv S, et al. A Cu(II)–ATP complex efficiently catalyses enantioselective Diels–Alder reactions. Nat Commun. 2020;11. doi: 10.1038/s41467-020-18554-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Shannon RD. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr Sect A. 1976;32:751–767. doi: 10.1107/S0567739476001551 [DOI] [Google Scholar]
  • 91.Scheller KH, Sigel H. A proton nuclear magnetic resonance study of purine and pyrimidine nucleoside 5’-diphosphates. Extent of macrochelate formation in monomeric metal ion complexes and promotion of self-stacking by metal ions. J Am Chem Soc. 1983;105 (18):5891–5900. doi: 10.1021/ja00356a029 [DOI] [Google Scholar]
  • 92.Sigel H. Interactions of metal ions with nucleotides and nucleic acids and their constituents. Chem Soc Rev. 1993;22:255–267. doi: 10.1039/CS9932200255 [DOI] [Google Scholar]
  • 93.Biaglow JE, Held KD, Manevich Y, Tuttle S, Kachur A, Uckun F. Role of Guanosine Triphosphate in Ferric Ion-Linked Fenton Chemistry. Radiation Research. 1996;145 (5):554–562. doi: 10.2307/3579273 [DOI] [PubMed] [Google Scholar]
  • 94.Di Sabato G, Jencks WP. Mechanism and Catalysis of Reactions of Acyl Phosphates. I. Nucleophilic Reactions. J Am Chem Soc. 1961;83:4400–4405. doi: 10.1021/ja01482a025 [DOI] [Google Scholar]
  • 95.Tauchert K, Jahn A, Oelze J. Control of diauxic growth of Azotobacter vinelandii on acetate and glucose. J Bacteriol. 1990;172:6447–6451. doi: 10.1128/jb.172.11.6447-6451.1990 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Reinscheid DJ, Schnicke S, Rittmann D, Zahnow U, Sahm H, Eikmanns BJ. Cloning, sequence analysis, expression and inactivation of the Corynebacterium glutamicum pta-ack operon encoding phosphotransacetylase and acetate kinase. Microbiology. 1999;145:503–513. doi: 10.1099/13500872-145-2-503 [DOI] [PubMed] [Google Scholar]
  • 97.Xu WJ, Wen H, Kim HS, Ko YJ, Dong SM, Park IS, et al. Observation of acetyl phosphate formation in mammalian mitochondria using real-time in-organelle NMR metabolomics. Proc Natl Acad Sci U S A. 2018;115:4152–4157. doi: 10.1073/pnas.1720908115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Wolfe AJ. Physiologically relevant small phosphodonors link metabolism to signal transduction. Curr Opin Microbiol. 2010;13:204–209. doi: 10.1016/j.mib.2010.01.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Jones ME, Spector L, Lipmann F. Carbamyl phosphate, the carbamyl donor in enzymatic citrulline synthesis. J Am Chem Soc. 1955;77:819–820. doi: 10.1021/ja01608a101 [DOI] [Google Scholar]
  • 100.Jones ME, Lipmann F. Chemical and enzymatic synthesis of carbamyl phosphate. Proc Natl Acad Sci. 1960;46:1194–1205. doi: 10.1073/pnas.46.9.1194 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Lohrmann R, Orgel LE. Urea-Inorganic Phosphate Mixtures as Prebiotic Phosphorylating Agents. Science (80-). 1971;171:490–494. doi: 10.1126/science.171.3970.490 [DOI] [PubMed] [Google Scholar]
  • 102.Yamagata Y. Non-enzymatic ATP synthesis by the phosphorylation of ADP with the assistance of cyanate and magnesium ion. Orig Life Evol Biosph. 1996;26:242–243. doi: 10.1007/bf02459735 [DOI] [Google Scholar]
  • 103.Yamagata Y. Prebiotic Formation of ADP and ATP From AMP, Calcium Phosphates and Cyanate in Aqueous Solution. Orig Life Evol Biosph. 1999;29:511–520. doi: 10.1023/a:1006672232730 [DOI] [PubMed] [Google Scholar]
  • 104.Pasek M. Thermodynamics of Prebiotic Phosphorylation. Chem Rev. 2020;120(11):4690–4706. doi: 10.1021/acs.chemrev.9b00492 [DOI] [PubMed] [Google Scholar]
  • 105.Maguire OR, Smokers IBA, Huck WTS. A physicochemical orthophosphate cycle via a kinetically stable thermodynamically activated intermediate enables mild prebiotic phosphorylations. Nat Commun. 2021;12:5517. doi: 10.1038/s41467-021-25555-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Saygin Ö. Nonenzymatic photophosphorylation with visible light—A possible mode of prebiotic ATP formation. Naturwissenschaften. 1981;68:617–619. doi: 10.1007/BF00398616 [DOI] [Google Scholar]
  • 107.Saygin Ö. Nonenzymatic phosphorylation of acetate by carbamyl phosphate. Orig Life Evol Biosph. 1983;13:43–48. [Google Scholar]
  • 108.Saygin Ö. Photochemical carbamylphosphate formation and metal ion catalysed transphosphorylations between carbamylphosphate and adenine nucleotides or carboxyl groups. Orig Life Evol Biosph. 1984;14:131–137. doi: 10.1007/BF00933649 [DOI] [Google Scholar]
  • 109.Ramón-Maiques S, Marina A, Guinot A, Gil-Ortiz F, Uriarte M, Fita I, et al. Substrate binding and catalysis in carbamate kinase ascertained by crystallographic and site-directed mutagenesis studies: Movements and significance of a unique globular subdomain of this key enzyme for fermentative ATP production in bacteria. J Mol Biol. 2010;397:1261–1275. doi: 10.1016/j.jmb.2010.02.038 [DOI] [PubMed] [Google Scholar]
  • 110.Schultheisz HL, Szymczyna BR, Scott LG, Williamson JR. Enzymatic de novo pyrimidine nucleotide synthesis. J Am Chem Soc. 2011;133:297–304. doi: 10.1021/ja1059685 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Shi D, Caldovic L, Tuchman M. Sources and Fates of Carbamyl Phosphate: A Labile Energy-Rich Molecule with Multiple Facets. Biology (Basel). 2018;7 (2):34. doi: 10.3390/biology7020034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Gibard C, Bhowmik S, Karki M, Kim EK, Krishnamurthy R. Phosphorylation, oligomerization and self-assembly in water under potential prebiotic conditions. Nature Chem. 2018;10 (SI). doi: 10.1038/nchem.2878 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Ducluzeau AL, van Lis R, Duval S, Schoepp-Cothenet B, Russell MJ, Nitschke W. Was nitric oxide the first deep electron sink? Trends Biochem Sci. 2009;34:9–15. doi: 10.1016/j.tibs.2008.10.005 [DOI] [PubMed] [Google Scholar]
  • 114.Bonfio C, Valer L, Scintilla S, Shah S, Evans DJ, Jin L, et al. UV-light-driven prebiotic synthesis of iron-sulfur clusters. Nat Chem. 2017;9(12):1229–1234. doi: 10.1038/nchem.2817 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Marteinsson VT, Kristjánsson JK, Kristmannsdóttir H, Dahlkvist M, Sæmundsson K, Hannington M, et al. Discovery and description of giant submarine smectite cones on the seafloor in Eyjafjordur, northern Iceland, and a novel thermal microbial habitat. Appl Environ Microbiol. 2001;67:827–833. doi: 10.1128/AEM.67.2.827-833.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Morse JW, Mackenzie FT. Hadean Ocean carbonate geochemistry. Aquat. Geochemistry. 1998;4:301–319. doi: 10.1023/a:1009632230875 [DOI] [Google Scholar]
  • 117.Shibuya T, Yoshizaki M, Masaki Y, Suzuki K, Takai K, Russell MJ. Reactions between basalt and CO2-rich seawater at 250 and 350 C, 500 bars: implications for the CO2 sequestration into the modern oceanic crust and the composition of hydrothermal vent fluid in the CO2-rich early ocean. Chem Geol. 2013;359:1–9. doi: 10.1016/j.chemgeo.2013.08.044 [DOI] [Google Scholar]
  • 118.Shibuya T, Russell MJ, Takai K. Free energy distribution and hydrothermal mineral precipitation in Hadean submarine alkaline vent systems: Importance of iron redox reactions under anoxic conditions. Geochim Cosmochim Acta. 2016;175:1–19. doi: 10.1016/j.gca.2015.11.021 [DOI] [Google Scholar]
  • 119.Camprubí E, Jordan SF, Vasiliadou R, Lane N. Iron catalysis at the origin of life. IUBMB Life. 2017;69:373–381. doi: 10.1002/iub.1632 [DOI] [PubMed] [Google Scholar]
  • 120.Muchowska KB, Varma SJ, Moran J. Synthesis and breakdown of universal metabolic precursors promoted by iron. Nature. 2019;569:104–107. doi: 10.1038/s41586-019-1151-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Branscomb E, Biancalani T, Goldenfeld N, Russell MJ. Escapement mechanisms and the conversion of disequilibria; the engines of creation. Phys Rep. 2017;677:1–60. doi: 10.1016/j.physrep.2017.02.001 [DOI] [Google Scholar]
  • 122.Martin W, Russell MJ. On the origins of cells: A hypothesis for the evolutionary transitions from abiotic geochemistry to chemoautotrophic prokaryotes, and from prokaryotes to nucleated cells. Philos Trans R Soc Lond B Biol Sci. 2003:59–85. doi: 10.1098/rstb.2002.1183 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Nitschke W, Russell MJ. Hydrothermal focusing of chemical and chemiosmotic energy, supported by delivery of catalytic Fe, Ni, Mo/W, Co, S and Se, forced life to emerge. J Mol Evol. 2009;69: 481–496. doi: 10.1007/s00239-009-9289-3 [DOI] [PubMed] [Google Scholar]
  • 124.Vasiliadou R, Dimov N, Szita N, Jordan SF, Lane N. Possible mechanisms of CO2 reduction by H2 via prebiotic vectorial electrochemistry. Interface Focus. 2019;9:20190073. doi: 10.1098/rsfs.2019.0073 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Jordan SF, Rammu H, Zheludev IN, Hartley AM, Maréchal A, Lane N. Promotion of protocell self-assembly from mixed amphiphiles at the origin of life. Nat Ecol Evol. 2019;2019:1–10. doi: 10.1038/s41559-019-1015-y [DOI] [PubMed] [Google Scholar]
  • 126.Jordan SF, Nee E, Lane N. Isoprenoids enhance the stability of fatty acid membranes at the emergence of life potentially leading to an early lipid divide. Interface Focus. 2019;Submitted. doi: 10.1098/rsfs.2019.0067 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Hudson R, de Graaf R, Rodin MS, Ohno A, Lane N, McGlynn SE, et al. CO2 reduction driven by a pH gradient. Proc Natl Acad Sci. 2020;117:22873–22879. doi: 10.1073/pnas.2002659117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Sojo V, Pomiankowski A, Lane N. A Bioenergetic Basis for Membrane Divergence in Archaea and Bacteria. PLoS Biol. 2014;13 (3):e1002102. doi: 10.1371/journal.pbio.1001926 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Nunes Palmeira R, Colnaghi M, Harrison SA, Pomiankowski A, Lane N. The Limits of Metabolic Heredity in Protocells. bioRxiv. 2022. doi: 10.1101/2022.01.28.477904 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Hulett HR. Non-enzymatic Hydrolysis of Adenosine Phosphates. Nature. 1970;225:1248–1249. doi: 10.1038/2251248a0 [DOI] [PubMed] [Google Scholar]
  • 131.Burcar BT, Cassidy LM, Moriarty EM, Joshi PC, Coari KM, McGown LB. Potential Pitfalls in MALDI-TOF MS Analysis of Abiotically Synthesized RNA Oligonucleotides. Orig Life Evol Biosph. 2013;43:247–261. doi: 10.1007/s11084-013-9334-5 [DOI] [PubMed] [Google Scholar]
  • 132.Jo S, Kim T, Iyer VG, Im W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J Comput Chem. 2008;29:1859–1865. doi: 10.1002/jcc.20945 [DOI] [PubMed] [Google Scholar]
  • 133.Kim S, Lee J, Jo S, Brooks CL 3rd, Lee HS, Im W. CHARMM-GUI ligand reader and modeler for CHARMM force field generation of small molecules. J Comput Chem. 2017;38(21):1879–1886. doi: 10.1002/jcc.24829 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Lee J, Cheng X, Swails JM, Yeom MS, Eastman PK, Lemkul JA, et al. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J Chem Theory Comput. 2016;12(1):405–413. doi: 10.1021/acs.jctc.5b00935 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Phillips JC, Hardy DJ, Maia JDC, Stone JE, Ribeiro JV, Bernardi RC, et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J Chem Phys. 2020;153(4):044130. doi: 10.1063/5.0014475 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Matsunaga Y. MDToolbox, version f8eed18. 2018. Available from: https://github.com/ymatsunaga/mdtoolbox.
  • 137.Spectrum MCH00020 for Adenosine 5’-triphosphate. In: MoNA—MassBank of North America [Internet]. 2016. Available from: https://mona.fiehnlab.ucdavis.edu/spectra/display/MCH00020.

Decision Letter 0

Roland G Roberts

30 Sep 2021

Dear Nick,

Thank you for submitting your manuscript entitled "A prebiotic basis for ATP as the universal energy currency" for consideration as a Research Article by PLOS Biology.

Your manuscript has now been evaluated by the PLOS Biology editorial staff, as well as by an academic editor with relevant expertise, and I'm writing to let you know that we would like to send your submission out for external peer review.

However, before we can send your manuscript to reviewers, we need you to complete your submission by providing the metadata that is required for full assessment. To this end, please login to Editorial Manager where you will find the paper in the 'Submissions Needing Revisions' folder on your homepage. Please click 'Revise Submission' from the Action Links and complete all additional questions in the submission questionnaire.

Please re-submit your manuscript within two working days, i.e. by Oct 04 2021 11:59PM.

Login to Editorial Manager here: https://www.editorialmanager.com/pbiology

During resubmission, you will be invited to opt-in to posting your pre-review manuscript as a bioRxiv preprint. Visit http://journals.plos.org/plosbiology/s/preprints for full details. If you consent to posting your current manuscript as a preprint, please upload a single Preprint PDF when you re-submit.

Once your full submission is complete, your paper will undergo a series of checks in preparation for peer review. Once your manuscript has passed all checks it will be sent out for review.

Given the disruptions resulting from the ongoing COVID-19 pandemic, please expect delays in the editorial process. We apologise in advance for any inconvenience caused and will do our best to minimize impact as far as possible.

Feel free to email us at plosbiology@plos.org if you have any queries relating to your submission.

Cheers,

Roli

Roland Roberts

Senior Editor

PLOS Biology

rroberts@plos.org

Decision Letter 1

Roland G Roberts

13 Dec 2021

Dear Nick,

Thank you for submitting your manuscript "A prebiotic basis for ATP as the universal energy currency" for consideration as a Research Article at PLOS Biology. Your manuscript has been evaluated by the PLOS Biology editors, an Academic Editor with relevant expertise, and by four independent reviewers. Thanks for your patience while we discussed the decision with the Academic Editor.

You'll see that while the reviewers are intrigued by your findings, they raise a series of significant concerns (some overlapping), which will need to be addressed before further consideration. The Academic Editor also kindly provided some additional advice, and I've included an edited version of his/her comments at the foot of this letter; these don't contain any further requests, but they do contextualise the requests from the reviewers, emphasising some, and explicitly excusing you one request (the EPR).

In light of the reviews (below), we will not be able to accept the current version of the manuscript, but we would welcome re-submission of a much-revised version that takes into account the reviewers' comments. We cannot make any decision about publication until we have seen the revised manuscript and your response to the reviewers' comments. Your revised manuscript is also likely to be sent for further evaluation by the reviewers.

We expect to receive your revised manuscript within 3 months.

Please email us (plosbiology@plos.org) if you have any questions or concerns, or would like to request an extension. At this stage, your manuscript remains formally under active consideration at our journal; please notify us by email if you do not intend to submit a revision so that we may end consideration of the manuscript at PLOS Biology.

**IMPORTANT - SUBMITTING YOUR REVISION**

Your revisions should address the specific points made by each reviewer. Please submit the following files along with your revised manuscript:

1. A 'Response to Reviewers' file - this should detail your responses to the editorial requests, present a point-by-point response to all of the reviewers' comments, and indicate the changes made to the manuscript.

*NOTE: In your point by point response to the reviewers, please provide the full context of each review. Do not selectively quote paragraphs or sentences to reply to. The entire set of reviewer comments should be present in full and each specific point should be responded to individually, point by point.

You should also cite any additional relevant literature that has been published since the original submission and mention any additional citations in your response.

2. In addition to a clean copy of the manuscript, please also upload a 'track-changes' version of your manuscript that specifies the edits made. This should be uploaded as a "Related" file type.

*Re-submission Checklist*

When you are ready to resubmit your revised manuscript, please refer to this re-submission checklist: https://plos.io/Biology_Checklist

To submit a revised version of your manuscript, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' where you will find your submission record.

Please make sure to read the following important policies and guidelines while preparing your revision:

*Published Peer Review*

Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

*PLOS Data Policy*

Please note that as a condition of publication PLOS' data policy (http://journals.plos.org/plosbiology/s/data-availability) requires that you make available all data used to draw the conclusions arrived at in your manuscript. If you have not already done so, you must include any data used in your manuscript either in appropriate repositories, within the body of the manuscript, or as supporting information (N.B. this includes any numerical values that were used to generate graphs, histograms etc.). For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5

*Blot and Gel Data Policy*

We require the original, uncropped and minimally adjusted images supporting all blot and gel results reported in an article's figures or Supporting Information files. We will require these files before a manuscript can be accepted so please prepare them now, if you have not already uploaded them. Please carefully read our guidelines for how to prepare and upload this data: https://journals.plos.org/plosbiology/s/figures#loc-blot-and-gel-reporting-requirements

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Thank you again for your submission to our journal. We hope that our editorial process has been constructive thus far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Roli

Roland Roberts

Senior Editor

PLOS Biology

rroberts@plos.org

*****************************************************

REVIEWERS' COMMENTS:

Reviewer #1:

[see attachment for fully formatted version]

The manuscript by Pinna et al. reports on the detailed characterization of the phosphorylation of ADP by acetyl phosphate (AcP) to form ATP in the presence of ferric ions. This previously observed reaction appears to be surprisingly specific as no other NDPs or transition metal ions can carry out equivalent reactions to any significant extent. ADP phosphorylation appears to be optimal at 30°C, mildly acidic pH and low salt. The authors provide explanations for their observations within the context of their already well-established hypothesis of life's origins in a hydrothermal vent. Besides the fact that, taken face value, the observations reported here would rather favor a surface pond or lake as the cradle of life, some of their propositions need to be further substantiated.

Page 3 "ATP in a prebiotic, monomeric world". This assumes that life originated in bulk water. But, if, instead, life first appeared on mineral surfaces then the needed energy could have been provided through the oxidation of H2 and CO by, for instance, catalytic FeNiS minerals. Only when more evolved protocells liberated themselves from those surfaces, and moved to bulk water, would have ATP become essential.

"Even if only one nucleotide triphosphate can be dominant, the implication of a frozen accident is not a satisfying explanation." Why not? There are different levels of "frozen accidents". This has been discussed by Brandon Carter who, in 1984, proposed that the evolution towards intelligent life required 6 "hard steps", the first one being abiogenesis. Each step was statistically postulated to last between 600 and 800 million years. Although these periods of time may be considered workable in astronomy and geology, in biology they do not make the same sense. In the latter case it is necessary to invoke the occurrence of extremely unlikely (highly contingent) events to explain that time duration. They can be indeed defined as "frozen accidents".

Page 13. "Such steep pH gradients could in principle operate across protocells as well as inorganic barriers." The authors propose that ATP would have been synthesized outside protocells, under acidic conditions, to be then transported across a proto-membrane into a cell where, under alkaline conditions, it would have phosphorylated some substrate(s). Have they considered the implications of such a mechanism in terms of ATP/ADP concentrations and diffusion? ATP would have to diffuse towards the protocell and ADP do it in the opposite direction, without risking to disappear into the bulk water. Another crucial point is membrane permeability. As F. H. Westheimer wrote in 1987: "Phosphoric acid is specially adapted for its role in nucleic acids because it can link two nucleotides and still ionize; the resulting negative charge serves both to stabilize the diesters against hydrolysis and to retain the molecules within a lipid membrane. A similar explanation for stability and retention also holds for phosphates that are intermediary metabolites and for phosphates that serve as energy sources." So, one very important role of even proto-membranes should have been to keep charged metabolites inside the cell to sustain metabolism. This, in turn, would have made ADP diffusion out of the protocell difficult to envision.

Page 15. "…enabling ATP to drive work even in a prebiotic monomer world." How are the authors defining 'work' in that setting? In contemporary biology work is i) performed mechanically as in muscle contraction and ii) against a gradient of chemical potential in active transport (see Jencks, 1989). This work depends on complex protein-based binding energy modulation and the coupling of intermediate reaction steps. Of course, this kind of work could not have existed prebiotically. At that time the most likely role of ATP (if it already existed) would have been , for instance, phosphorylation to render carbonyl carbons electrophilic enough to be attacked by a nucleophile such as ammonia. It is debatable that this process would qualify as work.

Page 31. Figure 6. Here and in the text, it is implied that the N6 -NH2 group interacts with Fe3+. In A-T base pairing N6 interacts with the O atom from the carbonyl group of T through a polarized delta+ proton. What would be the driving force in its postulated interaction with Fe3+? In b, the metal ion is hexacoordinated but the coordination geometry is not clear. In any case, it doesn't look octahedral. Maybe the authors should do some modeling to check whether the proposed arrangement is plausible. An experimental confirmation of their model would be even better.

Reviewer #2:

Pinna et al. present a study on the prebiotic formation of ATP from the Fe3+ catalysed reaction between ADP and acetyl phosphate. The authors attempt to answer a very important question in prebiotic chemistry which is how did ATP become the principle energy currency in Life and, in particular, why did ATP become dominant over the other nucleotide triphosphates (NTPs) for this purpose.

The authors extend the work of Kitani (Ref 32 and 33) who first demonstrate that ATP could be formed from a reaction between acetyl phosphate (AcP) and ADP in the presence of Fe3+ ions. The authors explore whether other metal ions can also catalyse this reaction but find that Fe3+ is unique in its ability to do so. The authors further explore the effect of pH, temperature, salt solutes of varying concentration and pressure on the reaction. The authors repeat the pH experiments carried out by Kitani (Ref 33) which affirm the reaction proceeds optimally at mildly acidic pHs. The temperature optimum was at 30 oC. Higher salt concentrations significantly affected the reaction and pressure had a negligible effect. The authors then proceed to test different prebiotically plausible phosphorylating reagents to see if there are others apart from AcP that can perform the reaction. The authors find that carbamoyl phosphate (CP) is the only other phosphorylating reagent which works. Next the authors explore whether Fe3+ catalysed phosphorylation of ADP by AcP also works for other NTPs. The authors find that the reaction exclusively phosphorylates ADP. Finally the authors attempt to unravel the mechanism behind why only ADP is phosphorylated. The authors attempt the phosphorylation of pyrophosphate to determine whether the nucleobase is important and conclude that it is. The authors conclude that a purine ring is important due to the lack of phosphorylation of the pyrimidines. Of the three purine rings explored only adenosine showed phosphorylation and therefore the authors conclude that the N6 amine is essential as both guanosine and inosine have a carbonyl group at the 6-position on the purine ring. The authors then determine that the optimal stoichiometry for ADP and Fe3+ is 1:1. Following Kitani they also find that addition of Mg2+ or Ca2+ increases the yield of the reaction. Lastly the authors conduct a MALDI-ToF experiment to ascertain whether Fe3+-ADP complex stacking is present in the reaction and conclude that it isn't. This all leads to propose the mechanism in Figure 6 whereby a macro chelate complex is formed between a single Fe3+ and a single ADP and the adenosine coordinates to the Fe3+ via the N6-amine and the pyrophosphate. AcP then coordinates to the Fe3+ and the phosphate is transferred and then Mg2+ displaces Fe3+ to promote catalyst turnover. Given that only ADP is phosphorylated to ATP the authors therefore conclude that this is the reason why Life uses ATP as the universal energy currency and that this has its origins in prebiotic chemistry.

I consider that the key result in this paper, namely that only ATP is phosphorylated and not the other NTPs, to be a very interesting result and this will be of great interest to the Origins of Life community. The paper is overall well written and well referenced. The experiments up to and including the demonstration that only ADP is phosphorylated by the reaction have been performed well, although I believe an additional control experiment is required to prove that ADP is being phosphorylated only by AcP in the presence of Fe3+. However, I have a number of serious concerns with both the experiments used to probe the mechanism of the reaction (the 31P NMR spectroscopy and the MALDI ToF experiments) and with the mechanistic interpretation of the results from these experiments which run contrary to what is known in the literature. The authors need to address these concerns in detail. Furthermore, the final conclusion of the paper where the authors claim that they have solved the reason why ATP is the universal energy currency of Life is not fully substantiated by the results presented in the paper. The authors have significantly overstretched themselves in making this claim and they need to tone down the conclusion of the paper substantially. I detail my reasoning behind this assessment in comments below.

[Page 6 Line 146]

In the Figure 2 caption the authors' provide rate constants for the reaction at different pHs and temperatures but no detail is provided as to what rate equation they used to determine these rate constants. Are these rate constants experimental first order rate constants?

[Page 7 Line 183]

The authors perform a control experiment whereby the reaction was performed in the absence of Fe3+ catalyst (Figure 2d). No ATP was detected in this control experiment which demonstrates that the formation of ATP requires Fe3+.

However, in order to prove the reaction requires acetyl phosphate an additional control experiment is required whereby the reaction is performed in the presence of Fe3+ but no acetyl phosphate. The reason this control experiment is required is because it is also possible that ATP could be produced from the reaction between two ADP molecules, i.e. ADP + ADP -> ATP + AMP, and this reaction could be catalysed by Fe3+. This reaction is known in the literature, see W. Huck Nat. Commun. 2021, 12, 5517.

It is notable that in the HPLC chromatogram in Figure 1b that a significant peak of AMP is also present. This could indicate that the ATP (at least partially) formed via a 'disproportionation' reaction between two ADP molecules.

The authors need to perform this control experiment to confirm whether or not acetyl phosphate is fully or only partially responsible for the phosphorylation of ADP.

[Page 8 Line 197] and [Page 14 Line 426]

The authors demonstrate that, in addition to acetyl phosphate, carbamoyl phosphate (CP) can also phosphorylate ADP to ATP in the presence of Fe3+ albeit in a lower yield than for AcP. This is a very nice result as it demonstrates that there are multiple prebiotic routes to the formation of ATP with Fe3+.

The authors correctly attribute the lower yields of ATP with CP to the decomposition of CP into Pi and cyanate. The citation of Jones and Lipmann (Ref 57 and Ref 74) is well received as they demonstrated that carbamoyl phosphate is formed in equilibrium with cyanate and Pi.

Given this good result with carbamoyl phosphate I am genuinely surprised at how dismissive the authors are of the importance of cyanate in prebiotic chemistry when it comes to the Discussion. The easiest prebiotically plausible route to carbamoyl phosphate is through a reaction between cyanate and Pi. While I appreciate that in extant metabolism carbamoyl phosphate is formed primarily through a glutamine deamination pathway which releases ammonia which then reacts with carboxy phosphate to form carbamoyl phosphate, this set of reactions is significantly less facile on the prebiotic Earth than the reaction with cyanate.

Furthermore the authors state [Page 15 Line 431]: 'That was important as it showed that biologically relevant condensations are possible in water, but differed from modern biochemistry in that cyanate does not feature in extant metabolism.'

This is not correct. Cyanate does feature in extant metabolism. As already mentioned above, and indeed cited by the authors, Jones and Lipmann (Ref 57 and Ref 74) demonstrated that in solution carbamoyl phosphate is in equilibrium with cyanate and Pi and therefore cyanate will be present in cells.

Ultimately, I believe that this dismissal of cyanate weakens the paper and undermines the authors' own promising result that Fe3+ catalyses the phosphorylation of ADP by carbamoyl phosphate. I consider that the demonstration of multiple pathways to the formation of ATP to be a strength of the paper and not a weakness.

One experiment the authors could try to raise the yield of ATP with carbamoyl phosphate is to perform the reaction in the presence of excess cyanate which will shift the equilibrium between CP and cyanate + Pi towards carbamoyl phosphate formation. Such a strategy was used in recent publication whereby excess cyanate enabled carbamoyl phosphate formation which in turn facilitated the formation of a phosphoramidate imidazole phosphate that was also demonstrated to phosphorylate ADP to ATP (W. Huck Nat. Commun. 2021, 12, 5517). A possible downside to the addition of excess cyanate here could be that coordination of cyanate to the Fe3+ could interfere with its ability to act as a catalyst. Nevertheless, I think it would be interesting reaction to try.

Another interesting set of reactions to try would be whether the Fe3+ catalysed CP phosphorylation also works with the other NDPs or whether it only works with ADP.

[Page 9, Line 230]

One strand of evidence that the authors claim for importance of the nucleoside moiety in the formation of ATP is that they perform an experiment with pyrophosphate and do not observe triphosphate formation. They analyse this reaction via 31P NMR spectroscopy and the results are shown in Figure 4f and SI Figure 4. The authors note that this analysis is complicated by the paramagnetism of Fe3+ which can reduce signals in the 31P NMR spectra but nevertheless describe this effect as 'cosmetic and does not conceal the absence of triphosphate in the reaction mixture'.

I have significant concerns with Figure 4f and SI Figure 4. In particular, I am concerned with the result for the reaction of acetyl phosphate with pyrophosphate in the presence of Fe3+ as from the 31P NMR spectra (SI Figure 4, green and purple traces) as no inorganic phosphate Pi is present after 5 h. I find it inconceivable that acetyl phosphate was not hydrolysed during this reaction as acetyl phosphate is shown to hydrolyse significantly in the absence of Fe3+ (SI Figure 4, orange and yellow traces). I am very sceptical that Fe3+ inhibited the hydrolysis reaction - if anything I would expect it to catalyse the hydrolysis. Given the glaring absence of Pi here, I do not accept the authors' description that the effect of Fe3+ is 'cosmetic'. If, for example, the Pi was lost during the solid-phase extractions then could this not also have happened to any triphosphate formed during the reaction? In summary, SI Figure 4 does not substantiate the authors claim that the adenosine is necessary to produce phosphorylation of a diphosphate.

I believe that the authors are correct in their assertion that Fe3+ does coordinate to the adenosine as it is well evidenced in the literature that metal ions coordinate to the adenosine nucleobase (H. Sigel Chem. Soc. Rev., 1993, 22, 255-267) but that the experiment performed here is inadequate to prove this.

Instead I would suggest that the authors consider using EPR spectroscopy measurements on the NDP-Fe3+ complexes to gain mechanistic insight into their very interesting results. EPR spectroscopy can be performed upon solutions and in conjunction with electronic structure modelling studies can be used to give mechanistic detail. As the authors are in the U.K. I would suggest contacting at Professor Victor Chechik at the University of York who I know has extensive experience in EPR spectroscopy and mechanistic studies.

[Page 9 Line 243] and [Page 10 Line 273] and [Page 11 Line 306]

The authors attribute only observing phosphorylation of ADP and not upon other NDPs to the coordination of the N-6 on adenosine to Fe3+ while other nucleobases lack an N-6 amine and therefore are unable to coordinate to Fe3+. This leads the authors to propose the formation of macro chelate complex between the N-6 on adenosine, Fe3+ and the pyrophosphate (Figure 6).

I strongly disagree with the coordination between the N-6 and Fe3+ and therefore I believe their proposed macro chelate complex structure in Figure 6 is incorrect. There is little to no evidence in the literature that the N-6 position of adenosine participates in the formation of a macro chelate complex that the authors describe

Coordination chemistry dictates that the N-6 amine on adenosine should be the second least able nitrogen on adenosine to coordinate to Fe3+ (after the N-9 which is bonded to the ribose). This is because the lone pair on the N-6 position can participate in the aromatic system on the adenosine ring and is therefore poorly available to coordinate. The lone pairs on N-1, N-3 and N-7 are orthogonal to the aromatic ring and therefore can effectively coordinate to metal centres.

Numerous studies have shown that in solution a macro chelate complex is formed between N-7 on adenosine, a metal ion and the pyrophosphate and not the N-6 position. See H. Sigel Chem. Soc. Rev., 1993, 22, 255-267 and references therein for an overview of this topic. For specific evidence of ADP-N-7 macro chelate complex with a metal ion, see H. Sigel J. Am. Chem. Soc. 1983, 105, 5891-5900. This paper also shows that a lower percentage of the M2+ ADP macro chelate complex is in the closed form than for GDP and IDP complexes (see Table V).

The authors dismiss the importance of the N-7 macro chelate complex to the phosphorylation chemistry stating [Page 9 Line 247]: 'We infer that the critical moiety in the adenosine ring for phosphorylation by AcP with Fe3+ as catalyst must be the N6-amino group of adenosine, as the IDP and GDP ring structures are equivalent elsewhere. In particular, from a mechanistic point of view, we note that the N7 is equivalent in all three purine rings, so although this might also interact with Fe3+, as suggested by others [60-63], it cannot be the critical moiety.'

However, the N-7 position on all three purine rings of adenosine, inosine and guanosine are not equivalent from a mechanistic point of view. In their thinking the authors have neglected to consider that replacing the amine with a carbonyl will affect the properties of other N's in the purine ring. Sigel has demonstrated that the basicity of the N-7 position varies significantly as the pKa of the N7-H+ conjugate acid is pKa(Adenosine) = - 0.2, pKa(Inosine) = 1.06 and pKa(Guanosine) = 2.11 (H. Sigel J. Am. Chem. Soc. 1994, 116, 7, 2958-2971). [While this study focused on NMPs I would expect a similar trend in the basicity of the N-7 position for NDPs and NTPs.] Note that this is a log scale so there is a 100-fold difference in the basicity between adenosine and guanosine. The pKa of the N-7 position has a significant effect on the formation of the macro chelate complex with higher basicities leading to higher proportions of the metal ion nucleotide complex being in the macro chelate form (see Figure 4 in H. Sigel J. Am. Chem. Soc. 1994, 116, 7, 2958-2971).

The differences in basicity between the N-7 positions on the purine nucleobases will affect the strength of the coordinate bond (and possibly the extent of backbonding between the metal centre and the aromatic system on the purine) which will therefore affect both the energies and spatial size of the orbitals on the Fe3+. The ability of Fe3+ ADP macro chelate complex to catalyse the phosphorylation of ADP will depend on its ability to stabilise the transition state. I would suggest to the authors that the reason they only observe phosphorylation on ADP and not GDP or IDP is because of these aforementioned differences in the Fe3+ orbitals that happen to be just right to stabilise the phosphorylation transition state with acetyl phosphate.

The lack of phosphorylation on CDP and UDP here can be attributed to CDP and UDP (as well as their NMPs and NTPs) not forming macro chelate complexes with metal ions and the only interaction in these cases is between the phosphate and the metal ion (see Sigel Chem. Soc. Rev., 1993, 22, 255-267). This is explained in part by the nucleobase more favourably adopting an anti conformation (as opposed to the syn conformation) in relation to the ribose which results in the N-3 facing away from the phosphate/pyrophosphate/triphosphate.

Later on in the Discussion [Page 11 Line 369], the authors cite a Mossbauer spectroscopy study for evidence of N-6 binding of Fe3+ to ADP (Ref 60 I. Rabinowitz J Am Chem Soc. 1966, 88, 4346-4354). I do not believe that this study gives adequate evidence for N-6 binding. First, this study was conducted in 1966 and all later studies contradicted the finding of N-6 binding and have shown that the binding is instead to N-7. Secondly, Rabinowitz et al. make the same mistake as the authors in comparing the results from inosine to adenosine to justify the choice of the N-6 position without considering how replacing the amine at the 6-position with a carbonyl will affect the basicity of other N in the purine nucleobase. Thirdly, Mossbauer spectroscopy is conducted on solid samples (in the Rabinowitz paper the pH of the sample was adjusted and then lyophilised to a powder). The change from solution to solid state means that the results are less reliable than the techniques in later studies which analysed the structure of the metal ion ADP complexes in solution.

In summary, I believe the weight of evidence is that an ADP N-7 macro chelate complex is formed in solution and not the N-6 macro chelate complex and consequently the authors need to adjust their text and Figure 6 to reflect this.

[Page 9 Line 257]

The authors show that the addition of low Mg2+ and Ca2+ concentrations to the reaction leads to higher yields of ATP. The authors justify this increased yield by stating that the Mg2+ and Ca2+ can stabilise the ATP to hydrolysis which is the explanation also provided by Kitani for this phenomenon. However, the authors go further and claim that the addition of Mg2+ and Ca2+ promote catalyst turnover by liberating the Fe3+. This would suggest that the rate determining step for the reaction is the liberation of the Fe3+. I am sceptical of this claim as I would have thought the phosphate transfer from AcP to ADP would be the more likely rate determining step. I would favour the explanation that the Mg2+ and Ca2+ stabilise ATP to hydrolysis. If the authors wish to prove that Mg2+ and Ca2+ are present in the rate determining step in the catalytic cycle then they would need to demonstrate that the reaction order is dependent on these ions.

An alternative explanation for the dependence on Mg2+ and Ca2+ is that they participate in the formation of a dimeric complex (e.g. FeMg[ADP2]) which is also catalytically active.

[Page 10 Line 261]

The authors make an assessment of whether Fe3+ is acting as a catalyst by measuring the rate of the reaction with different concentrations of the ADP substrate and then apply a Michalis Menten fitting to the data (Figure 5b). In these experiments the authors use [Fe3+] = 0.5 mM and they don't state the concentration of AcP but I presume it is 4 mM. The ADP concentration is then varied between approx. 0.1 mM - 4 mM. In these circumstances the application of a Michalis Menten kinetic analysis is technically speaking not appropriate as the Michalis Menten kinetic analysis assumes that the substrate concentration is in significant excess over the catalyst concentration (a 10-fold excess of substrate is usually regarded as the minimum viable). This is because the Michalis Menten kinetic analysis assumes pre-equilibrium formation of the enzyme substrate complex and rate determining kcat. In the experiments performed by the authors the concentration of ADP is either lower than Fe3+ or at most in an 8-fold excess. Thus, this Michalis Menten fitting is strictly speaking not valid although I do accept that the level off is indicative of catalysis.

Like for Figure 1, insufficient detail is provided by the authors as to how they measured the rates of reaction here. I appreciate the effort the authors have gone to in order to measure the rate of reaction and this makes a good addition to the study but I would like to see the kinetic data and equation used to determine the rate constant for the data in Figure 5b put into the Supplementary Information so that I can properly assess this aspect of the study.

[Page 10 Line 267] and [Page 11 Line 313]

The authors claim that a dimeric structure of ADP with Fe3+ (SI Figure 6) is not the catalytically competent species based upon a MALDI-TOF analysis which did not show the stacked species.

I would dismiss the validity of using MALDI-TOF to assess this claim on two criteria.

First, the MALDI-TOF is performed on a solid sample in a matrix and not upon a solution and thus the species present may be altered by the change in state.

Second, the MALDI analysis requires sample preparation steps that could have significantly interfered with this stacked dimer species. The main drive force behind why the stacking occurs in solution will be to minimise the hydrophobic surface area presented to the solvent by the nucleobases. During the MALDI sample prep the sample is washed with a 50:50 acetonitrile:water solution and acetonitrile is a significantly non-polar solvent it could disrupt the hydrophobic stacking and thereby break this stacked dimer apart.

I think the authors should remove this MALDI study from the paper as it is not possible that this experiment can give any effective insight into whether a dimer is formed during the reaction in solution.

Furthermore, the evidence from the literature that metal complexes can form dimeric structures (with intermolecular adenosine N-7 metal ion interactions) and can also stack is fairly substantial, see H. Sigel J. Am. Chem. Soc. 1983, 105, 5891-5900. I would therefore not dismiss the possibility that the catalytically competent species for the phosphorylation reaction is the dimeric form. Note too that the optimal 1: 1 stoichiometry for the reaction applies to both the monomeric complex (1:1) and the dimeric complex (2:2) and therefore cannot be used to distinguish between these mechanisms.

[Page 11 Line 301]

As acknowledged by the authors, there is not a link between Fe3+ and the phosphorylation of ADP to ATP in biology. I would agree with this as to the best of my knowledge ATP synthesising enzymes do not make use of Fe3+ as a cofactor. For example, the acetate kinases in Ref 41 use Mg2+ as a cofactor. The authors advocate a position on the origins of life whereby the chemistry out of which life emerges should align as closely as possible to the biochemistry of extant life. If Fe3+ were to play such a crucial role at the origins of life then why would this not have been retained in biology? Such a question is raised by the authors but they do not provide any adequate answer to it.

[Page 11 Line 296] and [Page 15 Line 447]

At the opening of the Discussion the authors state that: 'Taken together, these findings suggest that the pre-eminence of ATP in biology has its roots in aqueous prebiotic chemistry. … This implies that ATP became the universal energy currency of life not as the endpoint of genetic selection or some frozen accident, but for fundamental chemical reasons, and probably in a monomer world before the polymerization of RNA, DNA and proteins.'

And then end the Discussion stating: 'If so, then ATP became established as the universal energy currency for reasons of prebiotic chemistry, in a monomer world before the emergence of genetically encoded macromolecular engines.'

The authors claim that their results justify why ATP became the universal energy currency for life and moreover did so before the advent of enzyme catalysts. This is a bold claim and unfortunately the authors have significantly overextend themselves in making it as such a claim cannot be fully supported by their results.

My main objection here is that the authors have only demonstrated half of what is necessary to solve why ATP became Life's universal energy currency. To justify why ATP became Life's energy universal currency at the origins of Life requires consideration of both how ATP could have formed and how ATP came to be not just the dominant phosphorylating NTP but also the dominant phosphorylating reagent at the origins of Life.

The authors' results do provide a plausible explanation for why the formation of ATP became dominant over the other NTPs but this is as far as the authors can reasonably take their conclusions based upon their results.

The huge oversight by the authors here is that they do not seriously consider how ATP performed prebiotic phosphorylations. This receives scant attention in the paper and is only addressed in the penultimate sentence. The authors state [Page 15 Line 443]: 'Once formed, ATP would promote intermediary metabolism through phosphorylation'

This statement trivialises an enormous challenge in prebiotic chemistry which is how to perform prebiotic phosphorylation reactions. For example, a central challenge to prebiotic phosphorylations is how to overcome the high activity of water which results in the hydrolysis of the phosphorylating reagent rather than transfer of phosphate to an organic molecule. The authors show that solutions with a high water activity favour formation of ATP [Page 6 Line 166], however such conditions will be the least effective for the phosphorylation of organic compounds meaning that the authors will almost certainly need a very different set of conditions for the phosphorylation by ATP. Under these different conditions would ATP necessarily be a better phosphorylating reagent than the other NTPs? Furthermore, I would argue that ATP is a very poor choice as a prebiotic phosphorylating reagent. This is because ATP is a very stable compound that has a hydrolysis half-life of around 2 years at pH 8 and 25 oC in synthetic sea water (H. Hullet Nature 1970, 225, 1248). Given the high stability of ATP I would not expect its use in prebiotic phosphorylations until after the advent of macromolecular engines e.g. a ribozomyl or enzyme catalyst.

If the authors wish to demonstrate why ATP became Life's universal energy currency then they must consider questions such as the following and address these experimentally:

Why did life settle upon NTPs to perform reactions in prebiotic chemistry?

Why did ATP become dominant phosphorylating reagent over the other NTPs?

How did phosphorylation of organic compounds by ATP overcome the high activity of water?

If, as the authors show, ATP can be selectively formed be a metal ion catalysed reaction could it not also perform metal ion catalysed phosphorylations that other NTPs cannot?

Given the stability of ATP/NTPs Is it not possible that ATP is chosen 'late' in the origins of Life after the origin of macromolecular engines?

The experimental results presented in this paper do not address these questions and therefore do not tackle how ATP could perform prebiotic phosphorylations and thus for the authors to claim that their results show why ATP became Life's universal energy currency is not justified. The authors need to constrain their conclusions to what their experimental results can actually prove.

The authors must also consider how prebiotic chemistry also produced the other NTPs as Life has settled on the use of NTPs for synthesis of DNA and RNA. This requires an alternative route(s) to NTPs. If this alternative route(s) to NTPs produced all NTPs at a higher rate than the Fe3+ catalysed phosphorylation of ADP by acetyl phosphate then this would undermine the authors' case that ATP was formed in excess over other NTPs.

Minor considerations:

In the caption of Figure 4 it would be better to call it a nucleobase rather than base.

For Table 1, the authors may wish to cite a recent paper on the geological presence of cyclic trimetaphosphate: Britvin, S. N. et al. Cyclophosphates, a new class of native phosphorus compounds, and some insights into prebiotic phosphorylation on early Earth. Geology 49, 382-386 (2020) https://doi.org/10.1130/G48203.1

[Page 8 Line 207] with regards to cyanate being a condensing agent, in addition to ref 58 the authros may wish to also cite R. Pascal J. Am. Chem. Soc. 2006, 128, 7412-7413.

Mathew Pasek wrote a very good recent review of prebiotic phosphate chemistry that the authors may wish to cite: Chem. Rev. 2020, 120, 11, 4690-4706 10.1021/acs.chemrev.9b00492

Reviewer #3:

This is an interesting paper which relates to a chemical, non-enzymatic synthesis of ATP from ADP and Pi using acetyl phosphate and ferric ion. From a chemical perspective alone, the work is not necessarily original but it is more in the context of prebiotic chemistry and abiogenesis that the work is couched and consequently, it has been reviewed here in that context.

It is pertinent to comment first on the authors contextual history in the field which is significant. The authors have a long-standing presence in the sphere of prebiotic chemistry, specifically linked to the role of concentration and charge gradients as a source of energy to drive the emergence of biology.

The present contribution centres on the conversion of adenosine diphosphate (ADP) and inorganic phosphate (Pi) to adenosine triphosphate (ATP) using acetyl phosphate and ferric ion as co-factors, without the need to invoke multi-subunit protein complexes as catalysts. There a number of intriguing aspects to this work, including (i) the specificity of acetyl phosphate in this process (although carbamoyl phosphate also shows some activity) and (ii) the specificity of ferric ion. Both of these are sufficiently intriguing as to ask why these components appear to be so unique and what is the level of significance in this from a pre-biotic perspective.

From a practical methods perspective, these looks perfectly sound to me, chemically and analytically. Technically, relatively little discussion is devoted to the kinetic aspects of the study but the focus is more on thermodynamics. From a prebiotic energetic context this is certainly of value, however, within the context of dynamic kinetic behaviour as in biology, it tells only a part of the story.

In terms of comments, I have two to pass on to the authors for their consideration: (i) it is clear that the authors place great emphasis on the specificity of both acetyl phosphate and ferric ions in this chemistry, it is central to the actual conclusions. Figure 6 proposes a potential mechanism for how this might work. However, what is rather surprising missing is a more thorough molecular modelling analysis of just how reasonable this actually is. This is a significant omission and one that could logically be quite straight-forward to add. I would not suggest full molecular dynamics calculations be necessary at his stage but simple bench-top molecular mechanics modelling should be included. (ii) The second point is rather more contextual, but goes to the central importance of the present chemistry in primitive energy transduction. The authors emphasise the value of ATP as a primitive energetic "battery". This has some validity of course, but I feel it really does need to be nuanced a little more for the wider and especially biological audience. For sure, ATP can act as an energetic battery in terms of direct phosphorylation. However, the situation becomes more complex when it comes to linking ATP/ADP to proton gradient or indeed to enzymatic phosphorylation-dephosphorylation. The energy which results from ATP hydrolysis in biological systems is transduced via multi-subunit complex enzymes and converted to mechanical energy through a rotatory mechanism. Because the actual hydrolysis reaction of ATP is enzymatically very fast (femtosecond), it is generally believed that this chemical hydrolysis step itself is NOT the step in the whole process where energy is transduced, but rather the binding-de-binding of ADP/ATP components from the enzyme binding sites themselves. This is what allows the ATP hydrolysis to be connected to proton gradients. Non of this is made clear in the current analysis which makes it a little less clear just how significant pre-enzymatic energy transduction using ATP might have been and what could have acted as a pre-enzymatic catalyst and then also, how was that energy transduced from ATP to a gradient or other usable energy source. With a little more nuance, the authors might be able to place their work in a slightly more realistic context within the pre-biotic milieu.

Reviewer #4:

In "A prebiotic basis for ATP as the universal energy currency" the authors present an intriguing argument that the origin of ATP as the dominant energy carrier for prebiotic chemistry resides in the fact that acetylphosphate can phosphorylate ADP to form ATP (and ADP alone!) in the presence of catalytic Fe3+. Indeed, the authors demonstrate that all three of these components are uniquely situated to perform this task: changing any nucleobase, metal catalyst, or phosphorylating agent and the yields plummet. Given the specificity of adenosine in this task, the authors argue that ATP was predisposed to be the dominant energy currency of life.

In general, the experiments are well-thought out and easy to follow. The analytical tools appear reasonable, within the limitations of these instruments. The reasoning within the discussion is sound enough (in general) for the experiments that have been carried out.

I have several concerns regarding the conclusion, other phosphorylation agents that were not discussed in the text, the prebiotic geochemistry of Fe3+, and some of the methodology and reporting of results. These are explored below.

1. The main point of the paper is that ADP to ATP occurs readily only by phosphorylation by acetylphosphate with catalytic Fe3+, under a specific set of aqueous conditions (acidic pH, T <= 50°C, lacking too much salt or divalent cations). Thereby, the conclusion is that ATP as the dominant is not a "frozen accident" but is a consequence of the environment in which life originated. However, such a conclusion appears to be a circular argument: the specificity of conditions needed to generate ATP from ADP means that ATP was specifically chosen because of those conditions. If the conditions are not met on the prebiotic earth (e.g., no Fe3+, no acetylphosphate), then no ATP would be specific. There are other routes to generating prebiotic NTPs (detailed in points 2+3). If ATP is a "frozen accident" then these routes would be just as plausible. To some extent, this point is allayed by the discussion on acetylphosphate, as acetylphosphate is present in modern biology. However, if we presume the earth was somehow "different" with respect to its prebiotic geochemistry (see Benner et al. ChemSystemsChem 2020), and thereby other reactants not observed in geochemistry or biochemistry today would have been plausible (e.g., cyanate as discussed by the authors), then prebiotic chemistry need not follow modern biochemistry.

Prebiotic chemistry is often OK with "frozen accidents" as well, as the chirality of biomolecules is generally assumed to be a consequence of the possibly arbitrary higher abundance of molecules with the biological handedness. Also, some aspects of the genetic code (while highly evolved) were likely randomly frozen in as well.

2. The paper is lacking discussion of the amidophosphates, which have been explore by the Krishnamurthy lab for their ability to phosphorylate organics. Gibard et al. (Nat Chem 2018) demonstrated phosphorylation of NDPs to NTPs using diamidophosphate. Recently, Lin et al. (Ange. Chem 2021, very recently published, but building on Gibard et al. 2018) demonstrated NMPs are transformed into NTPs via reaction with diamidophosphate. Amidophosphates have a parallel in modern biochemistry in forms such as creatine phosphate.

3. Triphosphorylation of nucleosides via cyclic trimetaphosphate and a ribozyme has been reported by Akoopie et al. (Sci Adv 2021) and presents an alternative explanation for NTPs: post-RNA world (Moretti and Muller 2014). Although trimetaphosphate has its own issues with prebiotic provenance (e.g., Keefe and Miller 1995), some routes have been shown to these compounds (Yamagata et al. Nature 1991, Pasek et al. ACIE 2008). Kim and Benner (Astrobiology 2021) also showed NTPs from trimetaphosphate.

4. The prebiotic geochemistry and aqueous chemistry of Fe3+ is not adequately considered. Fe3+ is generally soluble only at low pH, well below the pH used by the authors. Iron redox diagrams (for instance, see Takeno 2005, Atlas of Eh-pH diagrams, Figure 47) show that soluble Fe3+ exists only at pH <2.3. Above a pH 2.3, Fe3+ reacts with water to produce Fe(OH) 2+, and then Fe(OH)2 +. These are not Fe3+. This chemistry is roughly reflected in the results presented by the authors. At the higher pH, Fe3+ precipitates out as Fe(OH)3 or FeOOH. The drop in yield at these higher pH values likely corresponds to a precipitation of Fe3+ or the production of neutral Fe species (such as Fe(OH)3 and FeOOH). The actual aqueous ionic chemistry of Fe3+ at the conditions being explored (Figure 6) is likely completely different than those being discussed here as Fe3+ (aq) isn't Fe3+ at a pH of 5.5.

To this end, it's also unclear if the pH of the solution was adjusted before or after the addition of Fe3+ (as Fe2(SO4)3). I presume after, but if before, Fe3+ is an acidic cation that will drive the pH to a lower value.

Furthermore, Fe3+ requires highly oxidizing conditions at low pH, well above the point of the H2S/SO42- redox couple (a full 0.6 V at pH 2), and given that thioesters are discussed in the context of this prebiotic chemistry, these highly disparate redox conditions themselves would introduce disequilibrium. The arguments for Fe3+ could also be expanded in the text, as Fe3+ also forms from the reaction of Fe2+ with H2O2, the latter formed by the photolysis of water (via Fenton chemistry).

In addition, Fe3+ is unstable in non-acidic water, eventually precipitating out as hematite (Fe2O3). The reaction 2Fe3+ + 3H2O = Fe2O3 + 6H+ has a reaction K of around 0.7 at 30°C, which means that at a pH of 5.5., the concentration of Fe3+ should be sparingly small at equilibrium. That's not to say solutions can't be supersaturated with respect to Fe3+, but that high concentrations of Fe3+ should be geologically ephemeral.

5. The chemistry of iron (III) is more closely replicated by aluminum (III) than by the other species investigated. The ionic size of Al3+ is similar to Fe3+ and the chemistry of these ions in water vs. pH is similar. It may be worth replicating these experiments with Al3+ to see if Fe3+ is still unique in its chemical potential for ADP to ATP, especially in comparison to the species chosen (Co3+, Cr3+, and Mo3+ are all quite different from Fe3+ in aqueous chemistry).

6. The results do not discuss the possibility of precipitates of the various components investigated. Does iron(III) form red FeOOH precipitates? Is there any precipitation of Mg-ADP? Is phosphate being precipitated by the cations? This is also a question of mass balance: Is the amount of phosphorus in ADP and AcP that was added to the beginning of the reaction the same as the total organophosphate that came out? If not, were there precipitates of some of these ions (perhaps ATP in some of the Mg/Ca experiments?)? If no precipitates were noticed, then this should be noted in the text. The Fe3+ should certainly provide precipitates, given the "gumminess" of this ion.

7. The study of the pyrophosphate phosphorylation (figure SI 4) could be improved significantly by changing the solution pH of the NMRs acquired. The spectra are quite broad, as acknowledged by the authors. However, given the multiplicity of the PPPi peak, and its expected low concentration (~20% of ADP was converted in these experiments) would a triphosphate peak even be visible if it was at ~5-10% yield in such conditions? The authors could attempt to precipitate Fe3+ using NaOH or Na2S to sharpen up the spectra and investigate for PPPi. As it is, the current spectra do not provide confidence in either its presence or absence.

COMMENTS FROM THE ACADEMIC EDITOR:

The manuscript by Pinna et al. characterizes a previously described reaction (as the authors note) whereby acetyl phosphate phosphorylates ADP to produce ATP, catalysed by Fe3+. However, the authors then go on to demonstrate the specificity of the reaction for ADP over other ribonucleotides. That result on its own makes the work publishable (generally speaking). The manuscript is written eloquently, and the message of the authors is clearly expressed. However, there are serious flaws in the interpretation of the data and in the placement of the data in a broader context. Therefore, in my opinion, the manuscript would need to be substantially revised for consideration.

All of the reviewers gave constructive criticisms that should be considered. Reviewer 2 provides a longer, extensive list of criticisms. I do not find the requests unreasonable, as several of the points seek to clarify and/or correct shortcomings. However, I would not push for the EPR analysis to be performed. EPR would be insightful, but that would take much time, and most of the criticisms can be addressed without EPR.

As the reviewers have done a thorough job, I will only briefly comment on a few of the points raised. The complex described in Figure 6 is questionable. If this were to be taken more seriously, then some type of modelling and additional experiments would be needed. As reviewers 1, 2 and 3 noted, and as is well known in coordination chemistry, the coordination of metal ions to the N6 position of adenosine is not reasonable since the electrons of N6 are distributed across the aromatic ring. Even when deprotonated at high pH, this position is a poor ligand.

It is also necessary to run a control demonstrating that ADP + ADP � ATP + AMP is not occurring, particularly since this reaction has been documented before (as reviewer 2 notes).

As reviewers 1 notes, the data are consistent with a “surface pond or lake” rather than a hydrothermal vent. It would be better if the authors let their own data guide them rather than to force interpretations onto previously espoused convictions. I can understand if the authors do not want to advocate for surface pond or lake conditions, but a more fair and balanced interpretation of the data are warranted and would contribute well to the field.

Regarding more fresh water conditions, on pages 4-5 the manuscript states “But the fact that substrate-level phosphorylation of ADP to ATP can be accomplished by AcP in water says nothing about whether his mechanism actually holds prebiotic relevance.” This is a fair point, but the conditions used by Pinna et al. are not necessarily anymore prebiotically plausible. Basically, extremely low levels of salt are necessary to allow for the reaction to occur. The fact that NaCl and MgCl2 greatly inhibits the reaction is alarming in terms of prebiotic relevance, especially if you are a proponent of hydrothermal vents. The fact that 1-2 mM Mg2+ promotes the reaction does not strengthen the arguments. Environmental concentrations of Mg2+ are much higher. Even the concentration inside of a cell is much higher. Unless one were to invoke freshwater conditions, the inhibition by Mg2+ (and Na+) suggests that the data are not relevant to prebiotic chemistry, as it is difficult to imagine how such chemistry could have been mediated on the prebiotic earth.

To help the reader, it would be good to indicate what pressures are typically found in hydrothermal vents. That way, the reader would understand what kind of environments the 80 bar data correspond to.

As reviewer 4 notes, it is alarming that some highly relevant work was overlooked, including work from the Huck group (Maguire et al. Nat Commun 2021, 12, 5517) and the Krishnamurthy group (Gibard et al. Nat Chem 10, 212–217). These are extremely relevant studies that explore prebiotic phosphorylation. If a fair and balanced perspective were to be taken, then the implications of such work should be taken into consideration.

Similarly, the neglect of the potential importance of cyanate further suggests biases. This is even more evident since the dismissal of cyanate is inconsistent with how the authors seem to view prebiotic chemistry, in that things that are not found in biology are not considered. Using the same logic, we should dismiss a role of Fe3+ in prebiotic phosphorylation, since that is not found in biology.

Reviewer 4 raises several points regarding the solubility and availability of Fe3+. Additionally, the photooxidation of Fe2+ to Fe3+ has been experimentally demonstrated (Bonfio et al. Nat Chem 9, 1229–1234).

Finally, as reviewer 1 notes, the speculation that ATP would be synthesized at acidic pH and then enter a protocell with an internal alkaline pH does not appear logical. Nucleotide triphosphates are not permeable across membranes that are thought to be prebiotically plausible. If membranes were stable to high concentrations of Mg2+, then perhaps enough Mg2+ would be bound to neutralize the charge, but such concentrations of Mg2+ would inhibit phosphorylation of ADP to ATP. It is difficult to imagine a scenario in which the gradients that the authors wish to invoke were established.

There is something interesting in the submitted manuscript. The specificity for ADP is new and definitely interesting. The details of the chemistry need to be improved, and a fairer representation of the implications of the work and how that fits into prior studies is needed.

Attachment

Submitted filename: The manuscript by Pinna et al.pdf

Decision Letter 2

Roland G Roberts

5 Aug 2022

Dear Nick,

Thank you for your patience while we considered your revised manuscript "A prebiotic basis for ATP as the universal energy currency" for publication as a Research Article at PLOS Biology. This revised version of your manuscript has been evaluated by the PLOS Biology editors, the Academic Editor, and three of the original reviewers. As intimated in my previous email, the original Academic Editor was not able to continue handling your paper, so we have assigned a new Academic Editor to help us with these final stages.

Based on the reviews, we are likely to accept this manuscript for publication, provided you satisfactorily address the remaining points raised by the reviewers. Please also make sure to address the following data and other policy-related requests.

IMPORTANT:

a) Please address the remaining concerns from the reviewers.

b) Just in case this is helpful for formulating your revisions, when I discussed the reviewers' comments with the Academic Editor, they said "R1 raises important clarification points on the charge of ADP, and how purine nucleotide synthesis could still be reasonably still occur on minerals presuming, for instance, the presence of nucleosides. The presence of nucleosides has been investigated for a substantial period of time, and though biologically it requires a rather large quantity of metabolic energy, there’s been a long history of thinking about nucleotide formation abiotically. From Oro’s early experiments showing adenine synthesis from HCN, to the formation of ribose from formaldehyde, to condensation and phosphorylation, there have been several routes to making and assembling nucleotides that do not require biology that have come from a prebiotic chemistry perspective. This would be a more heterotrophic view of the origin of life than suggested by the autotrophic view of the authors, and both should be viewed as reasonable schools of thought."

c) Please address my Data Policy requests below; specifically, we need you to supply the numerical values underlying Figs 1ABC, 2ABCD, 3ABCDEF, 4ABCDEF, 5ABC, 6ABCD, S1AB ,S2, S3, S4AB, S5, S7ABCD, S8, either as a supplementary data file or as a permanent DOI’d deposition like Zenodo, Dryad, Figshare, etc..

d) Please cite the location of the data clearly in all relevant main and supplementary Figure legends, e.g. “The data underlying this Figure can be found in S1 Data.”

As you address these items, please take this last chance to review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the cover letter that accompanies your revised manuscript.

We expect to receive your revised manuscript within two weeks.

To submit your revision, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' to find your submission record. Your revised submission must include the following:

- a cover letter that should detail your responses to any editorial requests, if applicable, and whether changes have been made to the reference list

- a Response to Reviewers file that provides a detailed response to the reviewers' comments (if applicable)

- a track-changes file indicating any changes that you have made to the manuscript.

NOTE: If Supporting Information files are included with your article, note that these are not copyedited and will be published as they are submitted. Please ensure that these files are legible and of high quality (at least 300 dpi) in an easily accessible file format. For this reason, please be aware that any references listed in an SI file will not be indexed. For more information, see our Supporting Information guidelines:

https://journals.plos.org/plosbiology/s/supporting-information

*Published Peer Review History*

Please note that you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

*Press*

Should you, your institution's press office or the journal office choose to press release your paper, please ensure you have opted out of Early Article Posting on the submission form. We ask that you notify us as soon as possible if you or your institution is planning to press release the article.

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Please do not hesitate to contact me should you have any questions.

Best wishes,

Roli

Roland Roberts, PhD

Senior Editor,

rroberts@plos.org,

PLOS Biology

-------------------------------------

DATA POLICY:

You may be aware of the PLOS Data Policy, which requires that all data be made available without restriction: http://journals.plos.org/plosbiology/s/data-availability. For more information, please also see this editorial: http://dx.doi.org/10.1371/journal.pbio.1001797

Note that we do not require all raw data. Rather, we ask that all individual quantitative observations that underlie the data summarized in the figures and results of your paper be made available in one of the following forms:

1) Supplementary files (e.g., excel). Please ensure that all data files are uploaded as 'Supporting Information' and are invariably referred to (in the manuscript, figure legends, and the Description field when uploading your files) using the following format verbatim: S1 Data, S2 Data, etc. Multiple panels of a single or even several figures can be included as multiple sheets in one excel file that is saved using exactly the following convention: S1_Data.xlsx (using an underscore).

2) Deposition in a publicly available repository. Please also provide the accession code or a reviewer link so that we may view your data before publication.

Regardless of the method selected, please ensure that you provide the individual numerical values that underlie the summary data displayed in the following figure panels as they are essential for readers to assess your analysis and to reproduce it: Figs 1ABC, 2ABCD, 3ABCDEF, 4ABCDEF, 5ABC, 6ABCD, S1AB ,S2, S3, S4AB, S5, S7ABCD, S8. NOTE: the numerical data provided should include all replicates AND the way in which the plotted mean and errors were derived (it should not present only the mean/average values).

IMPORTANT: Please also ensure that figure legends in your manuscript include information on where the underlying data can be found, and ensure your supplemental data file/s has a legend.

Please ensure that your Data Statement in the submission system accurately describes where your data can be found.

------------------------------------------------------------------------

DATA NOT SHOWN?

- Please note that per journal policy, we do not allow the mention of "data not shown", "personal communication", "manuscript in preparation" or other references to data that is not publicly available or contained within this manuscript. Please either remove mention of these data or provide figures presenting the results and the data underlying the figure(s).

------------------------------------------------------------------------

REVIEWERS' COMMENTS:

Reviewer #1:

[new comments, indicated by "R1':" are interdigitated]

R1: The manuscript by Pinna et al. reports on the detailed characterization of the phosphorylation of ADP by acetyl phosphate (AcP) to form ATP in the presence of ferric ions. This previously observed reaction appears to be surprisingly specific as no other NDPs or transition metal ions can carry out equivalent reactions to any significant extent. ADP phosphorylation appears to be optimal at 30°C, mildly acidic pH and low salt. The authors provide explanations for their observations within the context of their already well-established hypothesis of life's origins in a hydrothermal vent. Besides the fact that, taken face value, the observations reported here would rather favor a surface pond or lake as the cradle of life, some of their propositions need to be further substantiated.

R: We thank R1 for this clear synopsis. We agree that taken at face value our observations might seem to favour freshwater environments and we make this clearer in the revised version (p14 lines 401-408). We have also revised the order of the Discussion to bring up this passage before the section on work, as the paragraphs on work may have obscured this point in the original version.

R1': OK.

R1: Page 3 "ATP in a prebiotic, monomeric world". This assumes that life originated in bulk water. But, if, instead, life first appeared on mineral surfaces then the needed energy could have been provided through the oxidation of H2 and CO by, for instance, catalytic FeNiS minerals. Only when more evolved protocells liberated themselves from those surfaces, and moved to bulk water, would have ATP become essential.

R: We have revised the Introduction to make our reasoning clear here. We agree with R1 that the earliest forms of metabolism are indeed exergonic and could have occurred on mineral surfaces. But it is unlikely that this could account for purine nucleotide synthesis, for example, which has six phosphorylation steps. We lay out this perspective in the revised Introduction, specifically on pages 3-4, lines 69-83.

R1': In fact, there is no a priori reason to conclude that purine nucleotide synthesis could not have taken place on a mineral surface. The authors should check the paper by Deiana et al. (ChemCatChem 2013, 5, 2832-2834.) who argue that "… the interactions of carboxylate oxygen atoms with surface Ti4+ ions, which act as Lewis acid centers, is expected to withdraw electron density from the C atom, which becomes electrophilic enough to undergo nucleophilic attack by the nitrogen atom of the amine". This is exactly the effect of phosphorylation on the carbonyl carbons of reaction intermediates, which allows for the formation of 9 C-N bonds during purine synthesis. Although acetyl phosphate could have been involved in ATP synthesis other pathways are also possible (see Akouche et al. Angew. Chem. Int. Ed. 2017, 56, 7920-7923 for the remarkable one-pot synthesis of AMP from phosphate, adenine, and ribose on a fumed silica surface).

R1: "Even if only one nucleotide triphosphate can be dominant, the implication of a frozen accident is not a satisfying explanation." Why not? There are different levels of "frozen accidents". This has been discussed by Brandon Carter who, in 1984, proposed that the evolution towards intelligent life required 6 "hard steps", the first one being abiogenesis. Each step was statistically postulated to last between 600 and 800 million years. Although these periods of time may be considered workable in astronomy and geology, in biology they do not make the same sense. In the latter case it is necessary to invoke the occurrence of extremely unlikely (highly contingent) events to explain that time duration. They can be indeed defined as "frozen accidents".

R: We do not agree with R1 about frozen accidents, but this is a relatively trivial point in relation to the paper. We have de-emphasised frozen accidents in the revised Introduction, and instead stressed the centrality of other adenine nucleotide cofactors including NAD, FAD and coenzyme A, all of which point to the importance (and so availability) of adenosine or ATP in early metabolism.

R1': As the authors point out there is no doubt that adenine nucleotide must have been around quite early. A "frozen accident" in this context describes the -most likely- very low global probability for the required reactants to find themselves in the same location under productive conditions. Once that accident happens subsequent evolution will propagate the result in both time and space if it confers an advantage.

R1: Page 13. "Such steep pH gradients could in principle operate across protocells as well as inorganic barriers." The authors propose that ATP would have been synthesized outside protocells, under acidic conditions, to be then transported across a proto-membrane into a cell where, under alkaline conditions, it would have phosphorylated some substrate(s). Have they considered the implications of such a mechanism in terms of ATP/ADP concentrations and diffusion? ATP would have to diffuse towards the protocell and ADP do it in the opposite direction, without risking to disappear into the bulk water. Another crucial point is membrane permeability. As F. H. Westheimer wrote in 1987: "Phosphoric acid is specially adapted for its role in nucleic acids because it can link two nucleotides and still ionize; the resulting negative charge serves both to stabilize the diesters against hydrolysis and to retain the molecules within a lipid membrane. A similar explanation for stability and retention also holds for phosphates that are intermediary metabolites and for phosphates that serve as energy sources." So, one very important role of even proto-membranes should have been to keep charged metabolites inside the cell to sustain metabolism. This, in turn, would have made ADP diffusion out of the protocell difficult to envision.

R: This is a complete misunderstanding of what we had attempted to write, and we can only apologize for not being sufficiently clear. We agree with R1 (and Westheimer's classic paper) about phosphorylated intermediates, including especially triphosphates, not being able to cross cell membranes, even simple fatty acid bilayers. In fact, we envisaged ADP being phosphorylated inside the protocells immediately adjacent to the membrane, subject to proton influx from the acidic exterior. No phosphates ever cross the membrane in our model. We have made this clear in the revised Discussion (page 15, lines 435-443).

R1': I do not think that 'misunderstanding' is the right way to describe the problem in the original text (p 13). There, the authors wrote "…that thin inorganic barriers … can sustain proton gradients…". And later they concluded "… could promote the phosphorylation of ADP to ATP under locally acidic conditions close to the barriers, followed by hydrolysis linked to phosphorylation under more alkaline conditions in the cytosol of protocells". The question is, where are the "locally acidic conditions close to the barriers"? In the cytosol, as well? So, is there an intracellular gradient? This would be quite different from standard biological proton gradients.

The corrected text is fine.

R1: Page 15. "…enabling ATP to drive work even in a prebiotic monomer world." How are the authors defining 'work' in that setting? In contemporary biology work is i) performed mechanically as in muscle contraction and ii) against a gradient of chemical potential in active transport (see Jencks, 1989). This work depends on complex protein-based binding energy modulation and the coupling of intermediate reaction steps. Of course, this kind of work could not have existed prebiotically. At that time the most likely role of ATP (if it already existed) would have been, for instance, phosphorylation to render carbonyl carbons electrophilic enough to be attacked by a nucleophile such as ammonia. It is debatable that this process would qualify as work.

R: Our perspective is grounded in the idea of a monomer world, and we do not envisage any protein-based machinery here at all. We have made this clearer at several points in the revised MS and added a new paragraph to the end of the Discussion (page 15 lines 451- 464) where we discuss specifically what we mean by work in this context (essentially coupling exergonic to endergonic reactions, and in particular the polymerization of nucleotides and amino acids).

R1': OK

R1: Page 31. Figure 6. Here and in the text, it is implied that the N6 -NH2 group interacts with Fe3+. In A-T base pairing N6 interacts with the O atom from the carbonyl group of T through a polarized �+ proton. What would be the driving force in its postulated interaction with Fe3+? In b, the metal ion is hexacoordinated but the coordination geometry is not clear. In any case, it doesn't look octahedral. Maybe the authors should do some modeling to check whether the proposed arrangement is plausible. An experimental confirmation of their model would be even better.

R: We thank R1 for pointing out that base pairing in DNA and agree that our model as originally depicted was incorrect. We have also undertaken a series of molecular dynamic simulations, as suggested by R1, which do indeed strengthen our mechanistic interpretations. We discuss these new MD simulations at length in the Results section on page 10, lines 259-288, in two new figures (Fig. 6 and SI Fig. 7), and in the Discussion on page 12 lines 330-337. We use this new information to modify our proposed mechanism, specifically in relation to the N6 amine group and the N7, which is presented in Fig. 7. These new sections also respond to the concerns of other reviewers.

R1': I think there is a problem with the overall charges. In Fig. 7a ADP carries 3 negative charges, one on its alpha phosphate and two on its beta phosphate. In fact, one of the O- of the beta phosphate is the nucleophile that initiates ATP synthesis. However, in both the section about MD simulations and the legends of Figs. 6 and 7, ADP is described as having a 2- valence. Because the MD result is based on the electrostatic properties of the reactants one negative charge difference should have a major impact. This has to be clarified and revised.

In Fig. 7b the ferric ion might be tetrahedral but, again, the coordination sphere is not well defined. It may be out of the scope of this paper, but better modeling would come out of a QM/MM analysis of the proposed structures.

R: We thank R1 again for their helpful comments and hope that we have addressed them to their satisfaction. The revisions very clearly improve the paper.

R1': You are welcome and please fix the remaining problems.

Reviewer #3:

The authors have made considerable, significant and positive changes to their original submission. It is clear from the combined reviewers comments that many different features of this paper have benefitted from these changes.

From the perspective of this reviewer in particular, (i) the mechanistic aspects of ADP phosphorylation via AcP in the presence of Fe3+ has now been nuanced significantly, in line with accepted chemical precedence. (ii) Discussions around the "value" of mild, non-enzymatic ADP phosphorylation to a burgeoning prebiotic "organism" have also been modified and some additional caution introduced.

The question as to how, where and when nucleotide phosphorylation became a fundamental component of primitive biological organisms is a foundational one in origin-of-life studies and its discussion is relevant to the current paper. However, the experimental offerings here are not focused on addressing that question, but rather on questions of chemistry. In this regard, I'm happy that the authors have sufficiently strengthened these chemical features, whilst modulating the language around potential "value" judgements which are less well substantiated.

Reviewer #4:

The authors have taken into account prior recommendations and made changes as needed. Though there are different perspectives on this time of history (whether life was autotrophic early on is an open question), the authors have approached the subject reasonably.

Decision Letter 3

Roland G Roberts

30 Aug 2022

Dear Nick,

Thank you for the submission of your revised Research Article "A prebiotic basis for ATP as the universal energy currency" for publication in PLOS Biology. On behalf of my colleagues and the Academic Editor, Matthew Pasek, I am pleased to say that we can in principle accept your manuscript for publication, provided you address any remaining formatting and reporting issues. These will be detailed in an email you should receive within 2-3 business days from our colleagues in the journal operations team; no action is required from you until then. Please note that we will not be able to formally accept your manuscript and schedule it for publication until you have completed any requested changes.

Please take a minute to log into Editorial Manager at http://www.editorialmanager.com/pbiology/, click the "Update My Information" link at the top of the page, and update your user information to ensure an efficient production process.

PRESS: We frequently collaborate with press offices. If your institution or institutions have a press office, please notify them about your upcoming paper at this point, to enable them to help maximise its impact. If the press office is planning to promote your findings, we would be grateful if they could coordinate with biologypress@plos.org. If you have previously opted in to the early version process, we ask that you notify us immediately of any press plans so that we may opt out on your behalf.

We also ask that you take this opportunity to read our Embargo Policy regarding the discussion, promotion and media coverage of work that is yet to be published by PLOS. As your manuscript is not yet published, it is bound by the conditions of our Embargo Policy. Please be aware that this policy is in place both to ensure that any press coverage of your article is fully substantiated and to provide a direct link between such coverage and the published work. For full details of our Embargo Policy, please visit http://www.plos.org/about/media-inquiries/embargo-policy/.

Thank you again for choosing PLOS Biology for publication and supporting Open Access publishing. We look forward to publishing your study. 

Best wishes,

Roli

Roland G Roberts, PhD, PhD

Senior Editor

PLOS Biology

rroberts@plos.org

Associated Data

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

    Supplementary Materials

    S1 Results. Document reporting all Supporting information figures.

    (DOCX)

    S1 Data. Excel document detailing raw data for all analyses.

    (XLSX)

    Attachment

    Submitted filename: The manuscript by Pinna et al.pdf

    Attachment

    Submitted filename: Pinna et al PLOS Biol response to reviewers.pdf

    Attachment

    Submitted filename: Pinna et al PLOS Biol response to reviewers final.docx

    Data Availability Statement

    All relevant data are within the paper and its Supporting Information files.


    Articles from PLoS Biology are provided here courtesy of PLOS

    RESOURCES