Next Article in Journal
Limited Sequence Diversity Within a Population Supports Prebiotic RNA Reproduction
Next Article in Special Issue
The Origin of Prebiotic Information System in the Peptide/RNA World: A Simulation Model of the Evolution of Translation and the Genetic Code
Previous Article in Journal
Single-Frame, Multiple-Frame and Framing Motifs in Genes
Previous Article in Special Issue
Low-Digit and High-Digit Polymers in the Origin of Life
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Hypothesis

Making Molecules with Clay: Layered Double Hydroxides, Pentopyranose Nucleic Acids and the Origin of Life

by
Harold S. Bernhardt
Department of Chemistry, University of Otago, P.O. Box 56 Dunedin, New Zealand
Submission received: 2 December 2018 / Revised: 4 February 2019 / Accepted: 9 February 2019 / Published: 15 February 2019

Abstract

:
A mixture of sugar diphosphates is produced in reactions between small aldehyde phosphates catalysed by layered double hydroxide (LDH) clays under plausibly prebiotic conditions. A subset of these, pentose diphosphates, constitute the backbone subunits of nucleic acids capable of base pairing, which is not the case for the other products of these LDH-catalysed reactions. Not only that, but to date no other polymer found capable of base pairing—and therefore information transfer—has a backbone for which its monomer subunits have a plausible prebiotic synthesis, including the ribose-5-phosphate backbone subunit of RNA. Pentose diphosphates comprise the backbone monomers of pentopyranose nucleic acids, some of the strongest base pairing systems so far discovered. We have previously proposed that the first base pairing interactions were between purine nucleobase precursors, and that these were weaker and less specific than standard purine-pyrimidine interactions. We now propose that the inherently stronger pairing of pentopyranose nucleic acids would have compensated for these weaker interactions, and produced an informational polymer capable of undergoing nonenzymatic replication. LDH clays might also have catalysed the synthesis of the purine nucleobase precursors, and the polymerization of pentopyranose nucleotide monomers into oligonucleotides, as well as the formation of the first lipid bilayers.

1. Introduction

Demonstrating the potential utility of a unified ‘systems chemistry’ approach to the problem of the origin of life, Powner and colleagues [1,2,3] recently proposed a link between a plausible early metabolic pathway and amino acid synthesis, through reactions between the small aldehyde phosphates glycolaldehyde phosphate (GAP) and glyceraldehyde-2-phosphate (G2P). Previously, Krishnamurthy and colleagues [4] showed that these same two molecules react—at millimolar concentrations, low-to-moderate temperatures and near-neutral pH—in the presence of layered double hydroxide (LDH) clays to produce pentose-2,4-diphosphates in 7% yield. One of these—ribose-2,4-diphosphate—constitutes the backbone subunit of pyranosyl-RNA [5], a pentopyranose nucleic acid that differs from RNA only in the position of its phosphate groups on the ribose ring, and in the ribose ring being six-membered (as opposed to five-membered as it is in RNA) (Figure 1). Pentopyranosyl nucleic acid systems have been found to constitute stronger base-pairing systems than RNA, with α-arabinopyranose nucleic acids possessing especially remarkable pairing strength [6]. In addition, the four pentopyranose phosphate systems with 4′→2′-backbone connectivity (including pyranosyl-RNA and α-arabinopyranose nucleic acid) are able to base pair with each other; in contrast, none of the four form base pairs with RNA. LDH-catalyzed reactions between GAP and G2P produce a mixture of different phosphorylated sugars, as shown in Figure 1, including tetrose-2,4-diphosphates (Figure 1A), pentose-2,4-diphosphates (Figure 1B, boxed) and hexose-2,4,6-triphosphates (Figure 1C) [4]; of these, the pentose-2,4-diphosphates are unique in constituting the backbone subunit monomers of oligonucleotides capable of base pairing [7].
Layered double hydroxide (LDH) clays are mixed-valence metal hydroxides with positively charged layers that adsorb charge-balancing anions such as Cl, CO32– and PO43– within their aqueous interlayer [8,9]. These anions are exchangeable by diffusion with organic anions such as GAP and G2P, which bind through electrostatic interactions between the charged phosphate group and the charged layers, as well as through hydrogen bonding interactions with the LDH hydroxide groups. (A snapshot of a simulation of the interaction between a RNA 25-mer oligonucleotide and LDH [10] is shown in Figure 2). The distance between LDH sheets is remarkably flexible, with the LDH interlayer able to more than triple its width from ~7 Å to 24 Å in order to accommodate a polyanionic strand of DNA [11]. In catalysing the reaction between GAP and G2P, LDHs perform a number of key functions:
Concentrating GAP and G2P from millimolar—or even micromolar—concentrations in the bulk medium to ~10 molar within the interlayer [4,12].
Binding GAP and G2P molecules in close proximity, thereby promoting their reaction.
Stabilizing and sequestering the reaction products, with 95 % of the pentose-2,4-diphosphate products still being extractable from the LDH after three months [4].
In these roles, the LDH clay is functioning as both a proto-enzyme and quasi-compartment, with Arrhenius [12] stating that, “Like cells, they retain phosphate-charged reactants against high concentration gradients and exchange matter with the surroundings by controlled diffusion through the ‘pores’ provided by the opening of the interlayers at the crystal edges” (p. 1580). LDHs also catalyse reactions between cyanide (CN) anions to form diaminomaleonitrile, a precursor to the purine nucleobase adenine [13], and appear able to stabilize base pairs between RNA nucleotides in the adsorption of guanosine monophosphate (GMP) at low to moderate temperatures (Figure 3) [14]. Clay minerals such as LDHs are thought to have been present early in Earth’s history [15,16], with Arrhenius and colleagues suggesting that LDH clays such as hydrotalcite (a naturally-occurring Mg/Al-LDH clay) might have existed as “surface coatings on submerged weathering basalt, or in salt brine deposits in arid lakes” [17] p. 506, as occur on the Earth today. Another LDH mineral, green rust (an Fe2+/Fe3+-LDH), is thought to have been widespread on an early anoxic Earth, with the eventual rise in atmospheric oxygen causing its oxidation and precipitation from the ocean, giving rise to the ubiquitous banded iron formations. In relation to this, Russell has argued that, “the Hadean ocean crust…was likely thick, relatively cool, and covered in carbonate green rust” [18] p. 6.
Almost a decade ago, Powner and colleagues demonstrated a possible prebiotic synthesis of the pyrimidine nucleotides [19]; however, the search for a plausibly prebiotic synthesis of the purine nucleotides remains ongoing [20]. Kim and Benner have reported the synthesis of adenosine and inosine from reactions between ribose-1,2-cyclophosphate and the purine nucleobases [21], while Carell and colleagues have demonstrated synthesis of both ribopyranose and standard and non-standard RNA nucleotides starting with a derivatized pyrimidine as a molecular scaffold [22,23]. Powner and Szostak have also found a potential link with pyrimidine synthesis, discovering a branch point in Powner’s previously-discovered pyrimidine synthesis which leads to the non-standard 8-oxopurines, described as potential precursors to the standard purine nucleotides in the original report [24]. However, in subsequent work from the Szostak lab [25], the 8-oxopurines were found to be poor substrates in a nonenzymatic primer extension model system, which the authors now consider makes it unlikely that they played a role as purine precursors. They now believe that inosine—in contrast a good substrate in their model nonenzymatic system—is more likely to have played such a role [25]. We have previously proposed [26] that the prebiotic synthesis of RNA might have occurred through the progressive synthesis of the purine nucleobases on a pre-existing ribose-phosphate backbone, with the driving force for this synthesis being the increasing stabilization of the backbone through intermolecular interactions, including base pairing and—potentially—duplex formation. We also proposed that these nucleobase precursors were similar or identical to the intermediates of the modern de novo purine biosynthetic pathway, through which modern organisms synthesize purines [27], for example inosine and 5-aminoimidazole-4-carboxamide riboside (AICAR) (Figure 4). Furthermore, due to the plausibly prebiotic nature of many of the reactants in the modern purine biosynthetic pathway (such as glycine and CO2) [28], as well as the fact that two of the reactions also occur by alternative nonenzymatic reactions [29,30], we proposed that the biosynthetic pathway might have originated from a series of uncatalysed prebiotic reactions [26]. Could the first informational molecules have contained purines (and/or purine precursors) without pyrimidines? As discussed above, Szostak and colleagues have shown that the purine (precursor) inosine participates efficiently in nonenzymatic primer extension using a model system, leading them to conclude, in the words of their article title, “Inosine, but none of the 8-oxo-purines, is a plausible component of a primordial version of RNA.” [25] Crick [31] and Wachtershauser [32] have also previously argued in favour of an “all-purine precursor of nucleic acids”, although Wachtershauser has proposed a different sugar-nucleobase connectivity than exists in modern RNA.

2. Hypothesis

We propose that life has undergone (at least) two genetic transitions, the first from a mixed system of pentopyranose nucleic acids to RNA, and the second from RNA to a mixed RNA/DNA system. Because both transitions involved a decrease in the strength of base pairing, they must have been driven by selection for another property. RNA may have been selected for its catalytic ability. The relative flexibility of its backbone and consequent ability to adopt a variety of non-standard base-pairing interactions enables RNA to assume multiple complex 3D conformations, including binding motifs and catalytic sites [33,34]. In contrast, pentopyranose nucleic acids possess more rigid backbones and are relatively constrained in their base pairing interactions [7]; it therefore appears likely that these systems might possess a somewhat more modest catalytic repertoire than RNA. The expansion in the chemical landscape offered by RNA catalysis would see its ultimate achievement in the advent of coded peptide synthesis [35], which produced a virtual explosion in catalytic potential. The evolutionary driver from RNA to a mixed RNA/DNA system appears to have been DNA’s greater chemical stability, which allowed for long-term storage of genetic information as well as a massive increase in genome size [36]. Therefore, the selection criteria would have been different for each genetic transition. In contrast, initial selection of pentopyranose nucleic acids would have been due to the availability of the pentose-2,4-diphosphate backbone monomers through LDH-catalysed synthesis. In addition, pentopyranose nucleic acids appear able to undergo facile nonenzymatic replication, required prior to the advent of—RNA, or possibly pentopyranose nucleic acid—replicase enzymes able to catalyse these reactions. The large angle of inclination (~ 45º) between the base pairs and backbone in pentopyranose systems enables these oligonucleotides to undergo nonenzymatic replication using prebiotically-plausible 2′,3′-cyclic phosphate-activated monomers [7]. In contrast, in the case of RNA, the same 2′,3′-cyclic phosphate-activated monomers form unnatural 2′,5′-RNA phosphate linkages. It is unknown how a transition from a pentopyranose system to RNA would have occurred. It might have been directly, by—for example—the conversion of a six-membered ribopyranose ring to a five-membered RNA ring. However, if so, this would need to have occurred at the ribose phosphate level, as the transfer of the phosphate group from the 4′- to the 5′-OH first requires opening of the six-membered pyranose ring, which would have been prohibited at the nucleotide stage by the attachment of the nucleobase (precursor).
In addition to catalysing the synthesis of the monomer units of the pentopyranose–phosphate backbone, we propose that LDH clays played a role in the synthesis of the purine nucleobase precursors, either by catalysing reactions between cyanide anions [13] or through the promotion of reactions between prebiotically plausible molecules such as glycine, NH3 and CO2, similar to those which occur in the modern de novo purine biosynthetic pathway [26]. Greenwell and colleagues have demonstrated peptide bond synthesis catalysed by a ternary Mg/Cu/Al-LDH clay under wet–dry cycles [37], in an analogous reaction to the amide bond formation that occurs in the biosynthesis of GAR, the first stable intermediate in the purine biosynthetic pathway [26,27]. Krishnamurthy and colleagues have shown that LDH clays catalyse the formylation of GAP utilizing a sulfite–formaldehyde adduct [38], in a parallel to the two indirect formylation reactions that occur in the purine biosynthetic pathway. As described above, Gwak and colleagues [14] have produced evidence that GMP nucleotides in the presence of LDHs form non-standard base pairs at low to moderate temperatures (20–60 °C), whereas at higher temperatures (80–100 °C) only unpaired GMP is present (Figure 3). This suggests that the LDH-promoted polymerisation of pentopyranose nucleotide monomers to form oligomers might also be possible. In addition, montmorillonite, a phyllosilicate clay containing negatively charged sheets that bind Mg2+ and other cations, catalyses the polymerisation of RNA nucleotides to form up to 50-mer RNA oligos [39], suggesting such reactions might also be possible for LDH clays. Supporting this possibility, RNA nucleotides [14,40,41], oligonucleotides and duplexes [42] are strongly adsorbed by LDH clays through their charged phosphate backbones, as demonstrated both experimentally and in simulation experiments (Figure 2). Polymerisation of phosphorylated sugars lacking an attached nucleobase (precursor) may also be possible, as demonstrated by the putative polymerisation of fructose-1,6-bisphosphate (a sugar diphosphate with a similar structure to a pentose-2,4-diphosphate), in the presence of a Li+-LDH [43]. The high affinity of LDH clays for phosphate anions is rather striking, and suggests that these clays might have played a critical role in extracting and concentrating phosphate from low background levels in the prebiotic environment; however, this raises the question of how the strongly bound pentopyranose oligonucleotides might have been released from the LDH interlayer. Russell has posited that LDH clays (including green rust) initially formed in the high pH environment of alkaline deep-sea hydrothermal vents [18], which suggests the possibility that release from the LDH interlayer might have occurred in the increasingly acidic conditions distal from the vent, as LDH clays largely disintegrate at acidic pH [8]. Exchange of the phosphate moieties with divalent carbonate anions [8], possibly through the interaction with high atmospheric levels of CO2, would have been another possible mechanism for the eventual release of pentopyranose oligonucleotides from their LDH ‘cells’.
Conceivably, the high pairing strength of pentopyranose systems might be considered a disadvantage for nonenzymatic replication, as it could increase the chances of mispairing and make the system vulnerable to product inhibition (wherein the just-synthesized copy remains bound to the template sequence, preventing further replication) [44]. However, as described above, we previously proposed that the first nucleobases were precursors to the purine nucleobases [26]—perhaps the same or similar to the intermediates of the modern de novo purine biosynthetic pathway—and it would seem reasonable that weaker interactions between the precursors might have offset this high pairing strength. Two examples of possible precursors are inosine and AICAR (Figure 4). Inosine is a purine itself, as well as a precursor to adenosine and guanosine, and in fact plays a key role in genetic coding, occurring in the anticodon wobble position of alanine tRNA [45]. Interestingly, it is utilised due to its ability to form a purine–purine base pair (with adenosine) as well as more typical purine–pyrimidine base pairs with cytosine and uracil. When incorporated into RNA oligonucleotides, inosine and 1-(2-deoxy-β-d-ribofuranosyl)-imidazole-4-carboxamide (dICAR) (a 2′-deoxy analogue of AICAR) form base pair interactions with purines as well as pyrimidines [46,47,48]. The similarity in hydrogen bond-forming potential between adenosine, AICAR and dICAR is shown in Figure 4. However, these pairings between purine precursors are weaker—in some cases significantly—than standard purine–pyrimidine base pairs. In the case of dICAR, this is presumably due to a decrease in base-stacking stabilization due to AICAR possessing an imidazole ring as opposed to a purine double ring structure. Self-complementary RNA oligonucleotide 12-mers containing two inosine- or dICAR base pairs have melting temperatures ~20 °C lower than the comparable RNA duplexes containing standard purine–pyrimidine base pairs only [47], and it is likely that the presence of inosine and AICAR would similarly decrease the stability of pentopyranose duplexes, although this has not been shown experimentally. Nevertheless, we propose that this decrease in stability would have been offset by the greater strength of pentopyranosyl pairing interactions. In addition, it has been shown that pentopyranose duplexes accommodate purine–purine base pairs more easily than RNA duplexes [7], which a number of these interactions would constitute (or at least approximate). In fact, the two opposing effects—inherently more stable duplexes vs. weaker interactions between purine precursors—might have produced a system able to undergo nonenzymatic replication: strong enough for duplex formation and replication, but not so strong as to produce template-product inhibition [44].
The lack of cross pairing between pentopyranose and RNA systems—and the consequent inability to transfer sequence information in the proposed genetic transition—could be seen as a major weakness of our hypothesis. However, genetic transitions will as a rule result in a degree of information loss, due to the alteration in 3D structure (and therefore function) caused by the difference in backbone conformation between the preceding and succeeding informational polymers [7]. As Eschenmoser has pointed out, “Irrespective of its direct communication with the predecessor system, the successor system would have to evolve its phenotype (i.e., chemical catalytic functionality] de novo.” [7] If, however—as is the case with the non-cross pairing pentopyranose and RNA systems—there is zero possibility of information transfer, one might ask what is the point of proposing such a transition, and the need for a pentopyranose system to evolve first? Wouldn’t the principle of Occam’s razor favour the hypothesis that RNA arose de novo? As discussed above, advantages of the hypothesis are:
  • The demonstrated prebiotically plausible formation of the pentose–diphosphate backbone monomers; and, conversely, the absence of this for the RNA backbone subunit.
  • The structural features of pentopyranose systems, for example the large angle of inclination between base pairs and backbone, that make them structurally suited for nonenzymatic replication using plausibly prebiotic 2′,3′-cyclic phosphate-activated monomers.
  • The greater pairing strength of pentopyranose systems—important to counterbalance the initial weaker interactions between purine nucleobase precursors—to allow stable duplex formation and nonenzymatic replication.
  • The fact that a pentopyranose system provides a—structural and possibly also catalytic—stepping stone to RNA, removing some of the difficulties in the RNA backbone arising de novo.
Finally, it seems likely that LDH clays might also have played other roles in the origin of life, such as in the formation of the first lipid bilayers. The negatively charged carboxylate anions of CH3(CH2)16COONa+, the sodium salt of the long-chain fatty acid octadecanoic (or stearic) acid, adsorb to both sides of the LDH interlayer, forming a tilted version of a lipid bilayer, and increasing the interlayer width to 48 Å (Figure 5) [49]. In conclusion, what we have learned so far regarding LDH clays points to our having only just begun to scratch the surface in relation to this fascinating class of materials.

Acknowledgments

Thanks to Dave Larsen and Warren Tate for paying the article processing charge (APC), which enabled this article to be published. The author would like to thank Ram Krishnamurthy, Chris Greenwell, Dave Larsen and Jim McQuillan for helpful discussions, and to Ram Krishnamurthy for his comments on an earlier draft of this manuscript.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

AICAR5-Aminoimidazole-4-carboxamide riboside
dICAR1-(2-Deoxy-β-d-ribofuranosyl)-imidazole-4-carboxamide
G2PGlyceraldehyde-2-phosphate
GAPGlycolaldehyde phosphate
GMPGuanosine-5-monophosphate
LDHLayered double hydroxide

References

  1. Islam, S.; Powner, M.W. Prebiotic systems chemistry: Complexity overcoming clutter. Chem 2017, 2, 470–501. [Google Scholar] [CrossRef]
  2. Coggins, A.J.; Powner, M.W. Prebiotic synthesis of phosphoenol pyruvate by alpha-phosphorylation-controlled triose glycolysis. Nat. Chem. 2017, 9, 310–317. [Google Scholar] [CrossRef]
  3. Islam, S.; Bucar, D-K.; Powner, M.W. Prebiotic selection and assembly of proteinogenic amino acids and natural nucleotides from complex mixtures. Nat. Chem. 2017, 9, 584–589. [Google Scholar] [CrossRef] [Green Version]
  4. Krishnamurthy, R.; Pitsch, S.; Arrhenius, G. Mineral induced formation of pentose-2,4-bisphosphates. Orig. Life Evol. Biosph. 1999, 29, 139–152. [Google Scholar] [CrossRef]
  5. Pitsch, S.; Wendeborn, S.; Krishnamurthy, R.; Holzner, A.; Minton, M.; Bolli, M.; Miculca, C.; Windhab, N.; Micura, R.; Stanek, M.; et al. Pentopyranosyl oligonucleotide systems. 9th Communication. The β-D-ribopyranosyl-(2′→4’) oligonucleotide system (‘pyranosyl-RNA’): Synthesis and resumé of base-pairing properties. Helv. Chim. Acta 2003, 86, 4270–4363. [Google Scholar] [CrossRef]
  6. Jungmann, O.; Beier, M.; Luther, A.; Huynh, H.K.; Ebert, M.O.; Jaun, B.; Krishnamurthy, R.; Eschenmoser, A. Pentopyranosyl oligonucleotide systems. The α-L-arabinopyranosyl-(4’→2’)-oligonucleotide system: Synthesis and pairing properties. Helv. Chim. Acta 2003, 86, 1259–1308. [Google Scholar] [CrossRef]
  7. Eschenmoser, A. Etiology of potentially primordial biomolecular structures: From vitamin B12 to the nucleic acids and an inquiry into the chemistry of life’s origin: A retrospective. Angew. Chem. Int. Ed. 2011, 50, 12412–12472. [Google Scholar] [CrossRef] [PubMed]
  8. Mishra, G.; Dash, B.; Pandey, S. Layered double hydroxides: A brief review from fundamentals to application as evolving biomaterials. Appl. Clay Sci. 2018, 153, 172–186. [Google Scholar] [CrossRef]
  9. Kuma, K.; Paplawsky, W.; Gedulin, B.; Arrhenius, G. Mixed-valence hydroxides as bioorganic host minerals. Orig. Life Evol. Biosph. 1989, 19, 573–602. [Google Scholar] [CrossRef]
  10. Swadling, J.B.; Suter, J.L.; Greenwell, H.C.; Coveney, P.V. Influence of surface chemistry and charge on mineral–RNA interactions. Langmuir 2013, 29, 1573–1583. [Google Scholar] [CrossRef]
  11. Xu, Z.P.; Lu, G.Q. Layered double hydroxide nanomaterials as potential cellular drug delivery agents. Pure Appl. Chem. 2006, 78, 1771–1779. [Google Scholar] [CrossRef]
  12. Arrhenius, G.O. Crystals and life. Helv. Chim. Acta 2003, 86, 1569–1586. [Google Scholar] [CrossRef]
  13. Boclair, J.W.; Braterman, P.S.; Brister, B.D.; Jiang, J.; Lou, S.; Wang, Z.; Yarberry, F. Cyanide self-addition, controlled adsorption, and other processes at layered double hydroxides. Orig. Life Evol. Biosph. 2001, 31, 53–69. [Google Scholar] [CrossRef] [PubMed]
  14. Gwak, G.-H.; Kocsis, I.; Legrand, Y.-M.; Barboiu, M.; Oh, J.-M. Controlled supramolecular structure of guanosine monophosphate in the interlayer space of layered double hydroxide. Beilstein J. Nanotechnol. 2016, 7, 1928–1935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Hazen, R.M.; Sverjensky, D.A.; Azzolini, D.; Bish, D.L.; Elmore, S.C.; Hinnov, L.; Milliken, R.E. Clay mineral evolution. Am. Mineral 2013, 98, 2007–2029. [Google Scholar] [CrossRef]
  16. Hazen, R.M. Paleomineralogy of the Hadean eon: A preliminary species list. Am. J. Sci. 2013, 313, 807–843. [Google Scholar] [CrossRef]
  17. Arrhenius, G.; Sales, B.; Mojzsis, S.; Lee, T. Entropy and charge in molecular evolution – the case of phosphate. J. Theor. Biol. 1997, 187, 503–522. [Google Scholar] [CrossRef] [PubMed]
  18. Russell, M.J. Green rust: The simple organizing ’seed’ of all life? Life 2018, 8, E35. [Google Scholar] [CrossRef]
  19. Powner, M.W.; Gerland, B.; Sutherland, J.D. Synthesis of activated pyrimidine ribonucleotides in prebiotically plausible conditions. Nature 2009, 459, 239–242. [Google Scholar] [CrossRef]
  20. Powner, M.W.; Sutherland, J.D.; Szostak, J.W. Chemoselective multicomponent one-pot assembly of purine precursors in water. J. Am. Chem. Soc. 2010, 132, 16677–16688. [Google Scholar] [CrossRef]
  21. Kim, H.J.; Benner, S.A. Prebiotic stereoselective synthesis of purine and noncanonical pyrimidine nucleotide from nucleobases and phosphorylated carbohydrates. Proc. Natl Acad. Sci. USA 2017, 114, 11315–11320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Becker, S.; Thoma, I.; Deutsch, A.; Gehrke, T.; Mayer, P.; Zipse, H.; Carell, T. A high-yielding, strictly regioselective prebiotic purine nucleoside formation pathway. Science 2016, 352, 833–836. [Google Scholar] [CrossRef] [PubMed]
  23. Becker, S.; Schneider, C.; Okamura, H.; Crisp, A.; Amatov, T.; Dejmek, M.; Carell, T. Wet-dry cycles enable the parallel origin of canonical and non-canonical nucleosides by continuous synthesis. Nat. Commun. 2018, 9, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Stairs, S.; Nikmal, A.; Bučar, D.K.; Zheng, S.L.; Szostak, J.W.; Powner, M.W. Divergent prebiotic synthesis of pyrimidine and 8-oxo-purine ribonucleotides. Nat. Commun. 2017, 8, 15270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kim, S.C.; O’Flaherty, D.K.; Zhou, L.; Lelyveld, V.S.; Szostak, J.W. Inosine, but none of the 8-oxo-purines, is a plausible component of a primordial version of RNA. Proc. Natl. Acad. Sci. USA 2018, 115, 13318–13323. [Google Scholar] [CrossRef] [PubMed]
  26. Bernhardt, H.S.; Sandwick, R.K. Purine biosynthetic intermediate-containing ribose-phosphate polymers as evolutionary precursors to RNA. J. Mol. Evol. 2014, 79, 91–104. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, Y.; Morar, M.; Ealick, S.E. Structural biology of the purine biosynthetic pathway. Cell Mol. Life Sci. 2008, 65, 3699–3724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. de Duve, C. Blueprint for a Cell: The Nature and Origin of Life; Neil Patterson Publishers, Carolina Biological Supply Company: Burlington, NC, USA, 1991; pp. 133–134. [Google Scholar]
  29. Nierlich, D.P.; Magasanik, B. Phosphoribosylglycinamide synthetase of Aerobacter aerogenes. Purification and properties, and nonenzymatic formation of its substrate 5’-phosphoribosylamine. J. Biol. Chem. 1965, 240, 366–374. [Google Scholar] [PubMed]
  30. Mueller, E.J.; Meyer, E.; Rudolph, J.; Davisson, V.J.; Stubbe, J. N5-Carboxyaminoimidazole ribonucleotide: Evidence for a new intermediate and two new enzymatic activities in the de novo purine biosynthetic pathway of Escherichia coli. Biochemistry 1994, 33, 2269–2278. [Google Scholar] [CrossRef]
  31. Crick, F.H. The origin of the genetic code. J. Mol. Biol. 1968, 38, 367–379. [Google Scholar] [CrossRef]
  32. Wachtershauser, G. An all-purine precursor of nucleic acids. Proc. Natl. Acad. Sci. USA 1988, 85, 1134–1135. [Google Scholar] [CrossRef]
  33. Chandrasekhar, K.; Malathi, R. Non-Watson Crick base pairs might stabilize RNA structural motifs in ribozymes – a comparative study of group-I intron structures. J. Biosci. 2003, 28, 547–555. [Google Scholar] [CrossRef]
  34. Fedor, M.J.; Williamson, J.R. The catalytic diversity of RNAs. Nat. Rev. Mol. Cell Biol. 2005, 6, 399–412. [Google Scholar] [CrossRef]
  35. Bernhardt, H.S.; Tate, W.P. The transition from noncoded to coded protein synthesis: did coding mRNAs arise from stability-enhancing binding partners to tRNA? Biol. Direct 2010, 5, 16. [Google Scholar] [CrossRef] [Green Version]
  36. Krishnamurthy, R. On the emergence of RNA. Isr. J. Chem. 2015, 55, 837–850. [Google Scholar] [CrossRef]
  37. Grégoire, B.; Greenwell, H.C.; Fraser, D.G. Peptide formation on layered mineral surfaces: The key role of brucite-like minerals on the enhanced formation of alanine dipeptides. ACS Earth Space Chem. 2018, 2, 852–862. [Google Scholar] [CrossRef]
  38. Pitsch, S.; Krishnamurthy, R.; Arrhenius, G. Concentration of simple aldehydes by sulfite-containing double-layer hydroxide minerals: Implications for biopoesis. Helv. Chim. Acta 2000, 83, 2398–2411. [Google Scholar] [CrossRef]
  39. Ferris, J.P. Montmorillonite catalysis of 30-50 mer oligonucleotides: Laboratory demonstration of potential steps in the origin of the RNA world. Orig. Life Evol. Biosph. 2002, 32, 311–332. [Google Scholar] [CrossRef]
  40. Choy, J.-H.; Kwak, S.-Y.; Park, J.-S.; Jeong, Y.J.; Portier, J. Intercalative nanohybrids of nucleoside monophosphates and DNA in layered metal hydroxide. J. Am. Chem. Soc. 1999, 121, 1399–1400. [Google Scholar] [CrossRef]
  41. Aisawa, S.; Ohnuma, Y.; Hirose, K.; Takahashi, S.; Hirahara, H.; Narita, E. Intercalation of nucleotides into layered double hydroxides by ion-exchange reaction. Appl. Clay Sci. 2005, 28, 137–145. [Google Scholar] [CrossRef]
  42. Swadling, J.B.; Coveney, P.V.; Greenwell, H.C. Stability of free and mineral-protected nucleic acids: Implications for the RNA world. Geochim. Cosmochim. Acta 2012, 1, 360–378. [Google Scholar] [CrossRef]
  43. Jellicoe, T.C.; Fogg, A.M. Synthesis and characterization of layered double hydroxides intercalated with sugar phosphates. J. Phys. Chem. Solids 2012, 73, 1496–1499. [Google Scholar] [CrossRef]
  44. Engelhart, A.E.; Powner, M.W.; Szostak, J.W. Functional RNAs exhibit tolerance for non-heritable 2′-5′ versus 3′-5′ backbone heterogeneity. Nat. Chem. 2013, 5, 390–394. [Google Scholar] [CrossRef]
  45. Torres, A.G.; Piñeyro, D.; Filonava, L.; Stracker, T.H.; Batlle, E.; Ribas de Pouplana, L. A-to-I editing on tRNAs: Biochemical, biological and evolutionary implications. FEBS Lett. 2014, 588, 4279–4286. [Google Scholar] [CrossRef] [Green Version]
  46. Carter, R.J.; Baeyens, K.J.; SantaLucia, J.; Turner, D.H.; Holbrook, S.R. The crystal structure of an RNA oligomer incorporating tandem adenosine-inosine mismatches. Nucleic Acids Res. 1997, 25, 4117–4122. [Google Scholar] [CrossRef]
  47. Johnson, W.T.; Zhang, P.M.; Bergstrom, D.E. The synthesis and stability of oligodeoxyribonucleotides containing the deoxyadenosine mimic 1-(2′-deoxy-β-D-ribofuranosyl)imidazole-4-carboxamide. Nucleic Acids Res. 1997, 25, 559–567. [Google Scholar] [CrossRef]
  48. Sala, M.; Pezo, V.; Pochet, S.; Wain-Hobson, S. Ambiguous base pairing of the purine analogue 1-(2-deoxy-β-d-ribofuranosyl)-imidazole-4-carboxamide during PCR. Nucleic Acids Res. 1996, 24, 3302–3306. [Google Scholar] [CrossRef]
  49. Itoh, T.; Ohta, N.; Shichi, T.; Yui, T.; Takagi, K. The self-assembling properties of stearate ions in hydrotalcite clay. Langmuir 2003, 19, 9120–9126. [Google Scholar] [CrossRef]
Figure 1. Nucleic acid backbone subunits, showing position of phosphate groups and attached (nucleo)base. (AC) (minus base) are produced in reactions catalysed by layered double hydroxide (LDH) clay, as described in the text; of these, only (B) (boxed) (present in pentopyranose nucleic acids) is capable of base pairing. In contrast, the backbone subunits of RNA (D) and DNA (E) do not have a plausibly prebiotic synthesis. A = tetrooxetose nucleic acid, B = pentopyranose nucleic acid, C = hexopyranose nucleic acid, D = RNA, E = DNA. Numbering of carbohydrate carbon atoms is shown. Figure produced in ChemDraw® 18, PerkinElmer Informatics, Waltham, MA.
Figure 1. Nucleic acid backbone subunits, showing position of phosphate groups and attached (nucleo)base. (AC) (minus base) are produced in reactions catalysed by layered double hydroxide (LDH) clay, as described in the text; of these, only (B) (boxed) (present in pentopyranose nucleic acids) is capable of base pairing. In contrast, the backbone subunits of RNA (D) and DNA (E) do not have a plausibly prebiotic synthesis. A = tetrooxetose nucleic acid, B = pentopyranose nucleic acid, C = hexopyranose nucleic acid, D = RNA, E = DNA. Numbering of carbohydrate carbon atoms is shown. Figure produced in ChemDraw® 18, PerkinElmer Informatics, Waltham, MA.
Life 09 00019 g001
Figure 2. Computer simulation snapshot of the interaction between an RNA 25-mer and LDH interlayer surface, depicting the inner-sphere complexes between RNA phosphate oxygen atoms (red) and LDH metal hydroxide hydrogen atoms (white) in the absence of bridging water molecules. Numerical values shown in black indicate the distance between interaction donors and acceptors in angstroms. Color scheme: O (red), H (white), P (gold), C (grey), N (dark blue), Mg (cyan) and Al (pink). Reprinted with permission from [10]. Copyright 2013 American Chemical Society.
Figure 2. Computer simulation snapshot of the interaction between an RNA 25-mer and LDH interlayer surface, depicting the inner-sphere complexes between RNA phosphate oxygen atoms (red) and LDH metal hydroxide hydrogen atoms (white) in the absence of bridging water molecules. Numerical values shown in black indicate the distance between interaction donors and acceptors in angstroms. Color scheme: O (red), H (white), P (gold), C (grey), N (dark blue), Mg (cyan) and Al (pink). Reprinted with permission from [10]. Copyright 2013 American Chemical Society.
Life 09 00019 g002
Figure 3. Schematic illustrations for interlayer structure of GMP/LDH hybrids according to molecular arrangement of GMPs: (a) single molecule arrangement (GL-S) and (b) ribbon II arrangement (GL-R). GL-S: 12.6 Å (d-spacing) − 4.8 Å (LDH layer thickness) = 7.8 Å; GL-R: 17.7 Å (d-spacing) − 4.8 Å (LDH layer thickness) = 12.9 Å. Adapted with permission from [14].
Figure 3. Schematic illustrations for interlayer structure of GMP/LDH hybrids according to molecular arrangement of GMPs: (a) single molecule arrangement (GL-S) and (b) ribbon II arrangement (GL-R). GL-S: 12.6 Å (d-spacing) − 4.8 Å (LDH layer thickness) = 7.8 Å; GL-R: 17.7 Å (d-spacing) − 4.8 Å (LDH layer thickness) = 12.9 Å. Adapted with permission from [14].
Life 09 00019 g003
Figure 4. (Top) the potential of 5-aminoimidazole-4-carboxamide riboside (AICAR) (an intermediate in the modern de novo purine biosynthetic pathway) and its 2′-deoxy analogue 1-(2-deoxy-β-d-ribofuranosyl)-imidazole-4-carboxamide (dICAR) to form hydrogen-bonding interactions similar to those formed by adenosine. (Bottom) hydrogen-bonding (base pairing) interaction between adenosine and inosine (a later intermediate in the purine biosynthetic pathway), and proposed interaction between AICAR and inosine, examples of the weaker interactions between purine nucleobase precursors proposed to have preceded purine-pyrimidine base pairing. R = ribofuranose; dR = 2′-deoxyribofuranose, figure produced in ChemDraw® 18 (PerkinElmer Informatics, Waltham, MA).
Figure 4. (Top) the potential of 5-aminoimidazole-4-carboxamide riboside (AICAR) (an intermediate in the modern de novo purine biosynthetic pathway) and its 2′-deoxy analogue 1-(2-deoxy-β-d-ribofuranosyl)-imidazole-4-carboxamide (dICAR) to form hydrogen-bonding interactions similar to those formed by adenosine. (Bottom) hydrogen-bonding (base pairing) interaction between adenosine and inosine (a later intermediate in the purine biosynthetic pathway), and proposed interaction between AICAR and inosine, examples of the weaker interactions between purine nucleobase precursors proposed to have preceded purine-pyrimidine base pairing. R = ribofuranose; dR = 2′-deoxyribofuranose, figure produced in ChemDraw® 18 (PerkinElmer Informatics, Waltham, MA).
Life 09 00019 g004
Figure 5. Structural models of the stearate/LDH interaction showing (a) tilted stearate bilayer and (b) regular packing of stearate anions in the interlayer gallery. Reprinted with permission from [49]. Copyright 2003 American Chemical Society.
Figure 5. Structural models of the stearate/LDH interaction showing (a) tilted stearate bilayer and (b) regular packing of stearate anions in the interlayer gallery. Reprinted with permission from [49]. Copyright 2003 American Chemical Society.
Life 09 00019 g005

Share and Cite

MDPI and ACS Style

Bernhardt, H.S. Making Molecules with Clay: Layered Double Hydroxides, Pentopyranose Nucleic Acids and the Origin of Life. Life 2019, 9, 19. https://doi.org/10.3390/life9010019

AMA Style

Bernhardt HS. Making Molecules with Clay: Layered Double Hydroxides, Pentopyranose Nucleic Acids and the Origin of Life. Life. 2019; 9(1):19. https://doi.org/10.3390/life9010019

Chicago/Turabian Style

Bernhardt, Harold S. 2019. "Making Molecules with Clay: Layered Double Hydroxides, Pentopyranose Nucleic Acids and the Origin of Life" Life 9, no. 1: 19. https://doi.org/10.3390/life9010019

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop