Next Article in Journal
Physicochemical Foundations of Life that Direct Evolution: Chance and Natural Selection are not Evolutionary Driving Forces
Next Article in Special Issue
A Polyaddition Model for the Prebiotic Polymerization of RNA and RNA-Like Polymers
Previous Article in Journal
Life in The Context of Order and Complexity
Previous Article in Special Issue
Silica Precipitation in a Wet–Dry Cycling Hot Spring Simulation Chamber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Concept Paper

Polyesters as a Model System for Building Primitive Biologies from Non-Biological Prebiotic Chemistry

1
Space Science Center (ANGKASA), Institute of Climate Change, Level 3, Research Complex, National University of Malaysia, UKM Bangi, Selangor 43600, Malaysia
2
Department of Physical Chemistry, University of Chemistry and Technology, Prague, Technicka 5, Dejvice, 16628 Prague 6, Czech Republic
3
Earth-Life Science Institute, Tokyo Institute of Technology, 2-12-1-IE-1 Ookayama, Meguro-ku, Tokyo 152-8550, Japan
4
Blue Marble Space Institute for Science, 1001 4th Ave, Suite 3201, Seattle, WA 98154, USA
5
Institute for Advanced Study, Princeton, NJ 08540, USA
6
Center for Chemical Evolution, Georgia Institute of Technology, Atlanta, GA 30332, USA
*
Authors to whom correspondence should be addressed.
Submission received: 13 November 2019 / Revised: 22 December 2019 / Accepted: 10 January 2020 / Published: 19 January 2020
(This article belongs to the Special Issue Themed Issue Commemorating Prof. David Deamer's 80th Birthday)

Abstract

:
A variety of organic chemicals were likely available on prebiotic Earth. These derived from diverse processes including atmospheric and geochemical synthesis and extraterrestrial input, and were delivered to environments including oceans, lakes, and subaerial hot springs. Prebiotic chemistry generates both molecules used by modern organisms, such as proteinaceous amino acids, as well as many molecule types not used in biochemistry. As prebiotic chemical diversity was likely high, and the core of biochemistry uses a rather small set of common building blocks, the majority of prebiotically available organic compounds may not have been those used in modern biochemistry. Chemical evolution was unlikely to have been able to discriminate which molecules would eventually be used in biology, and instead, interactions among compounds were governed simply by abundance and chemical reactivity. Previous work has shown that likely prebiotically available α-hydroxy acids can combinatorially polymerize into polyesters that self-assemble to create new phases which are able to compartmentalize other molecule types. The unexpectedly rich complexity of hydroxy acid chemistry and the likely enormous structural diversity of prebiotic organic chemistry suggests chemical evolution could have been heavily influenced by molecules not used in contemporary biochemistry, and that there is a considerable amount of prebiotic chemistry which remains unexplored.

1. Introduction

How life emerged from inanimate chemistry is an unresolved scientific question and an area of active research [1,2,3]. Contemporary biochemistry uses a relatively small set of building blocks, e.g., only 20 coded amino acids, four ribonucleotides, four deoxyribonucleotides, etc., out of a large number of possibilities [4]. We here term the molecules used in contemporary biochemistry “biomolecules” to distinguish them from “non-biomolecules,” which are molecules not used for major macromolecular functions in contemporary biochemistry (or even not used at all in contemporary biochemistry), but could have played much more important roles in previous biological or proto-biological states. How, why, and at what stage of chemical and/or biological evolution the selection and canonicalization of the biomolecules occurred remains unclear. Much of origins of life (OoL) research is oriented towards synthesizing or reproducing biomolecules or processes under presumed primitive Earth conditions [2]. While such approaches are logical given the evidence available from contemporary biology, and appear to be a sound application of “Occam’s razor,” they are also often based on a number of assumptions. One such assumption is the direct continuity between biochemistry and prebiotic chemistry, which has given rise to OoL models variously focused on the primary roles of lipids [5], nucleic acids [6], metabolism [7], or proteins [8]. However, contemporary biomolecule-based hypotheses overlook the long-recognized challenges of robust prebiotic syntheses of biomolecules [9], and the fact that many abiotic syntheses produce a tremendous diversity of organic compounds.
For example, researchers often consider carbonaceous chondrite meteorites to be proxies that validate laboratory prebiotic chemistry simulations [10,11]. Carbonaceous meteorites contain a myriad of organic compounds divided between a soluble fraction and an insoluble organic matter (IOM) phase consisting of intractable complex macromolecular organic material. The soluble fraction is often incredibly diverse, for example 14,000–50,000 unique molecular formula compounds were identified in a methanolic extract of the Murchison meteorite, which it was estimated may represent millions of unique structural isomers [12] (Figure 1).
In fact, the number and diversity of compounds in the Murchison meteorite can only be crudely estimated due to analytical challenges [13], suggesting that the actual organic complexity may be far greater. The classes of molecules identified to date in the Murchison meteorite include amino acids [14], hydroxy acids [15], carboxylic acids [16], and nitrogen heterocycles [17]. Most of these classes are dominated by compounds not used in present-day biology [18]. In light of this, it seems plausible that a similarly complex compound suite was provided to primitive Earth (or existed on another planet [19]) from one or more sources (e.g., [20,21,22]), and indeed various prebiotic chemistry simulations similarly produce a complex array of organic products (Figure 2) [23].
Given the magnitude of prebiotic chemical diversity, the early environment likely included non-biomolecules along with biomolecules and quite possibly more of the former than the latter. However, relatively little effort has been expended towards understanding the identities or reactivity of the non-biomolecules generated in prebiotic syntheses [24], or the processes or novel chemical phenomena they might enable. It may be likely that during early chemical evolution, proto-biological functions involved different-possibly larger, smaller, or partially overlapping sets of chemical compounds compared to contemporary biochemistry.
Although several studies suggest that biomolecules conferred a significant advantage to primitive systems resulting in their dominance in contemporary biochemistry [25,26,27] it does not necessarily follow that life was originally based on the same biomolecules. Indeed, it is known that biomolecules represent a minute fraction of the molecular diversity produced by prebiotic chemistry (as mentioned above) and this would tend to lower the probability that chemical bonds formed exclusively among biomolecules particularly in a one-pot prebiotic synthesis [3,24,28]. The possible importance of non-biomolecules in the OoL has been suggested previously as a sort of “chemical opportunism” [29]. According to this notion whatever molecules conferred the greatest selective advantage (via function, persistence, availability, etc.) would be selected by primitive chemical systems. If non-biomolecules conferred a greater advantage to primitive chemical systems over biomolecules, then non-biomolecules would have dominated. A notional example of such “chemical opportunism” is the emergent systemic autocatalysis proposed by [30]. The idea that the first autocatalytic chemical systems could have been composed of non-biomolecules has also been proposed by Lancet and coworkers [19,29]. More nuanced examples of primitive non-biomolecular systems, include those incorporating purely in vitro-produced polymers (themselves produced from biomolecular components such as amino acids) which do not participate in modern biology, have also been explored including self-assemblies of lipid-like peptides [31]. Earlier examples of non-biomolecule-based OoL models include the so-called pre-RNA world [32,33,34,35] and polyester-based origins of life models [36].
Polyesters offer a particularly interesting model for studying chemical evolution due to their structural similarity to peptides, ease of polymerization into diverse polydisperse polyester mixtures (polyesters form easily during low-temperature wet-dry cycling of hydroxy acids [37,38]) (Figure 3), and ability to perform enzyme-like catalysis [39] and compartmentalization of molecules through microdroplet assembly [40]. Polyester-based oligomers are fundamentally different from contemporary biological macromolecular catalysts such as proteins. Polyester systems can interact due to hydrophilic, hydrophobic, and electrostatic interactions such as dipole-dipole interactions [41], but they lack the ability to form regular repeating intramolecular hydrogen bonding motifs of the sort that facilitate protein folding [42,43]. However, although polyesters cannot structurally fold in the same way as polypeptides, they might still be distant relatives of biochemistry and could have provided functions that promoted primitive molecular assembly and dynamics essential for the emergence of biology.
Simple non-biomolecule monomers, best experimentally demonstrated so far using α-hydroxy acids (αHAs) [38,39,40,44], but which could include many other prebiotic monomer types, can form diverse macromolecular assemblages with a wide range of physical properties and functions such as self-assembly and catalysis. Though these are often dismissively lumped together as “tars,” such assemblies could present opportunities to form extremely complex “ecosystems” of microcompartments, each with different chemical properties, in which different abiotic chemistries could have been facilitated and their products exchanged. These could represent a pre-genomic analog of Woese’s rampant horizontal genome transfer-characterized LUCA [45], also known as a progenote, and an important aspect of global chemical evolution. According to this model, prior to LUCA, a community of progenotes [45], as in primitive entities in the process of (but prior to) establishing the genotype-phenotype relationship, could have dominated the prebiotic world while allowing for genetic information transfer amongst themselves and the surrounding environment. These informational materials may not be limited to nucleic acid-based materials, but also could be composed of other types of genetic information, such as those based on the composome model [29]. Such a character could have arisen among such systems, and then spread throughout them, co-opting and replacing their more primitive modes of reproduction, such as autocatalytic chemical sets proposed by Kaufman [30], potentially eventually leading to systems with established genotype-phenotype relationships that could have evolved into modern biology as-we-know-it.

2. Polyesters as a Model System for Chemical Evolution

The free energy of the peptide bond is around +3.5 kcal mol−1 under physiological conditions [46]. In biochemistry, peptides are synthesized from amino acids through a complex system of enzyme-mediated, energy-consuming reactions. Abiotic peptide synthesis is often accomplished with varying degrees of success through activation (e.g., [47,48,49,50]), heating, and/or low water activity conditions (e.g., [51,52,53]). Alternatively, polyesters, which have a considerably lower energy for bond formation (~0 kcal mol−1 [54]), have been considered as possible ancestral precursors of peptides [36,55]. Such polyesters can be easily produced by drying hydroxy acids, the monomer precursors of polyesters. The loss of water, for example during evaporative drying, shifts the equilibrium towards polyester elongation [38] (Scheme 1). Hydroxy acids are equally as prebiotically plausible as amino acids, as hydroxy acids may be produced in the same model prebiotic reactions that produce amino acids [56,57], and are also common in meteorites [58].
Esters are common in modern biochemistry. For example, many cell membrane lipids are esters of glycerol, phosphate, and fatty acids. Another example is cutin, a polyester synthesized from ω-hydroxy acids that is the main component of the cuticle that covers plant aerial surfaces [59]. Polyhydroxyalkanoates are also produced by various bacteria as energy-storage molecules [60], and pores assembled from polyphosphate associations with β-hydroxy acid polymers have been isolated from E. coli [61].
While the relative lability of polyesters could be considered a disadvantage, the reversibility of ester formation provides an experimental model to study dynamic synthesis/hydrolysis cycling, potentially caused by environmental perturbations such as seasonal or diurnal cycling. This property makes polyester systems an abstract model of non-biomolecular prebiotic polymer formation suitable for the study of the environmental selection of different chemical properties. For example, the abiotic reactions between αHA and α-amino acid monomers have been studied in model wet-dry cycling experiments [44]. These co-polymerization reactions initially produce depsipeptide oligomers that contain both ester and amide linkages. Over time oligomers enriched in peptide bonds are generated through a combination of ester–amide bond exchange and ester bond hydrolysis reactions [44]. This is a good example of how a plausibly prebiotic non-biomolecule could help scaffold the formation of a biomolecule type in a primitive environment.
We have previously characterized the polymerization of L-malic acid, a dicarboxylic αHA, by wet-dry cycling simulating various geological settings that dry intermittently [37]. In this system, L-malic acid polymerizes during drying and hydrolyzes during rehydration. If the wet phase temperature in these reactions is sufficiently low, the polymer formed in the dry phase is kinetically trapped and undergoes only partial hydrolysis, suggesting that some components of the product mixture, i.e., the most stable/persistent products, could be carried over from cycle to cycle. The presence of carryover products thus raises the possibility of selective preservation and amplification of specific polymeric structures or functionalities, i.e., stability towards hydrolysis due to intramolecular folding or intermolecular aggregation.
Primitive polymerization processes should ideally survey large swathes of sequence space to effectively find the most stable or catalytic sequences. If a wide range of sequence space is not accessible, it is possible that the most stable and/or catalytic polymers cannot be produced. Reversible esterification offers a facile way to access polymer sequence space. Taking advantage of favorable polymerization thermodynamics, starting from a pool of 5 different αHA monomer types, a large diversity of oligomer sequences was achieved (Figure 3) [38]. The experiments in this study generated vast, likely sequence-complete libraries over a variety of reaction conditions (temperature, concentration, salinity, and presence of congeners) compatible with geochemical settings on primitive Earth and in other Solar System environments.
While many biopolymers, i.e., proteins and nucleic acids, are linear polymers (as was the model polyester system studied in [38], for purely analytical tractability purposes), some biopolymers, i.e., glycogen, amylopectin, and dextran, are branched. In abiotic settings, diverse pools of monomers likely oligomerized without enzymatic control, forming significant amounts of branched and crosslinked polymers, further increasing the chemical complexity above what could be present with purely linear polymers. Such branched polymers have been considered as possible precursors to globular proteins [36] and enzymes [39]. When branched polyesters are generated under wet/dry cycling or continuous drying conditions, different properties may be “imprinted” on the products [62]. For example, periodic wetting, during which partial hydrolysis of the polymer may occur, inhibits polymerization and helping keep the oligomer products water soluble which is presumably important for their function. Continuous drying, on the other hand, favors the production of insoluble high molecular weight crosslinked polyester structures [62].
While we have argued that the prebiotic milieu (for example, chemicals that are produced in model prebiotic reactions such as spark discharge experiments or those found in carbonaceous meteorites) is a “messy” chemical system with high molecular diversity, identifying each and every compound in such a system is still beyond the abilities of contemporary analytical chemistry [13]. The combinatorial hydroxy acid systems that have discussed here could be combinatorially complex, but likely still compositionally simple compared to what might be expected in a real environmental setting. Further advancement of analytical methods and techniques will allow future studies to examine systems including still greater chemical diversity, including both non-biomolecules and biomolecules.

3. Self-Assembled Polyester Microdroplet Compartments

Segregation and compartmentalization are fundamental cellular functions, and these can be achieved even with simple membraneless polyester droplets [40]. In the case of polyesters, drying aqueous solutions of mixed αHAs at a mild temperature (80 °C) produces a combinatorially diverse set of oligomers, which upon rehydration form gels that self-assemble into membraneless droplets (Figure 4). These droplets differentially segregate and compartmentalize various hydrophobic organic dyes and fluorescently labeled nucleic acids [40], suggesting a potential function as primitive compartments. Such droplets, or other proposed prebiotic compartments with similar functionality such as liquid-liquid phase-separated membraneless coacervate droplets [63,64,65] or lipid vesicles [5], may have served as protocells.
The polyester microdroplets generated in our studies likely assemble through hydrophobic interactions, and their surfaces and interiors are also likely non-polar, while their interiors contain less water than the surrounding aqueous environment [40]. Such microdroplets may have provided non-polar environments important for the function, synthesis, or accumulation of certain molecules. In a system with coexisting phases with differing chemical character, a given molecule could have a different affinity for each phase. The ratio of the amount of molecule that dissolves in one phase versus the other is known as the partition coefficient [66]. The partition coefficient largely depends on physicochemical properties of both the solute in question and the two phases it partitions into, and can be modulated by changing the charge, polarity, crowding, etc. of the components. In the case of non-polar polyester droplets in water, the partition coefficient determines how much of a molecule becomes sequestered in the polyester droplets. Non-polar molecules such as a hydrophobic fluorescent dye like thioflavin T tend to segregate within non-polar polyester droplets [40]. The amount of accumulation of thioflavin T within the droplets is correlated to the partition coefficient and inversely correlated to the relative volume of the droplet phase versus that of the aqueous phase, resulting in a specific concentration increase of the solute in the droplet phase. As another example, the structures of proteins change depending on solvent polarity, and would be expected to change within polyester droplets compared to polar solvents [67,68]. Likewise, primitive peptides containing large amounts of hydrophobic residues would likely have different structures (and by extension, functions) in an aqueous environment compared to a non-aqueous environment such as the droplet interiors. The interfaces between different polarity phases, such as those between polyester microdroplets or supercritical carbon dioxide droplets [69] and water, have also been shown to facilitate lipid assembly [40,70] and accumulation of amphiphilic peptides [70]. Such assembly or accumulation mechanisms could be extended to other possibly important early self-assembly chemistries [71].
The ability of polyester microdroplets to segregate and compartmentalize various molecules and reactions with different chemical characteristics could be useful for enabling higher-order chemical phenomena (Figure 5). As a hypothetical example, during template-directed non-enzymatic RNA polymerization in solution using activated nucleotide monomers [72,73], hydrolyzed, deactivated monomers accumulate over time, and these hinder template-directed polymerization by competitively occupying the complementary sites on the template strand [72]. Removal or reactivation of these oligomerization inhibitors is an important unresolved problem [74]. Potentially, conducting template-directed polymerization in the presence of droplet phases with an affinity for deactivated monomers could provide a sink for these inhibitory waste products.
Although polyester droplets could be engineered to display interesting chemical properties, whether the droplets themselves were direct or earlier transitional precursors to modern cells (and how this transition occurred) is an open question. Polyester droplets can scaffold the assembly of lipids around themselves, resulting in a molecularly crowded condensed phase with a lipid boundary [40]. Assembly of such a lipid encapsulated condensed-phases has been demonstrated with various membraneless droplet systems such as coacervates [63,64]. Encapsulating layers could also be composed of fatty acids, phospholipids, or even other molecules such as detergents [75], amphiphilic peptides [76], DNA liquid crystals [77], or even mineral particles [78].
Finally, polyester microdroplets can segregate fluorescently labeled nucleic acids and fluorescent proteins [40]. Primitive cells can, perhaps over-simplistically, be conceived of as molecularly crowded droplets containing genetic or functional polymers encased by a lipid boundary [79]. This is a reductionist view that obviously ignores the role lipids play in generating the cytosolic contents, and the role the cytosolic contents play in generating the cell membrane in modern biology. Nevertheless, that a cell-like lipid-enclosed droplet containing functional polymers [40] can be assembled using non-biomolecules suggests that much remains to be learned about the functional structures that may have been important for the emergence of life.
The study mentioned above [40] only included a small number of αHAs monomers, chosen for analytical tractability, which could only form linear polymers and was simply a proof-of-principle investigation that did not explore the effect of the presence of much greater molecular diversity [15]). However, by increasing the chemical diversity of αHAs such as by varying compositional ratios and introducing other non-amino-acid analog αHAs, the ability for a polymeric system to assemble into such droplets should still remain as long as the synthesized polyesters are of sufficient length (so they are insoluble in water) and that the initial monomer composition is not composed of a large quantity of very soluble species (such as polar monomers, charged monomers, or glycolic acid). In fact, by increasing the diversity of ester-bond forming monomers [18,80], a wider variety of higher-order chemistries could be obtained from droplets of this type. In addition to polyester droplets with non-polar interiors [40], polyester droplets with other interior properties, e.g., polar or charged interiors, could have also assembled depending on their chemical composition. Droplets could also co-assemble with other non-biomolecular chemistries or assemblies. Such complex polyester droplets could have acted as microreactors able to concentrate primitive chemical reactants. While we did not perform any functional searches for catalytic behavior within the droplets, incorporating hyperbranched structures [39], additional functional groups (e.g., charged groups), or other prebiotically available biomolecules (e.g., amino acids) into the droplet composition could result in potential catalytic systems and are a target of current investigations.

4. Prospective

After characterization of a more complex polyester system, the next step would be to increase chemical diversity to include other chemicals that would have likely existed in the same prebiotic milieu. Such realistic chemical diversity would likely result in systems which are difficult to characterize, except with respect to systemic functional or bulk properties. There are a few laboratory simulations of such stepwise increases in chemical diversity (examples include recent work with depsipeptides [44] and DNA/RNA chimeras [81]) but these studies are just the first step towards laboratory simulations of truly complex chemical systems that rival the actual complexity of the prebiotic milieu. It may actually be the case that such extremely chemically complex and heterogeneous systems will have properties which cannot be obtained in simpler and more analytically tractable model systems. Perhaps now that many “pure” systems in OoL theories have been initially characterized, researchers can focus more on the properties of more “messy” chemical systems, including those containing both biomolecules and non-biomolecules.
Some studies suggest the modern complement of biomolecules was at least partially derived via natural selection (e.g., [27,82,83,84]). If so, the question of when and how pre-biomolecular processes, structures, or systems were replaced by the current set of biomolecules would be an important topic of investigation. For example, ribosomes are also able to catalyze polyester formation using tRNAs charged with αHAs instead of α-amino acids, suggesting that biological translation machinery is compatible with such non-biomolecule monomer substrates [85]. Although the process to generate αHA-charged tRNA in this study was through an artificial chemical process, the fact that αHA-charged tRNA could be generated suggests that αHA could have participated in ribosome-catalyzed elongation at some point in history possibly contemporaneously with ribozyme-catalyzed peptide elongation. Co-existence of αHA-charged and α-amino acid-charged tRNA may have resulted in a fairly promiscuous translation system, but eventually, α-amino acids and peptides dominated and thus some mechanism to remove this promiscuity would have been necessary.
While the promiscuity of the ribosome and other examples from biochemistry may be interpreted as vestigial [86], it is possible many, if not most, transitional mechanisms in the evolution of biochemistry have left no easily discernible trace in modern biology. As an analogy, the language the present manuscript is written in is itself a mashup of at least two major previously existing spoken and written languages (Old English and French), which themselves had predecessors that were also cross-seeded before being melded into English. These two languages, by the time of their merging, had themselves already been adapted to using alphabets derived from prior “dead languages” (meaning no one living at the time spoke or wrote them) that those literate in Old English or French also could not have understood. The analogy between human culturally transmitted knowledge and templated or non-templated chemical information transfer is obviously imperfect, but it points to the gaps in our knowledge and perhaps our imagination as to how such transitions may have occurred. This is especially true when the up to one billion years [87] available for the origins of life is considered.
Biological evolution is not teleological [88], and early chemical evolution could not have been either. Rather, each new chemical circumstance that arose from previous non-directed reactions created conditions that allowed for the creation of new reaction pathways, perhaps driven by changes in the environment. While we don’t know which exact geological process would have a role in the Origins of Life, Earth’s geological environment has changed over time (e.g., [89]). Assuming a diverse chemical milieu earlier on, with each subsequent geological change (such as air temperature, water salinity, mineral chemistry, atmosphere composition, etc.), some molecules/polymers that persisted, evolved, and/or functioned in an earlier geological state could either degrade or fail to function in the new geological state. At the same time, a select group of molecules could still persist in the new geological state. As the geological state changes again and again over time, this results in even more “pruning” of the total number of functional molecules in the system to result in a small set of functional molecules which persisted throughout all cycles of geological change. These geological transitions, assuming that geological changes were the only transitions on early Earth before the advent of living systems (once living (or life-like) systems emerged, then geological changes would not have been the only potential selective force on early Earth) thus led the way for the transition from the diverse prebiotic milieu into more selected and functional early biochemistries. As an example, one reasonable possibility is that early in Earth’s history, there was a much higher pCO2 atmosphere which would have resulted in considerably more acidic surface waters [90]. As esters are more stable than amides under acidic conditions, it is possible that as the pCO2 was lowered over time, the pH of surface waters increased, leading to a relatively greater stability of amides vs. esters as we see in present day biology. Such environmental or geological changes could also have led to innovation of increasingly specific and functional catalysts from the initially diverse and chaotic prebiotic milieu [91,92]. It is unclear whether the earliest catalysts were composed exclusively of biomolecules, or what catalytic functions primitive polymers would need to accomplish. Though it is complicated to study the abiological emergence of functional polymers due to uncertainties about the relevant environmental conditions and reactant availability, various screening and dynamic combinatorial techniques [24,93,94,95,96,97] may be useful to probe the properties of complex chemical systems. Nevertheless, understanding the potential contribution of the likely diversity of prebiotic organic chemistry on early Earth to contribute to life’s emergence is a worthy enterprise.

Author Contributions

Conceptualization, K.C., H.J.C.II, T.Z.J.; methodology, K.C., I.M., H.J.C.II, T.Z.J.; writing—original draft preparation, K.C., I.M., H.J.C.II, T.Z.J.; writing—review and editing, K.C., I.M., H.J.C.II, T.Z.J. All authors have read and agree to the published version of the manuscript.

Funding

K.C. is supported by the European Structural and Investment Funds Operational Programme “Research, Development and Education”-funded project “ChemJets” (CZ 02.2.69/0.0/0.0/16_027/0008351) and Research Encouragement Fund UKM (GGP-2019-029). T.Z.J., I.M., and H.J.C.II are researchers at the Earth-Life Science Institute at Tokyo Institute of Technology, which is supported under the World Premier International Research Center (WPI) Initiative of the Japan Ministry of Education, Culture, Sports, Science and Technology. T.Z.J. is further supported by JSPS Kakenhi Grant-in-Aid JP18K14354 and Japan Astrobiology Center Project Grant AB311021, while I.M. is further supported by JSPS Kakenhi Grant-in-Aid JP17K01943. Initial discussion of this work occurred at the Earth-Life Science Institute, Tokyo in 2017, where K.C was supported by a travel grant from the ELSI Origins Network (EON) which was supported by a grant from the John Templeton Foundation. The opinions expressed in this publication are those of the authors and do not necessarily reflect the views of the John Templeton Foundation. Additionally, some crucial discussions for this paper occurred during the 2019 European Astrobiology Network Association (EANA) meeting in September 2019 in Orléans, France, and T.Z.J. was supported by a Programme Exploration France travel grant from the French Embassy in Japan and an early career travel grant from EANA.

Acknowledgments

We would like to thank André Brack (Centre National de la Recherche Scientifique, Orleans), Jan Heyda, Sergey Antipov, Ewa Wielogórska and Giyaullah Habibullah (all from the University of Chemistry and Technology, Prague), Ajith Harish (Uppsala University, Uppsala), Martina Preiner (Heinrich Heine University, Düsseldorf), and Niraja Bapat (Indian Institute of Science Education and Research, Pune) for their critical reading and comments of early versions of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scharf, C.; Virgo, N.; Cleaves, H.J.; Aono, M.; Aubert-Kato, N.; Aydinoglu, A.; Barahona, A.; Barge, L.M.; Benner, S.A.; Biehl, M.; et al. A strategy for origins of life research. Astrobiology 2015, 15, 1031–1042. [Google Scholar] [CrossRef] [Green Version]
  2. Mariscal, C.; Barahona, A.; Aubert-Kato, N.; Aydinoglu, A.U.; Bartlett, S.; Cárdenas, M.L.; Chandru, K.; Cleland, C.; Cocanougher, B.T.; Comfort, N.; et al. Hidden concepts in the history and philosophy of origins-of-life studies: A workshop report. Orig. Life Evol. Biosph. 2019, 49, 111–145. [Google Scholar] [CrossRef] [PubMed]
  3. Walker, S.I.; Packard, N.; Cody, G.D. Re-conceptualizing the origins of life. Philos. Trans. A Math. Phys. Eng. Sci. 2017, 375, 20160337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Meringer, M.; Cleaves, H.J. Exploring astrobiology using in silico molecular structure generation. Philos. Trans. A Math. Phys. Eng. Sci. 2017, 375, 20160344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Deamer, D. The role of lipid membranes in life’s origin. Life 2017, 7, 5. [Google Scholar] [CrossRef]
  6. Higgs, P.G.; Lehman, N. The RNA world: Molecular cooperation at the origins of life. Nat. Rev. Genet. 2015, 16, 7–17. [Google Scholar] [CrossRef]
  7. Smith, E.; Morowitz, H.J. Universality in intermediary metabolism. Proc. Natl. Acad. Sci. USA 2004, 101, 13168–13173. [Google Scholar] [CrossRef] [Green Version]
  8. Andras, P.; Andras, C. The origins of life—The “protein interaction world” hypothesis: Protein interactions were the first form of self-reproducing life and nucleic acids evolved later as memory molecules. Med. Hypotheses 2005, 64, 678–688. [Google Scholar] [CrossRef]
  9. Brack, A. The Molecular Origins of Life: Assembling Pieces of the Puzzle; Cambridge University Press: Cambridge, UK, 1998; ISBN 9780521564755. [Google Scholar]
  10. Pizzarello, S.; Shock, E. The organic composition of carbonaceous meteorites: The evolutionary story ahead of biochemistry. Cold Spring Harb. Perspect. Biol. 2010, 2, a002105. [Google Scholar] [CrossRef]
  11. Wolman, Y.; Haverland, W.J.; Miller, S.L. Nonprotein amino acids from spark discharges and their comparison with the Murchison meteorite amino acids. Proc. Natl. Acad. Sci. USA 1972, 69, 809–811. [Google Scholar] [CrossRef] [Green Version]
  12. Schmitt-Kopplin, P.; Gabelica, Z.; Gougeon, R.D.; Fekete, A.; Kanawati, B.; Harir, M.; Gebefuegi, I.; Eckel, G.; Hertkorn, N. High molecular diversity of extraterrestrial organic matter in Murchison meteorite revealed 40 years after its fall. Proc. Natl. Acad. Sci. USA 2010, 107, 2763–2768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Schmitt-Kopplin, P.; Hemmler, D.; Moritz, F.; Gougeon, R.D.; Lucio, M.; Meringer, M.; Müller, C.; Harir, M.; Hertkorn, N. Systems chemical analytics: Introduction to the challenges of chemical complexity analysis. Faraday Discuss. 2019, 218, 9–28. [Google Scholar] [CrossRef] [PubMed]
  14. Koga, T.; Naraoka, H. A new family of extraterrestrial amino acids in the Murchison meteorite. Sci. Rep. 2017, 7, 636. [Google Scholar] [CrossRef] [PubMed]
  15. Pizzarello, S.; Wang, Y.; Chaban, G.M. A comparative study of the hydroxy acids from the Murchison, GRA 95229 and LAP 02342 meteorites. Geochim. Cosmochim. Acta 2010, 74, 6206–6217. [Google Scholar] [CrossRef]
  16. Martins, Z.; Watson, J.S.; Sephton, M.A.; Botta, O.; Ehrenfreund, P.; Gilmour, I. Free dicarboxylic and aromatic acids in the carbonaceous chondrites Murchison and Orgueil. Meteorit. Planet. Sci. 2006, 41, 1073–1080. [Google Scholar] [CrossRef]
  17. Martins, Z. The nitrogen heterocycle content of meteorites and their significance for the origin of life. Life 2018, 8, 28. [Google Scholar] [CrossRef] [Green Version]
  18. Sephton, M.A. Organic compounds in carbonaceous meteorites. Nat. Prod. Rep. 2002, 19, 292–311. [Google Scholar] [CrossRef]
  19. Kahana, A.; Schmitt-Kopplin, P.; Lancet, D. Enceladus: First observed primordial soup could arbitrate origin-of-life debate. Astrobiology 2019, 19, 1263–1278. [Google Scholar] [CrossRef] [Green Version]
  20. Wollrab, E.; Scherer, S.; Aubriet, F.; Carré, V.; Carlomagno, T.; Codutti, L.; Ott, A. Chemical analysis of a “Miller-type” complex prebiotic broth: Part I: Chemical diversity, oxygen and nitrogen based polymers. Orig. Life Evol. Biosph. 2016, 46, 149–169. [Google Scholar] [CrossRef]
  21. Jiang, L.; Dziedzic, P.; Spacil, Z.; Zhao, G.-L.; Nilsson, L.; Ilag, L.L.; Córdova, A. Abiotic synthesis of amino acids and self-crystallization under prebiotic conditions. Sci. Rep. 2014, 4, 6769. [Google Scholar] [CrossRef] [Green Version]
  22. Amend, J.P.; LaRowe, D.E.; McCollom, T.M.; Shock, E.L. The energetics of organic synthesis inside and outside the cell. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20120255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Parker, E.T.; Cleaves, H.J.; Bada, J.L.; Fernández, F.M. Quantitation of α-hydroxy acids in complex prebiotic mixtures via liquid chromatography/tandem mass spectrometry. Rapid Commun. Mass Spectrom. 2016, 30, 2043–2051. [Google Scholar] [CrossRef] [PubMed]
  24. Guttenberg, N.; Virgo, N.; Chandru, K.; Scharf, C.; Mamajanov, I. Bulk measurements of messy chemistries are needed for a theory of the origins of life. Philos. Trans. A Math. Phys. Eng. Sci. 2017, 375, 20160347. [Google Scholar] [CrossRef] [PubMed]
  25. Frenkel-Pinter, M.; Haynes, J.W.; Martin, C.; Petrov, A.S.; Burcar, B.T.; Krishnamurthy, R.; Hud, N.V.; Leman, L.J.; Williams, L.D. Selective incorporation of proteinaceous over nonproteinaceous cationic amino acids in model prebiotic oligomerization reactions. Proc. Natl. Acad. Sci. USA 2019, 116, 16338–16346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ilardo, M.; Bose, R.; Meringer, M.; Rasulev, B.; Grefenstette, N.; Stephenson, J.; Freeland, S.; Gillams, R.J.; Butch, C.J.; Cleaves, H.J. Adaptive properties of the genetically encoded amino acid alphabet are inherited from its subsets. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef]
  27. Ilardo, M.; Meringer, M.; Freeland, S.; Rasulev, B.; Cleaves, H.J. Extraordinarily adaptive properties of the genetically encoded amino acids. Sci. Rep. 2015, 5, 9414. [Google Scholar] [CrossRef] [Green Version]
  28. Becker, S.; Feldmann, J.; Wiedemann, S.; Okamura, H.; Schneider, C.; Iwan, K.; Crisp, A.; Rossa, M.; Amatov, T.; Carell, T. Unified prebiotically plausible synthesis of pyrimidine and purine RNA ribonucleotides. Science 2019, 366, 76–82. [Google Scholar] [CrossRef] [Green Version]
  29. Lancet, D.; Zidovetzki, R.; Markovitch, O. Systems protobiology: Origin of life in lipid catalytic networks. J. R. Soc. Interface 2018, 15, 20180159. [Google Scholar] [CrossRef]
  30. Kauffman, S.A. Approaches to the origin of life on Earth. Life 2011, 1, 34–48. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, S. Lipid-like Self-Assembling Peptides. Acc. Chem. Res. 2012, 45, 2142–2150. [Google Scholar] [CrossRef]
  32. Joyce, G.F.; Schwartz, A.W.; Miller, S.L.; Orgel, L.E. The case for an ancestral genetic system involving simple analogues of the nucleotides. Proc. Natl. Acad. Sci. USA 1987, 84, 4398–4402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Joyce, G.F. RNA Evolution and the Origins of Life. Nature 1989, 338, 217–224. [Google Scholar] [CrossRef] [PubMed]
  34. Cleaves, H.J.; Bada, J.L. The prebiotic chemistry of alternative nucleic acids. In Genesis—In the Beginning: Precursors of Life, Chemical Models and Early Biological Evolution; Seckbach, J., Ed.; Springer: Dordrecht, The Netherlands, 2012; pp. 3–33. ISBN 9789400729414. [Google Scholar]
  35. Hud, N.V.; Cafferty, B.J.; Krishnamurthy, R.; Williams, L.D. The origin of RNA and “my grandfather’s axe”. Chem. Biol. 2013, 20, 466–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Mamajanov, I.; Callahan, M.P.; Dworkin, J.P.; Cody, G.D. Prebiotic alternatives to proteins: Structure and function of hyperbranched polyesters. Orig. Life Evol. Biosph. 2015, 45, 123–137. [Google Scholar] [CrossRef] [PubMed]
  37. Mamajanov, I.; MacDonald, P.J.; Ying, J.; Duncanson, D.M.; Dowdy, G.R.; Walker, C.A.; Engelhart, A.E.; Fernández, F.M.; Grover, M.A.; Hud, N.V.; et al. Ester formation and hydrolysis during wet–dry cycles: Generation of far-from-equilibrium polymers in a model prebiotic reaction. Macromolecules 2014, 47, 1334–1343. [Google Scholar] [CrossRef]
  38. Chandru, K.; Guttenberg, N.; Giri, C.; Hongo, Y.; Butch, C.; Mamajanov, I.; Cleaves, H.J. Simple prebiotic synthesis of high diversity dynamic combinatorial polyester libraries. Commun. Chem. 2018, 1, 30. [Google Scholar] [CrossRef]
  39. Mamajanov, I.; Cody, G.D. Protoenzymes: The case of hyperbranched polyesters. Philos. Trans. A Math. Phys. Eng. Sci. 2017, 375, 20160357. [Google Scholar] [CrossRef] [Green Version]
  40. Jia, T.Z.; Chandru, K.; Hongo, Y.; Afrin, R.; Usui, T.; Myojo, K.; Cleaves, H.J. Membraneless polyester microdroplets as primordial compartments at the origins of life. Proc. Natl. Acad. Sci. USA 2019, 116, 15830–15835. [Google Scholar] [CrossRef] [Green Version]
  41. Galbiati, E.; Zoppo, M.D.; Tieghi, G.; Zerbi, G. Dipole-dipole interactions in simple esters and in liquid-crystal polyesters. Polymer 1993, 34, 1806–1810. [Google Scholar] [CrossRef]
  42. Bordo, D.; Argos, P. The role of side-chain hydrogen bonds in the formation and stabilization of secondary structure in soluble proteins. J. Mol. Biol. 1994, 243, 504–519. [Google Scholar] [CrossRef]
  43. Myers, J.K.; Pace, C.N. Hydrogen bonding stabilizes globular proteins. Biophys. J. 1996, 71, 2033–2039. [Google Scholar] [CrossRef] [Green Version]
  44. Forsythe, J.G.; Yu, S.-S.; Mamajanov, I.; Grover, M.A.; Krishnamurthy, R.; Fernández, F.M.; Hud, N.V. Ester-mediated amide bond formation driven by wet-dry cycles: A possible path to polypeptides on the prebiotic Earth. Angew. Chem. Int. Ed. 2015, 54, 9871–9875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Woese, C.R. On the evolution of cells. Proc. Natl. Acad. Sci. USA 2002, 99, 8742–8747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Meggy, A.B. The free energy of formation of the amide bond in polyamides. J. Appl. Chem. 2007, 4, 154–159. [Google Scholar] [CrossRef]
  47. Hulshof, J.; Ponnamperuma, C. Prebiotic condensation reactions in an aqueous medium: A review of condensing agents. Orig. Life 1976, 7, 197–224. [Google Scholar] [CrossRef]
  48. Leman, L.; Orgel, L.; Ghadiri, M.R. Carbonyl sulfide-mediated prebiotic formation of peptides. Science 2004, 306, 283–286. [Google Scholar] [CrossRef] [Green Version]
  49. Danger, G.; Plasson, R.; Pascal, R. Pathways for the formation and evolution of peptides in prebiotic environments. Chem. Soc. Rev. 2012, 41, 5416–5429. [Google Scholar] [CrossRef]
  50. Leman, L.J.; Huang, Z.-Z.; Ghadiri, M.R. Peptide bond formation in water mediated by carbon disulfide. Astrobiology 2015, 15, 709–716. [Google Scholar] [CrossRef]
  51. Hill, A.R., Jr.; Böhler, C.; Orgel, L.E. Polymerization on the rocks: Negatively-charged alpha-amino acids. Orig. Life Evol. Biosph. 1998, 28, 235–243. [Google Scholar] [CrossRef]
  52. Liu, R.; Orgel, L.E. Polymerization on the rocks: Beta-amino acids and arginine. Orig. Life Evol. Biosph. 1998, 28, 245–257. [Google Scholar] [CrossRef]
  53. Rodriguez-Garcia, M.; Surman, A.J.; Cooper, G.J.T.; Suárez-Marina, I.; Hosni, Z.; Lee, M.P.; Cronin, L. Formation of oligopeptides in high yield under simple programmable conditions. Nat. Commun. 2015, 6, 8385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Compton, R.G.; Bamford, C.H.; Tipper†, C.F.H. Ester Formation and Hydrolysis and Related Reactions; Elsevier: Amsterdam, The Netherlands, 1972; ISBN 9780080868059. [Google Scholar]
  55. Orgel, L.E. Some consequences of the RNA world hypothesis. Orig. Life Evol. Biosph. 2003, 33, 211–218. [Google Scholar] [CrossRef] [PubMed]
  56. Barge, L.M.; Flores, E.; Baum, M.M.; VanderVelde, D.G.; Russell, M.J. Redox and pH gradients drive amino acid synthesis in iron oxyhydroxide mineral systems. Proc. Natl. Acad. Sci. USA 2019, 116, 4828–4833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bada, J.L. New insights into prebiotic chemistry from Stanley Miller’s spark discharge experiments. Chem. Soc. Rev. 2013, 42, 2186–2196. [Google Scholar] [CrossRef] [PubMed]
  58. Sephton, M.A. Organic matter in carbonaceous meteorites: Past, present and future research. Philos. Trans. A Math. Phys. Eng. Sci. 2005, 363, 2729–2742. [Google Scholar] [CrossRef] [PubMed]
  59. Graça, J.; Lamosa, P. Linear and branched poly(omega-hydroxyacid) esters in plant cutins. J. Agric. Food Chem. 2010, 58, 9666–9674. [Google Scholar] [CrossRef] [PubMed]
  60. Poirier, Y.; Nawrath, C.; Somerville, C. Production of polyhydroxyalkanoates, a family of biodegradable plastics and elastomers, in bacteria and plants. Biotechnology 1995, 13, 142–150. [Google Scholar] [CrossRef]
  61. Das, S.; Lengweiler, U.D.; Seebach, D.; Reusch, R.N. Proof for a nonproteinaceous calcium-selective channel in Escherichia coli by total synthesis from (R)-3-hydroxybutanoic acid and inorganic polyphosphate. Proc. Natl. Acad. Sci. USA 1997, 94, 9075–9079. [Google Scholar] [CrossRef] [Green Version]
  62. Mamajanov, I. Wet-dry cycling delays the gelation of hyperbranched polyesters: Implications to the origin of life. Life 2019, 9, 56. [Google Scholar] [CrossRef] [Green Version]
  63. Dora Tang, T.-Y.; Rohaida Che Hak, C.; Thompson, A.J.; Kuimova, M.K.; Williams, D.S.; Perriman, A.W.; Mann, S. Fatty acid membrane assembly on coacervate microdroplets as a step towards a hybrid protocell model. Nat. Chem. 2014, 6, 527–533. [Google Scholar] [CrossRef] [Green Version]
  64. Aumiller, W.M., Jr.; Pir Cakmak, F.; Davis, B.W.; Keating, C.D. RNA-based coacervates as a model for membraneless organelles: Formation, properties, and interfacial liposome assembly. Langmuir 2016, 32, 10042–10053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Jia, T.Z.; Hentrich, C.; Szostak, J.W. Rapid RNA exchange in aqueous two-phase system and coacervate droplets. Orig. Life Evol. Biosph. 2014, 44, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Tsai, W.-T. Environmental risk assessment of hydrofluoropolyethers (HFPEs). J. Hazard. Mater. 2007, 139, 185–192. [Google Scholar] [CrossRef] [PubMed]
  67. Soares, C.M.; Teixeira, V.H.; Baptista, A.M. Protein structure and dynamics in nonaqueous solvents: Insights from molecular dynamics simulation studies. Biophys. J. 2003, 84, 1628–1641. [Google Scholar] [CrossRef] [Green Version]
  68. Yasuda, S.; Oshima, H.; Kinoshita, M. Structural stability of proteins in aqueous and nonpolar environments. J. Chem. Phys. 2012, 137, 135103. [Google Scholar] [CrossRef] [Green Version]
  69. Mayer, C.; Schreiber, U.; Dávila, M.J. Periodic vesicle formation in tectonic fault zones—An ideal scenario for molecular evolution. Orig. Life Evol. Biosph. 2015, 45, 139–148. [Google Scholar] [CrossRef] [Green Version]
  70. Mayer, C.; Schreiber, U.; Dávila, M.J. Selection of prebiotic molecules in amphiphilic environments. Life 2017, 7, 3. [Google Scholar] [CrossRef] [Green Version]
  71. Frederix, P.W.J.; Scott, G.G.; Abul-Haija, Y.M.; Kalafatovic, D.; Pappas, C.G.; Javid, N.; Hunt, N.T.; Ulijn, R.V.; Tuttle, T. Exploring the sequence space for (tri-)peptide self-assembly to design and discover new hydrogels. Nat. Chem. 2014, 7, 30–37. [Google Scholar] [CrossRef] [Green Version]
  72. Prywes, N.; Blain, J.C.; Del Frate, F.; Szostak, J.W. Nonenzymatic copying of RNA templates containing all four letters is catalyzed by activated oligonucleotides. Elife 2016, 5, e17756. [Google Scholar] [CrossRef] [Green Version]
  73. Sosson, M.; Richert, C. Enzyme-free genetic copying of DNA and RNA sequences. Beilstein J. Org. Chem. 2018, 14, 603–617. [Google Scholar] [CrossRef] [Green Version]
  74. Szostak, J.W. The eightfold path to non-enzymatic RNA replication. J. Syst. Chem. 2012, 3, 2. [Google Scholar] [CrossRef] [Green Version]
  75. Hanczyc, M.M.; Monnard, P.-A. The origin of life and the potential role of soaps. Lipid Technol. 2016, 28, 88–92. [Google Scholar] [CrossRef]
  76. Dehsorkhi, A.; Castelletto, V.; Hamley, I.W. Self-assembling amphiphilic peptides. J. Pept. Sci. 2014, 20, 453–467. [Google Scholar] [CrossRef] [Green Version]
  77. Di Leo, S.; Todisco, M.; Bellini, T.; Fraccia, T.P. Phase separations, liquid crystal ordering and molecular partitioning in mixtures of PEG and DNA oligomers. Liq. Cryst. 2018, 45, 2306–2318. [Google Scholar] [CrossRef] [Green Version]
  78. Cakmak, F.P.; Keating, C.D. Combining catalytic microparticles with droplets formed by phase coexistence: Adsorption and activity of natural clays at the aqueous/aqueous interface. Sci. Rep. 2017, 7, 3215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Zwicker, D.; Seyboldt, R.; Weber, C.A.; Hyman, A.A.; Jülicher, F. Growth and division of active droplets provides a model for protocells. Nat. Phys. 2016, 13, 408–413. [Google Scholar] [CrossRef]
  80. Cronin, J.R.; Pizzarello, S.; Epstein, S.; Krishnamurthy, R.V. Molecular and isotopic analyses of the hydroxy acids, dicarboxylic acids, and hydroxydicarboxylic acids of the Murchison meteorite. Geochim. Cosmochim. Acta 1993, 57, 4745–4752. [Google Scholar] [CrossRef]
  81. Bhowmik, S.; Krishnamurthy, R. The role of sugar-backbone heterogeneity and chimeras in the simultaneous emergence of RNA and DNA. Nat. Chem. 2019, 11, 1009–1018. [Google Scholar] [CrossRef]
  82. Wong, J.T.-F.; Tze-Fei Wong, J. Coevolution theory of the genetic code at age thirty. BioEssays 2005, 27, 416–425. [Google Scholar] [CrossRef]
  83. Granold, M.; Hajieva, P.; Toşa, M.I.; Irimie, F.-D.; Moosmann, B. Modern diversification of the amino acid repertoire driven by oxygen. Proc. Natl. Acad. Sci. USA 2018, 115, 41–46. [Google Scholar] [CrossRef] [Green Version]
  84. Kim, K.M.; Qin, T.; Jiang, Y.-Y.; Chen, L.-L.; Xiong, M.; Caetano-Anollés, D.; Zhang, H.-Y.; Caetano-Anollés, G. Protein domain structure uncovers the origin of aerobic metabolism and the rise of planetary oxygen. Structure 2012, 20, 67–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fahnestock, S.; Neumann, H.; Rich, A. Assay of ester and polyester formation by the ribosomal peptidyltransferase. Methods Enzym. 1974, 30, 489–497. [Google Scholar]
  86. Benner, S.A.; Ellington, A.D.; Tauer, A. Modern metabolism as a palimpsest of the RNA world. Proc. Natl. Acad. Sci. USA 1989, 86, 7054–7058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Javaux, E.J. Challenges in evidencing the earliest traces of life. Nature 2019, 572, 451–460. [Google Scholar] [CrossRef] [PubMed]
  88. Mayr, E. This is Biology: The Science of the Living World; Harvard University Press: Cambridge, MA, USA, 1997; ISBN 9780674884694. [Google Scholar]
  89. Hazen, R.M.; Filley, T.R.; Goodfriend, G.A. Selective Adsorption of L- and D-Amino Acids on Calcite: Implications for Biochemical Homochirality. Proc. Natl. Acad. Sci. USA 2001, 98, 5487–5490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Halevy, I.; Bachan, A. The geologic history of seawater pH. Science 2017, 355, 1069–1071. [Google Scholar] [CrossRef]
  91. Khersonsky, O.; Tawfik, D.S. Enzyme promiscuity: A mechanistic and evolutionary perspective. Annu. Rev. Biochem. 2010, 79, 471–505. [Google Scholar]
  92. Carbonell, P.; Lecointre, G.; Faulon, J.-L. Origins of specificity and promiscuity in metabolic networks. J. Biol. Chem. 2011, 286, 43994–44004. [Google Scholar] [CrossRef] [Green Version]
  93. Reek, J.N.H.; Otto, S. Dynamic Combinatorial Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2010; ISBN 9783527629718. [Google Scholar]
  94. Tena-Solsona, M.; Wanzke, C.; Riess, B.; Bausch, A.R.; Boekhoven, J. Self-selection of dissipative assemblies driven by primitive chemical reaction networks. Nat. Commun. 2018, 9, 2044. [Google Scholar] [CrossRef]
  95. Pappas, C.G.; Shafi, R.; Sasselli, I.R.; Siccardi, H.; Wang, T.; Narang, V.; Abzalimov, R.; Wijerathne, N.; Ulijn, R.V. Dynamic peptide libraries for the discovery of supramolecular nanomaterials. Nat. Nanotechnol. 2016, 11, 960–967. [Google Scholar] [CrossRef]
  96. Surman, A.J.; Rodriguez-Garcia, M.; Abul-Haija, Y.M.; Cooper, G.J.T.; Gromski, P.S.; Turk-MacLeod, R.; Mullin, M.; Mathis, C.; Walker, S.I.; Cronin, L. Environmental control programs the emergence of distinct functional ensembles from unconstrained chemical reactions. Proc. Natl. Acad. Sci. USA 2019, 116, 5387–5392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Vincent, L.; Berg, M.; Krismer, M.; Saghafi, S.S.; Cosby, J.; Sankari, T.; Vetsigian, K.; Cleaves, H.J.; Baum, D.A. Chemical ecosystem selection on mineral surfaces reveals long-term dynamics consistent with the spontaneous emergence of mutual catalysis. Life 2019, 9, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Progressively zoomed in mass spectra (AD) of Murchison meteorite soluble organic extracts, each focusing on a more narrow region within the previous spectrum, showing the diversity of meteoritic organics (and analyses). In (D), 11 molecular assignments were made in a narrow range of 0.012 Da, which suggests that even in such a narrow range, a large diversity of isomers may have existed in primitive chemical milieu. Reprinted with permission from Schmitt-Kopplin, Gabelica, Gougeon, Fekete, Kanawati, Harir, Gebefuegi, Eckel, and Hertkorn, 2010. “High molecular diversity of extraterrestrial organic matter in Murchison meteorite revealed 40 years after its fall.” Proc. Nat. Acad. Sci. USA. 107(7):12763-2768. [12] Copyright Schmitt-Kopplin, et al., with an exclusive License to Publish to NAS.
Figure 1. Progressively zoomed in mass spectra (AD) of Murchison meteorite soluble organic extracts, each focusing on a more narrow region within the previous spectrum, showing the diversity of meteoritic organics (and analyses). In (D), 11 molecular assignments were made in a narrow range of 0.012 Da, which suggests that even in such a narrow range, a large diversity of isomers may have existed in primitive chemical milieu. Reprinted with permission from Schmitt-Kopplin, Gabelica, Gougeon, Fekete, Kanawati, Harir, Gebefuegi, Eckel, and Hertkorn, 2010. “High molecular diversity of extraterrestrial organic matter in Murchison meteorite revealed 40 years after its fall.” Proc. Nat. Acad. Sci. USA. 107(7):12763-2768. [12] Copyright Schmitt-Kopplin, et al., with an exclusive License to Publish to NAS.
Life 10 00006 g001
Figure 2. Mass spectrum of a complex product mixture from a laboratory simulation of an atmospheric discharge reaction consisting of liquid water, methane gas, and ammonium gas starting materials in the presence of cyanamide, showing the remarkable chemical diversity produced from a prebiotically plausible organic chemical reaction. Reprinted with permission from Parker, Cleaves, Bada, and Fernandez. “Quantitation of α-hydroxy acids in complex prebiotic mixtures via liquid chromatography/tandem mass spectrometry.” Rapid Commun. Mass Spectrom. 30(18):2043–2051. [23] Copyright John Wiley & Sons, Inc.
Figure 2. Mass spectrum of a complex product mixture from a laboratory simulation of an atmospheric discharge reaction consisting of liquid water, methane gas, and ammonium gas starting materials in the presence of cyanamide, showing the remarkable chemical diversity produced from a prebiotically plausible organic chemical reaction. Reprinted with permission from Parker, Cleaves, Bada, and Fernandez. “Quantitation of α-hydroxy acids in complex prebiotic mixtures via liquid chromatography/tandem mass spectrometry.” Rapid Commun. Mass Spectrom. 30(18):2043–2051. [23] Copyright John Wiley & Sons, Inc.
Life 10 00006 g002
Figure 3. Mass spectra showing the synthesis of polyesters from simple drying of solutions of the alpha hydroxy acids (αHAs) (A) glycolic acid (GA), (B) lactic acid (LA), and (C) a mixture of five different αHAs (LA, GA, DL-2-hydroxy-4-methylpentanoic acid, DL-2-hydroxy-4-(methylsulfanyl) butanoic acid, and DL-3-phenyllactic acid). Insets show higher m/z regions. Asterisks represent peaks showing water loss. Reprinted with permission from Chandru, Guttenberg, Giri, Hongo, Butch, Mamajanov, and Cleaves. 2018. “Simple prebiotic synthesis of high diversity dynamic combinatorial polyester libraries.” Commun. Chem. 1:30. [38] under a Creative Commons License.
Figure 3. Mass spectra showing the synthesis of polyesters from simple drying of solutions of the alpha hydroxy acids (αHAs) (A) glycolic acid (GA), (B) lactic acid (LA), and (C) a mixture of five different αHAs (LA, GA, DL-2-hydroxy-4-methylpentanoic acid, DL-2-hydroxy-4-(methylsulfanyl) butanoic acid, and DL-3-phenyllactic acid). Insets show higher m/z regions. Asterisks represent peaks showing water loss. Reprinted with permission from Chandru, Guttenberg, Giri, Hongo, Butch, Mamajanov, and Cleaves. 2018. “Simple prebiotic synthesis of high diversity dynamic combinatorial polyester libraries.” Commun. Chem. 1:30. [38] under a Creative Commons License.
Life 10 00006 g003
Scheme 1. Simple dehydration synthesis of an arbitrary number of αHAs with side-chains R1 and R2 (left), resulting in production of polyester polymers through condensation by loss of water molecules.
Scheme 1. Simple dehydration synthesis of an arbitrary number of αHAs with side-chains R1 and R2 (left), resulting in production of polyester polymers through condensation by loss of water molecules.
Life 10 00006 sch001
Figure 4. Rehydration of polyesters formed from wet–dry cycling of mixed αHA mixtures results in spontaneous assembly of microdroplets. LA = lactic acid, GA = glycolic acid, PA = phenyllactic acid, SA = 2-hydroxy-4-(methylsulfanyl)butanoic acid, MA = 2-hydroxy-4-methylpentanoic acid. Reprinted with permission from Jia, Chandru, Hongo, Afrin, Usui, Myojo, and Cleaves. 2019. “Membraneless polyester microdroplets as primordial compartments at the origins of life.” Proc. Nat. Acad. Sci. USA. 116(32):15830-15835. [40] Copyright Jia, T.Z., Chandru, K., et al., with an exclusive License to Publish to NAS.
Figure 4. Rehydration of polyesters formed from wet–dry cycling of mixed αHA mixtures results in spontaneous assembly of microdroplets. LA = lactic acid, GA = glycolic acid, PA = phenyllactic acid, SA = 2-hydroxy-4-(methylsulfanyl)butanoic acid, MA = 2-hydroxy-4-methylpentanoic acid. Reprinted with permission from Jia, Chandru, Hongo, Afrin, Usui, Myojo, and Cleaves. 2019. “Membraneless polyester microdroplets as primordial compartments at the origins of life.” Proc. Nat. Acad. Sci. USA. 116(32):15830-15835. [40] Copyright Jia, T.Z., Chandru, K., et al., with an exclusive License to Publish to NAS.
Life 10 00006 g004
Figure 5. Top: A prebiotic chemistry scenario in which a chemical, A, participates in a chemical reaction and produces a toxic byproduct, B, a molecule that inhibits reactions involving A. B accumulates and eventually results in the complete inhibition of reactions involving A (to produce B). Bottom: Chemical A in solution, coexisting with polyester droplets (blue). A produces the toxic molecule B as a byproduct. However, assuming A and B are sufficiently different chemically, B could preferentially partition into the droplets due to B’s affinity for that phase (due to polarity, electrostatic interactions, sterics, etc.), allowing A to continue reacting uninhibited by B. It should be stressed that this is but one possible scenario for the role such droplets could play in higher order prebiotic chemistry, one can easily enumerate numerous other, more complex systems, that could be explored experimentally.
Figure 5. Top: A prebiotic chemistry scenario in which a chemical, A, participates in a chemical reaction and produces a toxic byproduct, B, a molecule that inhibits reactions involving A. B accumulates and eventually results in the complete inhibition of reactions involving A (to produce B). Bottom: Chemical A in solution, coexisting with polyester droplets (blue). A produces the toxic molecule B as a byproduct. However, assuming A and B are sufficiently different chemically, B could preferentially partition into the droplets due to B’s affinity for that phase (due to polarity, electrostatic interactions, sterics, etc.), allowing A to continue reacting uninhibited by B. It should be stressed that this is but one possible scenario for the role such droplets could play in higher order prebiotic chemistry, one can easily enumerate numerous other, more complex systems, that could be explored experimentally.
Life 10 00006 g005

Share and Cite

MDPI and ACS Style

Chandru, K.; Mamajanov, I.; Cleaves, H.J., II; Jia, T.Z. Polyesters as a Model System for Building Primitive Biologies from Non-Biological Prebiotic Chemistry. Life 2020, 10, 6. https://doi.org/10.3390/life10010006

AMA Style

Chandru K, Mamajanov I, Cleaves HJ II, Jia TZ. Polyesters as a Model System for Building Primitive Biologies from Non-Biological Prebiotic Chemistry. Life. 2020; 10(1):6. https://doi.org/10.3390/life10010006

Chicago/Turabian Style

Chandru, Kuhan, Irena Mamajanov, H. James Cleaves, II, and Tony Z. Jia. 2020. "Polyesters as a Model System for Building Primitive Biologies from Non-Biological Prebiotic Chemistry" Life 10, no. 1: 6. https://doi.org/10.3390/life10010006

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