Next Article in Journal
Asymmetric Behavior in Ptyodactylus guttatus: Can a Digit Ratio Reflect Brain Laterality?
Next Article in Special Issue
A Few Experimental Suggestions Using Minerals to Obtain Peptides with a High Concentration of L-Amino Acids and Protein Amino Acids
Previous Article in Journal
Space Habitat Data Centers—For Future Computing
Previous Article in Special Issue
Racemic Phospholipids for Origin of Life Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Hypothesis

Symmetry Breaking of Phospholipids †

Institut de Chimie et de Biochimie Moléculaires et Supramoléculaires (UMR 5246), Université Claude Bernard Lyon 1, Université de Lyon, Bât. Edgar Lederer, 1 rue Victor Grignard, CEDEX, F-69622 Villeurbanne, France
*
Author to whom correspondence should be addressed.
In memory of Océane.
Symmetry 2020, 12(9), 1488; https://doi.org/10.3390/sym12091488
Submission received: 23 July 2020 / Revised: 4 September 2020 / Accepted: 8 September 2020 / Published: 10 September 2020
(This article belongs to the Special Issue Chirality and the Origin of Life)

Abstract

:
Either stereo reactants or stereo catalysis from achiral or chiral molecules are a prerequisite to obtain pure enantiomeric lipid derivatives. We reviewed a few plausibly organic syntheses of phospholipids under prebiotic conditions with special attention paid to the starting materials as pro-chiral dihydroxyacetone and dihydroxyacetone phosphate (DHAP), which are the key molecules to break symmetry in phospholipids. The advantages of homochiral membranes compared to those of heterochiral membranes were analysed in terms of specific recognition, optimal functions of enzymes, membrane fluidity and topological packing. All biological membranes contain enantiomerically pure lipids in modern bacteria, eukarya and archaea. The contemporary archaea, comprising of methanogens, halobacteria and thermoacidophiles, are living under extreme conditions reminiscent of primitive environment and may indicate the origin of one ancient evolution path of lipid biosynthesis. The analysis of the known lipid metabolism reveals that all modern cells including archaea synthetize enantiomerically pure lipid precursors from prochiral DHAP. Sn-glycerol-1-phosphate dehydrogenase (G1PDH), usually found in archaea, catalyses the formation of sn-glycerol-1-phosphate (G1P), while sn-glycerol-3-phosphate dehydrogenase (G3PDH) catalyses the formation of sn-glycerol-3-phosphate (G3P) in bacteria and eukarya. The selective enzymatic activity seems to be the main strategy that evolution retained to obtain enantiomerically pure lipids. The occurrence of two genes encoding for G1PDH and G3PDH served to build up an evolutionary tree being the basis of our hypothesis article focusing on the evolution of these two genes. Gene encoding for G3PDH in eukarya may originate from G3PDH gene found in rare archaea indicating that archaea appeared earlier in the evolutionary tree than eukarya. Archaea and bacteria evolved probably separately, due to their distinct respective genes coding for G1PDH and G3PDH. We propose that prochiral DHAP is an essential molecule since it provides a convergent link between G1DPH and G3PDH. The synthesis of enantiopure phospholipids from DHAP appeared probably firstly in the presence of chemical catalysts, before being catalysed by enzymes which were the products of later Darwinian selection. The enzymes were probably selected for their efficient catalytic activities during evolution from large libraries of vesicles containing amino acids, carbohydrates, nucleic acids, lipids, and meteorite components that induced symmetry imbalance.

Graphical Abstract

1. Introduction

This hypothesis, focusing on how phospholipid symmetry breaking occurs, was intended to complement our experimental paper on racemic phospholipids for the origin of life published in the Special Issue entitled “Chirality and the Origin of Life” [1]. Our hypothesis is based on the evolution of lipid synthesis from raw materials leading to racemic lipids toward the actual occurrence of chiral lipids in all living species. The background of our hypothesis is divided in four parts: (1) the advantage to be homochiral; (2) prebiotic scenarios for the symmetry imbalance of phospholipid precursors; (3) achiral and racemic amphiphiles; and (4) biological synthesis in archaea, bacteria and eukarya. We speculated that the biosynthesis of racemic lipids is less efficient than that of enantiomeric lipids. We will discuss critically the hypothesis by providing scientific evidence to support it.
Studies on the origin of life have been carried out in several directions including dynamic combinatorial chemistry [2], self-assembly and self-organization [3,4], prebiotic chemistry [5,6,7,8], minimal self-replicating molecules [9], autocatalytic systems [10], and the assembly of metabolic and non-metabolic networks [11,12]. The origin of chirality was considered only on a theoretical level with a few exceptions for the abiotic formation of nucleotides [13,14]. A few examples were reported for phospholipids and model membranes [15,16]. In evolved cells, enantiopure membranes are produced in living organisms, which are supramolecular chemical systems that maintain persistent structures and reaction networks through reproduction rather than thermodynamic stability [17]. Although enantiomorphism in crystals is one of the most supposed sources of homochirality of organic compounds on Earth [18], alternate theories have been proposed, such as the enantiomeric cross inhibition [19], for example.

2. The “Advantage” of Being Homochiral

Homochirality has an effect with respect to heterochirality, most strongly at the aggregate or polymer level [20]. Chemical and physical properties of homo- or heterochiral monomers are not sufficiently distinct from each other unless they form aggregates or polymers. The strongest effect is expected to be exerted by dense packing. Crystals can be enantiomorphic in 100% [18], exacerbating more distinct properties than in case of racemic crystals. Heterochiral and homochiral membranes have significant distinct properties in packaging organizations as in lipid rafts and in membrane permeability [1,21]. Phospholipids are aggregates organized as bilayer membranes [22]. Recent investigations showed that the homochirality packaging of phospholipids in prebiotic protocells [20] was not a necessary prerequisite to build up the first protocells. It was sufficient to have compartment property of the heterochiral membranes [1]. For example, bilayers and vesicles composed of heterochiral lipids have a useful permeability, since these bilayers are looser than the more compact homochiral bilayers [1,21]. Such a permeability property could serve to filter and select possible materials to build up primitive organic components, including carbohydrates, lipids, nucleic bases, amino acids and their derivatives. So why are biological membranes made of homochiral phospholipids? The “advantage” of membranes being homochiral rather than heterochiral is that it probably leads to optimal functions of lipophilic peptides and transmembrane proteins. Indeed, the fluidity of the membrane, the specific recognition of lipids with various ligands and lipid raft organization finely tune up enzymatic activities that are significantly different in homo and in heterochiral membranes [21,23]. Lipid composition can modulate and affect enzyme activity, even within homochiral membranes [24].
Several theories concerning membrane compositions in the primitive last common ancestor (LCA) were reviewed [25,26]. Among them, it was proposed that earlier life forms were dependent on the presence of membrane lipids with an isoprenoid hydrocarbon core, especially since fatty acid metabolism is underdeveloped in archaea. Other theories advocated that the most divergent feature is the glycerophosphate backbone [26,27]. Indeed, there is one chiral centre in the glycerol moiety leading to either G1P or G3P phospholipid derivatives, forming the basis of the “lipid divide” theory [26,27]. The key precursors for the biosynthesis of phospholipids in living species is the prochiral dihydroxy acetone phosphate (DHAP) [26,27,28]. Any type of oxidation from DHAP would lead to racemic species as well as to pure enantiomeric species.
Molecules react according to materials and conditions in their proximity. Modern cells are well evolved biochemical machines and the chemical processes are carried out by enzymes, which determine the path of the reactions. However, prebiotic mechanisms in LCA protocells [29,30,31,32], were probably not necessarily similar to the actual ones. System protobiology suggests that lipids played a fundamental role in the emergence of life. Thus, in a hypothetical racemic lipid word, life emerged thanks to the compartmentalization of simple lipophilic or poorly hydrophilic small proteins that showed a catalytic role together with auto-replicative functional nucleic acids [33,34].

3. Prebiotic Scenarios for the Symmetry Imbalance of Phospholipids Precursors

Speculations about where and how life emerged from a primordial soup of abiotic mixtures of molecules are extremely well reviewed and summarized including some aspects on the biological origin of chirality [5,35]. One main conclusion was that all chiral molecules can be formed in both enantiomeric types suggesting that a symmetry imbalance between the two possible stereoisomers occurred. Prebiotic symmetry breaking scenarios were depicted using mathematical models only [20].
Concerning the synthesis of life’s building blocks, Meierhenrich and co-workers showed that the exposure of circularly polarized light (CPL) in simulated interstellar media can induce a mirror symmetry breaking between the formation of l- or d-alanine where the imbalance depends from the wavelength of the incident CPL and sense of rotation [36,37]. Further investigations showed that glyceraldehyde (1, Scheme 1), the first chiral product of the “formose” reaction [38]—one of the chemical pathways for the synthesis of carbohydrates—is present in comets and other space bodies [37]. It is probable that the symmetry imbalance between the two possible stereoisomers of 1 occurred before seeding the Earth by asteroids’ or comets’ impact [39] creating the conditions for deracemization before the prebiotic polymerization of peptides and the formation of nucleic acids. Glyceraldehyde (1), dihydroxyacetone (2) and glycerol (3) together with their phosphate derivatives (4 and 5ac) are the most plausibly chemical precursors of glycero phospholipids such as phospholipid esters and ethers (Scheme 1).
In a well studied prebiotic scenario, Sutherland and co-workers, among others, showed that 1 can be one of the plausible precursors of 5a together with ribonucleosides and a few amino acids such as valine and leucine [40]. DHAP (4), instead, was hypothesized to be a key intermediate in the prebiotic synthesis 3-pentulose and racemic mixtures of erythrulose [41].
Glyceraldehyde (1) and its tautomer (2) (double arrow in Scheme 1) are precursors of glycerol (3) which is the reduced form of 2. The redox reactions (Equations (1)–(6)) that occur should be a key step for stereochemistry imbalance during the phosphorylation or oxidation of glycerol [42] (35a and 5c, Scheme 1). The synthesis of DHAP (4) under plausibly prebiotic reaction conditions is not reported, while the synthesis of glycerol (3) in interstellar ices was simulated instead [43], suggesting that glycerol is plausibly present in space bodies.
The oxidation reaction can be summarized as a loss of electrons in either chemistry or biochemistry. In several biological reactions, as in lipid beta oxidation, glycolysis and the Krebs cycle, the electrons are transferred via cofactor FAD or NAD(P)+. The reduction process is conducted via FADH2 or NAD(P)H which can recycle the cofactors for a next round of oxidation. These reactions can be found everywhere in archaea, bacteria and eukarya, suggesting that it was one of the most efficient oxidation or reduction mechanisms that evolution maintained. The oxidation and the reduction of 2 and 3 can be written in analogy with respect to NAD+/NADH processes.
Concerning the reduction of 2, where R1 and R2 are CH2OH, respectively:
R1-C=(O)-R2 + 2H+ + 2e = R1-CH(OH)-R2
NAD(P)H + H+ = NAD(P)+ + 2e +2H+
R1-C=(O)-R2 + NAD(P)H + H+ = R1-CH(OH)-R2 + NAD(P)+
Concerning the oxidation of 3, where R1 and R2 are CH2OH, respectively:
R1-CH(OH)-R2 = R1-C = (O)-R2 + 2H+ + 2e
NAD(P)+ + 2e- +2H+ = NAD(P)H + H+
R1-CH(OH)-R2 + NAD(P)+ = R1-C=(O)-R2 + NAD(P)H + H+
Tricyanocuprate [Cu(CN)3]2 and tetracyanocuprate [Cu(CN)4]2 are supposed to be a source of electrons for the oxidoreduction of glyceraldehyde (1, Scheme 1) in enzyme-free conditions [44,45,46]. The hydrogen cyanide–cyanocuprate photochemistry has been proven to be effective for synthesis in abiotic conditions of glyceraldehyde precursors starting the oxido-reduction of 1 into 2 then 3, respectively (Scheme 1). However, this system cannot lead to any symmetry imbalance of 1 in the absence of any chiral inductor even if deracemization or interconversion can occur using photocatalysis reaction conditions [47]. Iron (III)-sulfur-l-glutathione complexes are able to oxidize NADPH in catalytic networks across the membrane of model protocells made of (R)-POPC and oleic acid [48]. This suggests that simple but effective catalytic networks probably existed in protocells, before the advent of the LCA. Ferredoxins are one of the most known metallo-proteins and their sequences are well known as three of them were isolated from fermentative bacteria [49]. The presence of a high percentage (>64%) of plausibly prebiotic amino acids in their sequence [50] such as glycine, alanine, valine, proline, glutamic and aspartic acids together with cysteine [51], indicates that short hydrophobic peptides, able to complex iron (III) could have been precursors of ferredoxins in LCA. These peptides in the presence of iron (III) formed aggregates [52] that could perform redox reactions as those with NAD+/NADH in evolved cells. Such peptides may have been formed from the scalemic mixtures of amino acids due to the symmetry imbalance possibly induced by meteorite and comet seedings. The scalemic ratio between each d- and l-amino acid was plausibly improved by CPL [39] or from the presence of enantiomorphic crystals. Thus, amino acid sequences within peptides were selected on the basis of their emerging functions or properties. Their selections should have occurred in large libraries of vesicles containing various biopolymers during evolution (cf. Section 4). In our hypothesis, non-functional sequences were discarded in favour of enantiopure sequences probably due to their distinct structural properties. For example, homo peptides with either pure d- or l amino acid sequences more favourably induce alpha-helix structures than hetero peptides with alternate or stochastic d- and l-amino acids (ldld… or…ldll… or …ddld… etc.). The emblematic case is Gramicidin A, containing alternating l and d residues, which does not form alpha-helix structures but beta-helix structures with C–O moieties of the L residues parallel to helix axis, whereas for the d residues, they are antiparallel to it [53]. This is due to the fact that bulky side chains shall be positioned to the outside helix axis, otherwise bulky side-chain residues positioned inside the helix destabilize the helical structure. Not only is the structural topology of homo peptides different from that of hetero peptides, but their possibilities to interact with charged groups or to form hydrogen bonds are distinct due to the positions of polar groups. The enantiomeric excess in the peptide might have been amplified by autocatalytic pathways, gradually favouring the formation of a peptide containing the first dominating enantiomer [11,30,50,51,52] yielding chemical environments in which the predominance of one enantiopure sequence of peptides was preferred.

4. Achiral and Racemic Amphiphiles

4.1. Non-Chiral Amphiphiles

Obviously, the formation of large vesicles, precursors of protocells [29], occurred before the rise of full-fledged cells, since vesicles form spontaneously in aqueous solution from a variety of surfactants [54]. Closed membranes exert the confinement and protection of an internalised chemical network including reactions on their hydrophobic region [25,28,55]. According to the current view, early membranes were more likely formed from derivatives of alkanols, Ref. [56] fatty acids, Ref. [57] mono-alkyl phosphates, Ref. [58] and isoprenoids [59]. Most probably, they were composed of a mixture of components [60] (Figure 1).

4.2. Racemic Amphiphiles and Their Precursors

Several plausibly prebiotic syntheses were explored, however all the proposed pathways, carried out in enzyme free conditions from glycerol (3), yielded racemic phospholipids (68, Scheme 2A) [56,58,61,62,63,64,65] or racemic mixtures of glycerol phosphates (5ac and 910, Scheme 2B [40,66,67,68,69,70]. In addition, the symmetry imbalance between the R:S ratio of mono- and dialky phosphates (6 and 12, Scheme 2C) and cyclic glycerophosphates (cGP, 13, Scheme 2D) [71,72] from di-acyl glycerols 12 or glycerol 3, respectively, were not reported or investigated either. Remarkably, all the crude mixtures containing 68, 12 and 13 were able, using appropriate buffers, to form giant vesicles that per sizes and membrane properties are similar to those of the bilayer of modern cells.

5. Biological Synthesis in Archaea, Bacteria and Eukarya

5.1. Lipid Characteristics in Archaea, Bacteria and Eukarya

One essential characteristic of living species is their ability to create a compartmentalization of bioactive molecules [75,76,77]. The natural enantiomer of all phosphatidate derivatives, in eukarya and in most of bacteria, is d-diacylglycerol phosphate (Fischer convention), 1,2-diacyl-sn-glycerol-3-phosphate (sn-glycerol nomenclature) [78] or 2R—in the Cahn–Ingold–Prelog formalism (Figure 2) [79,80]. The opposite configuration is l-diacylglycerol phosphate, 2,3-diacyl-sn-glycerol-1-phosphate or (2S), which occurs mostly in archaea membranes. The archaea phospholipids usually contain isoprenoid glycerol ethers instead of hydrocarbon glycerol esters [81].
Last common ancestor (LCA) or Commonote Commonote (C. Commonote) [82], lived at around 3.5–3.8 Ga. There are still controversies about the environment where the LCA lived [83]. A sulphur-containing atmosphere [84,85,86] together with CO2, H2, N2 and CH4 [87] is the most probable. Contemporary archaea, comprising of methanogens (which generate actually around 85% of the methane in Earth’s atmosphere), halobacteria and thermoacidophiles are living under extreme conditions reminiscent of this primitive environment. These descendants are phylogenetically related to each other, while they share very little phylogenetic characteristics with bacteria and eukarya [25,26]. C. Commonote had archaeal and bacterial characteristics [88,89,90,91], while eukarya evolved from archaea [88,90,91]. There is an open debate between three domains of life, archaea, bacteria and eukarya which evolved separately from LCA versus Eocyte hypothesis where eukarya are descendent of prokaryotic Crenarchaeota [92] or other evolution models [26,91]. The origin of the controversy lies in the inconsistencies of the phylogenetic distributions and in the selection of appropriate genes to build up the phylogenetic tree [26,91,93]. Here, we focus on the phylogenetic tree based from the genes that encode sn-glycerol-1-phosphate dehydrogenase (G1DPH) or sn-glycerol-3-phosphate dehydrogenase (G3DPH), enzymes catalysing, respectively, sn-glycerol-1-phosphate (G1P) and sn-glycerol-3-phosphate (G3P) from pro-chiral DHAP. The reason to focus on the two genes for encoding G1DPH and G3DPH in this review is that G1P and G3P are key precursors of phospholipids and are essential to determine the mechanisms of symmetry breaking. The lipid composition in archaea is distinct from those in bacteria and eukarya [26,91]. Archaea membranes contain usually phospholipids having G1P moiety and isoprenoid hydrocarbon chains ether-linked to the G1P moiety [26], whereas membranes in bacteria and eukarya are usually composed of phospholipids derived from G3P and alkanoyl chains ester-linked to the G3P moiety (Figure 3) [25,26].

5.2. Appearance of Homochiral Membranes Based on Phylogenetic Analysis on Enzymes Forming sn-Glycerol-1-phosphate or sn-Glycerol-3-phosphate

One likely path of lipid synthesis at the appearance of the extremophile LCA is the geochemical production of racemic lipids via non-catalytic or catalytic, but enzyme-free pathways, giving rise to racemic membranes (Figure 3). Then, the appearance of homochiral membranes from, probably later in the evolution, in archaea, signals a selective catalytic activity that could have been initiated by non-enzymatic or enzymatic ways [77].
Pro-chiral DHAP, (4), is a starting material for the synthesis of lipids in all the three domains of life: archaea, bacteria and eukarya. The first step to obtain phospholipid precursors in archaea is the hydrogenation catalysed by G1PDH which gives G1P with NADH + H+ as proton donors (Scheme 3B). In bacteria and eukarya, the first step to obtain the phospholipid precursors is catalysed by a G3PDH giving rise to G3P (Scheme 3A) [25,26].
Generally, there are two biosynthetic pathways to obtain G3P in bacteria and in eukarya, (Scheme 3A) while there is only one to obtain G1P in archaea (Scheme 3B) [25,26]. G3P can be produced from glycerol and is catalysed by a glycerol kinase (GK) in bacteria and eukarya, while archaea lacks GK [26]. To the best of our knowledge, there is no GK producing G1P. To date, the catalytic activity of the one and the same enzyme producing both G1P and G3P from glycerol, if it existed, was not retained during evolution. The genes coding for G1PDH, G3PDH and GK in archaea and bacteria are used to construct phylogenetic trees [26,28,94,95]. This reveals possible evolutions of synthetic phospholipid pathways from a common ancestor [92]. Several models, based on the occurrence of G1P-lipids or G3P-lipids, were inferred from the presence of either G1PDH or G3PDH [25,26,96,97]. The separate evolution of G1PDH or G3PDH is supported from several phylogenetic analyses [26,28,94,95]. We summarize a few facts from these reports in this Section. According to the phylogenetic analyses, it seems that there is no any direct evidence that C. commonote could form G1P lipids via enzymatic reactions [26]. Among the Commonote ancestors having chiral membranes, Commonote archaea (C. archaea) were probably the first biological entities to be formed, since they could live with little concentration of O2. The early stage of the archaeal lineage, had G3DPH (GLpA/GlpD) leading to G3P lipid membranes instead of G1P lipid membranes (Figure 3) [26]. In the next stage of the evolution path, the archaeal lineage acquired G1PDH (Egs A) probably leading to a population of archaeal lineage mixed with GIPDH and G3PDH [26]. The archaeal descendants are phylogenetically related to each other, while they share very little phylogenetic characteristics with bacteria, Refs. [89,90,98,99] suggesting that both living organisms C. archaea and Commonote bacteria (C. bacteria) evolved separately (Figure 3). Commonote eukarya (C. eukarya) could have appeared much later than C. archaea since it was suggested that eukarya originated from archaea [88,90,100], consistent with the eocyte hypothesis [92]. Lokiarchaeta is closely related to eukarya because of the absence of gene for coding G1PDH, and the presence of a gene coding for G3PDH [100,101], which is rarely observed in archaea (Figure 3). This supports the evidence that eukarya originated from archaea.
During, the evolutionary path, there is the possibility that genes could have been horizontally exchanged (grey dashed lines between two arms of the evolution tree in Figure 3) [102]. This is supported by the fact that in certain bacterial lineages GIPDH (Egs A) having stereospecific synthesis of G1P from DHAP, was acquired as another lineage GIPDH (AraM) [26]. On the other hand, in certain archeal lineages G3PDH (GpsA), which is the major G3PDH of modern bacterial species, was acquired via horizontal transfer [26]. G3PDH homologs such as GlpA and GlpD, found in various eukaryotic cells, are involved in glycerol shuttle and not in the formation of G3P in cellular membrane [26]. Therefore, the horizontal transfer of genes coding for the enzymes could have resulted in a shift of enzymatic activity. G1PDH and G3PDH may coexist in C. bacteria or in C. archaea (Figure 3) [26]. Although it is rarely observed. These observations suggest that the choice of which kind of chirality, that of G1P or that of G3P, was not accidental but resulted from an efficient catalytic activity that was retained during the evolutionary process from LCA and C. commonote. Of interest, the evolution retained both catalytic activities. However, G1PDH and G3PDH genes are different, suggesting that archaea and bacteria evolved apart from one another (Figure 3). This is supported by the fact that G1PDH and G3PDH are not “mirror-image” enzymes since both had L amino acids. Indeed, G1PDH belongs to a larger structurally related superfamily comprising of NAD(P)H-dependent hydrogenases, including alcohol dehydrogenase, UDP-glucose 6-dehydrogenase, 3-hydroxyacyl-CoA dehydrogenase and dehydroquinate synthase, which are all unrelated to G3PDH [25,26]. Since G1PDH and G3PDH are not homologs, then one shall ask what is the evidence of convergent evolution? The convergence lies in the fact that both enzymes use the same substrate that is DHAP to yield either G1P or G3P phospholipid derivatives [26]. Furthermore, both enzymes have L amino acids indicating a common origin. Taken together, these facts indicate that G1PDH and G3PDH occurred much later in the evolution path than LCA and were retained during Darwinian evolution (Figure 3).

6. Conclusions

The question of why living species did not retain racemic lipids to form their membranes during the evolution path remains unanswered, as thus far only mathematical models have been used. We speculate that the biosynthesis of racemic lipids is less efficient than that of enantiomeric lipids. Indeed, only G1DPH and G3DPH enzymes leading to their respective enantiomeric G1P and G3P are actually observed in all living systems, while there are no enzymes producing a racemic mixture of G1P and G3P from DHAP. G1DPH and G3DPH evolved apart from one another since they are structurally different and are not “mirror image” enzymes. These enzymes, which appeared much later in the evolution path than the LCA, are essential, since they catalyse the formation of phospholipid precursors G1P and G3P from the pro-chiral DHAP. Indeed, DHAP is a key molecule for phospholipid metabolism. What is missing is why the Darwinian selection retained only these enzymes, G1DPH and G3PDH? To fill the gap between LCA, that probably possessed racemic membranes, and protocells with homochiral membranes, several hypothesises could be formulated. Apparently racemic membranes do not have the same properties as those in enantiomeric membranes due to their distinct ability to form lipid rafts, recognition process and packing organizations [21,23]. From a chemistry perspective, several aspects of this problem could be tackled. Firstly, organic synthesis from DHAP yielding racemic and enantiomeric lipid precursors under prebiotic conditions shall provide more insight into their mechanisms and efficiencies. Secondly, further analysis on the physico-chemical properties of vesicles made either from racemic or enantiomerically pure lipids may support the notion that the overall property of membranes made either by racemic and or enantiomeric lipids are distinct.
To conclude, our hypothesis speculates that the chemical evolution of proteins [103,104,105] allowed the biosynthesis of enantiomeric lipids [33]. Large libraries of vesicles containing biopolymers including amino acids, carbohydrates, nucleotides or other meteorite materials may have served as possible sources of symmetry imbalance (Figure 4).
Not only homochiral vesicles, but also vesicles made of racemic phospholipids or mixed achiral amphiphiles may contribute to the selection process of retaining the best enzymes able to catalyse a reaction from achiral DHAP to form enantiopure lipid precursors (Figure 4). The synthesis of enantiopure phospholipids occurred firstly in the presence of chemical catalysts, before being catalysed by enzymes which are the products of a later evolution stage. Symmetry imbalance in the deracemization of racemic mono- or di-alkyl glycerol could appear at different stages of the evolution process, driven by interaction on mineral surfaces [106] such as graphene. The growth and division of lipid boundaries [107] and the formation of enantiomerically pure vesicles drastically contributed to this selection process. Further studies on the symmetry breaking of phospholipids in protocell membranes can be carried out using synthetic protocells where the transmission of catalytic protein can be controlled under selection processes upon growth and division experiments [107,108,109,110,111]. The compartmentalization of primitive enzyme-free or enzymic molecular replicators, inside the organelles and/or protocells, was probably one of several strategies that evolution retained for the Darwinian selection processes.

Author Contributions

Both authors contribute equally to the conceptualization, resources, data curation, writing—original draft preparation, revision and review this hypothesis. Both authors contribute equally to the funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

M.F. dedicates this work to the memory of his daughter Océane (2015–2017). Pasquale Stano and Peter Strazewski are gratefully acknowledged for the useful discussions on the theme of symmetry breaking and for have reading the first draft of the manuscript. We also wish to thank Yannik Vallée (guest editor) that encouraged us to submit this hypothesis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Altamura, E.; Comte, A.; D’Onofrio, A.; Roussillon, C.; Fayolle, D.; Buchet, R.; Mavelli, F.; Stano, P.; Fiore, M.; Strazewski, P. Racemic Phospholipids for Origin of Life Studies. Symmetry 2020, 12, 1108. [Google Scholar] [CrossRef]
  2. Corbett, P.T.; Leclaire, J.; Vial, L.; West, K.R.; Wietor, J.-L.; Sanders, J.K.M.; Otto, S. Dynamic Combinatorial Chemistry. Chem. Rev. 2006, 106, 3652–3711. [Google Scholar] [CrossRef] [PubMed]
  3. Whitesides, G.M.; Boncheva, M. Beyond molecules: Self-assembly of mesoscopic and macroscopic components. Proc. Natl. Acad. Sci. USA 2002, 99, 4769–4774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lehn, J.-M. Toward Self-Organization and Complex Matter. Science 2002, 295, 2400–2403. [Google Scholar] [CrossRef] [Green Version]
  5. Ruiz-Mirazo, K.; Briones, C.; De La Escosura, A. Prebiotic systems chemistry: New perspectives for the origins of life. Chem. Rev. 2014, 114, 285–366. [Google Scholar] [CrossRef]
  6. Islam, S.; Powner, M.W. Prebiotic Systems Chemistry: Complexity Overcoming Clutter. Chem 2017, 2, 470–501. [Google Scholar] [CrossRef] [Green Version]
  7. Fiore, M.; Strazewski, P. Bringing Prebiotic Nucleosides and Nucleotides Down to Earth. Angew. Chem. 2016, 55, 13930–13933. [Google Scholar] [CrossRef]
  8. Fiore, M. The Origin and Early Evolution of Life: Prebiotic Chemistry. Life 2019, 9, 73. [Google Scholar] [CrossRef] [Green Version]
  9. Duim, H.; Otto, S. Towards open-ended evolution in self-replicating molecular systems. Beilstein J. Org. Chem. 2017, 13, 1189–1203. [Google Scholar] [CrossRef] [Green Version]
  10. Bissette, A.J.; Fletcher, S.P. Selecting complex behaviour. Nat. Chem. 2015, 7, 15–17. [Google Scholar] [CrossRef]
  11. Peretó, J. Out of fuzzy chemistry: From prebiotic chemistry to metabolic networks. Chem. Soc. Rev. 2012, 41, 5394–5403. [Google Scholar] [CrossRef] [PubMed]
  12. Dadon, Z.; Wagner, N.; Ashkenasy, G. The Road to Non-Enzymatic Molecular Networks. Angew. Chem. 2008, 47, 6128–6136. [Google Scholar] [CrossRef] [PubMed]
  13. 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]
  14. 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]
  15. Eghiaian, F.; Rico, F.; Colom, A.; Casuso, I.; Scheuring, S. High-speed atomic force microscopy: Imaging and force spectroscopy. FEBS Lett. 2014, 588, 3631–3638. [Google Scholar] [CrossRef]
  16. Böhm, C.; Möhwald, H.; Leiserowitz, L.; Als-Nielsen, J.; Kjaer, K. Influence of chirality on the structure of phospholipid monolayers. Biophys. J. 1993, 64, 553–559. [Google Scholar] [CrossRef] [Green Version]
  17. Eschenmoser, A.; Loewenthal, E. Chemistry of potentially prebiological natural products. Chem. Soc. Rev. 1992, 21, 1. [Google Scholar] [CrossRef]
  18. Klabunovskii, E.I. Can Enantiomorphic Crystals like Quartz Play a Role in the Origin of Homochirality on Earth? Astrobiology 2001, 1, 127–131. [Google Scholar] [CrossRef]
  19. 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] [Green Version]
  20. Subramanian, H.; Gatenby, R.A. Evolutionary advantage of directional symmetry breaking in self-replicating polymers. J. Theor. Biol. 2018, 446, 128–136. [Google Scholar] [CrossRef]
  21. Eghiaian, F. Lipid Chirality Revisited: A Change in Lipid Configuration Transforms Membrane-Bound Protein Domains. Biophys. J. 2015, 108, 2757–2758. [Google Scholar] [CrossRef] [Green Version]
  22. Altamura, E.; Stano, P.; Walde, P.; Mavelli, F. Giant vesicles as micro-sized enzymatic reactors: Perspectives and recent experimental advancements. Int. J. Unconv. Comput. 2015, 11, 5–21. [Google Scholar]
  23. Schreiber, A.; Huber, M.C.; Schiller, S.M. Prebiotic Protocell Model Based on Dynamic Protein Membranes Accommodating Anabolic Reactions. Langmuir 2019, 35, 9593–9610. [Google Scholar] [CrossRef] [PubMed]
  24. Spector, A.A.; Yorek, M.A. Membrane lipid composition and cellular function. J. Lipid Res. 1985, 26, 1015–1035. [Google Scholar] [PubMed]
  25. Koga, Y.; Kyuragi, T.; Nishihara, M.; Sone, N. Did Archaeal and Bacterial Cells Arise Independently from Noncellular Precursors? A Hypothesis Stating that the Advent of Membrane Phospholipid with Enantiomeric Glycerophosphate Backbones Caused the Separation of the Two Lines of Descent. J. Mol. Evol. 1998, 46, 54–63. [Google Scholar] [CrossRef] [PubMed]
  26. Yokobori, S.; Nakajima, Y.; Akanuma, S.; Yamagishi, A. Birth of Archaeal Cells: Molecular Phylogenetic Analyses of G1P Dehydrogenase, G3P Dehydrogenases, and Glycerol Kinase Suggest Derived Features of Archaeal Membranes Having G1P Polar Lipids. Archaea 2016, 2016, 1802675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Koga, S.; Williams, D.S.; Perriman, A.W.; Mann, S. Peptide-nucleotide microdroplets as a step towards a membrane-free protocell model. Nat. Chem. 2011, 3, 720–724. [Google Scholar] [CrossRef]
  28. Peretó, J.; López-García, P.; Moreira, D. Ancestral lipid biosynthesis and early membrane evolution. Trends Biochem. Sci. 2004, 29, 469–477. [Google Scholar] [CrossRef]
  29. Lopez, A.; Fiore, M. Investigating prebiotic protocells for a comprehensive understanding of the origins of life: A prebiotic systems chemistry perspective. Life 2019, 9, 49. [Google Scholar] [CrossRef] [Green Version]
  30. Stano, P.; Altamura, E.; Mavelli, F. Novel directions in molecular systems design: The case of light-transducing synthetic cells. Commun. Integr. Biol. 2017, 10, 3837–3842. [Google Scholar] [CrossRef]
  31. Altamura, E.; Milano, F.; Tangorra, R.R.; Trotta, M.; Omar, O.H.; Stano, P.; Mavelli, F. Highly oriented photosynthetic reaction centers generate a proton gradient in synthetic protocells. Proc. Natl. Acad. Sci. USA 2017, 114, 3837–3842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Altamura, E.; Albanese, P.; Marotta, R.; Milano, F.; Fiore, M.; Trotta, M.; Stano, P.; Mavelli, F. Light-driven ATP production promotes mRNA biosynthesis inside hybrid multi-compartment artificial protocells. bioRxiv 2020. [Google Scholar] [CrossRef]
  33. 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] [PubMed]
  34. Strazewski, P. Low-digit and high-digit polymers in the origin of life. Life 2019, 9, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Blackmond, D.G. The origin of biological homochirality. Cold Spring Harb. Perspect. Biol. 2010, 2, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Meinert, C.; Hoffmann, S.V.; Cassam-Chenaï, P.; Evans, A.C.; Giri, C.; Nahon, L.; Meierhenrich, U.J. Photonenergy-Controlled Symmetry Breaking with Circularly Polarized Light. Angew. Chem. 2014, 53, 210–214. [Google Scholar] [CrossRef]
  37. Garcia, A.; Meinert, C.; Sugahara, H.; Jones, N.; Hoffmann, S.; Meierhenrich, U. The Astrophysical Formation of Asymmetric Molecules and the Emergence of a Chiral Bias. Life 2019, 9, 29. [Google Scholar] [CrossRef] [Green Version]
  38. Benner, S.A.; Kim, H.J.; Biondi, E. Prebiotic chemistry that could not not have happened. Life 2019, 9, 84. [Google Scholar] [CrossRef] [Green Version]
  39. Myrgorodska, I.; Meinert, C.; Martins, Z.; d’Hendecourt, L.L.S.; Meierhenrich, U.J. Molecular Chirality in Meteorites and Interstellar Ices, and the Chirality Experiment on Board the ESA Cometary Rosetta Mission. Angew. Chem. 2015, 54, 1402–1412. [Google Scholar] [CrossRef]
  40. Patel, B.H.; Percivalle, C.; Ritson, D.J.; Duffy, C.D.; Sutherland, J.D. Common origins of RNA, protein and lipid precursors in a cyanosulfidic protometabolism. Nat. Chem. 2015, 7, 301–307. [Google Scholar] [CrossRef] [Green Version]
  41. Simonov, A.N.; Pestunova, O.P.; Matvienko, L.G.; Snytnikov, V.N.; Snytnikova, O.A.; Tsentalovich, Y.P.; Parmon, V.N. Possible prebiotic synthesis of monosaccharides from formaldehyde in presence of phosphates. Adv. Space Res. 2007, 40, 1634–1640. [Google Scholar] [CrossRef]
  42. Pasek, M.A. Thermodynamics of Prebiotic Phosphorylation. Chem. Rev. 2019. [Google Scholar] [CrossRef] [PubMed]
  43. Kaiser, R.I.; Maity, S.; Jones, B.M. Synthesis of Prebiotic Glycerol in Interstellar Ices. Angew. Chem. 2015, 127, 197–202. [Google Scholar] [CrossRef]
  44. Wachtershauser, G. Origin of Life: Life as We Don’t Know It. Science 2000, 289, 1307–1308. [Google Scholar] [CrossRef]
  45. Sutherland, J.D. The Origin of Life-Out of the Blue. Angew. Chem. 2016, 55, 104–121. [Google Scholar] [CrossRef]
  46. Katagiri, A.; Yoshimura, S.; Yoshizawa, S. Formation Constant of the Tetracyanocuprate(II) Ion and the Mechanism of Its Decomposition. Inorg. Chem. 1981, 20, 4143–4147. [Google Scholar] [CrossRef]
  47. Shi, Q.; Ye, J. Deracemization Enabled by Visible-Light Photocatalysis. Angew. Chem. 2020, 59, 4998–5001. [Google Scholar] [CrossRef]
  48. Bonfio, C.; Godino, E.; Corsini, M.; de Biani, F.F.; Guella, G.; Mansy, S.S. Prebiotic iron-sulfur peptide catalysts generate a pH gradient across model membranes of late protocells. Nat. Catal. 2018, 1, 616–623. [Google Scholar] [CrossRef]
  49. Hall, D.O.; Cammak, R.; Rao, K.K.; Cammack, R.; Rao, K.K. Role for Ferredoxins in the Origin of Life and Biological Evolution. Nature 1971, 233, 136–138. [Google Scholar] [CrossRef]
  50. Johnson, A.P.; Cleaves, H.J.; Dworkin, J.P.; Glavin, D.P.; Lazcano, A.; Bada, J.L. The Miller Volcanic Spark Discharge Experiment. Science 2008, 322, 404. [Google Scholar] [CrossRef] [Green Version]
  51. Parker, E.T.; Zhou, M.; Burton, A.S.; Glavin, D.P.; Dworkin, J.P.; Krishnamurthy, R.; Fernández, F.M.; Bada, J.L. A Plausible Simultaneous Synthesis of Amino Acids and Simple Peptides on the Primordial Earth. Angew. Chem. 2014, 53, 8132–8136. [Google Scholar] [CrossRef] [PubMed]
  52. Piotto, S.; Sessa, L.; Piotto, A.; Nardiello, A.; Concilio, S. Plausible Emergence of Autocatalytic Cycles under Prebiotic Conditions. Life 2019, 9, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Urry, D.W.; Goodall, M.C.; Glickson, J.D.; Mayers, D.F. The Gramicidin A Transmembrane Channel: Characteristics of Head-to-Head Dimerized (L,D) Helices. Proc. Natl. Acad. Sci. USA 1971, 68, 1907–1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Walde, P. Surfactant Assemblies and Their Various Possible Roles for the Origin(s) of Life. Orig. Life Evol. Biosph. 2006, 36, 109–150. [Google Scholar] [CrossRef] [PubMed]
  55. Douliez, J.-P.; Gaillard, C. Self-assembly of fatty acids: From foams to protocell vesicles. New J. Chem. 2014, 38, 5142–5148. [Google Scholar] [CrossRef]
  56. Fiore, M.; Strazewski, P. Prebiotic lipidic amphiphiles and condensing agents on the early earth. Life 2016, 6, 17. [Google Scholar] [CrossRef]
  57. Hargreaves, W.R.; Deamer, D.W. Liposomes from Ionic, Single-Chain Amphiphiles. Biochemistry 1978, 17, 3759–3768. [Google Scholar] [CrossRef]
  58. Fayolle, D.; Altamura, E.; D’Onofrio, A.; Madanamothoo, W.; Fenet, B.; Mavelli, F.; Buchet, R.; Stano, P.; Fiore, M.; Strazewski, P. Crude phosphorylation mixtures containing racemic lipid amphiphiles self-assemble to give stable primitive compartments. Sci. Rep. 2017, 7, 18106. [Google Scholar] [CrossRef] [Green Version]
  59. Ourisson, G.; Nakatani, Y. The terpenoid theory of the origin of cellular life: The evolution of terpenoids to cholesterol. Chem. Biol. 1994, 1, 11–23. [Google Scholar] [CrossRef]
  60. Jordan, S.F.; Rammu, H.; Zheludev, I.N.; Hartley, A.M.; Maréchal, A.; Lane, N. Promotion of protocell self-assembly from mixed amphiphiles at the origin of life. Nat. Ecol. Evol. 2019, 3, 1705–1714. [Google Scholar] [CrossRef]
  61. Hargreaves, W.R.; Mulvil, S.J.; Deamer, D.W. Synthesis of phospholipids and membranes in prebiotic conditions. Nature 1977, 266, 78–80. [Google Scholar] [CrossRef] [PubMed]
  62. Epps, D.E.; Sherwood, E.; Eichberg, J.; Oró, J. Cyanamide mediated syntheses under plausible primitive earth conditions. J. Mol. Evol. 1978, 11, 279–292. [Google Scholar] [CrossRef] [PubMed]
  63. Rao, M.; Eichberg, J.; Oró, J. Synthesis of phosphatidylcholine under possible primitive Earth conditions. J. Mol. Evol. 1982, 18, 196–202. [Google Scholar] [CrossRef] [PubMed]
  64. Rao, M.; Eichberg, J.; Oró, J. Synthesis of phosphatidylethanolamine under possible primitive earth conditions. J. Mol. Evol. 1987, 25, 1–6. [Google Scholar] [CrossRef]
  65. Rushdi, A.I.; Simoneit, B.R.T.T. Abiotic condensation synthesis of glyceride lipids and wax esters under simulated hydrothermal conditions. Orig. Life Evol. Biosph. 2006, 36, 93–108. [Google Scholar] [CrossRef]
  66. Maheen, G.; Tian, G.; Wang, Y.; He, C.; Shi, Z.; Yuan, H.; Feng, S. Resolving the enigma of prebiotic C-O-P bond formation: Prebiotic hydrothermal synthesis of important biological phosphate esters. Heteroat. Chem. 2010. [Google Scholar] [CrossRef]
  67. Bonfio, C.; Caumes, C.; Duffy, C.D.; Patel, B.H.; Percivalle, C.; Tsanakopoulou, M.; Sutherland, J.D. Length-Selective Synthesis of Acylglycerol-Phosphates through Energy-Dissipative Cycling. J. Am. Chem. Soc. 2019, 141, 3934–3939. [Google Scholar] [CrossRef] [Green Version]
  68. Pasek, M.A.; Gull, M.; Herschy, B. Phosphorylation on the early earth. Chem. Geol. 2017, 475, 149–170. [Google Scholar] [CrossRef]
  69. Xu, J.; Green, N.J.; Gibard, C.; Krishnamurthy, R.; Sutherland, J.D. Prebiotic phosphorylation of 2-thiouridine provides either nucleotides or DNA building blocks via photoreduction. Nat. Chem. 2019, 11, 457–462. [Google Scholar] [CrossRef]
  70. Pross, A. Toward a general theory of evolution: Extending Darwinian theory to inanimate matter. J. Syst. Chem. 2011, 2, 1. [Google Scholar] [CrossRef] [Green Version]
  71. Toparlak, D.; Karki, M.; Egas Ortuno, V.; Krishnamurthy, R.; Mansy, S.S. Cyclophospholipids Increase Protocellular Stability to Metal Ions. Small 2020, 16, 1903381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Gibard, C.; Bhowmik, S.; Karki, M.; Kim, E.-K.K.; Krishnamurthy, R. Phosphorylation, oligomerization and self-assembly in water under potential prebiotic conditions. Nat. Chem. 2018, 10, 212–217. [Google Scholar] [CrossRef] [PubMed]
  73. Burcar, B.; Pasek, M.; Gull, M.; Cafferty, B.J.; Velasco, F.; Hud, N.V.; Menor-Salván, C. Darwin’s Warm Little Pond: A One-Pot Reaction for Prebiotic Phosphorylation and the Mobilization of Phosphate from Minerals in a Urea-Based Solvent. Angew. Chem. 2016, 55, 13249–13253. [Google Scholar] [CrossRef] [PubMed]
  74. Menor-Salván, C.; Marín-Yaseli, M.R. Prebiotic chemistry in eutectic solutions at the water-ice matrix. Chem. Soc. Rev. 2012, 41, 5404–5415. [Google Scholar] [CrossRef] [PubMed]
  75. Gilbert, W. Origin of life: The RNA world. Nature 1986, 319, 618. [Google Scholar] [CrossRef]
  76. Johnston, W.K. RNA-Catalyzed RNA Polymerization: Accurate and General RNA-Templated Primer Extension. Science 2001, 292, 1319–1325. [Google Scholar] [CrossRef] [Green Version]
  77. Szostak, J.W. Systems chemistry on early Earth. Nature 2009, 459, 171–172. [Google Scholar] [CrossRef]
  78. IUPAC-IUD Commission on Biochemical Nomenclature. The nomenclature of lipids (Recommendations 1976) IUPAC-IUB Commission on Biochemical Nomenclature. Biochem. J. 1978, 171, 21–35. Available online: https://portlandpress.com/biochemj/article-abstract/171/1/21/11772/The-nomenclature-of-lipids-Recommendations-1976?redirectedFrom=fulltext (accessed on 22 July 2020). [CrossRef] [Green Version]
  79. Kennedy, E.P. The Biological Synthesis of Phosholipids. Can. J. Biochem. Physiol. 1956, 34, 334–348. [Google Scholar] [CrossRef]
  80. Weis, R.M.; McConnell, H.M. Two-dimensional chiral crystals of phospholipid. Nature 1984, 310, 47–49. [Google Scholar] [CrossRef]
  81. Lombard, J.; López-García, P.; Moreira, D. The early evolution of lipid membranes and the three domains of life. Nat. Rev. Microbiol. 2012, 10, 507–515. [Google Scholar] [CrossRef] [PubMed]
  82. Yamagishi, A.; Kon, T.; Takahashi, G.; Oshima, T. From the Common Ancestor of All Living Organisms to Protoeukaryotic Cell. In Thermophyles: The Keys to Molecular Evolution and the Origin of Life? Wiegel, J., Adams, M.W.W., Eds.; Rootledge Taylor & Francis: Abingdon, UK, 1998. [Google Scholar]
  83. Weiss, M.C.; Sousa, F.L.; Mrnjavac, N.; Neukirchen, S.; Roettger, M.; Nelson-Sathi, S.; Martin, W.F. The physiology and habitat of the last universal common ancestor. Nat. Microbiol. 2016, 1, 16116. [Google Scholar] [CrossRef] [PubMed]
  84. Glansdorff, N.; Xu, Y.; Labedan, B. The Last Universal Common Ancestor: Emergence, constitution and genetic legacy of an elusive forerunner. Biol. Direct 2008, 3, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Olson, K.R. Hydrogen sulfide, reactive sulfur species and coping with reactive oxygen species. Free Radic. Biol. Med. 2019, 140, 74–83. [Google Scholar] [CrossRef]
  86. Kasting, J.F.; Zahnle, K.J.; Pinto, J.P.; Young, A.T. Sulfur, ultraviolet radiation, and the early evolution of life. Orig. Life Evol. Biosph. 1989, 19, 252–253. [Google Scholar] [CrossRef]
  87. Catling, D.C.; Zahnle, K.J. The Archean atmosphere. Sci. Adv. 2020, 6, eaax1420. [Google Scholar] [CrossRef] [Green Version]
  88. Guy, L.; Ettema, T.J.G. The archaeal ‘TACK’ superphylum and the origin of eukaryotes. Trends Microbiol. 2011, 19, 580–587. [Google Scholar] [CrossRef]
  89. Eme, L.; Doolittle, W.F. Archaea. Curr. Biol. 2015, 25, R851–R855. [Google Scholar] [CrossRef] [Green Version]
  90. Eme, L.; Spang, A.; Lombard, J.; Stairs, C.W.; Ettema, T.J.G. Archaea and the origin of eukaryotes. Nat. Rev. Microbiol. 2017, 15, 711–723. [Google Scholar] [CrossRef]
  91. Akanuma, S.; Yokobori, S.; Yamagishi, A. Comparative Genomics of Thermophilic Bacteria and Archaea. In Thermophilic Microbes in Environmental and Industrial Biotechnology; Springer: Dordrecht, The Netherlands, 2013; pp. 331–349. [Google Scholar]
  92. Williams, T.A.; Foster, P.G.; Cox, C.J.; Embley, T.M. An archaeal origin of eukaryotes supports only two primary domains of life. Nature 2013, 504, 231–236. [Google Scholar] [CrossRef] [Green Version]
  93. Lake, J.A.; Sinsheimer, J.S. The Deep Roots of the Rings of Life. Genome Biol. Evol. 2013, 5, 2440–2448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Daiyasu, H.; Hiroike, T.; Koga, Y.; Toh, H. Analysis of membrane stereochemistry with homology modeling of sn-glycerol-1-phosphate dehydrogenase. Protein Eng. Des. Sel. 2002, 15, 987–995. [Google Scholar] [CrossRef] [PubMed]
  95. Carbone, V.; Schofield, L.R.; Zhang, Y.; Sang, C.; Dey, D.; Hannus, I.M.; Martin, W.F.; Sutherland-Smith, A.J.; Ronimus, R.S. Structure and Evolution of the Archaeal Lipid Synthesis Enzyme sn-Glycerol-1-phosphate Dehydrogenase. J. Biol. Chem. 2015, 290, 21690–21704. [Google Scholar] [CrossRef] [Green Version]
  96. Wächtershäuser, G. From pre-cells to Eukarya—A tale of two lipids. Mol. Microbiol. 2003, 47, 13–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kandler, O. Cell wall biochemistry in Archaea and its phylogenetic implications. J. Biol. Phys. 1995, 20, 165–169. [Google Scholar] [CrossRef]
  98. Woese, C.R.; Kandler, O.; Wheelis, M.L. Towards a natural system of organisms: Proposal for the domains Archaea, Bacteria, and Eucarya. Proc. Natl. Acad. Sci. USA 1990, 87, 4576–4579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Woese, C.R.; Fox, G.E. Phylogenetic structure of the prokaryotic domain: The primary kingdoms. Proc. Natl. Acad. Sci. USA 1977, 74, 5088–5090. [Google Scholar] [CrossRef] [Green Version]
  100. Saw, J.H.; Spang, A.; Zaremba-Niedzwiedzka, K.; Juzokaite, L.; Dodsworth, J.A.; Murugapiran, S.K.; Colman, D.R.; Takacs-Vesbach, C.; Hedlund, B.P.; Guy, L.; et al. Exploring microbial dark matter to resolve the deep archaeal ancestry of eukaryotes. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140328. [Google Scholar] [CrossRef] [Green Version]
  101. Spang, A.; Saw, J.H.; Jørgensen, S.L.; Zaremba-Niedzwiedzka, K.; Martijn, J.; Lind, A.E.; van Eijk, R.; Schleper, C.; Guy, L.; Ettema, T.J.G. Complex archaea that bridge the gap between prokaryotes and eukaryotes. Nature 2015, 521, 173–179. [Google Scholar] [CrossRef] [Green Version]
  102. Daubin, V.; Szöllősi, G.J. Horizontal Gene Transfer and the History of Life. Cold Spring Harb. Perspect. Biol. 2016, 8, a018036. [Google Scholar] [CrossRef] [Green Version]
  103. Carter, C.W. What RNA world? Why a peptide/rna partnership merits renewed experimental attention. Life 2015, 5, 294–320. [Google Scholar] [CrossRef] [PubMed]
  104. Carter, C.W. An Alternative to the RNA World. Nat. Hist. 2016, 125, 28–33. [Google Scholar] [PubMed]
  105. Carter, C.W.; Wills, P.R. Interdependence, Reflexivity, Fidelity, Impedance Matching, and the Evolution of Genetic Coding. Mol. Biol. Evol. 2018, 35, 269–286. [Google Scholar] [CrossRef] [PubMed]
  106. Escudero, C.; El-Hachemi, Z.; Díez-Pérez, I.; Crusats, J.; Ribó, J.M. Formation of an epitaxial monolayer on graphite upon short-time surface contact with highly diluted aqueous solutions of 1-monostearoylglycerol. Thin Solid Films 2007, 515, 5391–5394. [Google Scholar] [CrossRef]
  107. Fiore, M.; Maniti, O.; Girard-Egrot, A.; Monnard, P.-A.A.; Strazewski, P. Glass Microsphere-Supported Giant Vesicles for the Observation of Self-Reproduction of Lipid Boundaries. Angew. Chem. 2018, 57, 282–286. [Google Scholar] [CrossRef] [PubMed]
  108. Lopez, A.; Fayolle, D.; Fiore, M.; Strazewski, P. Chemical Analysis of Lipid Boundaries after Consecutive Growth of Supported Giant Vesicles. iScience 2020. accepted. [Google Scholar]
  109. Hanczyc, M.M.; Szostak, J.W. Replicating vesicles as models of primitive cell growth and division. Curr. Opin. Chem. Biol. 2004, 8, 660–664. [Google Scholar] [CrossRef] [PubMed]
  110. Terasawa, H.; Nishimura, K.; Suzuki, H.; Matsuura, T.; Yomo, T. Coupling of the fusion and budding of giant phospholipid vesicles containing macromolecules. Proc. Natl. Acad. Sci. USA 2012, 109, 5942–5947. [Google Scholar] [CrossRef] [Green Version]
  111. Miele, Y.; Medveczky, Z.; Holló, G.; Tegze, B.; Derényi, I.; Hórvölgyi, Z.; Altamura, E.; Lagzi, I.; Rossi, F. Self-division of giant vesicles driven by an internal enzymatic reaction. Chem. Sci. 2020, 11, 3228–3235. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Plausibly prebiotic pathways for glycerophosphates (5a5c), precursor of phospholipids, from glyceraldehyde, dihydroxyacetone or glycerol (13). Plausibly prebiotic pathway allowing DHAP (4) from glyceraldehyde (1) remains unexplored.
Scheme 1. Plausibly prebiotic pathways for glycerophosphates (5a5c), precursor of phospholipids, from glyceraldehyde, dihydroxyacetone or glycerol (13). Plausibly prebiotic pathway allowing DHAP (4) from glyceraldehyde (1) remains unexplored.
Symmetry 12 01488 sch001
Figure 1. Plausibly prebiotic lipid derivatives. Red colour is used to indicate the polar head group.
Figure 1. Plausibly prebiotic lipid derivatives. Red colour is used to indicate the polar head group.
Symmetry 12 01488 g001
Scheme 2. A few relevant prebiotic pathways that allow the formation of phospholipid esters and glycerol phosphates. (A) Summary of the prebiotic pathways explored during pioneering research (1977–1982); (B) phosphorylation of glycerol; (C) recent results obtained in phosphorylation of diacylglycerols and (D) concomitant acylation of glycerol in the presence of fatty acids and diamidophospahte. The asterisk (*) indicates the stereogenic carbon C2 of any phospholipid and phospholipid precursors; Pi stands for any phosphorous salt or plausibly phosphate-containing mineral able to promote the phosphorylation of primary or secondary alcohols [42]; ca, stands for any condensing agents [56]; Δ, stands for temperatures between 65 and 130 °C; DAP stands for diamidophosphate [72]; for eutectic conditions see the recent works of Menor–Salvan and Pasek [73,74].
Scheme 2. A few relevant prebiotic pathways that allow the formation of phospholipid esters and glycerol phosphates. (A) Summary of the prebiotic pathways explored during pioneering research (1977–1982); (B) phosphorylation of glycerol; (C) recent results obtained in phosphorylation of diacylglycerols and (D) concomitant acylation of glycerol in the presence of fatty acids and diamidophospahte. The asterisk (*) indicates the stereogenic carbon C2 of any phospholipid and phospholipid precursors; Pi stands for any phosphorous salt or plausibly phosphate-containing mineral able to promote the phosphorylation of primary or secondary alcohols [42]; ca, stands for any condensing agents [56]; Δ, stands for temperatures between 65 and 130 °C; DAP stands for diamidophosphate [72]; for eutectic conditions see the recent works of Menor–Salvan and Pasek [73,74].
Symmetry 12 01488 sch002
Figure 2. Phosphaditate enantiomers and their sn-glycerol and Cahn–Ingold–Prelog nomenclatures. 1,2-diacyl-sn-glycero-3-phosphate is the enantiomer of 2,3-diacyl-sn-1-glycerophosphate: the stereo numbering (sn-glycerol) is based on the position of the second oxygen of the glycerol moiety to the left side in the Fisher representation, with the top carbon numbered as one, second as two and the bottom carbon numbered as three, the enantiomer changes the order of numbers of glycerol moiety due to the opposite position of the second oxygen.
Figure 2. Phosphaditate enantiomers and their sn-glycerol and Cahn–Ingold–Prelog nomenclatures. 1,2-diacyl-sn-glycero-3-phosphate is the enantiomer of 2,3-diacyl-sn-1-glycerophosphate: the stereo numbering (sn-glycerol) is based on the position of the second oxygen of the glycerol moiety to the left side in the Fisher representation, with the top carbon numbered as one, second as two and the bottom carbon numbered as three, the enantiomer changes the order of numbers of glycerol moiety due to the opposite position of the second oxygen.
Symmetry 12 01488 g002
Scheme 3. Biosynthetic pathways leading to G3P (A) and G1P (B) from prochiral glycerol (3) or DHAP (4).
Scheme 3. Biosynthetic pathways leading to G3P (A) and G1P (B) from prochiral glycerol (3) or DHAP (4).
Symmetry 12 01488 sch003
Figure 3. Hypothetic phylogeny of the last common ancestor (LCA) and Commonnote Commonote and their evolution into Archaea, Bacteria and Eukarya or Euryarchaeotae from prebiotic pathways, adapted from [26]. The early stage of the archaeal lineage had G3PDH so that the ancestor C. commonote had a G3P polar lipid membrane rather than G1P lipid membranes, giving rise to C. archaea. Then, C archaea, had G1P lipids, probably mixed with G3P lipids [26]. C. bacteria appeared later than C archaea [26]. Hypothetical horizontal gene transfer (indicated by dashed grey arrows) may have occurred [102]. Eukarya was significantly distinct from bacteria and may have originated from Archaea [88,90,100].
Figure 3. Hypothetic phylogeny of the last common ancestor (LCA) and Commonnote Commonote and their evolution into Archaea, Bacteria and Eukarya or Euryarchaeotae from prebiotic pathways, adapted from [26]. The early stage of the archaeal lineage had G3PDH so that the ancestor C. commonote had a G3P polar lipid membrane rather than G1P lipid membranes, giving rise to C. archaea. Then, C archaea, had G1P lipids, probably mixed with G3P lipids [26]. C. bacteria appeared later than C archaea [26]. Hypothetical horizontal gene transfer (indicated by dashed grey arrows) may have occurred [102]. Eukarya was significantly distinct from bacteria and may have originated from Archaea [88,90,100].
Symmetry 12 01488 g003
Figure 4. A hypothetical pathway allowing the selection trough the formation of enantiopure phospholipids and deracemization of mixed protocell membranes upon the encapsulation of enantiopure biopolymers (geometrical forms) followed by the growth and division of membrane bilayers. Colour code is used to better highlight the symmetry imbalance from racemic (green) to enantiopure (violet).
Figure 4. A hypothetical pathway allowing the selection trough the formation of enantiopure phospholipids and deracemization of mixed protocell membranes upon the encapsulation of enantiopure biopolymers (geometrical forms) followed by the growth and division of membrane bilayers. Colour code is used to better highlight the symmetry imbalance from racemic (green) to enantiopure (violet).
Symmetry 12 01488 g004

Share and Cite

MDPI and ACS Style

Fiore, M.; Buchet, R. Symmetry Breaking of Phospholipids. Symmetry 2020, 12, 1488. https://doi.org/10.3390/sym12091488

AMA Style

Fiore M, Buchet R. Symmetry Breaking of Phospholipids. Symmetry. 2020; 12(9):1488. https://doi.org/10.3390/sym12091488

Chicago/Turabian Style

Fiore, Michele, and René Buchet. 2020. "Symmetry Breaking of Phospholipids" Symmetry 12, no. 9: 1488. https://doi.org/10.3390/sym12091488

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