Ascorbate Biosynthesis: A cross-kingdom history

  1. Alisdair R Fernie  Is a corresponding author
  2. Takayuki Tohge  Is a corresponding author
  1. Max Planck Institute of Molecular Plant Physiology, Germany

Between two and three billion years ago, living things evolved the ability to use oxygen to increase the amount of energy that could be released from carbon-rich molecules. However, these reactions produce harmful molecules called ‘reactive oxygen species’ as by-products, so antioxidant systems had to be developed in parallel to detoxify them (Noctor and Foyer, 1998; Gest et al., 2013).

A compound called ascorbate is an excellent antioxidant that is made by almost all living organisms. However, certain animals—including guinea pigs (Nishikimi et al., 1992), most bats, some fish and bird species, as well as humans and many other primates—have all have lost the ability to synthesise this compound for themselves (Mandl et al., 2009). Instead, these animals must obtain ascorbate (which is also known as ‘vitamin C’) from their diets. Moreover, all of these animals are unable to make ascorbate for the same reason: they have all lost a gene that encodes an enzyme called gulonolactone oxidase (or GULO). This enzyme catalyses the final step in the metabolic pathway that produces ascorbate in other animals.

Three major routes of ascorbate biosynthesis have been described across the kingdoms of life. The animal kingdom uses one of these pathways, and photosynthetic eukaryotes—such as plants, algae and some single-celled organisms—employ the other two. All three pathways proceed via different intermediate molecules, but the two from photosynthetic eukaryotes both use an enzyme called galactolactone dehydrogenase (or GLDH), instead of GULO, to catalyse the final step (Wheeler et al., 1998).

Now, in eLife, Glen Wheeler (Marine Biological Association), Takahiro Ishikawa (Shimane University), and Varissa Pornsaksit and Nicholas Smirnoff (both at the University of Exeter) have analysed the distribution of these two enzymes amongst the genomes of many different eukaryotes to determine the origins of ascorbate biosynthesis (Wheeler et al., 2015). They found that each genome either contained a gene for GULO or a gene for GLDH, but not both (Figure 1A). They also revealed that the ancestor of all animals and fungi (and their close relatives) likely used GULO, and that this ancestral trait has been lost in many later groups, including all insects.

Taxonomic distribution and evolutionary history of ascorbate biosynthesis.

(A) The phylogenetic tree depicts the currently accepted evolutionary relationships between the major lineages of eukaryotes. Black vertical lines indicate non-photosynthetic groups, whilst green lines indicate organisms that became photosynthetic after acquiring an ancestral cyanobacterium (via a so-called ‘primary endosymbiosis event’). Red lines indicate organisms that became photosynthetic after acquiring a red alga or green alga (via so-called ‘secondary endosymbiosis events’). Groups with a copy of the gene for an enzyme called GULO in their genome are marked with a closed blue circle, whilst those with a copy of the gene for an enzyme called GLDH are marked with a closed green circle. Empty circles indicate that the gene is absent from the genome. (B) The schematic model suggests how the primary routes of ascorbate biosynthesis arose in the various eukaryotic lineages and offers potential explanations based on Wheeler et al.'s findings. The presence and absence of GULO and GLDH are indicated as above, and chloroplasts (which are originally derived from the cyanobacterial ancestor) are indicated via green ovals. The wide arrows indicate the evolutionary relationships between groups. The thin arrows indicate the acquisition of a red or green alga by a non-photosynthetic eukaryote and dashed arrows represent events that may explain the gain or loss of genes involved in ascorbate biosynthesis.

Ascorbate biosynthesis via GULO also occurs in several other groups of eukaryotes, including a primitive, freshwater alga called Cyanophora paradoxa and two species in a single genus of red algae (called Galdieria). However, the gene for GULO is absent in the genomes of the vast majority of other algae and the land plants. Instead, Wheeler et al. revealed that almost all photosynthetic eukaryotes use the GLDH enzyme. Gene expression datasets confirmed that this enzyme is also found in non-photosynthetic organisms that are descended from photosynthetic ancestors—namely the water moulds (or oomycetes). Moreover, while several non-photosynthetic organisms (including the ciliates and an amboebozoan) lacked both genes, this was not the case for any photosynthetic eukaryote.

Wheeler et al. also provide strong evidence that the genes for GULO and GLDH have each only arisen once—and other evidence suggests that it is possible that the two enzymes are originally descended from a single gene. This implies two possible scenarios. Firstly, GULO and GLDH could result from a single gene that was duplicated in the last common ancestor of all eukaryotes, and different lineages have subsequently lost either one gene or the other. Secondly, the two genes might have evolved in different lineages and passed between species via ‘horizontal gene transfer’. However, whilst compelling arguments are put forward for both cases, it is not currently possible to exclude either of these explanations or, indeed, a combination of both in different lineages.

Nevertheless, their comprehensive analyses allowed Wheeler et al. to speculate on the selective pressures that might have driven the evolutionary path of ascorbate biosynthesis in eukaryotes (Figure 1B). Reactive oxygen species are produced when energy from sunlight is harvested during photosynthesis. This means that photosynthetic organisms experience ‘photo-oxidative stress’ and need more efficient antioxidant systems. As such, Wheeler et al. put forward an attractive hypothesis that GULO was lost in eukaryotes that became photosynthetic because this enzyme also produces some reactive oxygen species (namely hydrogen peroxide). Photosynthetic eukaryotes switched to use GLDH instead because this enzyme can synthesise ascorbate without producing reactive oxygen species as by-products. The fact that the Galdieria red algae have retained GULO can also be explained since it grows on or inside porous rocks in hot sulphur springs where photo-oxidative stress is believed to be minimal.

Studying other enzymes involved in the ascorbate antioxidant system revealed that C. paradoxa is the only photosynthetic organism to lack these enzymes. However, photosynthetic bacteria (i.e., cyanobacteria) don't appear to use ascorbate to protect against photo-oxidative stress. This suggests that the ‘photoprotective’ role of ascorbate may have emerged in photosynthetic eukaryotes after C. paradoxa (and other glaucophytes) diverged from the lineages that went on to be come the red algae, the green algae and the land plants.

Finally, the evolution of a photoprotective role of ascorbate would have increased the amount of this compound found in the leaves of land plants. As such the loss of GULO in several animal lineages may well reflect the increased availability of ascorbate in their largely plant-based diets.

Wheeler et al. thus provide major insight into the evolution of ascorbate metabolism in eukaryotes. In addition, when taken together with recent studies concerning the evolution of other antioxidants (Tohge et al., 2013) and plant hormones (Ju et al., 2015), the enormous potential resulting from the torrent of information provided by modern genome sequencing technologies becomes apparent (Tautz et al., 2010). In the coming years, one can anticipate that the widespread adoption of similar approaches will enable a far greater understanding of the evolution of both fundamental and specialised metabolic pathways in all kingdoms of life.

References

    1. Nishikimi M
    2. Kawai T
    3. Yagi K
    (1992)
    Guinea pigs possess a highly mutated gene for L-gulono-gamma-lactone oxidase, the key enzyme for L-ascorbic acid biosynthesis missing in this species
    Journal of Biological Chemistry 267:21967–21972.

Article and author information

Author details

  1. Alisdair R Fernie

    Max Planck Institute of Molecular Plant Physiology, Potsdam-Golm, Germany
    For correspondence
    fernie@mpimp-golm.mpg.de
    Competing interests
    The authors declare that no competing interests exist.
  2. Takayuki Tohge

    Max Planck Institute of Molecular Plant Physiology, Potsdam-Golm, Germany
    For correspondence
    tohge@mpimp-golm.mpg.de
    Competing interests
    The authors declare that no competing interests exist.

Publication history

  1. Version of Record published: April 15, 2015 (version 1)

Copyright

© 2015, Fernie and Tohge

This article is distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use and redistribution provided that the original author and source are credited.

Metrics

  • 1,446
    Page views
  • 232
    Downloads
  • 3
    Citations

Article citation count generated by polling the highest count across the following sources: Crossref, PubMed Central, Scopus.

Download links

A two-part list of links to download the article, or parts of the article, in various formats.

Downloads (link to download the article as PDF)

Open citations (links to open the citations from this article in various online reference manager services)

Cite this article (links to download the citations from this article in formats compatible with various reference manager tools)

  1. Alisdair R Fernie
  2. Takayuki Tohge
(2015)
Ascorbate Biosynthesis: A cross-kingdom history
eLife 4:e07527.
https://doi.org/10.7554/eLife.07527

Further reading

    1. Evolutionary Biology
    Alexei V Tkachenko, Sergei Maslov
    Research Article

    Life as we know it relies on the interplay between catalytic activity and information processing carried out by biological polymers. Here we present a plausible pathway by which a pool of prebiotic information-coding oligomers could acquire an early catalytic function, namely sequence-specific cleavage activity. Starting with a system capable of non-enzymatic templated replication, we demonstrate that even non-catalyzed spontaneous cleavage would promote proliferation by generating short fragments that act as primers. Furthermore, we show that catalytic cleavage function can naturally emerge and proliferate in this system. Specifically, a cooperative catalytic network with four subpopulations of oligomers is selected by the evolution in competition with chains lacking catalytic activity. The cooperative system emerges through the functional differentiation of oligomers into catalysts and their substrates. The model is inspired by the structure of the hammerhead RNA enzyme as well as other DNA- and RNA-based enzymes with cleavage activity that readily emerge through natural or artificial selection. We identify the conditions necessary for the emergence of the cooperative catalytic network. In particular, we show that it requires the catalytic rate enhancement over the spontaneous cleavage rate to be at least 102–103, a factor consistent with the existing experiments. The evolutionary pressure leads to a further increase in catalytic efficiency. The presented mechanism provides an escape route from a relatively simple pairwise replication of oligomers toward a more complex behavior involving catalytic function. This provides a bridge between the information-first origin of life scenarios and the paradigm of autocatalytic sets and hypercycles, albeit based on cleavage rather than synthesis of reactants.

    1. Cell Biology
    2. Evolutionary Biology
    Jonathan E Phillips, Duojia Pan
    Research Advance

    The genomes of close unicellular relatives of animals encode orthologs of many genes that regulate animal development. However, little is known about the function of such genes in unicellular organisms or the evolutionary process by which these genes came to function in multicellular development. The Hippo pathway, which regulates cell proliferation and tissue size in animals, is present in some of the closest unicellular relatives of animals, including the amoeboid organism Capsaspora owczarzaki. We previously showed that the Capsaspora ortholog of the Hippo pathway nuclear effector Yorkie/YAP/TAZ (coYki) regulates actin dynamics and the three-dimensional morphology of Capsaspora cell aggregates, but is dispensable for cell proliferation control (Phillips et al., 2022). However, the function of upstream Hippo pathway components, and whether and how they regulate coYki in Capsaspora, remained unknown. Here, we analyze the function of the upstream Hippo pathway kinases coHpo and coWts in Capsaspora by generating mutant lines for each gene. Loss of either kinase results in increased nuclear localization of coYki, indicating an ancient, premetazoan origin of this Hippo pathway regulatory mechanism. Strikingly, we find that loss of either kinase causes a contractile cell behavior and increased density of cell packing within Capsaspora aggregates. We further show that this increased cell density is not due to differences in proliferation, but rather actomyosin-dependent changes in the multicellular architecture of aggregates. Given its well-established role in cell density-regulated proliferation in animals, the increased density of cell packing in coHpo and coWts mutants suggests a shared and possibly ancient and conserved function of the Hippo pathway in cell density control. Together, these results implicate cytoskeletal regulation but not proliferation as an ancestral function of the Hippo pathway kinase cascade and uncover a novel role for Hippo signaling in regulating cell density in a proliferation-independent manner.