Elsevier

Gene

Volume 479, Issues 1–2, 15 June 2011, Pages 29-36
Gene

Methods paper
Evolutionary analysis of glycosyl hydrolase family 28 (GH28) suggests lineage-specific expansions in necrotrophic fungal pathogens

https://doi.org/10.1016/j.gene.2011.02.009Get rights and content

Abstract

Glycosyl hydrolase family 28 (GH28) is a set of structurally related enzymes that hydrolyze glycosidic bonds in pectin, and are important extracellular enzymes for both pathogenic and saprotrophic fungi. Yet, very little is understood about the evolutionary forces driving the diversification of GH28s in fungal genomes. We reconstructed the evolutionary history of family GH28 in fungi by examining the distribution of GH28 copy number across the phylogeny of fungi, and by reconstructing the phylogeny of GH28 genes. We also examined the relationship between lineage-specific GH28 expansions and fungal ecological strategy, testing the hypothesis that GH28 evolution in fungi is driven by ecological strategy (pathogenic vs. non-pathogenic) and pathogenic niche (necrotrophic vs. biotrophic). Our results showed that GH28 phylogeny of Ascomycota and Basidiomycota sequences was structured by specific biochemical function, with endo-polygalacturonases and endo-rhamnogalacturonases forming distinct, apparently ancient clades, while exo-polygalacturonases are more widely distributed. In contrast, Mucoromycotina and Stramenopile sequences formed taxonomically-distinct clades. Large, lineage-specific variation in GH28 copy number indicates that the evolution of this gene family is consistent with the birth-and-death model of gene family evolution, where diversity of GH28 loci within genomes was generated through multiple rounds of gene duplication followed by functional diversification and loss of some gene family members. Although GH28 copy number was correlated with genome size, our findings suggest that ecological strategy also plays an important role in determining the GH28 repertoire of fungi. Both necrotrophic and biotrophic fungi have larger genomes than non-pathogens, yet only necrotrophs possess more GH28 enzymes than non-pathogens. Hence, lineage-specific GH28 expansion is the result of both variation in genome size across fungal species and diversifying selection within the necrotrophic plant pathogen ecological niche. GH28 evolution among necrotrophs has likely been driven by a co-evolutionary arms race with plants, whereas the need to avoid plant immune responses has resulted in purifying selection within biotrophic fungi.

Introduction

Gene families are sets of genes that have arisen through the accumulation of duplicate copies of a single ancestral gene. Following duplication, the paralogous genes represent a genetic redundancy that may result in relaxation of selection pressure on one or both of the gene copies, allowing mutations to accumulate in one or both copies (Ohno, 1970, Prince and Pickett, 2002). Gene copies that incorporate deleterious mutations are often purged from the genome via purifying selection, while copies that acquire novel or enhanced functionality can become fixed in the population through positive or diversifying selection (Hughes, 2002, Lynch and Conery, 2003). This pattern of duplication and expansion of advantageous gene copies paired with the expulsion of dysfunctional gene copies has been described as evolution via the birth-and-death process (Nei et al., 1997, Nei and Rooney, 2005).

Glycosyl hydrolase family 28 (GH28) has interesting functional diversity and is variable in copy number among related organisms, making this gene family a likely candidate for birth-and-death evolution. GH28 enzymes are involved in degradation of pectin, a major structural constituent of the plant cell wall. Pectin is a long polysaccharide chain that is composed of α-linked galacturonic acid (GalA) monomers, with some regions of GalA alternating with rhamnose or branched xylose side-chains (Willats et al., 2006). Various enzymes in family GH28 degrade these bonds by catalyzing a series of functionally distinct reactions. Endo-glycosidases catalyze hydrolysis of internal glycosidic bonds at random locations within the polysaccharide, while exo-glycosidases catalyze hydrolysis of terminal bonds that attach individual sugars to the ends of the polysaccharides. GH28 enzymes are also categorized into polygalacturonases (PG), which hydrolyze GalA-GalA linkages (E.C.'s 3.2.1.15 [endo-PG] and 3.2.1.67 [exo-PG]), rhamnogalacturonases (RG), which hydrolyze GalA-rhamnose bonds (E.C. 3.2.1.-), and xylogalacturonases (XG), which hydrolyze GalA-xylose bonds (E.C. 3.2.1.-) (Markovič and Janeček, 2001).

Genes encoding GH28 enzymes have been identified in genomes of a wide range of plants, bacteria, and fungi. Fungal GH28 pectinases are degradative enzymes used by both saprotrophs growing on senesced leaf tissue (Kjøller and Struwe, 2002) and plant pathogens seeking to gain access to plant intracellular nutrients (Herron et al., 2000, Reignault et al., 2008). The contribution of pectinases to pathogenic infectivity remains unclear. Targeted gene disruption studies have demonstrated that various GH28 functional types do play a role in virulence (Garcia-Maceira et al., 2001, Kars et al., 2005, Shieh et al., 1997, ten Have et al., 1998), while several other studies have not found GH28 expression to be critically important during cell wall invasion (Scott-Craig et al., 1990, Gao et al., 1996). This disagreement may be due to lack of consideration of the differing strategies adopted by phytopathogenic fungi (Glazebrook, 2005). Necrotrophic pathogens acquire carbon and energy by extensively degrading plant tissue, often resulting in the host's death. In contrast, biotrophic fungi obtain nutrients from living plant tissues and depend on continued survival of their host (Oliver and Ipcho, 2004). Therefore, while necrotrophs indiscriminately destroy cell wall constituents, biotrophs colonize intact plant cells while avoiding immune responses. Extracellular enzymes like GH28s can elicit a wide array of plant immunological responses, such as the production of extracellular plant proteins that inhibit fungal PGs, known as polygalacturonase-inhibiting proteins or PGIPs (De Lorenzo et al., 2001, Federici et al., 2001). GH28 gene products also stimulate the PGIP-mediated hypersensitive immune response, or localized cell death, in host plants. Both PGs and PGIPs are highly polymorphic, resulting in a high level of PGIP specificity for particular PG isozymes (Casasoli et al., 2009, Cook et al., 1999, De Lorenzo et al., 2001, Di Matteo et al., 2006, Raiola et al., 2008). This molecular co-evolutionary arms race could be a source of strong diversifying selection on the GH28 repertoires of fungal necrotrophs. Conversely, plant immune system activity likely acts to reduce GH28 diversity in biotrophic fungi via purifying selection (Oliver and Ipcho, 2004). Furthermore, saprotrophic fungi that degrade dead plant material would not experience the strong diversifying selection associated with a co-evolutionary arms race, and should therefore also possess a low level of GH28 diversity.

Because of the importance of pectinase in liberating carbon and energy and the wide distribution of PGIPs, we hypothesize that the distribution of GH28 genes in fungi may be closely linked to the evolution of ecological strategy and pathogenic niche. In this study, we examine the pattern of GH28 gene family evolution by investigating its occurrence and distribution in fungal genomes and by comprehensively reconstructing the long-term evolutionary history of the GH28 family in fungi. Using GH28 genes from completely sequenced fungal genomes, we infer the ancestral gene copy number in most recent common ancestors (MRCA) of major taxonomic groups. Additionally, using a reconstructed phylogeny of fungal GH28 sequences, we infer the ancestral enzymatic mode(s) of action and subsequent functional diversification within this gene family. We then test the hypothesis that the occurrence of GH28s, or particular functional categories of GH28, is correlated with ecological strategy (pathogenic vs. non-pathogenic) and pathogenic niche (necrotrophic vs. biotrophic) against the null hypothesis that GH28 copy number is simply the result of variation in overall genome size.

Section snippets

Fungal SSU rRNA-based species tree and ecological characteristics

We first reconstructed a species tree using small subunit ribosomal RNA (SSU rRNA) sequences from 69 fully sequenced fungal genomes (Supplementary File 1), including 57 Ascomycetes, 8 Basidiomycetes, 3 Mucoromycotinas, and 1 Chytridiomycota, as well as an additional 2 Oomycetes (Stramenopiles) and 1 plant. Relevant ribosomal sequences were extracted from the SILVA database (Pruesse et al., 2007) and GenBank (Benson et al., 2005) and aligned using the MUSCLE sequence alignment tool using default

GH28 distribution and ancestral copy number

The survey of GH28 members revealed the presence of at least one GH28 homolog in 40 of the 69 completed fungal genome sequences examined, with GH28 copy number per genome ranging from 0 to 20 (Supplementary File 1). Such large variation is suggestive of a dynamic evolutionary history, and is consistent with evolution via the birth-and-death process, resulting in lineage-specific gene family expansions and contractions (Nei et al., 1997, Nei and Rooney, 2005).

We used a maximum parsimony approach

Conclusions

In this study, we comprehensively surveyed GH28 gene family distribution and diversity within the fungal kingdom, and characterized its evolution as being consistent with the birth-and-death model. The initial appearance of GH28s predates the evolution of fungi, which places the origin of this gene family at more than 1.5 billion years ago. Ancient gene families such as GH28 continue to radiate and expand into new molecular niches. Our analysis indicates that the earliest fungi already

Acknowledgments

The authors thank Bess Heidenreich for assistance in compilation of data. Financial support for the work was provided by Kent State University.

References (80)

  • A. Raiola et al.

    A single amino acid substitution in highly similar endo-PGs from Fusarium verticillioides and related Fusarium species affects PGIP inhibition

    Fung. Genet. Biol.

    (2008)
  • A.E. Todd et al.

    Plasticity of enzyme active sites

    Trends Biochem. Sci.

    (2002)
  • G. van Pouderoyen et al.

    Structural insights into the processivity of endopolygalacturonase I from Aspergillus niger

    FEBS Lett.

    (2003)
  • J.D. Watson et al.

    Predicting protein function from sequence and structural data

    Curr. Opin. Struct. Biol.

    (2005)
  • W.G.T. Willats et al.

    Pectin: new insights into an old polymer are starting to gel

    Trends Food Sci. Technol.

    (2006)
  • A.V. Wymelenberg et al.

    The Phanerochaete chrysosporium secretome: database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium

    J. Biotechnol.

    (2005)
  • J.O. Andersson

    Convergent evolution: gene sharing by eukaryotic plant pathogens

    Curr. Biol.

    (2006)
  • G.W. Beakes et al.

    The evolutionary phylogeny of oomycetes—insights gained from studies of holocarpic parasites of algae and invertegrates

  • L. Belbahri et al.

    Evolution of the cutinase gene family: evidence for lateral gene transfer of a candidate Phytophthora virulence factor

    Gene

    (2007)
  • D.A. Benson et al.

    GenBank

    Nucleic Acids Res.

    (2005)
  • J.E. Blair

    Fungi

  • H. Bussink et al.

    The polygalacturonases of Aspergillus niger are encoded by a family of diverged genes

    Eur. J. Biochem.

    (1992)
  • M. Casasoli et al.

    Integration of evolutionary and desolvation energy analysis identifies functional sites in a plant immunity protein

    Proc. Nat. Acad. Sci.

    (2009)
  • P.F. Cliften et al.

    After the duplication: gene loss and adaptation in Saccharomyces genomes

    Genetics

    (2006)
  • B.J. Cook et al.

    Fungal polygalacturonases exhibit different substrate degradation patterns and differ in their susceptibilities to polygalacturonase-inhibiting proteins

    Mol. Plant Microbe Interact.

    (1999)
  • G. De Lorenzo et al.

    The role of polygalacturonase-inhibiting proteins (PGIPs) in defense against pathogenic fungi

    Annu. Rev. Phytopathol.

    (2001)
  • R.P. de Vries et al.

    Aspergillus enzymes involved in degradation of plant cell wall polysaccharides

    Microbiol. Mol. Biol. Rev.

    (2001)
  • R.P. de Vries

    Regulation of Aspergillus genes encoding plant cell wall polysaccharide-degrading enzymes; relevance for industrial production

    Appl. Microbiol. Biotechnol.

    (2003)
  • Drummond, A.J., Ashton, B., Cheung, M., Heled, J., Kearse, M., Moir, R., Stones-Havas, S., Thierer, T., Wilson, A.,...
  • R.C. Edgar

    MUSCLE: multiple sequence alignment with high accuracy and high throughput

    Nucleic Acids Res.

    (2004)
  • L. Federici et al.

    Structural requirements of endopolygalacturonase for the interaction with PGIP (polygalacturonase-inhibiting protein)

    Proc. Nat. Acad. Sci.

    (2001)
  • J. Felsenstein

    Confidence limits on phylogenies: an approach using the bootstrap

    Evolution

    (1985)
  • J. Felsenstein

    Phylogenies and the comparative method

    Am. Nat.

    (1985)
  • S. Gao et al.

    Cloning and targeted disruption of enpg-1, encoding the major in vitro extracellular endopolygalacturonase of the chestnut blight fungus, Cryphonectria parasitica

    Appl. Environ. Microbiol.

    (1996)
  • F. Garcia-Maceira et al.

    Molecular characterization of an endopolygalacturonase from Fusarium oxysporum expressed during early stages of infection

    Appl. Environ. Microbiol.

    (2001)
  • L.Y. Geer et al.

    CDART: protein homology by domain architecture

    Genome Res.

    (2002)
  • D.M. Geiser et al.

    Eurotiomycetes: Eurotiomycetidae and Chaetothyriomycetidae

    Mycologia

    (2006)
  • J. Glazebrook

    Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens

    Annu. Rev. Phytopathol.

    (2005)
  • A. Gotesson et al.

    Characterization and evolutionary analysis of a large polygalacturonase gene family in the oomycete plant pathogen Phytophthora cinnamomi

    Mol. Plant Microbe Interact.

    (2002)
  • S. Guindon et al.

    A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood

    Sys. Biol.

    (2003)
  • Cited by (42)

    • Expression analysis of the NEP-1 and cell-wall degrading genes of Gilbertella persicaria during pathogenesis in papaya (Carica papaya L.) fruits

      2021, Physiological and Molecular Plant Pathology
      Citation Excerpt :

      On the other hand, the gene encoding a xyloglucanase (Xln) enzyme of the GH16 family included in the genome of the CBS 190.32 strain was selected for this study and although the xiloglucanase activity was observed in the in vitro assay; , no expression in vivo of this gene was observed throughout the infection process; for this reason, the activity of other enzymes can be deduced to be involved in the degradation of xyloglucan exists and should be considered in subsequent studies. The PG enzymes (family GH28) are important in the degradation of the pectin fraction and are widely distributed in the genome of diverse phytopathogenic fungi with necrotrophic lifestyles, such as Sclerotinia sclerotiorum, Phytophthora sojae, Botrytis cinerea, Mucor circinelloides and Rhizopus oryzae [9,36,37]. Furthermore, excessive softening of the pulp of papaya fruit correlates with solubilization, by the action of endopolygalacturonase enzymes, of high molecular weight galacturonans that make up the pectin fraction [38].

    • Comparison of pectin-degrading fungal communities in temperate forests using glycosyl hydrolase family 28 pectinase primers targeting Ascomycete fungi

      2016, Journal of Microbiological Methods
      Citation Excerpt :

      Primer design was conducted using a database of 293 fungal GH28 DNA sequences from 40 genome sequences (Sprockett et al., 2011). We focused primer design on GH28 phylogenetic clade F because of a high level of sequence conservation and good representation of major fungal lineages (Sprockett et al., 2011). Primers matching the maximum possible number of sequences were designed to have as few degeneracies as possible, avoid self-priming and primer–dimers, and other appropriate physiochemical properties (e.g., annealing temperatures) (Table 1).

    View all citing articles on Scopus
    View full text