quitinase

14
Evolutionary Bioinformatics 2008:4 47–60 47 ORGINAL RESEARCH Correspondence: Magnus Karlsson, Department of Forest Mycology and Pathology , Swedish University of Agricultural Sciences, P.O. 7026, SE-75007, Uppsala, Sweden. Tel: +46 18 671875; Fax: +46 18 673599; Email: magnus.karlsson@mykopat .slu.se Copyright in this article, its metadata, and any supplementary data is held by its author or authors. It is published under the Creative Commons Attribution By licence. For further information go to: http://creativecommons.org/licenses/by/3.0/. Comparative Evolutionary Histories of the Fungal Chitinase Gene Family Reveal Non-Random Size Expansions and Contractions due to Adaptive Natural Selection Magnus Karlsson and Jan Stenlid Department of Forest Mycology and Pathology, Swedish University of Agricultural Sciences, P .O. 7026, SE-75007, Uppsala, Sweden. Abstract: Gene duplication and loss play an importa nt role in the evolution of novel functions and for shaping an organism’s gene content. Recently, it was suggested that stress-related genes frequently are exposed to duplications and losses, while growth-related genes show selection against change in copy number. The fungal chitinase gene family constitutes an inter- esting case study of gene duplication and loss, as their biological roles include growth and development as well as more stress-responsive functions. We used genome sequence data to analyze the size of the chitinase gene family in different fungal taxa, which range from 1 in  Batrachochytrium dendrobatidi s and Schizosacchar omyces pombe to 20 in Hypocrea  jecorina and Emericella nidulans , and to infer their phylogenetic relationships. Novel chitinase subgroups are identi ed and their phylogenetic relationships with previously known chitinases are discussed. We also employ a stochastic birth and death model to show that the fungal chitinase gene family indeed evolves non-randomly, and we identify six fungal lineages where larger-than-expected expansions (Pezizomycotina,  H. jecorina, Gibberella zeae, Uncinocarpus reesii,  E. nidulans and  Rhizopus oryzae ), and two contractions ( Coccidioides immitis and S. pombe) potentially indicate the action of adaptive natural selection. The results indicate that antagonistic fungal-fungal interactions are an important process for soil borne ascomycetes, but not for fungal species that are pathogenic in humans. Unicellular growth is correlated with a reduction of chitinase gene copy numbers which emphasizes the requirement of the combined action of several chitinases for lamentous growth. Keywords: chitinases, gene family, fungi, evolution, phylogeny Introduction Chitin is a polymer which consists of N-acetylglucosamine monomers (GlcNAc), linked by β-1,4-glucosidic bonds. It is widely distributed in nature and it is a constituent of the exoskeleton of invertebrates, of zooplankton and of fungal cell walls. Chitinases (EC 3 .2.1.14) hydrolyze the bonds b etween GlcNAc residues releasing oligomeric, dimeric (chitobiose) or monomeric (GlcNAc) products. Chitobiose can be further cleaved by N-acetylhexosaminidases (EC 3.2.1.52) into GlcNAc (Keyhani and Roseman, 1999). Chitinases are divided into two different glycoside hydrolase families (18 and 19) based on amino acid sequence similarity (Henrissat, 1991; Henrissat and Bairoch, 1993). These two families share limited similarity at the amino acid level and have different three-dimen sional structures and modes of action (Is eli et al. 1996). These enzymes can display either exo- or endoactivity, depending on the structure of the catalytic site (Terwisscha van Scheltinga et al. 1994; van Aalten et al. 2000; van Aalten et al. 2001). Growth and morphological development of fungi makes cell wall remodelling a necessity. Cell expansion and division, spore germination, hyphal branching and septum formation all depend on the activities of hydrolytic enzymes intimately associ ated with the fungal cell wall, among them chitinases (Adams, 2004). Chitinases are also implied in autolysis and recycling of older parts of the fungal myce- lia (Duo-Chuan, 2006). Chitinases also have aggressive roles as fungal pathogenicity factors during infection of other fungi (mycoparasiti sm), insects and nematodes (Wattanalai et al. 2004; Duo-Chuan, 2006; Gan et al. 2007a). Furthermore, chitinases are involved in degradation of chitin for nutritional needs (Duo-Chuan, 2006). Lysis of the host cell wall and degradation of nematode egg shells are shown to be important steps in the mycoparasitic and nematophagous attack (Howell, 2003; Benitez et al. 2004; Gan et al. 2007a), and hence chitinases from various fungi used as biocontrol agents have been

Transcript of quitinase

Page 1: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 1/14

Evolutionary Bioinformatics 2008:4 47–60 47

ORGINAL RESEARCH

Correspondence: Magnus Karlsson, Department of Forest Mycology and Pathology, Swedish Universityof Agricultural Sciences, P.O. 7026, SE-75007, Uppsala, Sweden. Tel: +46 18 671875; Fax: +46 18 673599;Email: [email protected]

Copyright in this article, its metadata, and any supplementary data is held by its author or authors. It is published under the

Creative Commons Attribution By licence. For further information go to: http://creativecommons.org/licenses/by/3.0/.

Comparative Evolutionary Histories of the Fungal ChitinaseGene Family Reveal Non-Random Size Expansionsand Contractions due to Adaptive Natural Selection

Magnus Karlsson and Jan Stenlid

Department of Forest Mycology and Pathology, Swedish University of Agricultural Sciences, P.O. 7026,SE-75007, Uppsala, Sweden.

Abstract: Gene duplication and loss play an important role in the evolution of novel functions and for shaping an organism’sgene content. Recently, it was suggested that stress-related genes frequently are exposed to duplications and losses, whilegrowth-related genes show selection against change in copy number. The fungal chitinase gene family constitutes an inter-esting case study of gene duplication and loss, as their biological roles include growth and development as well as morestress-responsive functions. We used genome sequence data to analyze the size of the chitinase gene family in differentfungal taxa, which range from 1 in  Batrachochytrium dendrobatidis and Schizosaccharomyces pombe to 20 in Hypocrea

 jecorina and Emericella nidulans, and to infer their phylogenetic relationships. Novel chitinase subgroups are identifiedand their phylogenetic relationships with previously known chitinases are discussed. We also employ a stochastic birth anddeath model to show that the fungal chitinase gene family indeed evolves non-randomly, and we identify six fungal lineageswhere larger-than-expected expansions (Pezizomycotina, H. jecorina, Gibberella zeae, Uncinocarpus reesii, E. nidulans and Rhizopus oryzae), and two contractions (Coccidioides immitis and S. pombe) potentially indicate the action of adaptivenatural selection. The results indicate that antagonistic fungal-fungal interactions are an important process for soil borneascomycetes, but not for fungal species that are pathogenic in humans. Unicellular growth is correlated with a reduction of chitinase gene copy numbers which emphasizes the requirement of the combined action of several chitinases for filamentousgrowth.

Keywords: chitinases, gene family, fungi, evolution, phylogeny

IntroductionChitin is a polymer which consists of N-acetylglucosamine monomers (GlcNAc), linked byβ-1,4-glucosidic bonds. It is widely distributed in nature and it is a constituent of the exoskeleton of invertebrates, of zooplankton and of fungal cell walls. Chitinases (EC 3.2.1.14) hydrolyze the bonds betweenGlcNAc residues releasing oligomeric, dimeric (chitobiose) or monomeric (GlcNAc) products. Chitobiosecan be further cleaved by N-acetylhexosaminidases (EC 3.2.1.52) into GlcNAc (Keyhani and Roseman,1999). Chitinases are divided into two different glycoside hydrolase families (18 and 19) based on aminoacid sequence similarity (Henrissat, 1991; Henrissat and Bairoch, 1993). These two families share limitedsimilarity at the amino acid level and have different three-dimensional structures and modes of action (Iseliet al. 1996). These enzymes can display either exo- or endoactivity, depending on the structure of the catalyticsite (Terwisscha van Scheltinga et al. 1994; van Aalten et al. 2000; van Aalten et al. 2001).

Growth and morphological development of fungi makes cell wall remodelling a necessity. Cellexpansion and division, spore germination, hyphal branching and septum formation all depend on theactivities of hydrolytic enzymes intimately associated with the fungal cell wall, among them chitinases(Adams, 2004). Chitinases are also implied in autolysis and recycling of older parts of the fungal myce-lia (Duo-Chuan, 2006). Chitinases also have aggressive roles as fungal pathogenicity factors duringinfection of other fungi (mycoparasitism), insects and nematodes (Wattanalai et al. 2004; Duo-Chuan,

2006; Gan et al. 2007a). Furthermore, chitinases are involved in degradation of chitin for nutritionalneeds (Duo-Chuan, 2006). Lysis of the host cell wall and degradation of nematode egg shells are shownto be important steps in the mycoparasitic and nematophagous attack (Howell, 2003; Benitez et al.2004; Gan et al. 2007a), and hence chitinases from various fungi used as biocontrol agents have been

Page 2: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 2/14

48

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

cloned and characterised (Felse and Panda, 1999;Hoell et al. 2005; Klemsdal et al. 2006; Gan et al.2007a; Gan et al. 2007b; Dong et al. 2007).

The diversity of chitinase function during thefungal life cycle raises interesting questions regard-ing the evolution of this important gene family.Fungal chitinases belong to glycoside hydrolase

family 18 (GH18) and they consist of discretedomains, which are variously arranged in differentorders in different proteins (Gilkes et al. 1991;Warren, 1996; Henrissat and Davies, 2000). Besidesthe catalytic domain there is very often a substrate-  binding domain present. These substrate-bindingdomains are not necessary for chitinolytic activity,although they seem to enhance the ef ficiency of theenzymes (Suzuki et al. 1999; Limon et al. 2001).

There is a large variation in the number of GH18genes present in different fungal genomes, from 1in Schizosaccharomyces pombe to 20 in Hypocrea

 jecorina (Seidl et al. 2005). This implies that thesize of the fungal GH18 gene family has beenhighly dynamic throughout evolution. Gene dupli-cation is an important process that can contributeto the evolution of novel functions. However, themechanisms that govern the fate of duplicated genesare not very well understood. Recent progress sug-gests that stress-related genes frequently areexposed to duplications and losses, while growth-related genes show selection against change in copynumber (Wapinski et al. 2007). High copy-numbersof stress-responsive genes may be beneficial by

allowing adaptations to diverse ecological niches.Recent paralogues diversify most frequently at thelevel of regulation, and more rarely at the level of  biochemical function (Wapinski et al. 2007).

Identification of expansions as well as contractionsof protein families in fungi with diverse ecologicalroles can aid in understanding relationships betweenfunction and phylogeny. The fungal GH18 genefamily constitutes an interesting case study of geneduplication and loss, as their biological rolesincludes growth and development as well as morestress-responsive functions. Hence it is possible totest the hypothesis that growth-related genes displayselection against changes in copy number whilestress-related genes tolerate more duplications andlosses, within a single gene family. Here we presenta study where we use genome sequence data toanalyze the size of the GH18 gene family in differ-ent fungal taxa and to infer their phylogeneticrelationships. We also employ a stochastic birth anddeath model to test for non-random evolution of the

fungal GH18 gene family. We show that the fungalGH18 gene family indeed evolves non-randomlyand we identify fungal lineages where larger-than-expected expansions or contractions potentiallyindicate the action of adaptive natural selection.

Materials and Methods

Biomining of genome sequencesIn order to avoid sampling bias our study wasrestricted to fungal species where genome sequenceinformation and estimates of divergence times wereavailable.  In silico translated gene products fromindividual fungal genome sequences were screenedfor the presence of GH18s using BLASTP (Altschulet al. 1997) in an iterative process. Fungal genomeswere available at the homepages of the DOE JointGenome Institute (http://www.jgi.doe.gov/), theFungal Genome Initiative at the BROAD Institute(http://www.broad.mit.edu/annotation/fgi/) or atGénolevures (http://cbi.labri.u-bordeaux.fr/Genolev-ures/blast/index.php). The 18 published GH18s from  H. jecorina, Chi18-1 through Chi18-18, and twoadditional proteins annotated in the genome, Chi18-rel1 (Chi18-20) and Chi18-rel2 (Chi18-19), proteinID 65162 and 121355 in version 2.0 were used asstarting material, as this is so far the highest number of GH18s from a single fungal species (Seidl et al.2005). Later on the number of proteins used was

reduced to Chi18-1, Chi18-2, Chi18-12 and Chi18-19 from H. jecorina as these representatives providedthe same information as the larger set. In addition,the GH18s that were identified by the first round of similarity searches in a target genome was iterativelyused in a second round of BLAST searches againstthe same genome. The protein identifiers from therespective genome sequencing projects were usedduring subsequent analyses, except for when the protein was characterized and named.

Phylogenetic analysisAmino acid sequences of GH18 catalytic domainswere determined by InterProScan (Zdobnov andApweiler, 2001) or Conserved Domain Databasesearches (Marchler-Bauer et al. 2005). Sequenceswere manually trimmed and aligned with ClustalX (Thompson et al. 1997) and inspected usingBioEdit (Hall, 1999). Amino-acid similarity between sequences was calculated using MegAlign,implemented in the DNASTAR program package

Page 3: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 3/14

49

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

(DNASTAR, Madison, WI). Phylogenetic analysisof catalytic domains was performed using maximumlikelihood methods implemented in PhyML-aLRT1.1 (Guindon and Gascuel, 2003; Anisimova andGascuel, 2006). The JTT amino-acid substitutionmodel (Jones et al. 1992) was used, the proportionof invariable sites was set to 0, one category of 

substitution rate was used and gaps were treated asunknown characters. The starting tree to be refined  by the maximum likelihood algorithm was adistance-based BIONJ tree estimated by the program (Guindon and Gascuel, 2003). Statisticalsupport for phylogenetic grouping was assessed byapproximate likelihood-ratio tests based on aShimodaira-Hasegawa-like procedure (SH-aLRT)(Anisimova and Gascuel, 2006) and by bootstrapanalysis (500 resamplings).

Likelihood analysis of gene gain and lossIn order to statistically test whether the size of thefungal GH18 gene family is compatible with a

stochastic birth and death model we used the program CAFE (Computational Analysis of geneFamily Evolution) (De Bie et al. 2006), which is based on the probabilistic framework developed by Hahn et al. (2005). From a specified phyloge-netic tree and the gene family size in extant species,we inferred the most likely gene family size at

internal nodes, tested for accelerated rates of genefamily expansions or contractions and identifiedthe branches that are responsible for the non-random evolution.

The fungal GH18 gene family data in extantspecies that were used in the current analysis arefound in Table 1. Fungal GH18s can be dividedinto three major phylogenetic clusters, A, B andC, and further subdivisions within these clustersare made (Seidl et al. 2005). Our phylogeneticanalysis shows that cluster C GH18s can bemerged with cluster A and therefore we analysed

the data in three ways; cluster A GH18s separately,cluster B GH18s separately and all GH18s merged.CAFE assumes that the gene family under study

Table 1. Number of chitinase genes in different fungal species.

Species Class Cluster AGH18 genes

Cluster BGH18 genes

Total no. of GH18 genes

Batrachochytriumdendrobatidis

Chytridiomycetes 1 0 1

Rhizopus oryzae Mucormycotina 9 6 15

Schizosaccharomyces pombe

Schizosaccharomycetes 0 1 1

Yarrowia lipolytica Saccharomycetes 1 2 3

Candida albicans Saccharomycetes 2 3 5

Saccharomycescerevisiae

Saccharomycetes 1 1 2

Emericella nidulans Eurotiomycetes 17 3 20

 Ajellomycescapsulatus

Eurotiomycetes 7 2 9

Uncinocarpus reesii  Eurotiomycetes 11 3 14

Coccidioides immitis Eurotiomycetes 6 3 9

Gibberella zeae Sordariomycetes 16 3 19

Hypocrea jecorina Sordariomycetes 11 9 20

Magnaporthe grisea Sordariomycetes 11 4 15

Neurospora crassa Sordariomycetes 8 4 12

Ustilago maydis Ustilaginomycetes 2 1 3

Filobasidiellaneoformans

Tremellomycetes 3 1 4

Coprinopsis cinerea Agaricomycetes 7 1 8

Phanerochaetechrysosporium

Agaricomycetes 7 3 10

Page 4: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 4/14

50

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

is present in the most recent common ancestor of all taxa included in the analysis. Therefore  Batrachochytrium dendrobatidis was excludedfrom the analysis of cluster B GH18s. The phylogenetic relationships between the speciesthat were included in the analysis are shown inFigure 1, with branch lengths in millions of years.

Phylogenetic relationships and estimations of divergence times were taken from previous publications (Bowman et al. 1996; Kasuga et al.2002; Padovan et al. 2005; Taylor and Berbee,2006), assuming that the Devonian ascomycete Paleopyrenomycites devonicus(Taylor et al. 2005)represents Pezizomycotina (Taylor and Berbee,2006) which gives an estimated age of 923millions of years for the fungal phylum.

Alternative estimates of divergence times can be made by assuming that P. devonicus represents

Sordariomycetes (estimated age of the fungal phylum at 1630 millions of years) or Ascomycota(estimated age of the fungal phylum at 495 millionsof years) as outlined in Taylor and Berbee, (2006),although these alternative estimates resulted inmore improbable age estimates when comparedwith age estimates in other phyla. However, these

alternative estimates were included in the analysisalthough Coccidioides immitis, Uncinocarpusreesii, Ajellomyces capsulatus and B. dendrobatidis were excluded because of incompatibility of divergence estimates (Bowman et al. 1996; Kasugaet al. 2002; Padovan et al. 2005; Taylor and Berbee,2006). The four different phylogenetic trees usedin the current study including branch lengths arefound in Supplemental information (S1).

The birth and death parameter (λ ) was estimatedfrom the data (De Bie et al. 2006) and was 0.001

N crassa

M grisea

H jecorina +

G zeae +

C immitis -

U reesii +

 A capsulatus

E nidulans +

S cerevisiae

C albicans

Y lipolytica

S pombe -

P chrysosporium

C cinerea

F neoformans

U maydis

R oryzae +

B dendrobatidis

Pezizomycotina +

0100200300400500600700800

MY

14

911

9

11

20

12

12

12

15

20

19

15

15

16

5

5

5

55

5

4

4

4

2

5

3

1

10

8

8

3

15

1

Figure 1. Distribution of GH18 gain and loss among fungal lineages. Phylogenetic relationships among the fungal species used in the cur-rent study are shown, including divergence dates in millions of years (Taylor and Berbee, 2006). Circled numbers represent total number of GH18 genes in extant species and estimates of total number of GH18 genes for ancestral species. Boxed taxon names indicates a significant( p-values 0.05 or Likelihood ratios 50) expansion (+), or a significant contraction (-) of the GH18 gene family size.

Page 5: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 5/14

51

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

for all datasets.  p-values were computed using1000 re-samplings and identification of the branchthat was the most likely cause of deviations froma random model was determined by Viterbi,Branch-cutting and Likelihood ratio test procedures(De Bie et al. 2006). We considered  p-values  0.05 or likelihood ratios above 50 to be significant

for branch identification.

ResultsThere was considerable variation in the number of GH18 genes between different fungal species,ranging from 1 in B. dendrobatidis and S. pombe to 20 in  H. jecorina and  Emericella nidulans (Table 1). Because of the large variation in lengthand domain structure of fungal GH18s (Seidl et al.2005), only the GH18 catalytic domains wereused for phylogenetic analysis. Fungal GH18swere divided into two major clusters, A and B

(Figs. 2 and 3), each subdivided into groups.Cluster A sequences formed six separate groupsthat were equivalent with the A-II, A-III, A-IV,A-V, C-I and C-II groups described by Seidl et al.(2005) (Fig. 2). Cluster B sequences formed fiveseparate groups, here referred to as B-I throughB-V (Fig. 3). Groups B-I and B-II were described previously (Seidl et al. 2005). The average num- ber of cluster A GH18 genes in the included spe-cies was 6.6, ranging from 0 (S. pombe) and 17( E. nidulans). This average was higher than for cluster B GH18 genes where the average was 2.8

genes per species, ranging from 0 ( B. dendroba-tidis) and 9 ( H. jecorina). The different GH18subgroups were characterized by considerabledifferences in levels of amino-acid conservation.Subgroups A-II, A-IV, A-V and B-V showed thehighest levels of mean interspecific amino-acididentity between species in Eurotiales (Table 2).Members in subgroups C-I and C-II displayed thelowest levels of conservation while levels in B-Iand B-II were intermediate to the other groups(Table 2).

To investigate the evolutionary change of thenumber of GH18 genes in fungi, we estimated thenumber of genes in ancestral species and thenumber of gene gains and losses for each branchof the phylogenetic tree of the fungal species(Fig. 1). The analysis showed that the fungal GH18gene family, analysing both cluster A and Btogether, as well as cluster A GH18s alone haveevolved non-randomly (p 0.001, Table 3). Six  branches were identified as contributing to this

non-random pattern, including 5 expansions and 1contraction (Table 3). These branches included theancestor to the Pezizomycotina clade, as well asextant species C. immitis, U. reesii,  E. nidulans,Gibberella zeae and Rhizopus oryzae. Analysis of gene phylogenies of GH18 subgroups, identifiedsubgroups C-I and C-II as the likely targets

responsible for the non-random expansion seen in E. nidulans and U. reesii (Fig. 2). The contractionin C. immitis probably took place in subgroup C-II(Fig. 2). It was not possible to identify the targetsubgroup in G. zeae, although C-I and C-IIappeared to have expanded compared to other Sordariomycetes (Fig. 2, Seidl et al. 2005). Itshould be noted though, that an additional genewas also present in subgroups A-III and A-V, ascompared with the closely related   H. jecorina (Fig. 2, Seidl et al. 2005). The expansion seen incluster A GH18s in   R. oryzae took place in

subgroup A-III (Fig. 2). Although cluster B GH18sdid not display a non-random pattern of evolutionas a whole (p = 0.568), two branches were stillidentified where significant expansions took place(Table 3). For both H. jecorina and R. oryzae thisexpansion took place in the large B-I/II/III/IVcluster (Fig. 3). Analyzing all GH18s together using   R. oryzae as the most recent commonancestor of all taxa instead of   B. dendrobatidis resulted in a significant contraction in S. pombe (Table S2).

Taylor and Berbee, (2006) also published two

alternative estimates of divergence times of fungaltaxa, although these alternative estimates resultedin more improbable age estimates when comparedwith age estimates in other phyla. Analyzingthe evolution of the GH18 gene family using thesealternative estimates was performed to assess therobustness of the analysis to differences indivergence dates. The two more recent estimatesof divergence dates (estimated age of the fungal  phylum at 495 or 923 millions of years) bothshowed that the GH18 family have evolved non-randomly (p 0.001, Table S2), and identified thesame branches as contributing to this non-random pattern (Table S2). The oldest estimate (estimatedage of the fungal phylum at 1630 millions of years)gave no significant changes in size of the GH18gene family (Table S2).

The involvement of GH18 subgroups C-I andC-II in both expansions and contractions inascomycetes qualified them for further study. Basedon the genome sequence, C-I members in E. nidulans 

Page 6: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 6/14

52

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

EAA 74223 G. zeaeChi18 7 H. jecorina

100/1.0

MGG 0 7927 M. grisea04554 N. crassa

100/1.0

02795 C. immitis03042 U. reesii

100/1.0

 AN54 54 E. nidulans72/.97

 AN788 6 E. nidulans00203 U. reesii

03822 C. immitis06565 A. capsulatus05886 A. capsulatus

87/.9571/.94

-/.94

 AN48 71 E. nidulans

-/.97

04883 N. crassaMGG 0 0086 M. grisea

EAA 74986 G. zeaeEAA73155 G. zeae

Chi18 5 H. jecorina77/.87

95/.99

-/.79

MGG 0 1247 M. griseaChi18 6 H. jecorinaEAA 67655 G. zeae

MGG 05 533 M. grisea

75/.91

03209 N. crassa

93/1.0

89/.99

-/.97

01250 A. capsulatus05250 C. immitis

06375 U. reesii100/1.0

99/.99

-/1.0

UM04261 U. maydis134311 P. chrysosporium

09921. 1 C. cinerea100/1.0

124149 P. chrysosporium04870.1 C. cinerea

95/.99

-/.94

13635 R. oryzae

-/.82

03412 F. neoformans

-/.87

0F04532 Y. lipolyticaCTS2 S . c erevisiae

CHT4 C. albicans87/.98

-/.75

MGG 1 1231 M. grisea03026 N. crassa

81/.86

Chi18 3 H. jecorina98/1.0

 AN029 9 E. nidulans04742 A. capsulatus

04750 C. immitis06761 U. reesii

92/.9779/.85

98/.98

MGG 0 8458 M. grisea

EAA 70860 G. zeaeChi18 2 H. jecorina98/.98

99/1.0

95/1.0

-/.88

06156 B. dendrobatidis

76/.98

02598 F. neoformans02650 A. capsulatus

 AN02 21 E. nidulans77/.95

06020 N. crassaEAA 76014 G. zeae

Chi18 4 H. jecorina100/1.0

MGG 0 8054 M. grisea

88/.95

Chi18 11 H. jecorinaEAA 69039 G. zeae

MGG 0 6594 M. grisea74/.86

-/.79

05838 A. capsulatus AN944 7 E. nidulans

81/.84

100/1.0

93/.96

10252 R. oryzae07611 R. oryzae

89/.95

14659 R. oryzae05550 R. oryzae

99/1.0

98/.99

EAA 71245 G. zeaeMGG 0 4732 M. grisea

100/1.0

128098 P. chrysosporium138098 P. chrysosporium

05291. 1 C. cinerea05285.1 C. cinerea

93/1.0

-/.98

03252.1 C. cinerea01315.1 C. cinerea

-/.87

-/.85

39872 P. chrysosporium

-/.84

100/1.0

9211 P. chrysosporium01586.1 C. cinerea

99/1.0

UM06190 U. maydis129436 P. chrysosporium

04245 F. neoformans

88/.94

94/1.0

93/1.0

13934 R. oryzae01334 R. oryzae16170 R. oryzae100/1.0

100/1.0

-/.73

90/1.0

-/.80

07387 C. immitisEAA 75711 G. zeae

MGG 0 1336 M. grisea99/.97

-/.93

02055 U. reesii04323 A. capsulatus

100/.99

04761 C. immitis AN507 7 E. nidulans

94/1.0

 AN05 41 E. nidulans AN899 9 E. nidulans

 AN05 17 E. nidulans

100/1.085/.99

-/.91

 AN939 0 E. nidulansMGG 0 4534 M. grisea

07183 U. reesii100/1.0

EAA74768 G. zeae04769 U. reesii AN848 1 E. nidulans AN054 9 E. nidulans

100/1.0

78/.96

07035 N. crassaEAA 78168 G. zeae

83/.97

-/.93

-/.84

85/.96

19.1245 1 C. albicans01094 U. reesii

06727 U. reesiiChi18 1 H. jecorina

-/.76

EAA 68447 G. zeae05317 N. crassa

86/.89

 AN05 09 E. nidulans07484 N. crassa

100/1.0

-/.91

Chi18 9 H. jecorina

EAA 78214 G. zeae

-/.89

 AN76 13 E. nidulans AN666 2 E. nidulans

 AN397 8 E. nidulans

100/1.0-/.95

-/.96

Chi18 10 H. jecorina

-/.94

Chi18 8 H. jecorina04114 U. reesii

100/1.0

EAA 72565 G. zeaeEAA 77156 G. zeae

-/.76

-/.81

94/.94

97/1.0

84/.86

0.1

C-I

A-II

A-IV

A-V

C-II

A-III

Figure 2. Phylogenetic relationships of fungal cluster A family 18 glycoside hydrolase catalytic domains. Phylogenetic analyses wereperformed using maximum likelihood methods as implemented in PhyML-aLRT, based on an alignment of family 18 glycoside hydrolasecatalytic domain amino acid sequences. Branch support values (bootstrap proportions/approximate likelihood-ratio test probabilities) areassociated with nodes, with a dash indicating that the support was 70%/0.70. The bar marker indicates the number of amino-acid sub-stitutions. Protein identifiers include protein name, GenBank accession nos. or locus/protein ID from the respective genome projects. Groupnames are indicated, see text for reference. Cluster B GH18 Chi18-12 from H. jecorina was used as outgroup (not shown).

Page 7: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 7/14

53

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

 A N11059 E. nidulans073 48 C. immitis07888 U. reesii

100/.9780/.86

MGG 103 33 M. grisea04500 N. crassa

75/.91

Chi18 17 H. jecorinaChi18 12 H. jecorina

74/.92

-/.74

99/1.0

10112 R. oryzae

-/.79

Chi18 13 H. jecorina08153 R. oryzae

6412 P. chrysosporium40899 P. chrysosporium

044 07.1 C. cinerea77/.92

-/.88

2991 P. chrysosporium

70/-

-/.94

11152 R. oryzae04523 R. oryzae

96/.99

-/.80

-/.82

0D22 396 Y . lipolytica0A 018 70 Y . lipolyticaCHT2 C. albicans

CHT1 C. albicansCHT3 C. albicans

CTS1 S. cerevisiae74/.98

-/.96

-/.90

16099 R. oryzae

-/.84

NP 594 130 S. pombe

-/.79

MGG 035 99 M. grisea

-/.79

02184 N. crassaEAA 72615 G. zeae

Chi18 18 H. jecorina95/.97

87/.99

 AN8 241 E. nidulans003 48 C. immitis

00414 U. reesii98/.97

86/.95

Chi18 16 H. jecorinaChi18 14 H. jecorina

87/-

97/.99

Chi18 15 H. jecorinaMGG 018 76 M. grisea

08394 N. crassaChi18 20 H. jecorina

005 67 A . capsulatus065 95 C. immitis038 48 U. reesii

97/.9683/.91

 AN8 245 E. nidulansXP 386 145 G. zeae

MGG 040 73 M. grisea01393 N. crassa

Chi18 19 H. jecorina

-/.73

91/.93

XP 382 346 G. zeae

94/.97

98/.96

02351 F. neoformansUM027 58 U. maydis

100/1.0

94/.93

0. 1

B-II

B-III

B-IV

B-I

B-V

Figure 3. Phylogenetic relationships of fungal cluster B family 18 glycoside hydrolase catalytic domains. Phylogenetic analyses were per-formed using maximum likelihood methods as implemented in PhyML-aLRT, based on an alignment of family 18 glycoside hydrolase cata-lytic domain amino acid sequences. Branch support values (bootstrap proportions/approximate likelihood-ratio test probabilities) areassociated with nodes, with a dash indicating that the support was 70%/0.70. The bar marker indicates the number of amino-acid substi-

tutions. Protein identifiers include protein name, GenBank accession nos. or locus/protein ID from the respective genome projects. Groupnames are indicated, see text for reference. Cluster A GH18 06156 from B. dendrobatidis was used as outgroup (not shown).

were characterised by Chitin-binding type 1 domains(InterPro acc. no. IPR001002), while C-II membersin  E. nidulans contained LysM peptidoglycan binding domains (InterPro acc. no. IPR002482) inaddition to Chitin-binding type 1 domains. These  proteins showed considerable similarity withthe α-subunit of the yeast killer toxin from Kluyveromyces lactis (Fig. 4). The A-III membersin   R. oryzae were short and contained no other domains than the GH18 catalytic domain. Thedomain-structure of cluster B GH18s in H. jecorina has been described before (Seidl et al. 2005) andwere characterized by fungal cellulose bindingdomains (InterPro acc. no. IPR000254). The domain-structure of  R. oryzae cluster B members included both carbohydrate-binding family V/XII domains(InterPro acc. no. IPR003610) and GH18 carbohydrate binding domains (InterPro acc. no. IPR005089).

DiscussionAccurate estimates of divergence times for fungi arenotoriously hard to obtain due to a very limited fossilrecord. In the current study we have used threedifferent time estimates on our GH18 gene familydata. The differences relates to whether the Devonianascomycete   P. devonicus (Taylor et al. 2005)represents Sordariomycetes, Pezizomycotina or Ascomycota (Taylor and Berbee, 2006). Using theSordariomycete calibration result in no significantchanges in the GH18 gene family due to the estimatedancient origin of the fungal phylum (1630 millionsof years), more than three times the age of the firstfossil evidence of land plants (Taylor and Berbee,2006). The results when using the two divergenceestimates based on P. devonicus as a representativefor Pezizomycotina (origin of fungal phylum at 923millions of years) or Ascomycota (origin of fungal

Page 8: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 8/14

54

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

 phylum at 495 millions of years) are very similar,which show a certain level of robustness of theanalysis, even for large differences in divergenceestimates. Furthermore, the Pezizomycotina estimate

has been shown to result in least number of improbabletime estimates when compared with other phyla(Taylor and Berbee, 2006).

The number of GH18 genes in different fungalspecies varies considerably. An expansion in sizeof a particular gene family or subgroup within agene family, such as GH18s, suggests that this genefamily or subgroup has been important for thefitness of the species during evolution. Theobserved variation could possibly be attributed to

differences in morphology, growth patterns,nutrient acquisition or antagonistic ability betweenspecies. Our approach can be used to establish links between phylogeny of the GH18 gene family with

the ecological role of the species and to identifyspecific subgroups as important evolutionarytargets in specific fungal lineages.

Filamentous ascomycetes generally possesslarger number of GH18 genes as compared withother fungal groups. This larger GH18 gene familysize can possibly be attributed to a larger genecopy-number in certain subgroups, but moreimportantly to the presence of several GH18subgroups that appear to be unique for filamentous

Table 3. Non-randomly evolving branches in the fungal GH18 gene family.

Data set Branch ID  p-value1 Likelihood ratio1 Change2

All GH18 genes 0.001

Pezizomycotina 0.005 36 7

Coccidioides 0.032 48 −2

Uncinocarpus 0.009 31 3

Emericella 0.002 14 8

Rhizopus 0.010 1 10

Cluster A GH18 genes 0.001

Pezizomycotina 0.002 34 6

Gibberella 0.026 5 5

Coccidioides 0.002 161 −3

Uncinocarpus 0.053 57 2

Emericella 0.004 24 8

Rhizopus 0.043 1 6

Cluster B GH18 genes 0.568

Hypocrea 0.002 14 5

Rhizopus 0.018 1 5

1See De Bie et al. (2006) for reference.2Gene family size change as compared with the most recent ancestor.

Table 2. Mean intraspecific levels of % amino-acid identity for GH18 subgroups in E. nidulansand mean interspecificlevels of % amino-acid identity for GH18 subgroups between E. nidulans and other members of Eurotiales.

GH18 subgroup E. nidulans A. capsulatus U. reesii C. immitis

A-II 42.4 45.9 - -

A-IV * 52.5 56.8 57.1

A-V 58.3 56.4 52.4 54.0

C-I 22.8 12.0 15.8 19.3

C-II 25.3 - 27.1 -

B-I * - 35.7 35.3

B-II * - 29.2 30.2

B-V * 57.5 54.6 56.2

Abbreviations: *: Only one GH18 member in E. nidulans; -: no GH18 members present.

Page 9: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 9/14

55

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

ascomycetes (A-II, C-I and C-II). Subgroups C-Iand C-II are identified as the most likely target for 

the observed expansion in E. nidulans, U. reesii,G. zeae and the ancestor to Pezizomycotina. TheseGH18 genes share extensive homology with theα-subunit of the yeast killer toxin from K. lactis.This yeast killer toxin, zymocin, consists of threesubunits (α, β and γ ) where toxicity relies solelyon the γ -subunit and theα- and β-subunits functionin the delivery of γ inside the cell by permeabiliza-tion of the yeast cell wall and membrane (Stark et al. 1990; Magliani et al. 1997). This has led tothe suggestion that the C-I and C-II GH18s areinvolved in a similar mechanism in aggressivefungal-fungal interactions, by permeabilization of the cell wall and membrane to enable penetrationof antifungal molecules into the antagonist(Seidl et al. 2005). This role is supported byexpression data; the C-II member chi18-10 from H. atroviridis is only expressed during growth onfungal cell walls and during plate confrontationassays, but not by carbon starvation or chitin expo-sure (Seidl et al. 2005). The expansion of C-I andC-II GH18s suggests that interspecific interactionsare an important process for soil borne ascomyce-tes. It also supports the idea that genes involved instress-related functions can tolerate, or are evenunder selection for, increases in copy number (Wapinski et al. 2007).

Another intriguing result is the different evolu-tionary trajectories of the GH18 gene family between the human pathogen C. immitis and theclosely related U. reesii. An expansion of GH18subgroup C-II in saprotrophic U. reesii is incontrast to a contraction in the same subgroup in

the pathogenic C. immitis, although these speciesare very closely related (Bowman et al. 1996;

Kasuga et al. 2002). This difference should berelated to the different life-styles of the two fungi,and we hypothesize that the expansion of C-II seenin U. reesii is a consequence of the need for antagonistic interactions with other soil-dwellingfungi. The contraction in C. immitis may be reflect-ing adaptation to a pathogenic lifestyle, whereantagonistic interactions with other fungi areminimized. Another human pathogen, A. capsulatus ,also contains the same number of GH18 genes (9)as C. immitis and also lacks subgroup C-II mem- bers completely, which indicate that the function performed by C-II members of the GH18 proteinfamily is dispensable for the human pathogeniclifestyle.

The high interspecific sequence variability inC-I and C-II as compared with other GH18 sub-groups can be interpreted as the result of diversify-ing selection. Diversifying evolution due to positive selection is reported for plant chitinasesthat function as defence proteins against invadingfungal pathogens (Bishop et al. 2000; Tif fin, 2004)and in reproductive proteins, in animals and plants(Clark et al. 2006) as well as in fungi (Karlssonet al. 2008). It is possible that the high sequencevariability in C-I and C-II represents an adaptationtowards differences in cell wall composition inantagonistic species. The significant expansion of the fungal GH18 gene family in the ancestor of thePezizomycotina probably reflects the emergenceof the unique C-I and C-II GH18s in filamentousascomycete fungi, although the ecological factorsdriving the selection for these subgroups remain

Figure 4. Partial alignment of K. lactis zymocin with subgroups C-I and C-II members in E. nidulans. Identical residues in a column areindicated in white and boxed in black, two different residues in a column are indicated by gray boxes, gaps are indicated by dashes. Theconserved GH18 gene family active site residues are indicated by asterisks. AN10838 and AN0509 represents the C-II GH18 subgroup,AN5077, AN0517 and AN8481 represents the C-I subgroup, KT = Killer Toxin α-subunit.

*********

KT (K. lactis) SFGGWDFSTSPSTYTIFRNAVKTDQNRNTFANNLINFMNKYNLDGIDLDWEYPGAPDIPDIPA

AN10838 (E. nidulans) SFGGWGYSTEPETYDILRQAMSP.PNRKAFATNVAAFVTEHELDGVDFDWEYPGAVDIPGTPP

AN0509 (E. nidulans) SFGGWDFSTNPRTYNIFRQGTQP.ANRLKLATNIAKFVKDHGLDGVDIDWEYPGAPDIPGIPA

AN5077 (E. nidulans) ALGGWTFSDPGPWQAIFPTLASTAANRATFIQNLLGFMSEYGYDGVDFDWEYPGADDRGGSDS

AN0517 (E. nidulans) AVGGWAFSD.APTQHLWTQMARSHENRQTFINSVVKYLQDYHLDGIDIDWEYPSASDRGGAP-

AN8481 (E. nidulans) AVGGWTFNDPGPTATVFSDIASSLKNQRAFFKSLISFMSTYNFDGIDLDWEYPVADDRSGRP-

KT (K. lactis) DDSSSGSNYLTFLKLLKGKMP-SG--KTLSIAIPSSYWYLKNFPISDIQNTVDYMVYMTYDIH

AN10838 (E. nidulans) GLETDGPNYYKFLIVMRGQLA-EG--ISLSIAAPASYWYLKAFPIDLMAKELDYIVFMTYDLHAN0509 (E. nidulans) GSEDEGDNYLKFLVVLKNLLK-D---KSVSIAAPASYWYLRGFPINSISKVVDYIVFMTYDLH

AN5077 (E. nidulans) --MVDGENYTLLLKELQEAITASGRNYLVTFTAPTSYWYLRHFDLKAMMEYVDWVNLMSYDLH

AN0517 (E. nidulans) ---QDAANF-----N-------PG--WEISATLPTSYWYLRGFDVDRMQKYVDYFNLMSYDLH

AN8481 (E. nidulans) ---ADYENFPRFIANLKKALKGSGGRDGLSITLPASYWYLQHFDIINLQDHVDFFNIMSYDLH

Page 10: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 10/14

56

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

obscure. Although very speculative, the emergenceof C-I and C-II GH18s coincide with the estimatedtime of colonization of terrestrial environments by plants (Sanderson, 2003), which suggests the pos-sibility that the emergence of terrestrial plantscreated new ecological niches where filamentousascomycetes could expand into to compete for 

space and nutrients.Another expansion took place in cluster BGH18s in H. jecorina, which is closely related tomycoparasitic fungi such as H. lixii, H. virens and H. atroviridis. Seidl et al. (2005) reported thatcertain cluster B GH18 genes from   H. jecorina have high similarity to GH18s from entomopatho-gens, such as Metarhizium anisopliae, which sug-gests an aggressive role of these proteins in chitindegradation. Again, there is expression data thatsupport this role; chi18-13 from H. atroviridis isup-regulated during growth on fungal cell walls

and during plate confrontation assays (Seidl et al.2005). The fact that two GH18 subgroups that areimplied in aggressive fungal-fungal interactionshave expanded significantly during fungal evolu-tion suggests that interspecific antagonistic interac-tions are important determinants of fungalevolution, community development and function-ing. This result is in line with the idea that genesinvolved in stress-related functions can tolerate, or are even under selection for, increases in copynumber (Wapinski et al. 2007).

Species with yeast-like or monocentric growth

styles such as Saccharomyces cerevisiae, Yarrowialipolytica, S. pombe and B. dendrobatidis all havelow numbers of GH18 genes (2, 3, 1 and 1 genes)as compared with other fungi, even with closelyrelated species exhibiting filamentous growth e.g.Candida albicans (5 genes). This suggests thatfilamentous growth requires the combined actionof several GH18s as compared with a yeast-likegrowth style, even though the reduction in the non-filamentous species are only significant for  S. pombe (p = 0.023).

The analysis of the GH18 gene family size in basidiomycetes shows no conflict with a random  process. However, 5 of the 8 GH18 genes inCoprinopsis cinerea and 5 of the 10 GH18 genesin  Phanerochaete chrysosporium are found insubgroup A-III, compared to only one representa-tive from Ustilago maydis and  Filobasidiellaneoformans. Further studies of more closelyrelated species with shorter coalescence timeswill be needed to determine if the apparent

expansion of GH18 subgroup A-III in saprotrophic  basidiomycetes and the apparent contraction in  pathogenic basidiomycetes is a consequence of the different life-styles of the species. Based onthe observed expansion of GH18 subgroupsinvolved in interspecific interactions in ascomy-cetes, we can hypothesize about the involvement

of A-III GH18s in fungal-fungal interactions in basidiomycetes.Our analysis shows the usefulness of the com-

 bination of a stochastic birth and death model and phylogenetic information in a probabilistic frame-work for identification of lineages with unusuallyevolving gene families. However, the birth anddeath model assumes independence among indi-vidual genes. This means that any large-scalechromosome duplication, deletion or polyploidiza-tion that acts on several gene family members atonce violates the assumption of the model (Hahn

et al. 2005). Interpretation of gene family size dif-ferences between taxa that are separated by genomeduplications should be made with caution. Thereare indications of a recent polyploidization in  R.oryzae (Taylor and Berbee, 2006) which suggestthe possibility that the observed non-random GH18family size in this species may not be entirelyrelated to adaptive selection. On the other hand,the two GH18 genes in the duplicated S. cerevisiae genome (Kellis et al. 2004) belong to the funda-mentally different A and B clusters. This indicatethat the loss of one of the duplicated CTS1 andCTS2 paralogues have been selected for duringevolution after whole-genome duplication. Cts1and cts2 are involved in cell separation during budding and in sporulation (Kuranda and Robbins,1991; Giaever et al. 2002), which is in line withthe idea that genes involved in growth-relatedfunctions are under selection against changes incopy numbers (Wapinski et al. 2007).

Another violation against the assumption of independence among individual genes may be seenin Neurospora crassa. This species has the lowestnumber of GH18 genes (12) among the Sordari-omycetes, which may be attributed to the presenceof a wide array of genome defence mechanisms,including repeat-induced point mutations, greatlyslowing down the creation of new genes (Galaganet al. 2003). The low number of GH18 genes inthis species is not significantly violating a random process (p = 0.256), but interpretation of data for other gene families in  N. crassa should be donewith caution.

Page 11: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 11/14

57

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

In the current study we have used fungalgenome data in comparative work to infer phylo-genetic relationships in the fungal GH18 genefamily and to detect non-random expansions andcontractions. This approach can be used to establishlinks between phylogeny of the GH18 gene familywith the ecological role of the species, and identify

specific GH18 subgroups as important evolution-ary targets in specific fungal lineages. Within thefungal GH18 gene family we observe selectionagainst changes in copy number in GH18s involvedin growth and development as well as selection for increased copy number in GH18s involved instress-related functions, supporting the idea of a bipolar principle that governs tolerance to duplica-tions and losses (Wapinski et al. 2007). The resultsalso indicate that antagonistic fungal-fungal inter-actions constitute an important evolutionary forcein soil borne ascomycetes, but not for fungal spe-

cies that are pathogenic in humans.

AcknowledgementsThis work was supported by a grant from the Swed-ish Research Council for Environment, Agricul-tural Sciences and Spatial Planning (UppsalaMicrobiomics Center). Genome sequence datawere produced by the US Department of EnergyJoint Genome Institute (http://www.jgi.doe.gov/),The Fungal Genome Initiative at the BROADInstitute (http://www.broad.mit.edu/annotation/fungi/fgi/index.html) and Génolevures (http://cbi.labri.u-bordeaux.fr/Genolevures/blast/index.php).We thank Timothy James, Åke Olson and NilsHögberg for critical reading of the manuscript.

ReferencesAdams, D.J. 2004. Fungal cell wall chitinases and glucanases. Microbiol.,

150:2029–35.Altschul, S.F., Madden, T.L., Schäffer, A.A. et al. 1997. Gapped BLAST

and PSI-BLAST: a new generation of protein database search pro-grams. Nucleic Acids Res., 25:3389–402.

Anisimova, M. and Gascuel, O. 2006. Approximate likelihood-ratio test for 

  branches: A fast, accurate, and powerful alternative. Syst. Biol.,55:539–52.

Benitez, T., Rincon, A.M., Limon, M.C. et al . 2004. Biocontrol mechanismsof Trichoderma strains. Int. Microbiol., 7:249–60.

Bishop, J.G., Dean, A.M. and Mitchell-Olds, T. 2000. Rapid evolution in  plant chitinases: Molecular targets of selection in plant-pathogencoevolution. Proc. Natl. Acad. Sci. U.S.A., 97:5322–7.

Bowman, B.H., White, T.J. and Taylor, J.W. 1996. Human pathogeneicfungi and their close nonpathogenic relatives. Mol. Phyl. Evol.,6:89–96.

Clark, N.L., Aagaard, J.E. and Swanson, W.J. 2006. Evolution of reproduc-tive proteins from animals and plants. Reproduction, 131:11–22.

De Bie, T., Cristianini, N., Demuth, J.P. et al. 2006. CAFE: a computationaltool for the study of gene family evolution.  Bioinformatics ,22:1269–71.

Dong, L.Q., Yang, J.K. and Zhang, K.Q. 2007. Cloning and phylogeneticanalysis of the chitinase gene from the facultative pathogen Paecilo-myces lilacinus. J. Appl. Microbiol., 103:2476–88.

Duo-Chuan, L. 2006. Review of fungal chitinases. Mycopath., 161:345–60.Felse, P.A. and Panda, T. 1999. Regulation and cloning of microbial chitin-

ase genes. App. Microbiol. Biotech., 51:141–51.Galagan, J.E., Calvo, S.E., Borkovich, K.A. et al. 2003. The genome

sequence of the filamentous fungus Neurospora crassa.  Nature,422:859–68.

Gan, Z., Yang, J., Tao, N. et al. 2007a. Cloning of the gene Lecanicillium psalliotae chitinase Lpchi1 and identification of its potential role inthe biocontrol of root-knot nematode Meloidogyne incognita.

 Appl. Microbiol. Biotechnol., 76:1309–17.Gan, Z., Yang, J., Tao, N. et al. 2007b. Cloning and expression analysis of 

a chitinase gene Crchi1 from the mycoparasitic fungus Clonostachysrosea (syn. Gliocladium roseum). J. Microbiol., 45:422–30.

Giaever, G., Chu, A.M., Ni, L. et al . 2002. Functional profiling of the Sac-charomyces cerevisiae genome. Nature, 418:387–91.

Gilkes, N.R., Henrissat, B., Kilburn, D.G. et al. 1991. Domains in microbial beta-1,4-glycanases—Sequence conservation, function, and enzymefamilies. Microbiol. Rev., 55:303–15.

Guindon, S. and Gascuel, O. 2003. A simple, fast, and accurate algorithm

to estimate large phylogenies by maximum likelihood. Syst. Biol.,52:696–704.

Hahn, M.W., De Bie, T., Stajich, J.E. et al. 2005. Estimating the tempo andmode of gene family evolution from comparative genomic data.Genome Res., 15:1153–60.

Hall, T.A. 1999. BioEdit: a user-friendly biological sequence alignmenteditor and analysis program for Windows 95/98/NT.  Nucl. Acids

Symp. Ser., 41:95–8.Henrissat, B. 1991. A classification of glycosyl hydrolases based on amino

acid sequence similarities. Biochem. J., 280:309–16.Henrissat, B. and Bairoch, A. 1993. New families in the classification of 

glycosyl hydrolases based on amino acid sequence similarities. Biochem. J., 781–8.

Henrissat, B. and Davies, G.J. 2000. Glycoside hydrolases and glycosyl-transferases Families, modules, and implications for genomics.

 Plant Physiol., 124:1515–9.Hoell, I.A., Klemsdal, S.S., Vaaje-Kolstad, G. et al. 2005. Overexpression

and characterization of a novel chitinase from Trichoderma atroviridestrain P1. Biochim. Biophys. Acta. Prot. Proteom, 1748:180–90.

Howell, C.R. 2003. Mechanisms employed by Trichoderma species in the biological control of plant diseases: The history and evolution of current concepts. Plant Dis., 87:4–10.

Iseli, B., Armand, S., Boller, T. et al. 1996. Plant chitinases use two differ-ent hydrolytic mechanisms. FEBS Lett , 382:186–8.

Jones, D.T., Taylor, W.R. and Thornton, J.M. 1992. The rapid generation of mutation data matrices from protein sequences. Comp. Appl. Biosci.,8:275–82.

Karlsson, M., Nygren, K. and Johannesson, H. 2008. The evolution of the  pheromonal signal system and its potential role for reproductiveisolation in heterothallic Neurospora. Mol. Biol. Evol., 25:168–78.

Kasuga, T., White, T.J. and Taylor, J.W. 2002. Estimation of nucleotide sub-stitution rates in eurotiomycete fungi. Mol. Biol. Evol., 19:2318–24.

Kellis, M., Birren, B.W. and Lander, E.S. 2004. Proof and evolutionaryanalysis of ancient genome duplication in the yeast Saccharomycescerevisiae. Nature, 428:617–24.

Keyhani, N.O. and Roseman, S. 1999. Physiological aspects of chitincatabolism in marine bacteria.  Biochim. Biophys. Acta. Gen. Sub.,1473:108–22.

Klemsdal, S.S., Clarke, J.H.L, Hoell, I.A. et al. 2006. Molecular cloning,characterization, and expression studies of a novel chitinase gene(ech30) from the mycoparasite Trichoderma atroviride strain P1.

 FEMS Microbiol. Lett., 256:282–9.

Page 12: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 12/14

58

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

Kuranda, M.J. and Robbins, P.W. 1991. Chitinase is required for cell-separation during growth of Saccharomyces cerevisiae. J. Biol. Chem.,266:19758–67.

Limon, M.C., Margolles-Clark, E., Benitez, T. et al. 2001. Addition of substrate-binding domains increases substrate-binding capacity andspecific activity of a chitinase from Trichoderma harzianum. FEMS 

Microbiol. Lett., 198:57–63.Magliani, W., Conti, S., Gerloni, M. et al. 1997. Yeast killer systems. Clin.

Microbiol. Rev., 10:369–400.Marchler-Bauer, A., Anderson, J.B., Cherukuri, P.F. et al. 2005. CDD: a

conserved domain database for protein classification. Nucleic Acids

 Res., 33:D192–6.Padovan, A.C.B., Sanson, G.F.O, Brunstein, A. et al. 2005. Fungi evolution

revisited: Application of the penalized likelihood method to a bayes-ian fungal phylogeny provides a new perspective on phylogeneticrelationships and divergence dates of ascomycota groups.   J. Mol.

 Evol., 60:726–35.Sanderson, M.J. 2003. Molecular data from 27 proteins do not support a

Precambrian origin of land plants. Am. J. Bot., 90:954–956.Seidl, V., Huemer, B., Seiboth, B. et al. 2005. A complete survey of 

Trichoderma chitinases reveals three distinct subgroups of family 18chitinases. FEBS J., 272:5923–39.

Stark, M.J.R., Boyd, A., Mileham, A.J. et al. 1990. The plasmid-encodedkiller system of Kluyveromyces lactis—a review. Yeast , 6:1–29.

Suzuki, K., Taiyoji, M., Sugawara, N. et al. 1999. The third chitinase gene

(chiC) of Serratia marcescens 2170 and the relationship of its productto other bacterial chitinases. Biochem. J., 343:587–96.

Taylor, J.W. and Berbee, M.L. 2006. Dating divergences in the fungal treeof life: review and new analyses. Mycologia, 98:838–49.

Taylor, T.N., Hass, H., Kerp, H. et al. 2005. Perithecial ascomycetes fromthe 400 million year old Rhynie chert: an example of ancestral poly-morphism (vol 96, pg 1403, 2004). Mycologia, 97:269–85.

Terwisscha van Scheltinga, A.C.T, Kalk, K.H., Beintema, J.J. et al. 1994.Crystal-structures of hevamine, a plant defense protein with chitinaseand lysozyme activity, and its complex with an inhibitor. Structure,2:1181–9.

Thompson, J.D., Gibson, T.J., Plewniak, F. et al. 1997. The CLUSTAL Xwindows interface:flexible strategies for multiple sequence alignmentaided by quality analysis tools. Nucleic Acids Res., 25:4876–82.

Tif fin, P. 2004. Comparative evolutionary histories of chitinase genes in thegenus Zea and family Poaceae. Genetics, 167:1331–40.

van Aalten, D.M.F, Komander, D., Synstad, B. et al. 2001. Structural insightsinto the catalytic mechanism of a family 18 exo-chitinase. Proc. Natl.

 Acad. Sci. U.S.A., 98:8979–84.van Aalten, D.M.F, Synstad, B., Brurberg, M.B. et al. 2000. Structure of a

two-domain chitotriosidase from Serratia marcescens at 1.9-angstromresolution. Proc. Natl. Acad. Sci. U.S.A., 97:5842–7.

Wapinski, I., Pfeffer, A., Friedman, N. et al. 2007. Natural history and evo-lutionary principles of gene duplication in fungi. Nature, 449:54–61.

Warren, R.A.J. 1996. Microbial hydrolysis of polysaccharides.  Ann. Rev.

Microbiol., 50:183–212.Wattanalai, R., Boucias, D.G. and Tartar, A. 2004. Chitinase gene of the

dimorphic mycopathogen, Nomuraea rileyi.   J. Invertebr. Pathol.,85:54–57.

Zdobnov, E.M. and Apweiler, R. 2001. InterProScan—an integration  platform for the signature-recognition methods in InterPro. Bioinformatics, 17:847–8.

Page 13: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 13/14

59

Evolution of fungal chitinases

Evolutionary Bioinformatics 2008:4

Comparative Evolutionary Histories of the Fungal ChitinaseGene Family Reveal Non-Random Size Expansionsand Contractions due to Adaptive Natural Selection

Magnus Karlsson and Jan Stenlid

Department of Forest Mycology and Pathology, Swedish University of Agricultural Sciences, P.O. 7026,SE-75007, Uppsala, Sweden.

Supplemental InformationTree File S1. Phylogenetic trees with estimates of divergence times used in the current study.

Tree 1. Phylogenetic tree and estimates of divergence time in millions of years for all included fungi based on the Pezizomycotina calibration.((((((((N_crassa:200,M_grisea:200):10,(H_jecorina:200,G_zeae:200):10):190,(((C_immitis:38,U_ reesii:38):119,A_capsulatum:157):63,E_nidulans:220):180):200,((S_cerevisiae:145,C_ albicans:145):115,Y_lipolytica:260):340):50,S_pombe:650):70,(((P_chrysosporium:240,C_cinerea:240):245,F_ neoformans:485):60,U_maydis:545):175):72,R_oryzae:792):101,B_dendrobatidis:893);

Tree 2. Phylogenetic tree and estimates of divergence time in millions of years for a subset of theincluded fungi based on the Ascomycota calibration.(((((((N_crassa:120,M_grisea:120):10,(H_jecorina:120,G_zeae:120):10):90,(C_ immitis:130,E_nidulans:130):180):90,((S_cerevisiae:90,C_albicans:90):180,Y_lipolytica:270):90):140,S_  pombe:390):60,(((P_chrysosporium:150,C_cinerea:150):150,F_neoformans:300):40,U_ maydis:340):110):30,R_oryzae:480);

Tree 3. Phylogenetic tree a nd estimates of divergence time in millions of years for a subset of theincluded fungi based on the Pezizomycotina calibration.(((((((N_crassa:200,M_grisea:200):10,(H_jecorina:200,G_zeae:200):10):190,(C_ immitis:220,E_nidulans:220):180):200,((S_cerevisiae:145,C_albicans:145):115,Y_lipolytica:260):340):50,S_ 

pombe:650):70,(((P_chrysosporium:240,C_cinerea:240):245,F_ neoformans:485):60,U_ 

maydis:545):175):72,R_oryzae:792);Tree 4. Phylogenetic tree and estimates of divergence time in millions of years for a subset of theincluded fungi based on the Sordariomycete calibration.(((((((N_crassa:400,M_grisea:400):10,(H_jecorina:400,G_zeae:400):10):290,(C_ immitis:420,E_nidulans:420):280):480,((S_cerevisiae:280,C_albicans:280):590,Y_lipolytica:870):310):130,S_  pombe:1310):160,(((P_chrysosporium:500,C_cinerea:500):480,F_ neoformans:980):120,U_ maydis:1100):370):130,R_oryzae:1600);

Page 14: quitinase

8/6/2019 quitinase

http://slidepdf.com/reader/full/quitinase 14/14

60

Karlsson and stenlid

Evolutionary Bioinformatics 2008:4

   T  a   b   l  e   S   2 .

   N  o  n  -  r  a  n   d  o  m

   l  y  e  v  o   l  v   i  n  g   b  r  a  n  c   h  e  s   i  n   t   h  e   f  u  n  g  a   l   G   H   1   8  g  e  n  e   f  a  m   i   l  y  u  s   i  n

  g  a   l   t  e  r  n  a   t   i  v  e  c  a   l   i   b  r  a   t   i  o  n  p  o   i  n   t

  s   f  o  r   d  a   t   i  n  g   f  u  n  g  a   l

   d   i  v  e  r  g  e  n  c  e  s .

   D  a   t  a  s  e   t

   B  r  a  n  c   h   I   D

   S  o  r   d  a  r   i  o  m

  y  c  e   t  e  s

   1

   L   i   k  e   l   i   h  o  o   d

  r  a   t   i  o

   4

   P  e  z   i  z  o  m  y  c  o   t   i  n  a

   2

   L   i   k  e   l   i   h  o  o   d

  r  a   t   i  o

   4

   A  s  c  o  m

  y  c  o   t  a   3

   L   i   k  e   l   i   h  o  o   d

  r  a   t   i  o

   4

     p  -  v  a   l  u  e

   4

     p  -  v  a

   l  u  e

   4

     p  -  v  a

   l  u  e

   4

   A   l   l   G   H   1   8  g  e  n  e  s

   1

        0 .   0   0   1

        0 .   0   0   1

   P  e  z   i  z  o  m  y  c  o   t   i  n  a

  -

  -

   0 .   0   0   7

   2   2

   0 .   0   0   7

   2   2

   E  m  e  r   i  c  e   l   l  a

  -

  -

   0 .   0   0   6

   1   4

   0 .   0   0   6

   1   4

   R   h   i  z  o  p  u  s

  -

  -

   0 .   0   2   4

   1

   0 .   0   2   4

   1

   S

  c   h   i  z  o  s  a  c  c   h  a  r  o  m  y  c  e  s

  -

  -

   0 .   0   2   3

   3

   0 .   0   2   3

   3

   C   l  u  s   t  e  r   A   G   H   1   8  g  e  n  e  s

   1

   0 .   0   0   2

   0 .   0   0   2

   P  e  z   i  z  o  m  y  c  o   t   i  n  a

  -

  -

   0 .   0   0   7

   1   6

   0 .   0   0   7

   1   6

   G   i   b   b  e  r  e   l   l  a

  -

  -

   0 .   0   0   1

   5

   0 .   0   0   6

   5

   E  m  e  r   i  c  e   l   l  a

  -

  -

   0 .   0   0   4

   3   9

        0 .   0   0   1

   3   9

   R   h   i  z  o  p  u  s

  -

  -

   0 .   0   4   3

   1

   0 .   0   4   3

   1

   C   l  u  s   t  e  r   B   G   H   1   8  g  e  n  e  s

   1

   0 .   4   0   7

   0 .   0   3   3

   H  y  p  o  c  r  e  a

  -

  -

   0 .   0   0   2

   1   4

        0 .   0   0   1

   9   8

   R   h   i  z  o  p  u  s

  -

  -

   0 .   0   1   8

   1

   0 .   0   0   5

   2

   1   E  s   t   i  m  a   t  e  s  o   f   d   i  v  e  r  g  e  n  c  e   d  a   t  e  s  c  a   l   i   b  r  a   t  e   d  w   i   t   h   P .

   d  e  v  o  n   i  c  u  s  a  s  r  e  p  r  e  s  e  n   t   i  n  g

   S  o  r   d  a  r   i  o  m  y  c  e   t  e  s ,  g   i  v   i  n  g  a  n  e  s   t   i  m  a   t  e   f  o  r   t   h  e

   f  u  n  g  a   l  p   h  y   l  u  m   a

   t   1   6   3   0  m   i   l   l   i  o  n  s  o   f  y  e  a  r  s   (   T  a

  y   l  o  r  a  n   d   B  e  r   b  e  e ,

   2   0   0   6   ) .

   2   E  s   t   i  m  a   t  e  s  o   f   d   i  v  e  r  g  e  n  c  e   d  a   t  e  s  c  a   l   i   b  r  a   t  e   d  w   i   t   h   P .

   d  e  v  o  n   i  c  u  s  a  s  r  e  p  r  e  s  e  n   t   i  n  g

   P  e  z   i  z  o  m  y  c  o   t   i  n  a ,  g   i  v   i  n  g  a  n  e  s   t   i  m  a   t  e   f  o  r   t   h  e   f  u  n  g  a   l  p   h  y   l  u  m   a

   t   9   2   3  m   i   l   l   i  o  n  s  o   f  y  e  a  r  s   (   T  a  y   l  o  r  a  n   d   B  e  r   b  e  e ,

   2   0   0   6   ) .

   3   E  s   t   i  m  a   t  e  s  o   f   d   i  v  e  r  g  e  n  c  e   d  a   t  e  s  c  a   l   i   b  r  a   t  e   d  w   i   t   h   P .

   d  e  v  o  n   i  c  u  s  a  s  r  e  p  r  e  s  e  n   t   i  n  g   A  s  c  o  m  y  c  o   t  a ,  g   i  v   i  n  g  a  n  e  s   t   i  m  a   t  e   f  o  r   t   h  e   f  u  n  g

  a   l  p   h  y   l  u  m   a

   t   4   9   5  m   i   l   l   i  o  n  s  o   f  y  e  a  r  s   (   T  a  y   l  o  r  a

  n   d   B  e  r   b  e  e ,

   2   0   0   6   ) .

   4   D  e   B   i  e ,

   T . ,

   C  r   i  s   t   i  a  n   i  n   i ,   N . ,

   D  e  m  u   t

   h ,

   J .   P .  e   t  a   l .   2   0   0   6 .

   C   A   F   E  :  a  c  o  m  p  u   t  a   t   i  o  n  a   l   t

  o  o   l   f  o  r   t   h  e  s   t  u   d  y  o   f  g  e  n  e   f  a  m   i   l  y  e  v  o   l  u   t   i  o  n .   B

   i  o   i  n   f  o  r  m  a   t   i  c  s ,

   2   2  :   1   2   6   9  –   7   1 .

   A   b   b  r  e  v   i  a   t   i  o  n  :  -  :  n  o  s   i  g  n   i         fi  c  a  n   t   G   H

   1   8  g  e  n  e   f  a  m   i   l  y  e  x  p  a  n  s   i  o  n  o  r  c  o  n   t  r  a  c   t   i  o  n .