From green to red: horizontal gene transfer of phycoerythrin gene ...

1 downloads 0 Views 3MB Size Report
Aug 30, 2013 - The red and green appearance of Planktothrix strains is associated ... Our results show that all eight Planktothrix genomes are highly similar, ...
AEM Accepts, published online ahead of print on 30 August 2013 Appl. Environ. Microbiol. doi:10.1128/AEM.01455-13 Copyright © 2013, American Society for Microbiology. All Rights Reserved.

1

From green to red: horizontal gene transfer of phycoerythrin gene cluster between

2

Planktothrix strains

3

Running title: Transfer of phycoerythrin gene cluster

4

Ave Tooming-Klunderud1#, Hanne Sogge1,2, Trine Ballestad Rounge1,3, Alexander J.

5

Nederbragt1, Karin Lagesen1, Gernot Glöckner4,5, Paul Hayes6, Thomas Rohrlack7,8, Kjetill S.

6

Jakobsen1,2#

7

(1) University of Oslo, Centre for Ecological and Evolutionary Synthesis (CEES),

8

Department of Biosciences, P.O.Box 1066, 0316 Oslo, Norway

9

(2) University of Oslo, Microbial Evolution Research Group (MERG), Department of

10

Biosciences, P.O.Box, 1066, 0316 Oslo, Norway

11

(3) Cancer Registry of Norway, P.O. box 5313 Majorstuen, N-0304 Oslo, Norway

12

(4) Institute for Biochemistry I, Medical Faculty, University of Cologne, Joseph-Stelzmann-

13

Straße 52; D-50931 Köln, Germany

14

(5) Leibniz-Institute of Freshwater Ecology and Inland Fisheries, IGB, Müggelseedamm 301;

15

D-12587 Berlin, Germany

16

(6) Faculty of Science, University of Portsmouth, St Michael’s Building, White Swan Road,

17

Portsmouth, PO1 2DT, United Kingdom

18

(7) NIVA, Norwegian Institute for Water Research, 0411 Oslo, Norway

19

(8) Norwegian University of Life Sciences, Department of Plant and Environmental Sciences,

20

P.O.Box 5003, 1432 Ås, Norway

21

#

Corresponding authors:

1

22

Kjetill S. Jakobsen: [email protected]

23

Ave Tooming-Klunderud: [email protected]

24

Keywords: Cyanobacteria, Planktothrix, horizontal gene transfer, phycoerythrin gene cluster,

25

chemotype, recombination, phycobilisome

26 27

Abstract:

28

Horizontal gene transfer is common in cyanobacteria and transfer of large gene clusters may

29

lead to acquisition of new functions and conceivably niche adaption. In the present study, we

30

demonstrate that horizontal gene transfer between closely related Planktothrix strains can

31

explain the production of the same oligopeptide isoforms by strains of different color.

32

Comparison of the genomes of eight Planktothrix strains revealed that strains producing the

33

same oligopeptide isoforms are closely related, regardless of color. We have investigated

34

genes involved in the synthesis of the photosynthetic pigments phycocyanin and

35

phycoerythrin, which are responsible for green and red appearance, respectively. Sequence

36

comparisons suggest the transfer of a functional phycoerythrin gene cluster generating a red

37

phenotype in a strain that is otherwise more closely related to green strains. Our data show

38

that the insertion of a DNA fragment containing the 19.7 kb phycoerythrin gene cluster has

39

been facilitated by homologous recombination, also replacing a region of the phycocyanin

40

operon. These findings demonstrate that large DNA fragments spanning entire functional

41

gene clusters can be effectively transferred between closely related cyanobacterial strains and

42

result in a changed phenotype. Further, the results shed new light on the discussion of the role

43

of horizontal gene transfer in the sporadic distribution of large gene clusters in cyanobacteria,

44

as well as the appearance of red and green pigmented strains.

2

45 46

Introduction

47

Horizontal gene transfer (HGT), the exchange of genetic information between two

48

organisms that do not share a recent ancestor–descendant relationship, is now recognized as a

49

major force shaping the evolutionary history of prokaryotes (e.g. [1-4]). HGT is considered to

50

be common in cyanobacteria [5]. Through the availability of bacterial genome sequences it

51

has become clear that HGT can occur throughout the genome, and that a substantial fraction

52

of genes have been horizontally transferred [5,6]. The quantity of genetic material that can be

53

horizontally transferred may range from small gene fragments (e.g. [7-9]) to fragments

54

spanning complete genes (e.g. [10-12]) and whole operons encoding complex biochemical

55

pathways (e.g. [13-15]). As even the transfer of a single or a few genes can give recipient

56

organisms the opportunity to implement a new function and exploit new ecological niches,

57

HGT contributes to the rapid creation of biological novelty that otherwise, through mutations

58

and gene duplications, might have taken millions of years to appear.

59

According to Andam and co-workers [1], HGT is the norm and not the exception,

60

while others call the transfer of genes between bacteria ‘both rare and promiscuous’[4].

61

Successful HGT depends on transfer of genetic material to the cell (via transformation,

62

conjugation, transduction or gene transfer agents), survival of the DNA in the cell, integration

63

of foreign DNA via recombination and finally fixation of the integrated DNA in the

64

population (involving for example selection). Since the rate of recombination decreases with

65

increased sequence dissimilarity [16,17], HGT events are more common among close

66

relatives, as shown by a recent analysis of 657 sequenced prokaryotic genomes [18].

67

For fixation of a newly transferred gene in the population, it should provide a relevant

68

function and this function must operate within the native machinery of the host cell. Since

3

69

bacterial genomes are subject to deletional bias [19], genes that do not contribute to fitness of

70

the organism will eventually be removed from the genome. Integration of new genes into

71

existing cellular networks can be facilitated by acquisition of an operon containing all genes

72

and regulatory regions required for function [20]. For single-gene acquisitions, the fate of new

73

genes depends largely upon the existing genes in its new host. Experimental studies have

74

shown that most HGT events are deleterious [21,22]. However, rare HGT events and

75

mutations can be selected for under particular conditions and are thus contribute to bacterial

76

adaptation and evolution [23-25].

77

Horizontal gene transfer events have also been demonstrated for the filamentous

78

cyanobacterium Planktothrix (e.g., [26-29]), which occurs in deep and stratified lakes in

79

temperate regions of the northern hemisphere. Traditionally, Planktothrix isolated from

80

different lakes have been classified into species according to morphological characteristics,

81

such as cell dimension and pigmentation. Following the first description of the genus

82

Planktothrix including 14 distinct species by Anagnostidis and Komárek [30], the number of

83

different species has been heavily disputed. Studies based on molecular data like sequences of

84

gas vesicle genes and 16S rRNA have suggested that whole Planktothrix genus is

85

monospecific [31,32] while Suda and co-workers [33] described four Planktothrix species

86

based on several genetic and phenotypic properties.

87

Planktothrix strains isolated from Norwegian lakes and classified as distinct species at

88

Algal Culture Collection of the Norwegian Institute for Water Research cannot be separated

89

by 16S rRNA. Recently, Rohrlack and co-workers [34,35] reported that strains of

90

Planktothrix showing >99% 16S rRNA gene sequence similarity may produce distinct

91

cellular patterns of oligopeptides, bioactive secondary metabolites synthesized mostly by non-

92

ribosomal peptide synthetases. Using the oligopeptide profiles produced by each strain as

93

markers, they grouped strains into distinct chemotypes (Cht). Based on field studies of the 4

94

Norwegian Lake Steinsfjorden, four coexisting Planktothrix chemotypes differing

95

considerably in seasonal dynamics, depth distribution and participation in loss processes,

96

were identified [34]. Since the production of oligopeptides is facilitated by several large and

97

independently evolving operons [36,37], strains associated with a distinct chemotype are

98

assumed to be more closely related. This hypothesis is also supported by data showing that

99

Planktothrix strains associated with same chemotype generally have same color, either red or

100

green [35]. However, in Lake Steinsfjorden, one chemotype was shown to comprise both red

101

and green strains [34,35]. The red and green appearance of Planktothrix strains is associated

102

with the content of accessory light harvesting pigments, the phycobiliproteins, involved in the

103

photoautotrophic machinery. Phycobilisomes, the macromolecular complexes formed from

104

phycobiliproteins, have an allophycocyanin core that links to the photosystems, and

105

peripheral light harvesting rods that comprise either phycocyanin or phycocyanin and

106

phycoerythrin (for review, see e.g. [38]). Phycocyanin, common to all cyanobacteria, imparts

107

a green appearance to the cell and absorbs red light (620-630 nm). Phycoerythrin absorbs

108

green light (560-570 nm) and imparts a dominant red color when present. The co-existence of

109

red and green strains within the same chemotype can be explained by acquisition or loss of

110

genes coding for phycoerythrin as suggested earlier for Synechococcus and other

111

picocyanobacteria [39-41].

112

The aim of this study was to investigate the genome arrangements leading to the co-

113

occurrence of red and green strains within the same oligopeptide chemotype. For that

114

purpose, the genomes of eight different Planktothrix strains classified as four different species

115

were sequenced, four red and four green strains, including a red and two green strains from

116

the same chemotype. We address the following questions: (1) how similar are the genomes of

117

closely related Planktothrix strains and is there any evidence for genetic substructuring

118

according to color or chemotype; (2) is the structure and chromosomal location of genes 5

119

encoding phycocyanin and phycoerythrin pigments the same in all strains; (3) in the light of

120

results from the two first questions, can the co-occurrence of red and green strains within the

121

same chemotype be explained by altered phycoerythrin genes and is this because of a) an

122

acquisition of phycoerythrin gene cluster by the red strain, or b) mutations leading to non-

123

functional phycoerythrin genes in two green strains.

124

Our results show that all eight Planktothrix genomes are highly similar, and that

125

strains associated with the same chemotype are the most closely related, regardless of color.

126

Furthermore, we reveal that a red strain from a chemotype dominated by green strains has

127

acquired the 19.7 kb phycoerythrin gene cluster. Our data indicate that the DNA fragment

128

containing phycoerythrin operon originated from a strain associated with a “red” chemotype.

129 130

Materials and Methods

131

Planktothrix strains and DNA isolation

132

Eight Planktothrix strains isolated from lakes Steinsfjorden and Kolbotnvatnet

133

(Norway) have been investigated. All strains have been kept in continuous, non-axenic

134

cultures at the Algal Culture Collection of the Norwegian Institute for Water Research

135

(NIVA), in Z8 medium and light at a photon flux density of 10 μmol m-2s-1, and a light-dark

136

cycle of 12:12 hours. Prior to genomic DNA isolation, cells were centrifuged and re-

137

suspended in TE buffer. DNA was extracted by the following procedure: cells were treated

138

with lysozyme (final concentration 15 mg/ml), followed by RNaseA and Proteinase K, (5

139

mg/ml, 1 % LDS) treatment. Samples were then incubated at 60°C (shaking at 300 rpm) for

140

60 min. In cases where the solution did not clear after 60 minutes, the incubation time was

141

prolonged with additional Proteinase K. Subsequently, 1 volume phenol: chloroform:

142

isoamylalcohol [25:24:1 (v/v)]) was added. The solution was mixed by inversion at 37°C for 6

143

30 min to remove pigments and proteins. After centrifugation, the upper layer was treated

144

twice with 1 volume chloroform–isoamyl alcohol [24:1 (v/v)]. DNA was precipitated using

145

0.1 volume 3 M sodium acetate and 2.5 volume ice-cold 96 % ethanol on ice for 1 hour. The

146

DNA pellet was washed twice with ice-cold 70 % ethanol, dried at room temperature and

147

dissolved in Tris-HCl buffer (pH 8.0). All DNA samples were purified using Amicon Ultra-

148

0.5mL 50K Centrifugal columns to ensure high-quality DNA for paired-end library

149

preparation using the 454 Life Sciences protocols. DNA was concentrated using

150

manufacture’s instructions and washed twice using Tris-HCl buffer, pH 8.0.

151

Sequencing

152

Seven out of eight Planktothrix genomes were sequenced using 454 pyrosequencing at

153

the Norwegian Sequencing Centre (http://www.sequencing.uio.no/). Strain NIVA-CYA 34

154

was initially sequenced using Sanger sequencing at Max Planck Institute for Chemical

155

Ecology. DNA sample of NIVA-CYA 34 was amplified using the REPLI-g® kit (Qiagen).

156

The resulting DNA was randomly sheared and the fragment size range from 2500 to 3000 kb

157

was selected for cloning into pUC18 standard vectors. The resulting clones were sequenced

158

from both ends with the standard sequencing primers on an ABI 3700 machine generating 20

159

066 sequencing reads comprising 16 Mb. An additional 454 shotgun library was prepared and

160

sequenced at the Norwegian Sequencing Centre to ensure satisfactory quality of the NIVA-

161

CYA 34 genome assembly. For the remaining seven strains, both shotgun and paired-end

162

libraries were prepared and sequenced to 23 – 41x coverage (see Supplementary Table 1 for

163

information about libraries and number of reads/bases sequenced for each strain).

164

Assembly and annotation

165

The newbler program (gsAssembler, Roche-454, Branford, USA) was used to

166

assemble the 454 reads into scaffolds and contigs (newbler version used for assembly of each

7

167

strain together with assembly statistics are shown in Supplementary Table 1). Since all strains

168

had been cultivated non-axenically, reads from co-cultured bacteria were present in the

169

dataset. To decrease the chance of co-assembling these contaminating reads with the genome

170

of the strain of interest, stringent overlap settings with a minimum of 98 % overlap identity

171

and a minimum of 60 bp overlap length were used. However, the assemblies of most strains

172

still contained some short scaffolds with low average read depth (below 5-10 x) and/or GC

173

percentages that diverged from high read depth scaffolds. These scaffolds were considered to

174

be derived from co-cultured bacteria present in the sample used for DNA extraction [42]. The

175

low read depth scaffolds were compared to the non-redundant NCBI protein database using

176

BLASTX, and all non-cyanobacterial matching scaffolds were removed before annotation.

177

Assembly of the NIVA-CYA 34 genome was done using both the Sanger and 454-

178

reads. Newbler was given a trimming file (-vt option) to remove pUC18 plasmid vector

179

sequences from the Sanger reads. The ‘-stopjoin’ and ‘-repfill’ options were used for

180

assembly. One low-coverage non-cyanobacterial scaffold was removed from the newbler

181

assembly. Four of the unscaffolded contigs were identified as containing cyanobacterial

182

sequences; these were added to the scaffolds before annotation.

183

According to classification of Genomes Sequence Standards [43], we consider the

184

assemblies as improved high-quality drafts. Annotation of all genome sequences was

185

performed using the U.S. Department of Energy (DOE) Joint Genome Institute (JGI)

186

integrated microbial genomes database and comparative analysis system (IMG) [44].

187

Comparison of genomes

188

The IMG system was used for pair wise comparison of genomes and calculation of

189

Pearson correlation coefficients of COG (Cluster of Orthologous Groups) profiles.

8

190

Homologous genes were defined as genes having a minimum of 80% sequence identity and

191

identified using BLASTP.

192

Hierarchical clustering by COG profiles was performed by IMG system using the tool

193

Cluster (http://www.falw.vu/~huik/cluster.htm). For construction of the pairwise hierarchical

194

tree, a function profile vector (gene counts per COG) was generated for each genome. The

195

distance metrics between these profile vectors was calculated by means of un-centered

196

correlation. Pairwise single-linkage clustering was performed, grouping the two closest

197

profile vectors first to form a group, then grouping the next pair of closest groups or vectors

198

until all genomes were incorporated into the hierarchical tree.

199

Phylogenetic analyses of genomes were performed using a set of core genes. Core

200

genes were defined as genes present in all genomes and having a minimum of 90% sequence

201

identity, resulting in data set of 3 690 genes. The final data set (after removing 100% identical

202

genes in all genomes together with transposases, oligopeptide synthetases, retron type reverse

203

transcriptases, phage-associated proteins and proteins shorter than 50 aa) contained 2 914

204

genes. Ten sub-sets of genes for generating phylogenies were created by random sampling of

205

20 genes from the core gene set (2914 genes) repeated 10 times using R (R development core

206

team 2012. R: A language and environment for statistical computing, reference index version

207

2.10.1. R Foundation for Statistical Computing, Vienna, Austria. ISBN 3-900051-07-0, URL

208

http://www.R-project.org). Sequences were aligned using Mega 5 software [45] and

209

concatenated for each of the ten data sets. List of genes for each data set is shown in

210

Supplementary Table 2. Maximum likelihood (ML) analyses were carried out using RAxML

211

[46] with 1000 bootstrapped resamplings with GTRCAT model. The resulting individual

212

phylogenetic trees are shown in Fig. S1. Finally, the aligned sequences of all 200 genes were

213

concatenated

resulting

in

217

546

bp

long

alignment

(available

through

9

214

http://dx.doi.org/10.6084/m9.figshare.719100) and Maximum likelihood (ML) analyses were

215

carried out using RAxML [46] with 1000 bootstrapped resamplings with GTRCAT model.

216

Analyses of phycobilisome genes

217

All gene sequences analysed (phycoerythrin, phycocyanin and flanking genes) were

218

downloaded from IMG. Sequences were aligned using Mega 5 software [45]. Nucleotide

219

diversity was analyzed using the computer program DNA Sequence Polymorphism (DnaSP)

220

[47]. Maximum likelihood (ML) analyses of cpc genes were carried out using RAxML [46]

221

with 1000 bootstrapped resamplings. GTRCAT was determined to be the best evolutionary

222

model using ModelTest [48].

223

Detection of recombination events

224

Recombination events were detected by visual analyses of informative sites (variable

225

sites where each variant occurs in at least two sequences) as described by Rudi and co-

226

workers [7]. In order to detect the recombination breakpoints, concatenated sequences of

227

cpcBA (reverse-complement) and CHAP domain genes were analyzed using the RDP 4

228

software [49]. Sequences from strains NIVA-CYA 406 and NIVA-CYA 15 were discarded as

229

these are identical with sequences from strains NIVA-CYA 98 and NIVA-CYA 34,

230

respectively. Recombination signals were accepted if at least three different methods detected

231

statistically significant (P