AEM Accepted Manuscript Posted Online 10 April 2015 Appl. Environ. Microbiol. doi:10.1128/AEM.00402-15 Copyright © 2015, American Society for Microbiology. All Rights Reserved.

1

Title: Validation of reference genes for transcriptional analyses in Pleurotus ostreatus using RT-qPCR.

2 3

Running title: A reference gene set for P. ostreatus RT-qPCR analyses

4 5

Authors: Raúl Castanera1&, Leticia López-Varas1&, Antonio G. Pisabarro1, Lucía Ramírez1#

6

Affiliations:

7

1

8 9 10 11

Genetics and Microbiology Research Group, Department of Agrarian Production, Public University of Navarre, Pamplona, Navarre, Spain

&

Equally contributing authors

# Corresponding author E-mail: [email protected]

12

ABSTRACT

13

Recently, the lignin-degrading basidiomycete Pleurotus ostreatus has become a widely used model for

14

fungal genomics and transcriptomics. The increasing interest in this species has led to an increasing

15

number of studies analyzing the transcriptional regulation of multigene families that encode

16

extracellular enzymes. Reverse transcription followed by real-time PCR is the most suitable technique

17

for analyzing the expression of gene sets under multiple culture conditions. In this work, we have tested

18

the suitability of 13 candidate genes for their use as reference genes in P. ostreatus time-course cultures

19

for enzyme production. We have applied three different statistical algorithms and obtained a

20

combination of stable reference genes for an optimal normalization of RT-qPCR assays. This reference

21

index can be used for future transcriptomic analyses and validations of RNA-seq or microarray data.

22

Moreover, we have analyzed the expression patterns of a laccase and a manganese peroxidase (lacc10

23

and mnp3, respectively) in lignocellulose and glucose-based media using submerged, semi-solid and

24

solid-state fermentation. By testing different normalization strategies, we demonstrate that the use of

25

non-validated reference genes as internal controls leads to biased results and misinterpretations of the

26

biological responses underlying expression changes

27 28

Key words: RT-qPCR, Reference Gene, Pleurotus ostreatus, Laccase, Manganese peroxidase,

29

Transcription

30

2

31

INTRODUCTION

32

The basidiomycete Pleurotus ostreatus is an efficient producer of extracellular enzymes such as

33

laccases (Lac; EC 1.10.3.2) and manganese peroxidases (MnP; EC 1.11.1.13), which are notable for

34

their application in multiple industrial and biotechnological processes (1, 2). It has a special relevance

35

in agroindustry due to its world-wide cultivation for mushroom production, and it has also been studied

36

for its nutritional and medicinal value (3). Currently, two genomic sequences of P. ostreatus are

37

publicly available, and it has become a very popular model for the study of fungal genetics and

38

comparative genomics (4-6). Additionally, many studies have recently been published on the

39

production of ligninolytic enzymes, with a focus on culture optimization for enzyme production. Most

40

of these enzymes are encoded by genes organized in gene families, occasionally arranged into clusters

41

as a result of gene duplications. Despite the information uncovered by sequencing efforts, there is still

42

substantial work to be performed on the functional characterization of gene family members. In this

43

sense, transcriptomic analyses are powerful tools to understand gene regulation and its presumptive

44

function. Despite the high productivity and low cost of RNA-seq, reverse transcription followed by

45

real-time PCR (RT-qPCR) is likely the best option for analyzing expression patterns of certain genes

46

under multiple conditions. Additionally, it is the most reliable technique for validating RNA-seq and

47

microarray data because of its specificity, reproducibility, and capacity for detecting and measuring tiny

48

amounts of nucleic acids in a wide quantification range (7). Since its first appearance in 1993 (8), RT-

49

qPCR has evolved into a well-established technology. Many studies have been performed to improve

50

RT-qPCR accuracy in terms of new quantification models (9), PCR efficiency calculations (10, 11) or

51

baseline determinations (12). Several types of software that implement these approaches are currently

52

available (13, 14). Within the two possible quantification strategies (relative and absolute), relative

53

quantification is a very common strategy for analyzing the expression levels of a target gene in multiple

54

conditions or samples. In this method of quantification, the mRNA levels of the target gene are

55

compared to the expression of a reference gene that passes through all the steps along with the target 3

56

gene. The use of an internal standard is essential to compensate for the variance introduced in every

57

step of the process, such as RNA quantification, reverse-transcription or pipetting (9, 15-17). The ideal

58

reference gene should have constant expression regardless of the sample, condition, cell or treatment

59

used (16, 18). Housekeeping genes such as 18S rRNA, β-actin, β-tubulin or glyceraldehyde 3-

60

phosphate dehydrogenase have been commonly used as reference genes. Unfortunately, there is an

61

overwhelming amount of studies showing that such genes are often unacceptable for RT-qPCR

62

normalization due to unstable expression (19, 20). Additionally, there is strong evidence that the use of

63

more than one reference gene is more reliable than the use of only one (7, 16, 18). The choice of

64

adequate internal standards is crucial to obtain a successful and unbiased determination of expression,

65

which is essential to analyze the expression stability of reference genes in validation assays. Several

66

methods have been proposed to analyze reference gene panels (16, 18, 20-22), and studies using these

67

methodologies in different species are becoming more frequent. In this work, we have tested the

68

suitability of 13 candidate genes for their use as reference genes in P. ostreatus using different culture

69

conditions and growing stages. We have applied three different statistical algorithms and obtained a

70

combination of reference genes that can be used for future P. ostreatus transcriptomic studies.

71

Moreover, we have analyzed the relative expression of laccase and manganese peroxidase genes

72

(lacc10 and mnp3, respectively) in lignocellulose and glucose-based media using different

73

normalization strategies. Our results demonstrate that the use of unstably expressed genes as internal

74

controls leads to biased data and misinterpretations of the biological responses underlying expression

75

changes.

76

4

77

MATHERIALS AND METHODS

78 79

Fungal strain

80

In this study, the dikaryotic strain N001 of Pleurotus ostreatus var. florida was used. The

81

genomic sequences of its monokaryotic protoclones PC15 and PC9 (23) were obtained from the JGI

82

genome portal (http://genome.jgi.doe.gov/).

83 84

Culture media and growing conditions

85

Mycelia were grown on Petri dishes with solid PDA medium (Scharlab, S.L, Barcelona, Spain)

86

and maintained in darkness at 24 °C. Inocula were prepared in Erlenmeyers containing 150 ml of SMY

87

medium (10 g of sucrose, 10 g of malt extract and 4 g of yeast extract per liter) and 4 mycelium pieces

88

of approximately 0.25 cm2. These cultures were maintained in darkness at 24 °C for 6 days in orbital

89

shaking (125 rpm) and homogenized in an omnimixer. 15 ml of the homogenized inocula were added

90

to the final media. Regarding the experimental conditions, three media were used: i) Submerged

91

Fermentation cultures (SmF) set on Erlenmeyers with 135 ml of SMY medium; ii) Semi-Solid State

92

Fermentation cultures (s-SSF) containing 10 g of milled wheat straw and 70 ml of distilled water set in

93

glass flasks with a surface to volume ratio of 1 cm-1 and iii) Solid-State Fermentation cultures (SSF) in

94

polypropylene plastic bags containing 15 g of poplar chips of approximately 200 x 50 mm and 22.5 ml

95

of distilled water. All cultures were inoculated with 15 ml of inoculum and incubated in darkness at 24

96

°C. SmF cultures were kept in orbital shaking (125 rpm), whereas s-SSF and SSF cultures were

97

incubated in static conditions.

98 99

Experimental conditions

100

To include different growth stages in our analyses, the sampling times were based on previous

101

studies analyzing variations observed during P. ostreatus growth (24). Three independent biological 5

102

replicates were sampled for each culture medium. In this sense, we sampled SmF cultures on days 1, 3,

103

7, 11 and 15; s-SSF cultures on days 3, 7, 11, 15 and 30; and SSF cultures on days 3, 7, 11, 15 and 30.

104 105

Nucleic acid extraction and real-time qPCR

106

Mycelia were harvested, frozen and ground in a sterile mortar with liquid nitrogen. Total RNA

107

was extracted from 200 mg of deep frozen tissue using Fungal RNA E.Z.N.A Kit (Omega Bio-Tek,

108

Norcross, GA, USA); its integrity was estimated by denaturing electrophoresis on 1% (w/v) agarose

109

gels. Nucleic acid concentrations were measured using a NanodropTM 2000 (Thermo Scientific,

110

Wilmington, DE, USA), and the purity of the total RNA was estimated by the 260/280 nm absorbance

111

ratio. Samples were DNase treated using 1 U of RQ1 DNase (Promega, Madison, WI, USA) per µg of

112

RNA. Subsequently, samples were purified using RNeasy MinElute Cleanup (Qiagen Iberia S.L.,

113

Madrid, Spain) and set to a final concentration of 400 ng/μl. Total RNA (800 ng per sample) was

114

reverse-transcribed into cDNA in a 20 μl volume using the iScript cDNA Synthesis kit (Bio-Rad,

115

Alcobendas, Spain). Reverse-transcription (RT) was carried out in a Thermal Cycler (MJ Research,

116

Inc.) and the protocol consisted of 5 min at 25°C, 30 min at 42°C and 5 min at 85°C.

117

RT-qPCR reactions were performed in a CFX96TM Real-Time System C1000TM (Bio-Rad

118

Laboratories, S.A.) using SYBR Green dye to detect product amplification. Primers shown in Table 1

119

were used for the amplification of the transcripts from the reference genes and the of lacc10 and mnp3

120

genes. Each reaction was set to a final volume of 20 µl containing 10 μl iQTM SYBR Green Supermix

121

(Bio-Rad Laboratories, S.A.), 2 μl of 3 μM stock forward and reverse primers (Table 1), 1 μl of a 1:20

122

dilution of RT product and 5 μl of sterile water. The amplification program consisted of 5 min at 95°C,

123

40 cycles of 15 s at 95 °C, 30 s at 63 °C and 15 s at 72 °C, followed by a final melting-curve with

124

increments of 0.5 °C every 5 s in a linear gradient of 65-95 °C. The specificity of each reaction was

125

confirmed by the inspection of melt curve profiles. Each reaction was performed in triplicate, and Non

126

Template Controls (NTCs) were included for each primer set. An experimentally validated Interplate 6

127

Calibrator (IPC) was used to compensate inter-plate variation. Crossing-point values (Cp) and RFUs

128

were recorded, and the latter were used to calculate amplification efficiencies by linear regression using

129

the LinReg program (25).

130 131

Reference gene panel

132

We selected 13 genes of different functional classes as candidates (Table 2). Our gene panel

133

contained i) housekeeping genes such as tubulin (tub), actin (actin1) or cytochrome-c, (cyt-c); ii) genes

134

chosen based on stable expression in several RNA-seq and qPCR experiments such as lipase (lip) or

135

methylthioadenosine phosphorylase (phos) (26, 27); and iii) the sar1 gene, described as the ideal

136

reference gene in the ascomycetes Aspergillus niger

137

candidates were classified according to their suitability as internal standards using the geNorm (18),

138

NormFinder (16) and BestKeeper (21) algorithms.

(28) and Trichoderma reesei (29). The 13

139 140

Data normalization and quantification strategy

141

Data pre-processing was carried out using GenEx software (http://www.multid.se) and included

142

interplate-calibration, efficiency correction, and normalization using different sets of reference genes

143

and the calculation of relative quantities (RQ, formula 1) as well as the relative expression ratios

144

(Expression, Formula 2). SmF samples harvested at day 1 were used as control samples in the latter

145

quantification.

146

Formula 1

147

Formula 2

148

where CpE is the efficiency corrected crossing point. ΔCpE in Formula 1 represents the difference

149

between the CpE of the target gene and CpE of the mean CpE of the reference genes. The expression

150

value of Formula 2 represents the classical ΔΔCt method (15).

RQ = 2 − ΔCp E

Expression = 2 − ( ΔΔ Cp E )

7

151

To analyze the impact of reference gene selection in the relative expression calculations, we

152

obtained the expression of the laccase lacc10 and the manganese-peroxidase mnp3 using different

153

normalization strategies: i) The mean expression of 12 genes, ii) the mean expression of the top three

154

reference genes from the ranking, iii) the expression of the most stable reference gene, iv) the

155

expression of the least stable reference gene, and v) the amount of total RNA (no reference genes).

156 157

Statistical analyses

158

Heatmap and hierarchical clustering were performed on autoscaled expression data using Wards'

159

algorithm and the Euclidean distance. A Mann Whitney test was used to uncover significant differences

160

in the relative expression ratios calculated using the different normalization strategies. All statistical

161

analyses were performed using GenEx.

162

8

163

RESULTS AND DISCUSSION

164 165

The expression level of the 13 genes was measured in each of the 15 experimental conditions

166

(see Materials and Methods). The mean crossing point values (Cp) ranged from 18 to 25 cycles (Fig 1).

167

The standard deviation among the 15 different conditions was clearly variable, ranging from 0.64

168

cycles in vsn to 1.66 cycles in lip. The gene chs showed the lowest expression (highest Cp value, 28.04

169

cycles), whereas actin1 showed the highest expression (lowest Cp value, 16.71 cycles). Other

170

housekeeping genes such as tub, cyt or gapdh1 also showed high levels of transcription in concordance

171

with the important roles that they play in most cellular processes.

172 173

Stability of reference genes: NormFinder, geNorm and Bestkeeper analyses.

174

The NormFinder algorithm identified sar1 as the most stable reference gene based on its lower

175

SDN (NormFinder standard deviation) in comparison to the other candidates (0.27 cycles). Interestingly,

176

the gene showing lower inter-sample variation in the exploratory analysis of Cp values (vsn gene, Fig

177

1) was ranked 10th, evidence that raw Cp inter-sample variation itself is not a reliable indicator of

178

expression stability. The accumulated standard deviation (Acc.SDN) was explored to determine the

179

optimal number of reference genes to be used. Our results showed that a reference index composed of

180

12 genes (sar1, phos, pep, lip, cyph, cyt, gapdh1, actin1, chs, vsn, pas and tub) would be the most

181

appropriate for data normalization according to its low Acc.SDN (0.23 cycles) (Fig 2B). Nevertheless,

182

the difference between this index and another composed of the best three reference genes (sar1, phos,

183

pep) was only 0.03 cycles. Thus, the use of three reference genes for normalization (0.26 Acc.SDN) is

184

reasonable and avoids the increased cost and the loss of efficiency of using 12 genes. To obtain more

185

robust results, data were analyzed again taking into account the intra and inter-group variations in gene

186

expression among the three culture mediums (Tables S1 and S2). Genes showing high inter-group

187

variation (cutoff SDN of 0.65 cycles) were discarded because including unstable genes in the analysis 9

188

would lead to a less reliable ranking of expression stability. After removing tub, vsn, pas and actin1,

189

NormFinder was run again, considering all of the samples to form one group (Fig 2C). The results

190

showed a very similar profile in comparison to the first run. The gene sar1 was again the best and was

191

classified with an SDn of 0.29 cycles, followed by phos and pep; cyc was the least stable. Nevertheless,

192

the analysis of accumulated SDn showed that in this instance, the ideal reference index was composed

193

of three genes, displaying an Acc.SDN of 0.24 cycles (Fig 2D). This deviation is useful for evaluating

194

the systematic error introduced when normalizing RT-PCR data. In our case, this value falls in the

195

range of variability that is often found between technical replicates. Thus, we can conclude that the

196

systematic error (Acc.SDN) introduced by our proposed reference index (average of sar1, phos and

197

pep) is admissible for an accurate expression analysis. To give more strength to the results, we analyzed

198

the same dataset with the geNorm and Bestkeeper programs. The geNorm tool identified the

199

combination of sar1 and pep as the most stable gene-pair (M value of 0.62), and cyc was again

200

identified as the least stable candidate (Fig 3). Because Bestkeeper can analyze up to ten genes, we

201

chose the top ten candidates from the NormFinder classification as input data (80% of them were also

202

within the top ten in geNorm ranking). According to the descriptive statistics released by Bestkeeper,

203

sar1 and phos displayed the lowest coefficient of variance (3.97% y 3.67%, respectively) and standard

204

deviation (0.80 cycles and 0.87 cycles, respectively) (Table 3). Following the authors´ considerations

205

(21), genes showing a standard deviation greater than one cycle were considered inconsistent and

206

removed from the analysis; the remainder were used in further analyses. The Bestkeeper algorithm

207

calculates the pairwise correlations between Cp values and the Bestkeeper Index (BI) based on the

208

geometric mean of stable genes (SD < 1). Thus, genes showing a higher correlation (r) and a significant

209

p value (p < 0.05) were considered the most stable. According to this, sar1 was the first to be classified

210

in the ranking (r = 0.94, p < 0.05). The high correlation between this gene and the BI undermines its

211

stable expression and thus its suitability as a reference gene.

212 10

213

Reference genes´ stability ranking

214

According to the four methodologies used, sar1 was the most stable candidate gene (Table 4).

215

The reference gene pep appeared at the top three positions in all the rankings, similar to phos, with the

216

exception of geNorm results where it occupied the 5th place. When comparing the NormFinder top six

217

reference genes with the other methodologies, five of the six genes were the same in the Bestkeeper

218

results, whereas four of the six were the same in the geNorm results. Small discrepancies between the

219

three algorithms are common (30-33), nevertheless, results are usually consistent within the best

220

classified genes, as found in this study. Interestingly, several housekeeping genes such actin1, gapdh1

221

or tub are among the less stable reference genes tested. These genes have been widely used as internal

222

standards for normalization of RT-PCR and RT-qPCR data. During the last decade, many studies have

223

revealed the instability of housekeeping genes´ expression (31, 34, 35), although they can also be

224

acceptable internal standards under certain conditions. This fact is evidence that gene expression

225

stability varies with the gene, species, tissue and culture conditions used (36, 37). In summary, our

226

results reinforce the fact that reference genes should never be chosen on the unique basis of suitability

227

in other organisms. Therefore, if no data are available relative to a given species, experimental

228

validation should always be the first step in RT-qPCR expression studies. In this sense, geNorm,

229

NormFinder and Bestkeeper are the three methods commonly used for analyzing the stability of

230

reference gene panels. Other statistical methods have been developed with the same aim (20, 22, 38)

231

but are used less frequently, most likely due to their more complex application. Bestkeeper has a

232

limitation in the number of genes that can be analyzed. This is inconvenient when attempting to

233

analyze larger (and thus more robust) reference gene panels. The geNorm tool calculates the best pair

234

of genes based on the variation with the other genes tested. In our case, its results were the most

235

discordant. In general, the three methods are susceptible to producing the wrong results when

236

introducing co-regulated genes. The choice of the most appropriate tool is a difficult task, and due to

237

the lack of a gold standard at this issue, the use of several algorithms represents a reasonable way to 11

238

proceed. Normalization of RT-qPCR data based on multiple reference genes allows for a more precise

239

evaluation of gene expression than single-gene correction (16, 18, 21). In this sense, the balance

240

between costs and benefits plays an important role in the number of genes selected for normalization.

241

The analysis of the optimal number of reference genes to be used revealed that the difference in

242

Acc.SDN is not very large when comparing three versus one gene (0.24 vs. 0.29 cycles, Fig 2D).

243

Nevertheless, the difference between using one versus three genes can be decisive in case of pipetting

244

errors while setting up the reference gene reaction. Taking into account the results obtained using the

245

three algorithms, a reference index formed by sar1, pep and phos could be considered the ideal internal

246

standard for normalizing RT-qPCR data in P. ostreatus species. Interestingly, the best reference gene

247

detected in our study was sar1, an orthologue of the sar1 GTPase of Trichoderma reesei, a species in

248

which it was described as an ideal reference gene (29). According to the conserved domains of this

249

protein, sar1 participates in important biological processes such as GTP binding, intracellular transport,

250

protein secretion and signal transduction. Thus, it seems reasonable for the organism to maintain its

251

constitutive transcription. This gene has also been validated as stable in the ascomycete Aspergillus

252

niger (28). According to its exceptional intra-group stability (SDn < 0.1 in every medium, Table S1), its

253

use as a single reference gene could be justified (although not recommended) when comparing the

254

expression of time-course samples from the same culture medium. In order to extend our results to

255

other fungal species we performed an exploratory analysis using publicly available microarray data

256

(Supplemental methods, Fig S1). The level of expression of sar1, pep and phos compared to the mean

257

expression of all genes evidenced the apparent expression stability of our candidates (especially of

258

sar1) in the fungal models Phanerochaete chrysosporium, Coprinopsis cinerea and Postia placenta.

259

These results should be taken with caution; they are not a direct validation of their general suitability

260

for normalization of RT-qPCR data, as the exploratory approach used in these fungal models lacks of

261

sufficient statistical robustness (i.e. sufficient samples of different media or developmental stages).

262

Nevertheless, according to the results obtained we propose that genes pep, phos, and specially sar1 are 12

263

good candidates to be included in RT-qPCR reference gene panels to be further validated in these

264

species.

265 266

Impact of normalization strategy on gene expression quantification.

267

To uncover the impact of the selection of reference genes on gene expression calculations, we

268

analyzed the expression of lacc10 and mnp3 in all of the samples using five normalization strategies

269

(Table 5). We obtained time-course expression ratios (Formula 2) for use as reference samples in

270

mycelium cultured on SmF medium harvested at day 1. The results showed clear evidence that the

271

normalization strategy strongly influences expression results. In fact, in some cases, the expression

272

ratios showed totally opposite patterns when normalized with validated versus non-validated reference

273

genes. This was particularly evident in the expression of both genes in s-SSF and SSF media (Fig 4,

274

Fig 5). For example, at day 15 of s-SSF culture, lacc10 showed an overexpression of 3.1 log2-fold

275

when using alternative 2 (the optimal reference index); nevertheless, the use of the single-gene cyc

276

(alternative 4) yielded a repression of -1.25 log2-fold (Fig 4). This represents a very important

277

difference between both normalizations using the same input data, which can lead to entirely incorrect

278

conclusions. In general, normalization with unstable reference indexes (alternatives three to five) led to

279

an underestimation of the expression values. As expected, the results obtained with the first two

280

alternatives yielded strong correlations of 0.978 in mnp3 and 0.995 in lacc10, which substantially

281

decreased when using indexes of lower quality (Table S3, Table S4, Supplementary information). The

282

five gene expression sets (one for each normalization strategy) were further analyzed using hierarchical

283

clustering combined with a heatmap (Fig 6). The exploratory analysis separated two main clusters

284

corresponding to the lacc10 and mnp3 genes, regardless of the normalization strategy. Within each

285

group, alternatives I and II always clustered separate from alternatives IV and V. The Mann-Whitney

286

test confirmed that there were not significant differences (p

Validation of Reference Genes for Transcriptional Analyses in Pleurotus ostreatus by Using Reverse Transcription-Quantitative PCR.

Recently, the lignin-degrading basidiomycete Pleurotus ostreatus has become a widely used model organism for fungal genomic and transcriptomic analyse...
2MB Sizes 1 Downloads 9 Views