Introduction

 

The Ganoderma fungi are responsible for considerable yield losses in oil palm (Elaeis guineensis), coconut (Cocos nucifera), rubber (Hevea Brasiliensis) and palmyra palm (Borassus flabellifer) (Chee 2005; Sankaran et al. 2005; Kandan et al. 2010; Bejo and Vong 2014). Initially, Ganoderma disease had been thought to be caused by the Ganoderma boninense alone; however, in recent studies, several species have been reported to be responsible for the basal stem rot disease in oil palm trees, namely, G. zonatum, G. boninense, G. sinense and G. miniatocinctum (Zhao et al. 2007; Rashid et al. 2014). G. lucidum and G. applanatum were the most aggressive pathogen that caused the basal stem rot in coconut trees (Bhaskaran 2000; Rajendran et al. 2009). G. boninense also occurs on coconut trees, reportedly due to a saprophyte (Pilotti et al. 2004). On the other hand, G. pseudoferreum and G. philiplii are the most common species to infect rubber trees (Zakaria et al. 2009; Ogbebor et al. 2010).

Pilotti (2005) reported the physiological characteristics of the G. boninense causing basal stem rot in the oil palm. The genetic diversity of G. boninense has been reported based on species diversity and genetic heterogeneity in oil palm (Miller et al. 1999; Pilotti et al. 2003). The basidiospores of G. boninense implicated in distribution and genetic diversity have been found to cause basal stem rot (BSR) and upper stem rot (USR) (Rees et al. 2012). In particular, the availability of new molecular biology tools has made it possible to develop a large number of genetic markers. Microsatellites marker remains the most widely and powerful used tool for studying population genetics and the geographical spread of plant pathogens (Schoebel et al. 2013a, b). Purba et al. (2019) have reported thirteen species of G. boninense in oil palm in Indonesia, which they discovered through a sequence analysis. Moreover, Hapuarachchi et al. (2019) also reported twenty species of G. sinense in China found through sequence analysis.

Furthermore, the bioinformatics method provided more information on genetics (Pop and Salzberg 2008; Horner et al. 2009). G. boninense has a role in the diversity and the physical and chemical characteristics that are distinguishable by amino acids (Basyuni et al. 2018a). Moreover, Zhu et al. (2015) have reported more than 30 genes of a cluster of G. sinense, which they classified using the bioinformatics approach. This study extended our previous work and aimed to determine the genetic diversity within G. Boninense populations in some palm trees through sequence analysis and the bioinformatics method.

 

Materials ad Methods

 

Sample collection

 

A total of 49 isolates of the Ganoderma sp. belong to Socfindo derived from C. nucifera (3), H. brasiliensis (6), E. guineensis (37), and B. flabellifer (3) were collected in some parts of Indonesia (Fig. 1), as follows: Bah-Lias, Simalungun (SL) 3°15' N 99°18' E, Faculty Agricultural University Sumatera Utara (FA) 3°33' N 98˚ 39' E, Gambus Land-Batu Bara (GL) 3°09' N 99°26' E, Sirandorung-Central Tapanuli (CP) 2°02' N 98°21' E, South Sulawesi (SSul) 8°32' S 120°11' E, Dolok-Batu Bara (DB) 3°12' N 99°29' E, Bangun Bandar-Serdang Bedagai (BB) 3°15' N 99°41' E, Sei Dadap-Asahan (SD) 2°57' N 99°41' E, North Labuhan Batu (LB) 2°28' N 99°53' E, and South Sumatera (SSum) 3°19' S 104°00' E.

 

Growth of G. boninense in the PDA

 

The fungi were maintained on Potato Dextrose Agar (PDA) and the mycelial growth for a minimum of ten days’ subculture (Nasreen et al. 2005).

 

DNA extraction

 

Total DNA was extracted using cetyl trimethyl ammonium bromide (CTAB) method (Brandfass and Karlovsky 2008). The quantity DNA test was carried out following the nanophotometer method (Gallagher and Desjardins 2006) and the quality test with agarose gel (1–2%) and quantified using the UV-Tex method (Lee et al. 2012).

 

Polymerase chain reaction

 

A set of specific primers SSR selected is shown in Table 1 (Mercière et al. 2015). The amplification reaction for the PCR product was performed in 10 μL of a total volume containing 3 μL of DNA templates mixed with 2.5 μL Gotaq master, 0.5 μL primer forward and 0.5 μL primer reverse, and 3.5 μL ddH2O. The amplification was operated for 35 cycles (30 sec at 95°C 30 sec at 60°C, 40 sec at 78°C) and after the final to extensions of 8 min at 72°C. The PCR product was analyzed with the electrophoresis agarose gel stained with GelRed® and visualized with Ultraviolet Translumination (Voytas 2000).

 

Microsatellite analysis

 

The polymorphism data for each population and locus was assessed by calculating the mean for allele frequency correlations between individuals in the subpopulation (Fis), allele frequency correlations between subpopulations (Fst), allele frequencies in the population caused by both factors (Fit), the total of migrant (Nm), the number of different alleles (N), the number of different alleles frequency > 0.5% (Na), the number of active alleles (Ne) and the Shannon of information Index (I).

The genetic polymorphism for each population and locus was assessed by calculating the mean observed heterozygosity (Ho), expected heterozygosity (He), and the fixation index (F) (Nei 1978). The polymorphic information content (PIC) was determined by Avval (2017). Genetic structure analyzed was calculated using the analysis of molecular variance (AMOVA) package GenAlEx ver. 6.502 (Peakal and Smouse 2012).

 

DNA sequencing analysis

 

The polymerase chain reaction (PCR) products were purified and sequenced. The nucleotides to confirm our microsatellite data were selected, and other sequences, which are available in the NCBI database (https://blast.ncbi.nlm.nih.gov/) at BLASTX program (Altschul et al. 1997), were searched for sequence similarity. The sequence of the comparable database with BLASTX score were an E-value <10-4, considered to have significant similarities (Bohnert et al. 2001).

 

Physical and chemical characteristics

 

The structure and physico-chemical characteristics of selected data on G. sinense were analyzed analysis with Protparam after an online search (web.expasy.org/protparam/). The calculated factors designated of the length of genes, molecular weight, theoretical isoelectric point values, total number of atoms, extinction coefficients, half-life period, instability coefficient, aliphatic index, and grand average of hydropathicity (Basyuni et al. 2018b).

 

Possible peptide transfer and subcellular localization

 

The transit peptide of selected data on G. sinense was analyzed by accessing the online P1.1 target server (www.cbs.dtu.dk/services/targetp/). This position corresponds to the estimated presence of one of the chloroplast pre-sequential N-terminal terminals of transit peptide (cTP), mitochondrial targeting peptide (mTP), and also the peptide signal (SP) peptide pathway. The prediction tool for subcellular localization proteins (PSORT) was used to access online predictions with psort.hgc.jp/form.html to control the protein-induced of subcellular determinations (Basyuni and Wati 2017).

 

Phylogenetic analysis

 

Table 1: The primer specific for G. boninense

 

Primer

Primer Sequences (‘5-‘3)

Amplification (bp)

KT124397

F: CGCCATGCCCACCACCAGAG

R: GACCCGGCTGCCCGAATGAG

283-325

KT124403

F: GGCGACGAGGGCACGAGAGA

R: CCGCACTTTCGCCAACCACC

293-297

KT124399

F: GCACAGGCACAAGCGCAAGG

R: CGACGACCGCCCCAAAGGAT

204-267

KT124394

F: CGGGAAGTGGTGAACGGT

R: GGGTGGCTTGACAGCGGCAT

234-243

 

 

Fig. 1: Map of location of Ganoderma sampling

 

Phylogenetic analysis was done based on the location and grouping analysis of the phylogenetic tree of G. boninense strains NJ3. To extend our knowledge on the relationship of Ganoderma, a total of 12 locations and 49 samples were selected. Furthermore, they were analyzed using the Unweighted Pair Group Method with Arithmetic Mean (UPGMA) by MVSP ver. 3.22 software (Basyuni et al. 2018c). Fig. 2 was constructed based on the length of the base pair nucleotide of the location Ganoderma isolates. On the other hand, Fig. 3 was constructed based on the DNA sequences that were obtained using the FASTA ver 3.4t26 software (Pearson and Lipman 1988) from the Bank of Japan Data DNA (Mishima, Shizuoka, Japan); this was carried out by CLUSTAL W ver 1.83 program (Thompson et al. 1994) based on the neighbor-join method. The bootstrap analysis with 1000 replications was used to assess the strength of nodes (Felsenstein 1985).

 

Result

 

Genetic structure

 

Forty nine samples were investigated for their locus (KT124397, KT124403, KT124399 and KT124394), and the additional descriptions of microsatellite loci have reported the success rate of the specific primers. After an interpretation of the genetic variation and the interaction of locus frequency between alleles, the mean of allele frequency in collaboration with individual alleles was found to be 0.57. The mean frequencies of the entire Ganoderma population differing in alleles were suspected to be Fit 0.66.

On the other hand, the test aimed at the relationship between heterogeneity per locus. The summary of the genetic differentiation between groups obtained was Fst 0.21. It has been suggested that the value depends on the allele frequency at loci, and it exhibited a variety of properties related to genetic diversity. The estimated number of migrants obtained on average was Nm 0.97. The mean value of the polymorphic information content (PIC) was 0.441, which was the lowest value of the polymorphism allele on the loci KT124394 (0.172) and the loci KT124399 (0.372), while the other two loci have PIC values more than 50% (Table 3).

The microsatellites observed in six populations showed the value of alleles (N) to be 6.75 with differences in alleles in the frequency of 5%. This was found to be 7.21 (Na), and the number of active alleles (Ne) was found to be 6.38. The highest population variation index found in basal stem rot, E. guineensis (BSR – EG), was 2.82 (I). Furthermore, the lowest value in basal stem rot, H. brasiliensis (BSR – HB), was 1.03. This value was the same as the genetic diversity (He). The value of heterozygosity (Ho) showed the number of genes or allele distribution in the population and locus, and this value was lower than the expected heterozygosity (He). The mean of Ho was found to be 0.75, higher than that of Ho 0.32. The fixation index value in the upper stem rot, E. guineensis (USR – EG) was found to be 0.92 (F), and the lowest value in the population of rubber wood block (RWB) was found to be 0.19; however, the value of heterozygosity had different comparisons (Table 2).

We observed statistically significant differences in variation between individual isolates, and the expression differences between individuals increased the variation. Molecular of variance analysis (AMOVA) summarized based on the characteristics of the genetic diversity revealed 85% among individuals and 15% within the individual (Table 4). It was not the genetic diversity detection in the population.

 

Distribution of the BLASTX

 

The BLASTX of the 40 sequences analysis indicated previously as in G. boninense had a closing similarity as Transcriptional factor proteins to G. sinense. However, only eight sequences were identified as G. sinense strain ZZ0214-1 (Table 5). The overall results of that species with varied E-value ranges, and the total scores obtained and identified were 49.3–97.4 and 64–96%, respectively. The predictive approaches were other fungal proteins such as hypothetical proteins for Hebeloma cylindrosporum, Stylophora pistillata, Bostrobasidium botryosum, Aspergillus cristatus, Mixia osmundae, Pseudogymnoascus sp, Lomentospora prolificans, Trametes versicolor, Trichoderma gamsii, and Acidomyces richmondensis, with different strains. However, an identified hypothetical protein, G. sinense, only had a total score of 33.9 with a large E-value of seven despite identifying a 68% similarity. On the other hand, the proteins were other sequences, including Alpha/Beta hydrolase for Leucosporidium creatinivorum, Probable to GDP/GTP for Ustilago hordei, Utp-14 domain for Trametes coccinea, putative glicyne-tRNA ligase for Tolypocladium paradoxum, Neurexin-3b for Rhinocerous sinocyclocheilus, and non-identity of sequence analysis for 11 samples.

 

 

Fig. 2: Cluster analysis of Ganoderma pathogen from 12 locations. BB = Bangun Bandar

DB = Dolok-Batu Bara, CP = Central Tapanuli, SSum = South Sumatera, SD = Sei Dadap-Asahan, LB = Labuhan Batu, GL = Gambus Land, SL = Bah Lias Simalungun, SSul = South Sulawesi, FA = Faculty of Agricultural USU. BF = Borassus flabellifer, EG = Elaeis guineensis, HB = Hevea brasiliensis, RWB = rubber wood block

 

 

Fig. 3: Phylogenetic tree of Ganoderma isolates. DNA sequences were collected by the neighbor-joining method, the Clustel W, with scale indication corresponding to 0.1; the DNA sequences substitution were per site and indicated bootstrap 1000 replicates. BF = Borassus flabellifer, EG = Elaeis guineensis, HB = Hevea brasiliensis, CN = Cocos nucifera. RWB = rubber wood block. G. boninense data of sequence obtained from our previous research (Purba et al., 2019)

 

Physical and chemical properties of the G. sinense

 

The bioinformatics information of the Ganoderma sequence genes for G. sinense, including several physical and chemical parameters are described in Table 6. The database identified eight population sequences of G. sinense. It has the value of the length from 252 – 264 bp. The molecular weight was shown in the range of value 21,118 to 22,508. The theoretical isoelectric point values were 5.09 to 5.24, suggesting it has a minor variation. B. flabellifer (BF) and H. brasiliensis (HB) have a stable total number of atoms. The highest extinction coefficient in E. guineensis (EG) was 6000, with the smallest half-life period (1.2 h). Instability coefficient has been established for isolates of G. sinense from 59.29 to 78.11. The aliphatic index EG was significantly higher than that of BF, HB, CN, and RWB. The grand average of hydropathicity was unstable from 0.900 to 1.085.

 

Potential transit of peptide and subcellular localization of G. sinense

 

The DNA spread into the chloroplast transit peptide (cTP) Text Box: Table 2: The profile of microsatellite loci for all population of Ganoderma isolates

Population	N	Na	Ne	I	Ho	He	F
BSR-CN	2.25 ± 0.25	3.25 ± 0.25	2.90 ± 0.23	1.11 ± 0.07	0.46 ± 0.04	0.65 ± 0.02	0.29 ± 0.09
BSR-BF	2.75 ± 0.25	3.25 ± 0.25	3.07 ± 0.19	1.14 ± 0.06	0.21 ± 0.13	0.67 ± 0.02	0.69 ± 0.20
BSR-EG	19.50 ± 1.85	18.25 ± 1.31	15.44 ± 0.91	2.82 ± 0.06	0.10 ± 0.05	0.94 ± 0.00	0.90 ± 0.05
USR-EG	10.75 ± 0.25	11.00 ± 0.41	10.22 ± 0.44	2.36 ± 0.04	0.07 ± 0.02	0.90 ± 0.00	0.92 ± 0.03
BSR-HB	2.25 ± 0.25	3.00 ± 0.41	2.73 ± 0.33	1.03 ± 0.13	0.33 ± 0.12	0.62 ± 0.05	0.49 ± 0.19
RWB	3.00 ± 0.00	4.50 ± 0.29	3.90 ± 0.37	1.42 ± 0.08	0.75 ± 0.16	0.74 ± 0.03	0.00 ± 0.19
Mean	6.75 ± 1.37	7.21 ± 1.20	6.38 ± 1.02	1.65 ± 0.15	0.32 ± 0.06	0.75 ± 0.03	0.54 ± 0.09
Number of different alleles (N), number of different alleles frequency > 0.5% (Na), number of effectife alleles (Ne) and shannon of information index (I), observed heterozigosity (Ho), expected heterozigosity (He), fixation index (F). BSR – CN (basal stem rot – Cocos nucifera), BSR – BF (basal stem rot – Borassus flabellifer), BSR – EG (basal stem rot – Elaeis guineensis), USR – EG (upper stem rot – E. guineensis), BSR – HB (basal stem rot – Hevea brasiliensis), RWB = rubber wood block

Table 3: F-Statistics and Estimates of Nm overall population for each locus

Locus	Fis	Fit	Fst	Nm	   PIC
KT124397	0.65	0.71	0.19	1.08	   0.61
KT124403	0.57	0.66	0.19	1.07	   0.61
KT124399	0.61	0.70	0.23	0.85	   0.37
KT124394	0.47	0.59	0.23	0.86	   0.17
Mean	0.57	0.66	0.21	0.97	   0.44
SE	0.04	0.03	0.01	0.06	
Allele frequency correlations between the individuals in the subpopulation (Fis), allele frequency correlations between subpopulation (Fst), allele frequency in the population caused by both factors above (Fit) and the total number of migran (Nm), polimorfic information content (PIC)

Table 4: Summary analysis of molecular variance (AMOVA)

Source	Df	SS	MS	Est. var.	% Var
Among pops	5	18.00	3.60	0.00	0
Among indiv	43	152.42	3.55	1.63	85
Within indiv	49	14	0.29	0.29	15
Total	97	184.42		1.92	100
Df = degree of freedom, SS = source of variation, MS = mean squares, Est. Var = estimation of variant, Var = varian
and mitochondria transit peptide (mTP), was described in Table 7. Estimation of the total number of different proteins occupant in the chloroplast varied among the G. sinense of EG, RWB, and CN except BF and HB, which had a value of 0.119. Table 8 showed the most significant subcellular localization of G. sinense found in the cytoplasm (Cyto) and mytochondria (Myto), wherein it reported in Cyto 2.0 to 8.5 with Myto 2.0 to 6.0.

 

Phylogenetic analysis

 

We identified the genetic variation from six populations of BSR – CN (basal stem rot – C. nucifera), BSR – BF (basal stem rot – B. flabellifer), BSR – EG (basal stem rot – E. guineensis), USR – EG (upper stem rot – E. guineensis), BSR – HB (basal stem rot – H. Brasiliensis) and RWB (rubber wood block). It showed 12 locations in Fig. 2. The dendrogram (UPGMA) showed two large groups in the population. The first group consisting of location was Bangun Bandar and Dolok – Batu Bara. The second group consisted mainly only eight sequences from Central Tapanuli, South Sumatra, Sei Dadap – Asahan, Labuhan Batu, Gambus Land, Bah Lias – Simalungun, South Sulawesi and the Faculty of Agriculture. To extend our knowledge of the relationship between G. boninense and G. sinense, 13 sequences were used to draw the dendrogram. Eight relationships of G. sinense (Rubber Wood Block, E. guineensis, C. nucifera, H. brasiliensis, B. flabellifer, C. nucifera) were selected to be displayed as a dendrogram. Fig. 3 showed two large groups separated from G. boninense strain NJ3 with original isolates from the E. guineensis, but G. sinense strain ZZ0214-1 from various plant sources have a close genetic diversity. They were identified as G. sinense scattered over two form the same group.

 

Discussion

 

The genetic analysis of the population structure of Indonesian species of Ganoderma supported the taxonomic distinction among the isolates studied, and five distinct species were identified. The allele of frequency can be classified according to the Buchert et al. (1997), which can be divided into four categories. The alleles with a frequency of ≥ 0.75 at particular loci were categorized as high, and alleles with a frequency of 0.75 > P ≥ 0.25 were categorized as medium. The alleles of frequency 0.25 > P ≥ 0.01 were taken as low alleles and frequencies < 0.01 as rare alleles.

On the other hand, according to Marshall and Brown (1975), the alleles categorized were as general alleles if they had a frequency ≥ 0.05 and special alleles if the frequency was < 0.05. The population and loci of Ganoderma have a mean frequency with value 0.54. According to Mercière et al. (2015), and our SSR data for the PIC value showed high polymorphic information content (PIC) with only two loci, KT124397 and KT12440. Both had a value of 0.61. The specific primers were capable of producing several alleles in several genotypes in specific loci.

Table 5: Distribution of the BLASTX for Ganoderma isolates of sequence gene Accession

 

Accession

Description

Sequences

Identify (%)

Total score

E-value

PIL35715.1

Transcription factor (Ganoderma sinense ZZ0214-1)

8

64-96

49.3-97.4

(1e-04)- (9e-23)

KIM39402.1

Hypothetical protein M413DRAFT_29551 (Hebeloma cylindrosporum h7)

1

62

33.5

2.5

ORY85318.1

Alpha/Beta hydrolase protein (Leucosporidium creatinivorum)

1

33

32

9.4

PFX24695.1

Hypothetical protein AWC38_SpisGene10691 (Stylophora pistillata)

2

46

37.7

0.59-0.61

D05693.1

Hypothetical protein BOTBODRAFT_182311 (Botryobasidium botryosum FD-172SS1)

1

35

118

9e-28

CCF48258.1

Probable to GDP/GTP exchange factor Rom2p (Ustilago hordei)

1

55

32.3

6.8

SD00840.1

Utp14-domain-containing protein (Trametes coccinea BRFM310)

1

58

34.3

4.2

ODM20124.1

Hypothetical protein (Aspergillus cristatus)

1

33

35.8

1.7

XP_014568041.1

Hypothetical protein L969DRAFT_48519 (Mixia osmundae IAM 14324)

1

32

35

2.7

PIL26871.1

Hypothetical protein GSI_11051 (Ganoderma sinense ZZ0214-1)

1

68

33.9

7

POR37917.1

Putative glycine tRNA ligase (Tolypocladium paradoxum)

1

71

58.2

2e-8

F47594.1

Hypothetical protein 495_01898 (Pseudogymnoascus sp. M F 4514 FW-929)

1

72

31.6

6.8

PKS09315.1

Hypothetical protein 003929 (Lemontospora prolificans)

1

54

34.7

4.5

XP_008042932.1

Hypothetical protein TRAVEDRAFT 74270 (Trametes versicolor FP 101664 SS1)

1

40

32

8.1

XO_024400.1

Hypothetical protein TGAM01_v203499 (Trichoderma gamsii)

1

51

35.4

6.9

XP_016393967.1

Neurexin-3b-like (Sinocyclocheilus rhinocerous)

1

46

36.6

3.9

KYG40648.1

Hypothetical protein M433DRAFT_160121 (Acidomyces richmondensis BFW)

2

48

32

4.4 - 4.6

KFY47594.1

Hypothetical protein V495_01898 (Pseudogymnoascus sp. VKM F-4514 FW-929)

2

78

35-37

0.08 - 0.45

KYG406113.1

Hypothetical protein M433DRAFT_8635 (Acidomyces richmondensis BFW)

1

48

32

4.4

UNKNOWN

Not Detection

11

nd

nd

nd

 

Table 6: Physical and chemical properties of Ganoderma sinense

 

Plant spesies

BF

EG

HB

RWB

RWB

RWB

CN

EG

Length of genes/bp

252

264

257

253

255

252

255

260

Molecular weight

21118

22508

21297

21357

21491

21206

21544

22037

Theoretical isoelectric point values

5.24

5.22

5.23

5.24

5.22

5.24

5.21

5.09

Total number of atoms

2582

2753

2582

2609

2619

2596

2612

2725

Extinction coefficient

5500

6000

5875

5625

5750

5500

6125

5250

Half-life period

4.4h

1.2h

4.4h

4.4h

1.2h

4.4h

1.2h

7.2h

Instability coefficient

62.57

77.17

66.14

64.12

66.15

59.29

78.11

65.82

Aliphatic index

21.83

21.59

20.62

20.55

18.43

21.83

20.78

19.62

Grand average of hydropathicity

1.054

1.085

1.074

1.042

0.998

1.052

1.126

0.900

BF = Borassus flabellifer, EG = Elaeis guineensis, HB = Hevea brasiliensis, RWB = rubber wood  block, CN = Cocos nucifera

The genetic diversity of the G. sinense isolates from different sources in this study provided information on C. nucifera, H. brasiliensis, E. guineensis, and B. flabellifer plants, as they were similar strains of the G. sinense. The nucleotide sequences from specific regions depicted the phylogeny at various taxonomic levels. All populations were detected as randomly grouped to this study, and they do not cause geographical distribution. The isolates from the North Sulawesi province have a genetic relationship close to the North Sumatra Province.

Furthermore, the sequence analysis with BLASTX does not reflect the numerous same on G. boninense. However, the conjunction morphology among the different isolates was similar. The comparison of the genes sequences indicated some translocations and duplications among two G. sinense and G. boninense fungal species and also revealed many non-homology genes. It suggested rapid divergent evolution consequent speciations. Zhu et al. (2015) reported the genome sequence of G. sinense, one of the most valuable medicinal fungi such as G. lucidum.

The newly identified conserved DNA of G. boninense in E. guineensis have been predicted by the homology with some diverse plant groups (Purba et al. 2019). The phylogeny is indicated as G. boninense. They did not cluster together with the identified G. sinense. The present study implied high intra species diversity or the presence of the gene. The data generated in this study could be helpful for researchers in studying the conservation of both species. Mercière et al. (2015) reported that the mean value identified for N was eight for isolates from Indonesia. It was not much different in this study. The phylogenetic analysis of the SSR sequence separated the Ganoderma isolates. It showed several isolates maybe had been misclassification (Smith and Sivasithamparam 2000). The bioinformatics method of the DNA identification isolate the family G. sinense for the species research community is reported.

We have identified the physical and chemical properties conserving the DNA belonging to the G. sinense sequences (strain ZZ0214-1, GenBank Accession number PIL35715.1; PIL26871.1). Experiments confirm the powerfulness of bioinformatic prediction of DNA. These findings will be helpful to comprehend the DNA life processes in the Ganoderma pathogen. Recently, G. lucidum and G. sinense were registered as Lingzhi in Chinese Pharmacopoeia, which are used to relieve cough and dyspnea (Liu et al. 2009). Many studies reported G. lucidum as an antitumor activity, but few reports were only focused on G. sinense (Lin and Zhang 2004; Grace et al. 2006; Nonaka et al. 2006; Cheng et al. 2007; Pang et al. 2007).

Table 7: Possibility of the potential transit peptide of Ganoderma sinense

 

Nucleotide ID

Reliability

Chloroplast transit peptide

Mitochondrial target peptide

Signal peptide of secretory pathway

Reliability prediction

BF

0.119

0.089

0.057

3

EG

0.074

0.109

0.088

3

HB

0.119

0.116

0.052

3

RWB

0.102

0.080

0.082

3

RWB

0.110

0.075

0.079

4

RWB

CN

EG

0.085

0.055

0.096

0.090

0.069

0.097

0.091

0.379

0.072

3

5

3

BF = Borassus flabellifer, EG = Elaeis guineensis, HB = Hevea brasiliensis, RWB = rubber wood block, CN = Cocos nucifera

 

Table 8: Subcellular localization of the predicted Ganoderma sinense

 

Protein ID

Golg

Cyto

Plasm

Perox

Cyto-Mito

Cyto- Nucl

Myto

Vac

ER

Nucl

Secr

BF

nd

8.5

nd

nd

6.6

12.3

3.5

nd

nd

14

1.0

EG

nd

8.5

nd

2.0

nd

5.5

3.0

2.0

5.0

1.5

5.0

HB

2.0

5.5

1.0

9.0

nd

6.0

4.0

nd

nd

5.5

nd

RWB

nd

8.0

2.0

nd

nd

nd

6.0

nd

nd

4.0

7.0

RWB

nd

6.0

4.0

nd

nd

nd

4.0

nd

nd

9.0

4.0

CN

EG

nd

nd

2.0

6.0

1.0

1.0

nd

nd

nd

nd

nd

nd

2.0

2.0

4.0

3.0

1.0

nd

nd

1.0

17

14

BF = Borassus flabellifer, EG = Elaeis guineensis, HB = Hevea brasiliensis, RWB = rubber wood block, CN = Cocos nucifera, Golg = golgi body, Cyto = cytoplasm, Plasm = plasma membrane, Perox = peroxixom, Cyto – Myto = cytoplasm – mythochondria, Cyto – Nucl = cytoplasm – nuclear, Myto = mythochondria, Vac = vacuolar, ER = endoplasmic reticulum, Nucl = nucleolar, Secr = secretary

 

Eight G. sinense genes were assessed using bioinformatics analysis, and the clusters were expected to produce in chloroplast transit peptide (cTP) or mitochondrial transit peptide (mTP) and were identified from this fungus. An upregulation of ferredoxin may reflect its direct participation in pathogen defense wherein the chlorophyll of several photosynthetic proteins is affected by the plant-fungi interaction in the chloroplast (Konishi et al. 2001; Castillejo et al. 2004). Bioactive compounds such as the Ganoderic Acid T (GA-T) were predicted in the mitochondrial of G. lucidum, wherein a triterpenoid exerted cytotoxicity on various human carcinoma cell (Tang et al. 2006).

 

Conclusion

 

Thirteen G. boninense strains were chosen in the present study in order to identify eight G. sinense strains clustered into two groups. It was deduced that G. sinense had genetic diversity in four loci. The data obtained in this study demonstrated SSR is very sensitive and practical tool to identify the Ganoderma.

 

Acknowledgment

 

A Master Education towards Doctoral Research supported a part of this work [No. 152/SP2H/LT/DPRM/2018] with approval from the Directorate for Research and Community Service, Ministry of Research, Technology and Higher Education, Republic of Indonesia.

 

References

 

Altschul SF, TL Madden, AA Schäffer, J Zhang, Z Zhang, W Miller, DJ Lipman (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucl Acids Res 25:3389–3402

Avval SE (2017). Assessing polymorphism information content (PIC) using SSR molecular markers on local species of Citrullus colocynthis. Case study: Iran, Sistan-Balouchestan province. J Mol Biol Res 7:42–49

Basyuni M, R Wati (2017). Bioinformatics analysis of the oxidosqualene cyclase gene and the amino acid sequence in mangrove plants. J Phys Conf Ser 801; Article 012011

Basyuni M, R Wati, H Sagami, H Oku, S Baba (2018a). Bioinformatics approach of three partial polyprenol reductase genes in Kandelia obovata. J Phys Conf Ser 978; Article 012044

Basyuni M, A Purba, LAP Putri, R Hayati, D Chalil, I Syahputra (2018b). Bioinformatics analysis of predicted Ganoderma boninense from oil palm (Elaeis guineensis). J Phys Conf Ser 1235; Article 012071

Basyuni M, R Wati, I Deni, AR Tia, ES Siregar, I Syahputra (2018c). Cluster analysis of polyisoprenoid in oil palm (Elaeis guineensis) leaves in different land-uses to find the possible cause of yield gap from planting materials. Biodiversitas 19:1492–1501

Bejo KS, CN Vong (2014). Detection of basal stem rot (BSR) infected oil palm tree using laser scanning data. Agric Sci Proc 2:156–164

Bhaskaran R (2000). Management of Basal Stem Rot Disease of Coconut Caused by Ganoderma Iucidum, pp: 121–128. Ganoderma Diseases of Perennial Crops, CABI Publishing, London, UK

Bohnert HJ, P Ayoubi, C Borchert, RA Bressan, RL Burnap, JC Cushman, MA Cushman, M Deyholos, R Fischer, DW Galbraith, PM Hasegawa, M Jenks, S Kawasaki, H Koiwa, SK Eda, BH Lee, CB Michalowski, E Misawa, M Nomura, N Ozturk, B Postier, R Prade, CP Song, Y Tanaka, H Wang, JK Zhu (2001). A genomics approach towards salt stress tolerance. Plant Physiol Biochem 39:295–311

Brandfass C, P Karlovsky (2008). Upscaled CTAB-based DNA extraction and real-time PCR assays for Fusarium culmorum and F. graminearum DNA in plant material with reduced sampling error. Intl J Mol Sci 9:2306–2321

Buchert GP, OP Rajora, JV Hood, BP Dancik (1997). Effects of harvesting on genetic diversity in old growth eastern white pine in Ontario. Conserv Biol 11:747–758

Castillejo MA, N Amiour, ED Gaudot, D Rubiales, MA Castillejo, N Amiour, E Dumas-Gaudot, E Rubiales, JV Jorrı́n (2004). A proteomic approach to studying plant response to crenate broomrape (Orobanche crenata) in pea (Pisum sativum). Phytochemistry 65:1817–1828

Chee BJ (2005). Ganoderma boninense: Fungus with blood sugar lowering property. J Trop Med Plants 6:159-164

Cheng KC, HC Huang, JH Chen, JW Hsu, HC Cheng, CH Ou, WB Yang, ST Chen, CH Wong, HF Juan (2007). Ganoderma lucidum polysaccharides in human monocytic leukemia cells: from gene expression to network construction. BMC Genomics 9:411-428

Felsenstein J (1985). Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39:783–791

Gallagher SR, PR Desjardins (2006). Quantitation of DNA and RNA with absorption and fluorescence spectroscopy. Curr Protoc Mol Biol 93:1-21

Grace GL, KP Fung, GM Tse, PC Leung, CB Lau (2006). Comparative studies of various Ganoderma species and their different parts with regard to their antitumor and immunomodulating activities in vitro. J Altern Compl Med 12:777–789

Hapuarachchi KK, SC Karunarathna, EHC McKenzie, XL Wu, P Kakumyan, KD Hyde, TC Wen (2019). High phenotypic plasticity of Ganoderma sinense (Ganodermataceae, Polyporales) in China. Asian J Mycol 2:1–47

Horner DS, G Pavesi, T Castrignano, PDOD Meo, S Liuni, M Sammeth, E Picardi, G Pesole (2009). Bioinformatics approaches for genomics and post genomics applications of next-generation sequencing. Brief Bioinform 11:181–197

Kandan A, R Bhaskaran, R Samiyappan (2010). Ganoderma–a basal stem rot disease of coconut palm in south Asia and Asia pacific regions. Arch Phypathol 43:1445–1449

Konishi H, H Ishiguro, S Komatsu (2001). A proteomics approach towards understanding blast fungus infection of rice grown under different levels of nitrogen fertilization. Proteomics 1:1162–1171

Lee PY, J Costumbrado, CY Hsu, YH Kim (2012). Agarose gel electrophoresis for the separation of DNA fragments. J Visual Exp 62; Article e3923

Lin ZB, HN Zhang (2004). Anti-tumor and immunoregulatory activities of Ganoderma lucidum and its posssible mechanisms. Acta Pharmacol Sin 25:1387–1395

Liu YW, JL Gao, J Guan, ZM Qian, K Feng, SP Li (2009). Evaluation of antiproliferative activities and action mechanisms of extracts from two species of Ganoderma on tumor cell lines. J Agric Food Chem 57:3087–3093

Marshall DR, AHD Brown (1975). Optimum Sampling Strategies in Genetic Conservation, pp: 53–80. Cambridge University Press, Cambridge, UK

Mercière M, A Laybats, CC Lacombe, JS Tan, C Klopp, TD Gasselin, SSRS Alwee, LC Kulandaivelu, F Breton (2015). Identification and development of new polymorphic microsatellite markers using genome assembly for Ganoderma boninense, causal agent of oil palm basal stem rot disease. Mycol Progr 14:103-114

Miller RNG, M Holderness, PD Bridge, GF Chung, MH Zakaria (1999). Genetic diversity of Ganoderma in oil palm plantings. Plant Pathol 48:595–603

Nasreen Z, T Kausar, M Nadeem, R Bajwa (2005). Study of different growth parameters in Ganoderma lucidum. Micol Appl Intl 17:5–8

Nei M (1978). Estimation of average heterozygosity and genetic distance from a small number of individuals. Genetics 89:583–590

Nonaka Y, H Shibata, M Nakai, H Kurihara, H Ishibashi, Y Kiso, T Tanaka, H Yamaguchi, S Abe (2006). Anti-tumor activities of the antlered form of Ganoderma lucidum in allogeneic and syngeneic tumor-bearing mice. Biosci Biotechnol Biochem 70:2028–2034

Ogbebor N, A Adekunle, N Eghafona, A Ogboghodo (2010). Ganoderma psuedoferreum: biological control possibilities with microorganisms isolated from soils of rubber plantations in Nigeria. Afr J Agric Res 6:301–305

Pang X, Z Chen, X Gao, W Liu, M Slavin, W Yao, LL Yu (2007). Potential of a novel polysaccharide preparation (GLPP) from Anhui-grown Ganoderma lucidum in tumor treatment and immunostimulation. J Food Sci 72:435–442

Peakall R, PE Smouse (2012). GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research-an update. Bioinformatics 28:2537-2539

Pearson WR, DJ Lipman (1988). Improved tools for biological sequence comparison. Proc Natl Acad Sci USA 85:2444–2448

Pilotti CA (2005). Stem rots of oil palm caused by Ganoderma boninense: Pathogen biology and epidemiology. Mycopathologia 159:129–137

Pilotti CA, FR Sanderson, EA Aitken, W Armstrong (2004). Morphological variation and host range of two Ganoderma species from Papua New Guinea. Mycopathologia 158:251–265

Pilotti CA, FR Sanderson, EAB Aitken (2003). Genetic structure of a population of Ganoderma boninense on oil palm. Plant Pathol 52:455–463

Pop M, SL Salzberg (2008). Bioinformatics challenges of new sequencing technology. Trends Genet 24:142–149

Purba A, M Basyuni, LAP Putri, D Chalil, R Hayati, D Arifiyanto, I Syahputra (2019). Sequence analysis of Ganoderma boninense isolates from oil palm. IOP Conf Ser Earth Environ Sci 260; Article 012172

Rajendran L, A Kandan, G Karthikeyan, T Raguchander, R Samiyappan (2009). Early detection of Ganoderma causing basal stem rot disease in coconut plantations. J Oil Palm Res 21:627–635

Rashid M, MR Choon-Fah, J Bong, A Khairulmazmi, AS Idris (2014). Genetic and morphological diversity of Ganoderma species isolated from infected oil palms (Elaeis guineensis). Intl J Agric Biol 16:691-699

Rees RW, J Flood, Y Hasan, MA Wills, RM Cooper (2012). Ganoderma boninense basidiospores in oil palm plantations: evaluation of their possible role in stem rots of Elaeis guineensis. Plant Pathol 61:567–578

Sankaran KV, PD Bridge, C Gokulapalan (2005). Ganoderma diseases of perennial crops in India–an overview. Mycopathologia 159:143–152

Schoebel CN, S Brodbeck, D Buehler, C Cornejo, J Gajurel, H Hartikainen, D Keller, M Leys, S Ricanova, G Segelbacher, S Werth, D Csencsics (2013a). Lessons learned from microsatellite development for nonmodel organisms using 454 pyrosequencing. J Evol Biol 26:600–611

Schoebel CN, E Jung, S Prospero (2013b). Development of new polymorphic microsatellite markers for three closely related plant-pathogenic Phytophthora species using 454-pyrosequencing and their potential applications. Phytopathology 103:1020–1027

Smith BJ, K Sivasithamparam (2000). Internal transcribed spacer ribosomal DNA sequence of five species of Ganoderma from Australia. Mycol Res 104:943–951

Tang W, JW Liu, WM Zhao, DZ Wei, JJ Zhong (2006). Ganoderic acid T from Ganoderma lucidum mycelia induces mitochondria mediated apoptosis in lung cancer cells. Life Sci 80:205–211

Thompson JD, DG Higgins, TJ Gibson (1994). CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucl Acids Res 22:4673–4680

Voytas D (2000). Agarose gel electrophoresis. Curr Protoc Mol Biol 51:2–5

Zakaria L, NS Ali, B Salleh, M Zakaria (2009). Molecular analysis of Ganoderma species from different hosts in Peninsula Malaysia. J Biol Sci 9:12–20

Zhao M, W Liang, D Zhang, N Wang, C Wang, Y Pan (2007). Cloning and characterization of squalene synthase (SQS) gene from Ganoderma lucidum. J Microbiol Biotechnol 17:1106–1112

Zhu Y, J Xu, C Sun, S Zhou, H Xu, DR Nelson, J Qian, J Song, H Luo, L Xiang, Y Li, Z Xu, A Ji, L Wang, S Lu, A Hayward, W Sun, X Li, DC Schwartz, Y Wang, S Chen (2015). Chromosome-level genome map provides insights into diverse defense mechanisms in the medicinal fungus Ganoderma sinense. Sci Rep 5; Article 11087