Skip to main content

Intraspecific and interspecific variations in the synonymous codon usage in mitochondrial genomes of 8 pleurotus strains

Abstract

In this study, we investigated the codon bias of twelve mitochondrial core protein coding genes (PCGs) in eight Pleurotus strains, two of which are from the same species. The results revealed that the codons of all Pleurotus strains had a preference for ending in A/T. Furthermore, the correlation between codon base compositions and codon adaptation index (CAI), codon bias index (CBI) and frequency of optimal codons (FOP) indices was also detected, implying the influence of base composition on codon bias. The two P. ostreatus species were found to have differences in various base bias indicators. The average effective number of codons (ENC) of mitochondrial core PCGs of Pleurotus was found to be less than 35, indicating strong codon preference of mitochondrial core PCGs of Pleurotus. The neutrality plot analysis and PR2-Bias plot analysis further suggested that natural selection plays an important role in Pleurotus codon bias. Additionally, six to ten optimal codons (ΔRSCU > 0.08 and RSCU > 1) were identified in eight Pleurotus strains, with UGU and ACU being the most widely used optimal codons in Pleurotus. Finally, based on the combined mitochondrial sequence and RSCU value, the genetic relationship between different Pleurotus strains was deduced, showing large variations between them. This research has improved our understanding of synonymous codon usage characteristics and evolution of this important fungal group.

Peer Review reports

Introduction

Codon bias indicates the non-uniform or biased usage of synonymous codons that encode the same amino acid in a gene or genome [1]. The genetic information contained in DNA is transferred to the sequence of 20 amino acids through transcription and translation steps [2]. Among the 64 triplet codon arrangements contained in DNA, 61 triplets can encode 20 standard amino acids, while the other three are translation termination codons. Among the 20 amino acids encoded, 18 amino acids are encoded by multiple different codons, while tryptophan and methionine are encoded by only one codon in most species. The degeneration of the genetic code allows the same amino acid to be encoded by synonymous codons or different codons [3, 4]. However, in most cases, the probability of synonymous codons being used is not random or equal. This common phenomenon is called codon usage bias (CUB) [5,6,7]. This phenomenon of synonymous codons appearing with different frequencies is often observed in different genes, different organisms, or even the same gene from different species [8,9,10]. CUB is mainly caused by mutations in the gene coding region, especially mutations in the second or third nucleotides of the codon in the gene coding region [11,12,13]. A synonymy mutation or “silent mutation” will lead to the variability of synonymous codons in organisms during evolution [14, 15]. Since some codons are more prone to mutation than others, selection can sustain this bias [16]. As a result of GC heterogeneity and GC biased gene transformation (gBGC), codon usage bias may also be a result of local recombination rate-based codon usage bias [17,18,19]. Consequently, synonymous codons evolve through a combination of mutation, natural selection, and genetic drift of gene translation efficiency, which may play a significant role in genome evolution [20, 21]. There is a mutation mechanism that explains the interspecific differences in codon usage by explaining codon bias by the rate or repair of nucleotide bias or point mutations [22, 23]. Furthermore, the theory of natural selection assumes that synonymous mutations that affect biological adaptability will be favored or suppressed throughout the evolutionary process, leading to changes in the use of codons in genomes or genes [24, 25].

Several cellular processes can be affected by codon bias, including transcription, translation efficiency and accuracy, mRNA stability, protein expression, structure, function, and folding during cotranslation [26,27,28]. Codon bias affects transcription by altering chromatin structure and mRNA folding, which then affects translation efficiency by affecting translation elongation rate [29, 30]. Therefore, codon bias arises as the result of genome adaptation to transcription and translation mechanisms. The study of molecular evolution of genes benefits from selecting genes that do not change amino acids. The codon bias analysis can reveal evolutionary relationships between closely related organisms because codons are used similarly by closely related organisms [31, 32]. Most highly expressed proteins are encoded by genes with the best codons. With the rapid development of high-throughput sequencing technology, codon bias analysis is now essential for understanding species evolution, environmental adaptation, and genetics, etc [33,34,35]. In fungal species, particularly in large higher fungi, however, genetic characteristics of codon bias remain unknown [36].

Pleurotus is one of the largest cultivated edible fungi in the world, which has rich species diversity [37,38,39,40]. Some species of Pleurotus are delicious edible fungi, which are widely welcomed by consumers [37, 38, 41]. In addition, Pleurotus species also contain a variety of bioactive ingredients, with anti-tumor, antioxidant, anti-inflammatory, anti-virus and other effects [42,43,44,45]. Mitochondrial genome, known as the second genome of eukaryotes, plays an important role in maintaining the energy supply of eukaryotic cells [46]. Most fungi have 15 core protein coding genes (PCGs), including atp6, atp8, atp9, cob, cox1, cox2, cox3, nad1, nad2, nad3, nad4, nad4L, nad5, nad6, and rps3 [47, 48]. The variation of mitochondrial genome has an important impact on the homeostasis, stress resistance and tolerance, development of eukaryotic cells [49,50,51]. Our previous research found that the mitochondrial genomes of different Pleurotus species had undergone large-scale gene rearrangement, indicating that Pleurotus species had undergone significant genetic differentiation [52]. However, the codon bias, genetic characteristics, and evolution of the mitochondrial core PCGs of Pleurotus within and between species are still unknown.

In this study, we analyzed and compared the usage characteristics of synonymous codons of mitochondrial core PCGs within and between 8 Pleurotus strains, including P. citrinopileatus, P. cornucopiae, P. eryngii, P. giganteus, P. ostreatus P51, P. ostreatus, P. platypus, and P. pulmonarius. We also deduced the phylogenetic relationship of different Pleurotus strains based on relative synonymous codon usage (RSCU) data and compared it with the phylogenetic relationship based on mitochondrial genome sequence inference. This study is the first report to analyze the intraspecific and interspecific synonymous codon usage characteristics of important cultivated edible fungi, which will promote the understanding of the evolution, genetics, and species differentiation of Pleurotus species and other related species.

Materials and methods

Sequence processing

A total of 8 complete Pleurotus mitochondrial genomes have been published in the National Center for Biotechnology Information (NCBI) database, 2 of which were reported by our previous studies [52]. The 8 Pleurotus mitochondrial genomes were first downloaded from the NCBI database under the accession numbers NC_036998, NC_038091, NC_033533, NC_062374, OX344747, NC_009905, NC_036999, and NC_061177 [53,54,55,56,57]. We further obtained the core protein coding sequence of the mitochondrial genomes of 8 Pleurotus strains. Those core protein coding genes whose sequence length is less than 300 bp were excluded from subsequent analysis [14]. Finally, we obtained 12 core protein coding genes in each Pleurotus strains for subsequent analysis, including atp6, cob, cox1, cox2, cox3, nad1, nad2, nad3, nad4, nad5, nad6, and rps3.

Codon usage indices

The GC3s parameter is used to measure the amount of codons with guanine and cytosine at the third synonymous position, with the exception of Met, Trp, and termination codons [58]. The third base of a codon, which is often the least conserved and most variable position. The codon adaptation index (CAI) is a measure of the bias towards codons that are commonly found in highly expressed genes [59]. CAI reflects the adaptation of a gene’s codon usage to the tRNA pool of the organism, which affects translational efficiency. It is a numerical value ranging from 0 to 1.0, with larger values indicating a greater frequency of synonymous codon usage. The Codon Bias Index (CBI) is a metric for evaluating gene expression, which quantifies the deviation from a random or uniform distribution of codons encoding the same amino acid. The Frequency of Optimal Codons (FOP) is determined by dividing the amount of optimal codons by the total number of synonymous codons in a gene, which provides a direct measure of how often a gene uses the “best” or most efficiently translated codons. The Effective Number of Codons (ENC) is a measure of the number of codons used in a gene, ranging from 20 to 61. A value of 20 indicates that only one codon is used for each amino acid, while 61 indicates that each codon is used on average. A low ENC value (below 35) indicates a strong codon usage preference, while a higher value (above 35) indicates a weak preference. The Relative Synonymous Codon Usage (RSCU) value is calculated by dividing the amino acids encoded by the same codons and their probability of appearing in the same codons, which provides a direct comparison of codon usage across genes or species, accounting for differences in codon composition due to amino acid composition. A value greater than 1 indicates a positive codon bias, while a value less than 1 indicates a negative codon bias. The General Average Hydropathicity (GRAVY) value is determined by summing the hydropathy values of all of the amino acids in the polymerase gene sequences and multiplying them by the number of residues in the gene sequences, which provides insights into the potential membrane-spanning or intracellular localization of a protein. GRAVY values range from − 2 to 2, with positive and negative values representing hydrophobic and hydrophilic proteins, respectively. The Aromaticity (AROMO) value is an indicator of the frequency of aromatic amino acids (Phe, Tyr, and Trp). Aromatic amino acids have a unique chemical structure that confers stability and specific interactions with other molecules. The aromaticity of a protein can affect its structure, function, and interactions with other molecules. GRAVY and AROMO values are also indicators of amino acid usage, and changes in amino acid composition will also affect the results of codon usage analysis. All of these codon usage indicators can be calculated using CodonW1.4.2 [60] or CAIcal server [61].

Neutrality plot analysis

The neutrality plot (GC12 vs. GC3) can be used to analyze the balance between mutation and selection when codon bias is formed. GC12 represents the average GC content in the first and second positions of the codon (GC1 and GC2), while GC3 represents the GC content in the third position. Neutral evolution theory assumes that mutations occur randomly and have no effect on the fitness of the organism. However, selection pressure can introduce biases in the observed mutation frequencies, leading to deviations from neutrality [62]. A strong statistical correlation between GC12 and GC3 indicates that the species is mainly driven by mutation, whereas a lack of correlation implies the main driving force is natural selection.

ENC-GC3s plot analysis

The ENC-GC3s plot (ENC vs. GC3s) is typically employed to assess whether the codon usage of a particular gene is impacted solely by mutation or other factors, such as natural selection. This diagram consists of the ordinate ENC value and abscissa GC3s value, with an expected curve calculated via a specific formula [63]. If the corresponding points are distributed around the expected curve, mutation pressure is an independent force in the formation of codon bias. However, if the points are significantly lower or distant from the expected curve, some other factors, such as natural selection, likely play a key role in the formation of codon bias.

$${\rm{ENC}}{\;_{{\rm{exp}}}}\; = \;2\; + \;{\rm{GC}}{3{\rm{s}}}\; + \;{{29} \over {{\rm{GC}}_{\rm{s}}^2\; + \;{{{\rm{(}}1\; - \;{\rm{G}}{{\rm{C}}{\rm{s}}}{\rm{)}}}^2}}}$$

The ENCRatio value reflects the variation range between the expected value and the actual value of ENC.

$${\rm{ENC}}{\;_{{\rm{Ratio}}}}\; = \;\;{{{\rm{EN}}{{\rm{C}}_{{\rm{exp}}}}\; - \;{\rm{EN}}{{\rm{C}}_{{\rm{obs}}}}} \over {{\rm{EN}}{{\rm{C}}_{{\rm{exp}}}}}}$$

PR2-Bias plot analysis

Additionally, the Parity Rule 2 bias (PR2-Bias) plot analysis based on [A3/(A3 + U3) vs. G3/(G3 + C3)] can be utilized to determine the degree and direction of the gene bias. The center point in the plot is A = T and C = G, meaning the codon has no usage bias.

Correspondence analysis

Correspondence analysis (COA) is a widely accepted multivariate statistical analysis method used to identify codon usage patterns. All genes were placed in a 59-dimensional hyperspace, taking into account the 59 sense codons (Met and Trp excluded). This method can detect the main trends in codon usage in the core CDS of Pleurotus and arrange codons along the axis according to the RSCU value.

Determination of optimal codons

The genes were ordered from highest to lowest expression according to the ENC value, and 10% of the genes from the front and rear ends were selected to form a high- and low-expression gene dataset. The D-value between the RSCU of the two datasets (ΔRSCU) was then calculated, with ΔRSCU values greater than 0.08 being defined as codons with high expression. Codons with RSCU values greater than 1 were considered high-frequency codons. A codon with ΔRSCU > 0.08 and RSCU > 1 was defined as the optimal codon.

Phylogenetic analysis

The phylogenetic relationships of Pleurotus strains were compared between codon usage-based and mitochondrial sequence-based methods. Using the RSCU values of the 8 Pleurotus strains, SPSS v19.0 software was employed to generate a hierarchical clustering method to illustrate the relationship tree between the different species. We employed the method described in our previous studies [48, 64] to construct phylogenetic trees of the 8 Pleurotus strains using the combined mitochondrial gene datasets. To do this, individual mitochondrial genes were aligned using MAFFT v7.037 [65], and then the aligned sequences were combined into a single set using Sequence Matrix v1.7.8 [66]. Potential phylogenetic conflicts between different mitochondrial genes were identified through a partition homogeneity test. Partition Finder 2.1.1 [67] was used to determine the most suitable model of partitioning and evolution for the combined mitochondrial gene set. The phylogenetic tree was constructed using the Bayesian inference (BI) method with MrBayes v3.2.6 [68]. Two independent runs with four chains (three heated and one cold) were conducted for 2 × 106 generations, with samples taken every 100 generations. The first 25% of samples were discarded as burn-in, and the remaining trees were used to calculate Bayesian posterior probabilities (BPP) in a 50% majority-rule consensus tree. Ganoderma lingzhi was set as the outgroup [69, 70].

Results

Nucleotide composition of Pleurotus core PCGs

The codon usage analysis of 12 mitochondrial core PCGs from 8 Pleurotus strains revealed that the average length of these genes ranged from 370 bp to 2262 bp, with the nad3 gene having the shortest average length and the rps3 gene having the longest. Out of these 12 core PCGs, 10 genes had varying sequence lengths among the different Pleurotus species, while the cox2 and nad6 genes had the same gene length across all 8 strains. The rps3 gene showed the greatest length variation, with a maximum difference of 318 bp. Different Pleurotus species show great differences in base composition, even between the same species (P. ostreatus). The base composition of these 12 core PCGs was found to be rich in T base, with an average content of 41.80%, followed by A base, with an average content of 31.70%. The G and C base contents were relatively low, with an average of 13.59% and 12.92%, respectively. The average GC content of core PCGs ranged from 19.36 to 33.57%, with the rps3 gene having the lowest GC content and the cox1 gene having the highest.

Codon usage analysis

The GC1, GC2 and GC3 contents of the 12 core PCGs in the 8 Pleurotus strains were 34.23%, 34.31% and 11.08%, respectively (Fig. 1). The average GC3s value of these 12 PCGs was 9.44%, indicating that the mitochondrial core PCGs of Pleurotus tend to end with an A or T base. Additionally, the indices of A3s, T3s, G3s, and C3s of the 12 core PCGs of Pleurotus species showed that the codons were more likely to end with A, followed by T, C and G, with values of 54.80%, 54.65%, 7.53%, and 2.27%, respectively. We conducted an analysis of the codon bias of 12 core PCGs in 8 Pleurotus strains. The CAI values of the core PCGs ranged from 0.12 to 0.20, with nad2 having the lowest value and nad3 having the highest. P. giganteus had the highest CAI value, while P. citrinopileatus and P. pulmonarius had the lowest, indicating that they had a strong codon bias. The CBI values of the 8 Pleurotus strains ranged from − 0.164 to -0.173, with P. giganteus having the lowest value and P. ostreatus having the highest. The average FOP values of the 12 core PCGs ranged from 0.25 to 0.37, with nad1 having the lowest value and nad3 having the highest. P. ostreatus and P. platypus had the lowest FOP value, while P. giganteus had the largest. The GRAVY values of the 12 core PCGs were mostly positive, indicating that they were likely hydrophobic proteins, with the exception of rps3, which was considered hydrophilic. The AROMO values of the PCGs ranged from 0.08 to 0.17, with rps3 having the highest value and cox3 having the lowest. The AROMO values of the 8 Pleurotus strains were relatively similar, with an average of 0.14. The two P. ostreatus species showed differences in various base bias indicators, among which P. ostreatus P51 had high CBI, FOP, ENC, GC3s and Aromo values, while P. ostreatus had higher CAI and Gravy indicators, indicating that the frequency of base synonymous codon usage also changed within Pleurotus species.

Fig. 1
figure 1

Codon usage indicators of 12 mitochondrial core protein coding genes in different Pleurotus strains

Codon usage correlation analysis

A significant correlation was observed between the GC1 content of mitochondrial codons and GC2, GC3, GC3s, and AROMO values in all eight Pleurotus strains (P < 0.05) (Fig. 2). Furthermore, a significant correlation was found between the GC2 content and GC content and AROMO values (P < 0.05). GC3 content was significantly correlated with GC3s and GC content (P < 0.05), and it was also found to affect codon bias in two Pleurotus species (P. citrinopileatus, and P. giganteus). GC3s and GC content were significantly correlated in all eight Pleurotus strains (P < 0.05). Additionally, the GC content was found to be significantly correlated with the AROMO values in all Pleurotus strains (P < 0.01). Furthermore, the CAI index of mitochondrial codons was significantly correlated with the FOP index and CBI index in seven out of eight Pleurotus strains (P < 0.05). Lastly, a negative correlation was observed between the ENC value and GRAVY value in P. giganteus (P < 0.01).

Fig. 2
figure 2

Pearson’s correlation analysis heatmap of different codon usage indicators of 8 Pleurotus strains. The color of the color block changes from green to red, indicating that the correlation index is increasing. One asterisk indicates a significant correlation between the two indicators at the P < 0.05 level, while two asterisks indicate a significant correlation between the two indicators at the P < 0.01 level. The 8 Pleurotus species are P. citrinopileatus, P. cornucopiae, P. eryngii, P. giganteus, P. ostreatus P51, P. ostreatus, P. platypus, and P. pulmonarius, from left to right and from top to bottom

Neutrality plot analysis

We calculated the relationships between GC12 and GC3 based on neutrality plot analysis (Fig. 3). The GC12 content varied from 23.29 to 41.25%, and the GC3 content varied from 6.37 to 20.53%. The analysis between GC12 and GC3 content in mitochondrial codons of Pleurotus revealed a weak positive correlation, with the regression coefficient ranging from 0.55 to 0.95 and the R2 value ranging from 0.2219 to 0.4458. Statistical analysis showed that there was no significant correlation between GC12 and GC3 values (P > 0.05), indicating that natural selection played a major role in codon bias of Pleurotus.

Fig. 3
figure 3

Neutrality plot analysis of GC12 and the third codon position (GC3) for the entire coding DNA sequence of 8 Pleurotus strains. a, P. citrinopileatus; b, P. cornucopiae; c, P. eryngii; d, P. giganteus; e, P. ostreatus P51; f, P. ostreatus; g, P. platypus; h, P. pulmonarius

ENC-GC3s plot analysis

The average ENC value of all 12 core PCGs detected was found to be 29.86, which is lower than 35, indicating a strong codon usage preference (Fig. 1). Moreover, the ENC values of 8 Pleurotus strains ranged from 29.58 to 30.74, further confirming the strong codon usage preference of Pleurotus species. The ENC plot showed that all Pleurotus genes were below the expected ENC-plot curve (Fig. 4), indicating that factors other than mutation pressure, such as natural selection, play a role in codon bias formation.

Additionally, the ENCRatio values for all core PCGs ranged from 18.59 to 20.55%, indicating that the expected values were greater than the actual values (Fig. S1). This demonstrates that GC3s have an important influence on the formation of codon bias. In conclusion, it can be inferred that natural selection is a major factor determining the formation of Pleurotus codon bias.

Fig. 4
figure 4

ENC-GC3 plot analysis of 12 core PCGs in 8 Pleurotus strains. The solid line represents the expected curve when codon usage bias is affected only by mutation pressure. a, P. citrinopileatus; b, P. cornucopiae; c, P. eryngii; d, P. giganteus; e, P. ostreatus P51; f, P. ostreatus; g, P. platypus; h, P. pulmonarius

PR2-Bias plot analysis

We conducted a Parity Rule 2 (PR2) plot analysis to investigate whether Pleurotus mitochondrial genes have any biases (Fig. 5). Both axes were centered on 0.5 to divide the plot into four quadrants. The results showed that the third base of the mitochondrial codon of Pleurotus had a strong preference for T over A and C over G. Most of the dots were found to be distributed in the third quadrant, while six out of the eight strains were not distributed in the fourth quadrant (preferring A to T and C to G), with P. citrinopileatus and P. giganteus being the exception. All the 8 Pleurotus strains were not distributed in the first quadrant (preferring A to T and G to C). This suggests that strong codon usage preference exist in Pleurotus species.

Fig. 5
figure 5

Parity Rule 2 (PR2) plot analysis of 12 core PCGs in 8 Pleurotus strains. a, P. citrinopileatus; b, P. cornucopiae; c, P. eryngii; d, P. giganteus; e, P. ostreatus P51; f, P. ostreatus; g, P. platypus; h, P. pulmonarius

Correspondence analysis

To further analyze codon biases in Pleurotus, we conducted a correspondence analysis (COA) based on the RSCU values of mitochondrial genes from the 8 Pleurotus strains (Fig. 6). Axis 1, Axis 2, Axis 3 and Axis 4 are the main contributors to variance, with average contribution rates of 45.61%, 15.62%, 8.24% and 6.30%, respectively. The results showed that Axis 1 was the largest contributor to variance. Pearson correlation analysis showed that Axis 1 had significant correlation with CAI and ENC values. Additionally, we observed large variation in the rps3 gene and other core PCGs, indicating the differentiation of synonymous codon usage of core PCGs.

Fig. 6
figure 6

Correspondence analysis (COA) based on the relative synonymous codon usage (RSCU) values of 12 mitochondrial genes from 8 Pleurotus strains. Purple represents the cox gene, red represents the nad gene, green represents the atp6 gene, blue represents the cob gene, and yellow represents the rps3 gene. a, P. citrinopileatus; b, P. cornucopiae; c, P. eryngii; d, P. giganteus; e, P. ostreatus P51; f, P. ostreatus; g, P. platypus; h, P. pulmonarius

Optimal codon analysis

Analysis of the Relative Synonymous Codon Usage (RSCU) of eight Pleurotus strains revealed 27 high-frequency codons in six species (P. citrinopileatus, P. cornucopiae, P. eryngii, P. ostreatus P51, P. platypus and P. pulmonarius), with P. giganteus containing 26 and P. ostreatus containing 28 (Fig. 7). AUA was found to be used at a low frequency in P. citrinopileatus and P. giganteus, but used at a high frequency in other Pleurotus species. Of the 28 frequently used codons, 15 ended in T, 11 in A, and only 2 in G, indicating a preference for codons ending in A/T. In addition, 22, 15, 23, 30, 19, 28, 28, and 20 highly expressed codons (ΔRSCU > 0.08) were identified in the 8 Pleurotus strains, including P. citrinopileatus, P. cornucopiae, P. eryngii, P. giganteus, P. ostreatus P51, P. ostreatus P. platypus, and P. pulmonarius, respectively (Fig. 8). Comparative analysis revealed that 6, 6, 7, 7, 10, 9, 8, and 6 optimal codons (ΔRSCU > 0.08 and RSCU > 1) were found in P. citrinopileatus, P. cornucopiae, P. eryngii, P. giganteus, P. ostreatus P51, P. ostreatus P. platypus, and P. pulmonarius, respectively. All of these optimal codons ended with A/T, with UGU and ACU being the most widely used, followed by GGA, AUU, and UUU, which were used as the optimal codons of five strains. GCA, GCU, AAU, UCU, UAA, and ACA were each used as the optimal codons of one species. Furthermore, P. ostreatus P51 and P. ostreatus showed great differences in the use of optimal codons. AAU, GGU, AUA, UUA, UAA, and GUA were used as the optimal codons in P. ostreatus P51, while GCA, GGA, AUU, CCU, and GUU were used as the optimal codons in P. ostreatus.

Fig. 7
figure 7

Relative synonymous codon usage (RSCU) analysis of 12 mitochondrial genes from 8 Pleurotus strains. The color blocks with different colors on the bottom vertical axis represent different codons in the image above. a, P. citrinopileatus; b, P. cornucopiae; c, P. eryngii; d, P. giganteus; e, P. ostreatus P51; f, P. ostreatus; g, P. platypus; h, P. pulmonarius

Fig. 8
figure 8

Optimal codons of 8 Pleurotus strains (ΔRSCU > 0.08 and RSCU > 1), which are marked in purple. Highly expressed codons (ΔRSCU > 0.08) were marked in yellow and high-frequency codons (RSCU > 1) were marked in green

Phylogenetic analysis

The Bayesian inference (BI) method was employed to construct phylogenetic trees of 8 Pleurotus strains based on the combined mitochondrial gene set (Fig. 9). The results demonstrated that P. giganteus and P. citrinopileatus had diverged from the Pleurotus population earlier. P. cornucopiae was identified as the sister species of P. platypus. Furthermore, two P. ostreatus strains were grouped in the same evolutionary clade, which indicated their close phylogenetic relationship. In contrast to the phylogenetic relationship inferred from sequences, the species relationship inferred from RSCU had some discrepancies, such as the phylogenetic status of P. ostreatus, P. eryngii, and P. pulmonarius. Nevertheless, the RSCU-based species relationship also clearly revealed the close relationship between P. platypus and P. cornucopiae, as well as the early divergence of P. giganteus and P. citrinopileatus from the Pleurotus population.

Fig. 9
figure 9

Relationship inference of different Pleurotus strains based on the Bayesian inference (BI) (a) method and relative synonymous codon usage (RSCU) hierarchical clustering (b)

Discussion

The development of sequencing technology has enabled researchers to gain access to the genetic sequences of various species and types of genomes, including the nuclear genome, chloroplast genome and mitochondrial genome [71,72,73]. Through the analysis of genetic information, it has been observed that the usage of synonymous codons varies among different species, with some codons being used more frequently than others [74]. This codon usage bias is mainly affected by several factors, such as gene base composition, gene length, gene expression level, tRNA abundance, amino acid hydrophobicity, aromaticity, mutation, and selection, with mutation and selection being the most influential [75, 76]. Examining the codon bias characteristics of different species can help to understand the genetic structure and evolution trend of species [77, 78]. However, the codon usage of important organelle genomes of higher fungi has not been thoroughly studied.

The mitochondrial genome is often referred to as the ‘second genome’ of eukaryotes. In this study, it was found that the length and base composition of mitochondrial core PCGs of different Pleurotus strains varied significantly, even within the same Pleurotus species, indicating the differentiation of Pleurotus mitochondrial genes. The differences in synonymous codons were mainly reflected in the third codon. Additionally, it was observed that all core PCGs of Pleurotus species tend to end with A/T, which is in line with the rule of mitochondrial codon usage in many eukaryotes [79, 80]. The majority of high-frequency codons parsed by RSCU also end with A/T, further confirming the tendency of using the third codon of Pleurotus. Moreover, variations in base usage were observed among different species and genes. The two P. ostreatus species also showed differences in various base bias indicators, including CAI, CBI, FOP, ENC, and GC3s values, indicating that the frequency of base synonymous codon usage also changed in the within Pleurotus species. Furthermore, correlations were detected between codon base composition and GC3s, CAI, CBI, and FOP, suggesting the influence of base composition on codon bias. An ENC value lower than 35 indicates a strong codon preference [81, 82]. The average ENC value of the mitochondrial core PCGs of Pleurotus was found to be 29.86, which indicates strong codon preference. Furthermore, the expected and actual ENC values showed significant differences (18.59-20.55%). Neutrality plot analysis and PR2-Bias plot analysis also showed evidence of natural selection in Pleurotus codon bias. This is consistent with the results seen in the mitochondrial genomes of other species [83,84,85]. The findings of this study revealed that, despite some discrepancies in codon usage indicators between different Pleurotus species, they all experienced strong natural selection on their mitochondrial PCGs.

Mitochondria are believed to have been obtained from bacteria by the ancestors of eukaryotes [86], and most mitochondrial genes have since been transferred to the nuclear genome [87]. While most eukaryotes still retain some core PCGs, some tRNA genes and rRNA genes for energy metabolism [88, 89], which can be used as a molecular marker for phylogeny. As such, the mitochondrial genome is considered a useful tool for inferring phylogenetic relationships of species [90,91,92]. In this study, the genetic relationship of different Pleurotus species was analyzed based on a combined mitochondrial gene set and high support rates were found for each evolutionary clade. Additionally, the relationship between different Pleurotus species was determined based on their RSCU values, which differed from the sequence-based relationships. The phylogenetic tree constructed with RSCU values can serve as a supplement and reference for constructing mitochondrial gene phylogenetic trees, which agreed with previous research [93, 94]. Codon bias, the non-uniform usage of synonymous codons, plays a role in species biodiversity, physiology, morphology, and nutrition of fungi. It can contribute to species-specific genetic signatures, influence translational efficiency and protein expression levels, potentially affect protein structure and function related to morphology, and influence the ability of a species to utilize different nutrients. However, the precise mechanisms and causal relationships between codon bias and these biological characteristics remain incompletely understood [95, 96]. Consequently, this research enhanced the comprehension of codon usage characteristics and genetic evolution of this higher fungal group.

Data availability

All data generated or analyzed during this study are included in this published article [and its supplementary information files].

References

  1. Kurland CG. Codon bias and gene expression. FEBS Lett. 1991;285(2):165–9.

    Article  CAS  PubMed  Google Scholar 

  2. Parvathy ST, Udayasuriyan V, Bhadana V. Codon usage bias. Mol Biol Rep. 2022;49(1):539–65.

    Article  CAS  PubMed  Google Scholar 

  3. Holm L. Codon usage and gene expression. Nucleic Acids Res. 1986;14(7):3075–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Iriarte A, Lamolle G, Musto H. Codon usage Bias: an endless tale. J Mol Evol. 2021;89(9–10):589–93.

    Article  CAS  PubMed  Google Scholar 

  5. Chakraborty S, et al. Codon usage trend in genes associated with obesity. Biotechnol Lett. 2020;42(10):1865–75.

    Article  CAS  PubMed  Google Scholar 

  6. Deng Y, et al. Hidden patterns of codon usage bias across kingdoms. J R Soc Interface. 2020;17(163):20190819.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. He Z et al. Synonymous Codon usage analysis of three Narcissus potyviruses. Viruses, 2022. 14(5).

  8. Ahn I, Jeong BJ, Son HS. Comparative study of synonymous codon usage variations between the nucleocapsid and spike genes of coronavirus, and C-type lectin domain genes of human and mouse. Exp Mol Med. 2009;41(10):746–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Brandão PE. The evolution of codon usage in structural and non-structural viral genes: the case of avian coronavirus and its natural host Gallus gallus. Virus Res. 2013;178(2):264–71.

    Article  PubMed  Google Scholar 

  10. Jia R, et al. Analysis of synonymous codon usage in the UL24 gene of duck enteritis virus. Virus Genes. 2009;38(1):96–103.

    Article  CAS  PubMed  Google Scholar 

  11. Kokate PP, Techtmann SM, Werner T. Codon usage bias and dinucleotide preference in 29 Drosophila species G3 (Bethesda), 2021. 11(8).

  12. Li G, Zhang L, Du N. Relative synonymous codon usage of ORF1ab in SARS-CoV-2 and SARS-CoV. Genes Genomics. 2021;43(11):1351–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Long S, et al. Analysis of compositional bias and codon usage pattern of the coding sequence in Banna virus genome. Virus Res. 2018;258:68–72.

    Article  CAS  PubMed  Google Scholar 

  14. Bu Y, et al. Codon usage bias predicts the functional MYB10 gene in Populus. J Plant Physiol. 2021;265:153491.

    Article  CAS  PubMed  Google Scholar 

  15. LaBella AL, et al. Signatures of optimal codon usage in metabolic genes inform budding yeast ecology. PLoS Biol. 2021;19(4):e3001185.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Arella D, Dilucca M, Giansanti A. Codon usage bias and environmental adaptation in microbial organisms. Mol Genet Genomics. 2021;296(3):751–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Delport W, Scheffler K, Seoighe C. Models of coding sequence evolution. Brief Bioinform. 2009;10(1):97–109.

    Article  CAS  PubMed  Google Scholar 

  18. Hugaboom M et al. Evolution and codon usage bias of mitochondrial and nuclear genomes in Aspergillus section Flavi G3 (Bethesda), 2022.

  19. Liu Y. A code within the genetic code: codon usage regulates co-translational protein folding. Cell Commun Signal. 2020;18(1):145.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Plotkin JB, et al. Codon usage and selection on proteins. J Mol Evol. 2006;63(5):635–53.

    Article  CAS  PubMed  Google Scholar 

  21. Shackelton LA, Parrish CR, Holmes EC. Evolutionary basis of codon usage and nucleotide composition bias in vertebrate DNA viruses. J Mol Evol. 2006;62(5):551–63.

    Article  CAS  PubMed  Google Scholar 

  22. Schmid P, Flegel WA. Codon usage in vertebrates is associated with a low risk of acquiring nonsense mutations. J Transl Med. 2011;9:87.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Stoletzki N, Eyre-Walker A. Synonymous codon usage in Escherichia coli: selection for translational accuracy. Mol Biol Evol. 2007;24(2):374–81.

    Article  CAS  PubMed  Google Scholar 

  24. Dhindsa RS, et al. Natural selection shapes codon usage in the Human Genome. Am J Hum Genet. 2020;107(1):83–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Dilucca M, et al. Co-evolution between codon usage and protein-protein interaction in bacteria. Gene. 2021;778:145475.

    Article  CAS  PubMed  Google Scholar 

  26. Wang Q et al. Codon usage provides insights into the adaptive evolution of mycoviruses in their Associated Fungi host. Int J Mol Sci, 2022. 23(13).

  27. Wang YY, et al. Optimized codon usage enhances the expression and immunogenicity of DNA vaccine encoding Taenia solium oncosphere TSOL18 gene. Mol Med Rep. 2015;12(1):281–8.

    Article  CAS  PubMed  Google Scholar 

  28. Xu Y, et al. Codon usage bias regulates gene expression and protein conformation in yeast expression system P. pastoris. Microb Cell Fact. 2021;20(1):91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Yang Q, et al. Effects of codon usage on gene expression are promoter context dependent. Nucleic Acids Res. 2021;49(2):818–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Zhao F et al. Genome-wide role of codon usage on transcription and identification of potential regulators. Proc Natl Acad Sci U S A, 2021. 118(6).

  31. Tuller T. Codon bias, tRNA pools and horizontal gene transfer. Mob Genet Elem. 2011;1(1):75–7.

    Article  Google Scholar 

  32. Tuller T, et al. Association between translation efficiency and horizontal gene transfer within microbial communities. Nucleic Acids Res. 2011;39(11):4743–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Callens M et al. Read between the lines: diversity of Nontranslational Selection pressures on local Codon usage. Genome Biol Evol, 2021. 13(9).

  34. Ma J, et al. METTL1/WDR4-mediated m(7)G tRNA modifications and m(7)G codon usage promote mRNA translation and lung cancer progression. Mol Ther. 2021;29(12):3422–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. López JL et al. Codon Usage Optimization in the Prokaryotic Tree of Life: how synonymous codons are differentially selected in sequence domains with different expression levels and degrees of Conservation. mBio, 2020. 11(4).

  36. Jiang W, et al. Analysis of synonymous codon usage patterns in the edible fungus volvariella volvacea. Biotechnol Appl Biochem. 2017;64(2):218–24.

    Article  PubMed  Google Scholar 

  37. Otieno OD, et al. Genetic diversity of Kenyan native oyster mushroom (Pleurotus). Mycologia. 2015;107(1):32–8.

    Article  PubMed  Google Scholar 

  38. Lin P et al. Genetic and Chemical Diversity of Edible Mushroom Pleurotus Species Biomed Res Int, 2022. 2022: p. 6068185.

  39. Flores GA, et al. Diversity of Pleurotus spp. (Agaricomycetes) and their metabolites of Nutraceutical and Therapeutic Importance. Int J Med Mushrooms. 2023;25(6):1–20.

    Article  PubMed  Google Scholar 

  40. Hasan HA et al. Assessment of genetic diversity among Pleurotus Spp. Isolates from Jordan. J Fungi (Basel), 2018. 4(2).

  41. Sánchez C. Cultivation of Pleurotus ostreatus and other edible mushrooms. Appl Microbiol Biotechnol. 2010;85(5):1321–37.

    Article  PubMed  Google Scholar 

  42. Krakowska A, et al. Selected edible medicinal mushrooms from Pleurotus Genus as an answer for human civilization diseases. Food Chem. 2020;327:127084.

    Article  CAS  PubMed  Google Scholar 

  43. Dos Reis EE, Schenkel PC, Camassola M. Effects of bioactive compounds from Pleurotus mushrooms on COVID-19 risk factors associated with the cardiovascular system. J Integr Med. 2022;20(5):385–95.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Cohen R, Persky L, Hadar Y. Biotechnological applications and potential of wood-degrading mushrooms of the genus Pleurotus. Appl Microbiol Biotechnol. 2002;58(5):582–94.

    Article  CAS  PubMed  Google Scholar 

  45. Li L, et al. Effect of Pleurotus Eryngii mycelial fermentation on the composition and antioxidant properties of tartary buckwheat. Heliyon. 2024;10(4):e25980.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Gray MW, Burger G, Lang BF. The origin and early evolution of mitochondria. Genome Biol. 2001;2(6):REVIEWS1018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Li Q, et al. Comparative Mitogenome Analysis reveals mitochondrial genome differentiation in Ectomycorrhizal and Asymbiotic Amanita Species. Front Microbiol. 2020;11:1382.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Li Q, et al. First two mitochondrial genomes for the order Filobasidiales reveal novel gene rearrangements and intron dynamics of Tremellomycetes. IMA Fungus. 2022;13(1):7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Cai N, et al. Mitochondrial DNA variants modulate N-formylmethionine, proteostasis and risk of late-onset human diseases. Nat Med. 2021;27(9):1564–75.

    Article  CAS  PubMed  Google Scholar 

  50. Chen K et al. Mitochondrial mutations and mitoepigenetics: focus on regulation of oxidative stress-induced responses in breast cancers. Semin Cancer Biol, 2020.

  51. Hayashi JI, et al. Mutations in mitochondrial DNA regulate mitochondrial diseases and metastasis but do not regulate aging. Curr Opin Genet Dev. 2016;38:63–7.

    Article  CAS  PubMed  Google Scholar 

  52. Li Q, et al. Comparative mitogenomics reveals large-scale gene rearrangements in the mitochondrial genome of two Pleurotus species. Appl Microbiol Biotechnol. 2018;102(14):6143–53.

    Article  CAS  PubMed  Google Scholar 

  53. Liu Z, et al. The complete mitochondrial genome of the edible mushroom Pleurotus Giganteus (Agaricales, Pleurotus) and insights into its phylogeny. Mitochondrial DNA B Resour. 2022;7(7):1313–5.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Wang Y, et al. The mitochondrial genome of the Basidiomycete fungus Pleurotus Ostreatus (oyster mushroom). FEMS Microbiol Lett. 2008;280(1):34–41.

    Article  CAS  PubMed  Google Scholar 

  55. Xu LM, Hinsinger DD, Jiang GF. The complete mitochondrial genome of the Basidiomycete fungus Pleurotus cornucopiae (Paulet) Rolland. Mitochondrial DNA B Resour. 2018;3(1):73–5.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Yang R, et al. The complete mitochondrial genome of the Basidiomycete edible fungus Pleurotus Eryngii. Mitochondrial DNA B Resour. 2016;1(1):772–4.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Ye LY, et al. Mitochondrial genome and diverse inheritance patterns in Pleurotus Pulmonarius. J Microbiol. 2020;58(2):142–52.

    Article  CAS  PubMed  Google Scholar 

  58. Huo X, et al. Analysis of synonymous codon usage of transcriptome database in Rheum palmatum. PeerJ. 2021;9:e10450.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Sharp PM, Li WH. The codon Adaptation Index–a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 1987;15(3):1281–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Peden JF. Analysis of codon usage PhD Thesis, University of Nottingham, 1999. UK.

  61. Puigbo P, Bravo IG, Garcia-Vallve S. CAIcal: a combined set of tools to assess codon usage adaptation. Biol Direct. 2008;3:38.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Orr HA. The evolutionary genetics of adaptation: a simulation study. Genet Res. 1999;74(3):207–14.

    Article  CAS  PubMed  Google Scholar 

  63. Wright F. The ‘effective number of codons’ used in a gene. Gene. 1990;87(1):23–9.

    Article  CAS  PubMed  Google Scholar 

  64. Li Q, et al. The first two mitochondrial genomes for the genus Ramaria reveal mitochondrial genome evolution of Ramaria and phylogeny of Basidiomycota. IMA Fungus. 2022;13(1):16.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Katoh K, Rozewicki J, Yamada KD. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Brief Bioinform; 2017.

  66. Vaidya G, Lohman DL, Meier R. SequenceMatrix: concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics. 2011;27(2):171–80.

    Article  PubMed  Google Scholar 

  67. Lanfear R, et al. PartitionFinder 2: New methods for selecting Partitioned models of Evolution for Molecular and Morphological phylogenetic analyses. Mol Biol Evol. 2017;34(3):772–3.

    CAS  PubMed  Google Scholar 

  68. Ronquist F, et al. MrBayes 3.2: efficient bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61(3):539–42.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Li Q et al. Comparative mitogenomic analysis reveals Intraspecific, Interspecific Variations and Genetic Diversity of Medical Fungus Ganoderma. J Fungi (Basel), 2022. 8(8).

  70. Wu P, et al. Comprehensive analysis of codon bias in 13 Ganoderma mitochondrial genomes. Front Microbiol. 2023;14:1170790.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Chakraborty S, Yengkhom S, Uddin A. Analysis of codon usage bias of chloroplast genes in Oryza species: Codon usage of chloroplast genes in Oryza species. Planta. 2020;252(4):67.

    Article  CAS  PubMed  Google Scholar 

  72. Barbhuiya PA, Uddin A, Chakraborty S. Analysis of compositional properties and codon usage bias of mitochondrial CYB gene in anura, urodela and gymnophiona. Gene. 2020;751:144762.

    Article  CAS  PubMed  Google Scholar 

  73. Xu W, et al. Synonymous codon usage bias in plant mitochondrial genes is associated with intron number and mirrors species evolution. PLoS ONE. 2015;10(6):e0131508.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Liu H et al. Codon usage by chloroplast gene is bias in Hemiptelea Davidii. J Genet, 2020. 99.

  75. Chen H, et al. Mutation and selection cause codon usage and bias in mitochondrial genomes of ribbon worms (Nemertea). PLoS ONE. 2014;9(1):e85631.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Trotta E. Selection on codon bias in yeast: a transcriptional hypothesis. Nucleic Acids Res. 2013;41(20):9382–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Yang C, et al. Comparative analysis of genomic and transcriptome sequences reveals divergent patterns of Codon Bias in Wheat and its ancestor species. Front Genet. 2021;12:732432.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Yu X, et al. Comprehensive Analysis of Synonymous Codon usage Bias for Complete genomes and E2 gene of atypical Porcine Pestivirus. Biochem Genet. 2021;59(3):799–812.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Mazumder GA, Uddin A, Chakraborty S. Analysis of codon usage bias in mitochondrial CO gene among platyhelminthes. Mol Biochem Parasitol. 2021;245:111410.

    Article  CAS  PubMed  Google Scholar 

  80. Montana-Lozano P, Balaguera-Reina SA, Prada-Quiroga CF. Comp Anal Codon Usage Mitochondrial Genomes Provides Evolutionary Insights into Reptiles Gene. 2022;851:146999.

    Google Scholar 

  81. Pepe D, K DEK. Codon bias analyses on thyroid carcinoma genes. Minerva Endocrinol. 2020;45(4):295–305.

    PubMed  Google Scholar 

  82. Prabha R, et al. Genome-wide comparative analysis of codon usage bias and codon context patterns among cyanobacterial genomes. Mar Genomics. 2017;32:31–9.

    Article  PubMed  Google Scholar 

  83. Barbhuiya PA, Uddin A, Chakraborty S. Understanding the codon usage patterns of mitochondrial CO genes among amphibians. Gene. 2021;777:145462.

    Article  CAS  PubMed  Google Scholar 

  84. Barbhuiya PA, Uddin A, Chakraborty S. Codon usage pattern and evolutionary forces of mitochondrial ND genes among orders of class Amphibia. J Cell Physiol. 2021;236(4):2850–68.

    Article  CAS  PubMed  Google Scholar 

  85. Li Q, et al. Analysis of synonymous codon usage patterns in mitochondrial genomes of nine Amanita species. Front Microbiol. 2023;14:1134228.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Lang BF, Gray MW, Burger G. Mitochondrial genome evolution and the origin of eukaryotes. Annu Rev Genet. 1999;33:351–97.

    Article  CAS  PubMed  Google Scholar 

  87. Adams KL, Palmer JD. Evolution of mitochondrial gene content: gene loss and transfer to the nucleus. Mol Phylogenet Evol. 2003;29(3):380–95.

    Article  CAS  PubMed  Google Scholar 

  88. Costa GG, et al. The mitochondrial genome of Moniliophthora Roreri, the frosty pod rot pathogen of cacao. Fungal Biol. 2012;116(5):551–62.

    Article  CAS  PubMed  Google Scholar 

  89. Bullerwell CE, Burger G, Lang BF. A novel motif for identifying rps3 homologs in fungal mitochondrial genomes. Trends Biochem Sci. 2000;25(8):363–5.

    Article  CAS  PubMed  Google Scholar 

  90. Johri P, et al. Population genetics of Paramecium mitochondrial genomes: recombination, mutation spectrum, and efficacy of selection. Genome Biol Evol; 2019.

  91. Li H, et al. Higher-level phylogeny of paraneopteran insects inferred from mitochondrial genome sequences. Sci Rep. 2015;5:8527.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Li Q, et al. The first two mitochondrial genomes from Apiotrichum reveal mitochondrial evolution and different taxonomic assignment of Trichosporonales. IMA Fungus. 2023;14(1):7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Crane A et al. Phylogenetic relationships and codon usage bias amongst cluster K mycobacteriophages G3 (Bethesda), 2021. 11(11).

  94. Gupta S, Paul K, Roy A. Codon usage signatures in the genus Cryptococcus: a complex interplay of gene expression, translational selection and compositional bias. Genomics. 2021;113(1 Pt 2):821–30.

    Article  CAS  PubMed  Google Scholar 

  95. Hershberg R, Petrov DA. Selection on codon bias. Annu Rev Genet. 2008;42:287–99.

    Article  CAS  PubMed  Google Scholar 

  96. Gustafsson C, Govindarajan S, Minshull J. Codon bias and heterologous protein expression. Trends Biotechnol. 2004;22(7):346–53.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by National Natural Science Foundation of China (No. 82102738).

Author information

Authors and Affiliations

Authors

Contributions

Q.L and W.G designed the experiment; X.C., J.H., A.S., Y.L., W.X., and Z.X. analyzed the data; Q.L and W.G wrote and review the manuscript; Q.L. managed the project. All authors reviewed the manuscript.

Corresponding author

Correspondence to Qiang Li.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Adherence to national and international regulations

Not applicable.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gao, W., Chen, X., He, J. et al. Intraspecific and interspecific variations in the synonymous codon usage in mitochondrial genomes of 8 pleurotus strains. BMC Genomics 25, 456 (2024). https://doi.org/10.1186/s12864-024-10374-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-024-10374-3

Keywords