Next Article in Journal
Synovial Extracellular Vesicles: Structure and Role in Synovial Fluid Tribological Performances
Next Article in Special Issue
Single-Cell DNA Methylation Analysis of Chicken Lampbrush Chromosomes
Previous Article in Journal
Plasma Campesterol Is Positively Associated with Carotid Plaques in Asymptomatic Subjects
Previous Article in Special Issue
Evolutionary Invariant of the Structure of DNA Double Helix in RNAP II Core Promoters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bioinformatic Assessment of Factors Affecting the Correlation between Protein Abundance and Elongation Efficiency in Prokaryotes

by
Aleksandra E. Korenskaia
1,2,3,*,
Yury G. Matushkin
2,3,
Sergey A. Lashin
1,2,3 and
Alexandra I. Klimenko
1,2
1
Kurchatov Genomics Center, Institute of Cytology and Genetics, Siberian Branch of the Russian Academy of Science, Lavrentiev Avenue 10, 630090 Novosibirsk, Russia
2
Institute of Cytology and Genetics, Siberian Branch of the Russian Academy of Science, Lavrentiev Avenue 10, 630090 Novosibirsk, Russia
3
Department of Natural Sciences, Novosibirsk National Research State University, Pirogova St. 1, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11996; https://doi.org/10.3390/ijms231911996
Submission received: 24 August 2022 / Revised: 23 September 2022 / Accepted: 30 September 2022 / Published: 9 October 2022
(This article belongs to the Special Issue Bioinformatics of Gene Regulations and Structure - 2022)

Abstract

:
Protein abundance is crucial for the majority of genetically regulated cell functions to act properly in prokaryotic organisms. Therefore, developing bioinformatic methods for assessing the efficiency of different stages of gene expression is of great importance for predicting the actual protein abundance. One of these steps is the evaluation of translation elongation efficiency based on mRNA sequence features, such as codon usage bias and mRNA secondary structure properties. In this study, we have evaluated correlation coefficients between experimentally measured protein abundance and predicted elongation efficiency characteristics for 26 prokaryotes, including non-model organisms, belonging to diverse taxonomic groups The algorithm for assessing elongation efficiency takes into account not only codon bias, but also number and energy of secondary structures in mRNA if those demonstrate an impact on predicted elongation efficiency of the ribosomal protein genes. The results show that, for a number of organisms, secondary structures are a better predictor of protein abundance than codon usage bias. The bioinformatic analysis has revealed several factors associated with the value of the correlation coefficient. The first factor is the elongation efficiency optimization type—the organisms whose genomes are optimized for codon usage only have significantly higher correlation coefficients. The second factor is taxonomical identity—bacteria that belong to the class Bacilli tend to have higher correlation coefficients among the analyzed set. The third is growth rate, which is shown to be higher for the organisms with higher correlation coefficients between protein abundance and predicted translation elongation efficiency. The obtained results can be useful for further improvement of methods for protein abundance prediction.

1. Introduction

It is well-known that proteins are the key elements that provide cell function, hence many physiological processes are controlled by the efficient allocation of the cellular proteome [1]. That is why quantification of protein abundance is of great importance for medical and biological studies. Experimental methods for measuring the amount of protein are expensive and labor-intensive; therefore, the problem of predicting the amount of protein based on genetic data is urgent.

1.1. Protein Abundance Prediction Tools

There are several approaches for prediction of protein abundance and tools that are used to calculate protein abundance for a particular organism based on mRNA abundance data and parameters of mRNA sequence. Many of them are species-specific, such as the tool that was developed for predicting protein abundance of Saccharomyces cerevisiae and Schizosaccharomyces pombe [2]. The calculations are based on experimentally obtained mRNA abundance, codon usage, the mRNA folding energy, and proteins’ half-life, which were obtained for both organisms and used as constants for each protein. The correlation between predicted and measured protein abundance is 0.77 and 0.74 for S. cerevisiae and S. pombe, respectively. The authors underline the factors that highly impact prediction accuracy: mRNA abundance, codon usage, and energy of mRNA fold. Another example is a tool that uses mRNA data for the prediction of protein abundance for immune cells of humans and mice [3]. The correlation between predicted and measured protein abundance is about 0.79–0.94.
Also, a few tools working without mRNA abundance data exist. There is an algorithm that uses only mRNA sequence features data to predict highly and lowly expressed proteins of S. cerevisiae. This algorithm uses 91 features including various estimates of codon bias calculated in various ways, CG content, codon-pair frequencies, etc. It shows a high correlation coefficient (0.75) between predicted and measured data for the organisms under study; however, the application of this model to Escherichia coli shows a modest (0.5) correlation between measured and predicted data [4]. These algorithms show high accuracy on organisms that have been used for training. However, predicting basal protein levels in a more general case for non-model organisms still remains challenging. Moreover, even for such a well-studied model organism as Escherichia coli, different studies emphasize different factors contributing to actual protein abundance. According to [5], the major part of protein abundance (53%) is determined by transcript level and at least 12% of protein abundance is determined by effectors of translation elongation. On the other hand, translational initiation determines about 1% of protein abundance. However, there is other evidence suggesting that translation initiation also might play a significant role, especially, when the expression levels of individual genes are under focus rather than global translation efficiency and cellular fitness [6]. Therefore, bioinformatic estimation of factors contributing to protein abundance, such as elongation efficiency, is an important step towards in silico prediction of protein abundance levels. Here we investigate the capability to predict protein abundance for phylogenetically diverse organisms by using the EloE (Elongation Efficiency) tool [7,8,9,10]. This tool calculates parameters that impact elongation efficiency of the translation stage: codon bias, number of secondary structures in mRNA, and energy of secondary structures in mRNA.

1.2. Codon Usage Bias Impacts Elongation Efficiency

The codon usage bias is the unequal proportion of synonymous codons occurrence throughout a genome. The fact that codon usage bias is associated with gene expression level was shown by Sharp, Tuohy, and Mosurski [11], who demonstrated on yeast that the highly expressed genes consist of common codons and lowly expressed genes contain infrequent codons. The experiment with replacement of common codons to infrequent codons into E. coli genes showed that infrequent codons could increase translation time [12]. Per-codon elongation rates are crucially dependent on the tRNA pool [13]. Codons with low abundance of the corresponding tRNA require more time for the correct accommodation of the corresponding tRNA in the A-site of the ribosome, which makes them non-optimal codons. Therefore, the content of optimal and non-optimal codons across mRNA affects gene elongation efficiency. The fact that codon bias is associated with the level of gene expression has been shown in many studies and reviews [14,15,16,17,18,19,20,21,22,23]. The great impact of codon bias on translation efficiency was shown on E. coli [16,20], S. cerevisiae, and Trichomonas vaginalis [14]. The codon usage bias has proven to be a good predictor of gene expression for S. cerevisiae [4] and Trypanosoma brucei [19]; however, it proves to be insufficient for a number of other organisms including Rickettsia, Ehrlichia, Buchnera, Mycoplasma, Micrococcus, Helicobacter, and some spirochaetes [24,25,26,27,28,29,30,31].
There is a classical approach to quantify codon usage bias known as the Codon Adaptation Index (CAI) [32]. CAI is widely used to assess the codon usage bias in various organisms and is implemented in several programming packages [33,34,35].
The codon usage bias might tune gene expression according to gene function. As it has been shown for several S. cerevisiae genes, mRNAs of housekeeping genes, such as those involved in glycolysis, are uniformly enriched in high-optimality codons, whereas proteins involved in transient responses to stimuli, such as the pheromone response, are enriched in non-optimal codons [36]. Genes that control circadian rhythms in various organisms, including cyanobacteria and Neurospora crassa, are also deficient in optimal codons. The authors suggest that enrichment in non-optimal codons helps to provide low levels of such regulatory gene products to allow quick elimination of the product after stimulus has disappeared [36]. Frumkin with colleagues [18] recoded genes in E. coli to demonstrate that codon usage patterns not only tune the elongation rates of particular genes, but also affect the global protein translation efficiency. Optimal codon composition of highly expressed genes increases the efficiency of translation and, consequently, reduces the number of ribosomes required for expression of these genes. This allows to indirectly increase the rate of translation initiation for other transcripts due to an increase in the pool of free ribosomes. Similar results have been shown on S. cerevisiae [37]. Shah et al. also demonstrated on S. cerevisiae that codon bias strongly affects protein abundance in genes with high mRNA level, whereas the effect of codon bias on protein abundance in genes with low mRNA abundance (<1% of transcriptome) is much lower but still significant [38]. The latter might be accounted for by the fact that translation of low-expressed genes slightly contributes into the pool of free ribosomes.
Besides the factors listed above, codon usage bias is associated with a number of other factors. It can be related to the gene function, e.g., hydrophobic loci of encoded proteins are associated with specific codon usage and signal peptides demonstrate non-optimal codons enrichment, or to the location of gene on the chromosome strand (leading or lagging) [22].
In conclusion, optimization by codon content is crucially important for the highly expressed genes, especially the genes expressed constitutively, which are supposed to be optimized by codon content, but it might be less critical for other genes, especially those with low expression and under special regulation. Therefore, it is proven that codon bias is a good predictor of efficiency of translation elongation and gene expression levels for several organisms, which provides opportunities for prediction of gene expression at the constitutive level based on the codon bias of genes.

1.3. Secondary Structures Impact Elongation Efficiency

It has been shown that the translation elongation rate is tuned not only by the codon usage bias but also by mRNA secondary structures [39,40,41]. Strong mRNA secondary structures reduce the speed of translating ribosomes [37,38,42], as it requires time for unweaving [43]. It is provided by the helicase activity of the ribosome using the active mechanical unwinding mechanism [43,44,45,46]. In this mechanism, the ribosome is translocated by applying force to the closed state of the mRNA duplex, which requires additional energy consumption. This mechanism, revealed in the data obtained from E. coli, affects the basal rate of translocation in a prokaryotic cell [43]. Moreover, it is known that different ORFs on an intra-operon level translate differentially varying in rates as much as 100-fold, which was demonstrated for E. coli [47]. Apparently, minimization of the abundance and energy of secondary structures in mRNA is supposed to increase the translation elongation rate.
However, it should be noted that the abundance and the energy of mRNA secondary structures tend to be higher over coding regions compared to untranslated regions, as was shown in yeast and E. coli [48]. Secondary structures in mRNA perform various functions [39,40], including modulation of folding for some proteins [49], regulation of mRNA half-life [36,50,51,52], regulation of translational frameshifting [53,54], termination–insulation and re-initiation control [55], whereby secondary structures can influence other stages of gene expression. This can introduce uncertainty about the effect of mRNA secondary structures on protein abundance. However, since for the abovementioned reasons, the number and stability of mRNA structures negatively affect translation efficiency, and the impact of secondary structures in the implementation of the listed functions may vary among different taxonomic groups (for example, there are many other mechanisms for regulation of mRNA decay in bacteria [52]), we can assume that the number and energy of secondary structures in mRNA can be a good predictor of the protein abundance for some organisms.
It is interesting that codon usage might be associated with the stability of secondary structures. It has been shown for E. coli and S. cerevisiae that the regions of high secondary structure content are preferentially enriched in high-optimal codons while non-optimal codons are located in low structured regions. Authors suggest that this pattern allows compensation for their independent effects on translation, helping to smooth overall translational speed and reducing the chance of potentially detrimental points of excessively slow or fast ribosome movement [56]. Moreover, genes tend to have significant codon bias in the regions of extremely high and low levels of secondary structure, which is found across all domains of life [57]. As has been shown in yeast, both codon usage bias and mRNA structural stability positively regulate mRNA expression levels and, moreover, highly structured and stable mRNA is selected [58]. It seems that codon bias and secondary structures in mRNA tend to be balanced to ensure optimal level of gene expression.

1.4. Summary

In conclusion, codon bias and secondary structures greatly impact translation elongation efficiency and contribute to gene expression. Therefore, a prediction of protein abundance based on these parameters seems to be a useful perspective. Here we analyze the capability to predict protein abundance using the EloE tool that calculates elongation efficiency indexes (EEI) based on these parameters. Previously, this tool was applied to show a significant correlation between EEI and gene expression for S. cerevisiae (0.79) and for Helicobacter pylori (0.28) [10]. As demonstrated for H. pylori, the correlation between gene expression and EEI increases with gene length, showing a maximum correlation (0.58) at a gene length of about 2200 bp [59]. In this work, we assess the correlation between EEI and gene expression at the protein level for various prokaryotes with diverse lifestyles, including archaea, obligate and opportunistic pathogenic bacteria, cyanobacteria, and species adapted to harsh environments (in particular, extremely acidophilic bacteria Acidithiobacillus ferrooxidans, and halophilic archaea Halobacterium salinarum).

2. Results

We have analyzed the correlation between protein abundance and base elongation efficiency index (EEI) value for various groups of microorganisms (see the details in the Materials and Methods section) and have investigated how this correlation depends on the following factors:
  • Base EEI type, i.e., the mode of evolutionary optimization of translation exhibited by a particular genome;
  • Taxonomical identity of an analyzed genome;
  • Cell doubling time, i.e., microorganism’s reproduction rate;
  • Mean (M) and standard deviation (R) of ranks of ribosomal protein genes measured on the base EEI scale.
Taking into account these factors allows us to study the structure of the sample, disentangling their impact on the correlation coefficient value between protein abundance and EEI.
Different genomic features in association with the obtained correlation coefficients (corr(PA|EEI)) between base EEI and protein abundance have also been analyzed. Neither genome length (r = −0.004, p = 0.85) nor number of genes (r = 0.01, p = 0.84) nor number of tRNAs (r = 0.36, p = 0.37) correlate significantly with the correlation coefficient between protein abundance and EEI. At the same time, such characteristics as number of ribosomal protein genes (r = 0.488, p = 1 × 10−16) and GC content (r = −0.394, p = 0.02) demonstrate significant correlation with the corr(PA|EEI). The number of ribosomal genes also correlates with the minimal doubling time of a microbe (Spearman’s correlation coefficient r = −0.428, p = 0.046).
To understand the representativeness of using proteomic data, we have calculated proteome coverage. Proteome coverage, which is a percentage of protein-coding genes presented in proteomic data, varies among samples. The median coverage per studying organism is 50.8 with the standard deviation 24.1. This means that for most of the analyzed organisms, the data used for analysis do not characterize the entire proteome, but do cover at least a significant part of it.
The correlation, coverage, minimal doubling time, EEI type, and mean (M) and standard deviation (R) values for each organism are demonstrated in Table 1.
Overall, the mean Spearman’s correlation coefficient between protein abundance and EEI calculated for the whole sample equals to 0.4 (the boxplot depicting corresponding descriptive statistics is shown in Figure 1). The majority of analyzed organisms, with the exception of Neisseria meningitidis, have shown a significant correlation between protein abundance and base EEI values. However, the correlation coefficient values vary greatly among the organisms.
This result means that predicting protein abundance solely based on elongation translation characteristics, such as those calculated by EloE, will have good accuracy for some organisms and poor accuracy for others. Further analysis aims to reveal the parameters that contribute to the correlation coefficients’ values.

2.1. Dependence of Correlation between Protein Abundance and the EEI from EEI Type

To determine the patterns of the correlation coefficients’ distribution among the organisms depending on their mode of evolutionary optimization of translation, we split the sample into several subsamples according to the genome’s base EEI type established by EloE (see Figure 2).
The highest correlation was obtained for the organisms belonging to the EEI1 type, which relies primarily on codon usage optimization for efficient translation. The correlation coefficients for organisms which were assigned to the EEI2 and the EEI4 types are significantly lower. The optimization of elongation efficiency types for these organisms were based on the optimization of number of secondary structures in mRNA for the EEI2 type, and codon usage and the optimization of number of secondary structures in mRNA for the EEI4 type.
It is important to note that the organisms belonging to the types other than the codon usage bias optimization only type (the EEI1 type) do not demonstrate higher correlation coefficients if elongation efficiency indices are calculated taking into account codon usage bias only, i.e., using the EEI1 formula (see Figure 3 and Table A1). The correlation coefficients between the EEI1 indices and protein abundance are significantly lower (p = 0.02, Welch’s t-test) than the correlation coefficients between the base EEI type and protein abundance for the organisms belonging to the type which minimizes the number of secondary structures (EEI2) (Figure 3a). They are also lower for Pseudomonas aeruginosa, which belongs to the type that considers only energy of secondary structures (EEI3, see Figure 3b), though we do not have enough sample size to deduce any extrapolations from here. Finally, the type that considers the codon usage bias and the number of secondary structures in mRNA (EEI4, Figure 3c) demonstrates higher corr(PA|EEI4) values than corr(PA|EEI1) at a trend level (p = 0.24). Thus, applying the approach that considers different elongation efficiency types allows improvement of the accuracy of predictions for those organisms that do not demonstrate a clear codon usage optimization pattern.

2.2. Dependence of Correlation between Protein Abundance and the EEI from Phylogeny

Phylogenetically distant organisms can have significant differences in the regulation of gene expression. Therefore, the significance of the effect of translation elongation factors on the overall level of gene expression may also differ among phylogenetically diverse organisms. In this regard, the ability to predict protein abundance based on the elongation translation characteristics can vary greatly for different phylogenic groups.
Below, we have mapped the analyzed strains onto a phylogenetic tree in order to reflect the diversity of phylogenetic groups represented in the analysis and to determine for which phylogenetic groups the prediction of protein abundance by EloE provides the most accurate results, which is demonstrated in Figure 4 rendered using iTol [60].
As one can see, the tree includes both species known for codon usage bias being a reliable measure of their translation elongation efficiency (such as E. coli), and those who have been shown to contravene that pattern (such as H. pylori and the representatives of Mycoplasma genus). Accordingly, the former belong to the EEI1 optimization type, while the latter are distributed to the other elongation efficiency optimization types, which take into account the effect of secondary structures in mRNA. Moreover, there are a number of new species that have not been studied in this regard before, and which, therefore, present a special interest.
The most represented taxa are phylum Firmicutes, namely, class Bacilli, and phylum Proteobacteria, in particular, class Gammaproteobacteria. Also, the mean correlation coefficients among the studied organisms of these taxa are 0.59 and 0.46, respectively, which are higher than the mean correlation coefficient for the entire dataset (0.4). Notably, most of the bacteria belonging to these classes belong to the EEI1 type, which show higher correlation coefficients. However, this difference is significant only for class Bacilli, compared with the other microorganisms from the analyzed set (Welch test, p = 2.2 × 10−5). Other taxa are represented by only a couple of species, if any, and their correlation coefficients corr(PA|EEI) are highly varied. The differences among correlation coefficients probably occur due to the different extents of influence of the codon usage bias and secondary structures on gene expression among species.
Thus, one can use elongation efficiency indices for a theoretical assessment of expected protein expression profile in the case of absence of proteomic data for a particular representative of one of those classes that demonstrate relatively high correlation between protein abundance and their base EEI, though biological implications of belonging to a specific elongation efficiency optimization type might vary depending on the particular taxa.

2.3. Dependence of Correlation between Protein Abundance and the EEI from Minimal Doubling Time

Doubling time as a characteristic reflecting reproduction rate varies greatly, both among various bacterial species and inside the same species if it grows in different conditions [61]. It is known that bacterial growth rates are correlated with ribosome abundance [62], and therefore it correlates with the entire translation rate due to reduction in active ribosome fraction during slow growth [63]. However, translation elongation maintains a significant rate even in poor nutrient conditions with slow bacterial growth [63], which enables cells to produce proteins crucial for surviving in harsh environments in a timely manner.
The prediction of protein abundance using elongation efficiency indices assumes that coding sequences of highly expressed genes, such as ribosomal protein genes, are heavily optimized compared to the genes with low level of basal expression. This means that if elongation efficiency is more evenly optimized because it is a less essential step in determining protein abundance than, for instance, a gene regulation, such an organism can demonstrate a reduced quality of protein abundance prediction. Higgs and Ran [64] found a low correlation between tRNA gene abundance and codon usage for most bacteria with high doubling time. They supposed that, although the translation is the limiting factor of division in fast-growing organisms, this is not the case for slow-multiplying organisms. Although their results could also be explained by the high impact of mRNA secondary structures in translation, this aspect is still worth being tested.
Also, it was demonstrated [65] that a prokaryotic growth rate is highly correlated with the codon usage bias. In fast-growing organisms, codon usage bias is more pronounced due to codon usage optimization, which is crucial since the tRNA pool becomes limiting at very high growth rates. Based on the codon usage bias of ribosomal protein, Weissman, Hou, and Fuhrman have predicted [66] the minimum doubling time for about 200,000 prokaryotes. Such an estimation of the growth rate divides prokaryotes into two groups, which fits their ecological roles. The first one is copiotrophs, consisting of fast-growing microbes that grow in nutrient-rich environments. The other is oligotrophs, represented by microbes that are adapted to low levels of nutrients and tend to have slow growth rates. Based on these results, authors have defined oligotroph as an organism for which a selection for rapid maximal growth is weak enough so that translation efficiency is not optimized by selection for optimized codon usage.
In the light of the listed above, a hypothesis can be formulated that protein abundance predictions will be less efficient for prokaryotes with the high minimal doubling time.
Indeed, one can notice (Figure 5a) an increase in the corr(PA|EEI) with a decrease in the minimum doubling time (DT), although bacteria with fast growth and a low correlation coefficient also exist. The Pearson correlation coefficient between corr(PA|EEI) and minimal doubling time for 25 organisms is r = −0.446 (p = 0.025). No relationship was found between the base EEI type and the doubling time. However, it is worth noting that slowly growing bacteria (with the DT ≥ 5 h) are mostly represented by EEI types which consider secondary structures (only one out of seven organisms belong to the EEI1 type). Consistent with previous studies, codon usage bias slightly reflects the gene expression profile for those six organisms, which is demonstrated by calculation of corr(PA|EEI) for each of the five EEI types (see Table A1). Considering the secondary structures enables us to reach higher (but still quite low) correlation coefficients.
We hypothesize that some prokaryotic species living in harsh environments could demonstrate a similar level of translation efficiency optimization throughout the genome. Such organisms are supposed to show a high minimum doubling time and lower translation elongation efficiency for ribosomal genes than fast-growing species. As mean elongation efficiency of ribosomal proteins is reflected by M values, we have compared them for fast-growing and slow-growing prokaryotes (Figure 5b).
The Welch test between M values of fast-growing organisms (with the minimal doubling time no more than two hours) and slow-growing organisms (with the minimal doubling time higher than five hours) has shown a significant difference (p-value = 7.059 × 10−6). The comparison of medium-growing organisms (with the minimal doubling time between two and five hours) and slow-growing organisms also has shown a significant difference for M values (p = 0.0002).
Notably, the lower correlation between protein abundance and elongation efficiency for organisms with higher minimum doubling time cannot be explained only by a weaker optimization of ribosomal protein genes in favor of other genes. If we do not consider elongation efficiency of ribosomal protein genes during the selection of the base EEI type by selection of the EEI type that shows higher correlation coefficients between protein abundance and elongation efficiency, which simulates the usage of the optimal group of highly optimized genes, the correlation coefficients do not necessarily rise. In particular, changing EEI type greatly increases (from 0.12 to 0.34 for Acidithiobacillus ferrooxidans, and from 0.36 to 0.46 for Leptospira interrogans) the correlation coefficient only for two of seven slow-growing organisms under study (see Table A1). In summary, the prediction of protein abundance is less efficient for slow-growing organisms, which can be explained by less pronounced differences in elongation efficiency optimization throughout the genomes of these organisms. In other words, translation elongation efficiency does not appear to be a limiting factor in determining protein abundance for slow-growing microorganisms.

2.4. Dependence of Correlation between Protein Abundance and the EEI from Elongation Efficiency of Ribosomal Protein Genes

As mentioned earlier, the ranks of ribosomal gene proteins, which contribute to the M (mean) and R (standard deviation) parameters, are used to determine a genome’s base elongation efficiency index type, which describes the mode of evolutionary optimization of translation in a particular genome in the most accurate way. Here we have examined how the correlation coefficient between the EEI and protein abundance depends on the M and R values for the base EEI type (see Figure 6).
The Pearson correlation coefficient between M and corr(PA|EEI) for 25 organisms is 0.7344 (p = 2.9 × 10−5). The Pearson correlation coefficient between R and corr(PA|EEI) is −0.454 (p = 0.022). This reassuring result indicates that the strategy of maximizing M and minimizing R that is used to determine the base EEI type in the EloE is the right way, which not only has a theoretical basis but also is substantiated by experimental data.
As these parameters are highly correlated with corr(PA|EEI), they could be used for estimating prediction potential (correlation coefficient between the EEI and protein abundance) for an organism under study. Also, these parameters are calculated by the algorithm itself and do not require the involvement of additional data, which makes them convenient enough to assess the efficiency of the algorithm.
For this purpose, a linear regression model has been built. The independent variable is represented by the M parameter only, since M and R parameters are highly correlated.
C o r r . c o e f = 0.0432 + 0.0054 M
The determination coefficient (R2) equals 0.35, and the mean squared error (MSE) equals 0.011. The test for significance of regression shows F > F-critical (10.36 > 4.2793), p = 0.038, which means that the regression model is statistically significant. In summary, the statistics shows that the model has a prediction power.
Using this formula with caution, and taking into account the observed range of M values, one could predict the expected correlation coefficient for another organism, which does not have enough data covering its protein expression profile.
In summary, we can use the EloE for a rough prediction of gene expression at the protein level. Taking into account the EEI type, doubling time, taxonomic identity, as well as the M and R parameters, allows us to derive an approximate estimate of the expected correlation coefficient between base EEI values and actual protein abundance.

3. Discussion

The gene expression is a multi-level process including various regulation on a transcriptional and translational level. The protein abundance reflects the overall effect of all the factors contributing to the gene expression, whereas each of these factors has its own particular share in this cumulative effect. One of the intriguing questions within this context is the problem of predicting the basal gene expression based on only partial information available, in particular, the genomic sequence data. This study focuses on investigating correlation between the translation elongation characteristics and proteomic data. As our analysis indicates, the mean correlation coefficient between protein abundance and base elongation efficiency index (EEI) calculated for the whole sample is not high, which was expected, since we are trying to predict the protein abundance based on the elongation efficiency, while the protein yield is also influenced by other stages, including the stage of transcription, translation initiation [6,15,67], and other factors such as half-life values of the respective protein and mRNA [15,50,52], as well as the protein’s structure and its resistance to proteases [68,69,70]. To the best of our knowledge, this is the first time such an analysis of correlation between protein abundance and different elongation efficiency measures has been performed based on the proteomics data for the prokaryotes belonging to such a range of taxonomic groups including non-model organisms and the organisms which are known for codon usage being an ineffective measure of translation elongation efficiency of their genes.
However, the correlation coefficients between protein abundance and the EEI values vary greatly among the organisms. The bioinformatic assessment of the factors affecting the correlation between protein abundance and elongation efficiency in prokaryotes has shown that there are several factors associated with the value of the correlation coefficient. The first is the EEI type—organisms that correspond to the EEI1 type, which takes into account codon bias only, have significantly higher correlation coefficients. Such a difference between these types could be explained by ambiguous [71] contributions of secondary structures to protein abundance. Although secondary structures in mRNA decrease ribosome velocity, they can protect mRNA from ribonucleases and, therefore, increase mRNA abundance. As a result, protein abundance could both decrease and increase under the influence of secondary structures. Thus, we should expect a lower prediction accuracy for organisms belonging to optimization types, for which secondary structures play a significant role in determining the protein abundance (EEI2, EEI3, EEI4, and, probably, EEI5). Unfortunately, among the organisms with available protein profiles, Neisseria meningitidis, the only one belonging to the EEI5 base type, do not show a significant correlation between protein abundance and EEI values—not only base ones but any EEI values, including classic codon usage bias. Therefore, we refrain from making any decisive conclusions about that particular optimization type. It is worth noting, however, that for those organisms under study, which fall into one of the optimization types (EEI2, EEI3, and EEI4) characterized by the role of mRNA secondary structures, applying their base elongation efficiency index allows us to reach higher correlation coefficients than if using EEI1, which represents classic codon usage bias. We believe that this indicates the complex nature and the role of translation elongation efficiency in determining protein abundance in these classes of organisms.
The second factor is taxonomic identity of an organism under study—such a class as Bacilli is among those characterized by the highest correlation coefficient between EEI and protein abundance. Using this information to derive estimates of expected correlation coefficients for the organisms that lack proteomic profiles seems to be a promising approach, though we definitely need more data to be able to improve the quality of such an assessment. The third factor is the microorganism’s reproduction rate. We observe an increase in the correlation coefficient between the EEI and protein abundance with a decrease in the minimum doubling time, that is, fast-growing prokaryotes tend to have a high correlation coefficient. The latter might be associated with the similar level of elongation efficiency across the genome in slow-growing species, which is reflected in ribosomal protein coding genes being not the most highly optimized group of genes among them. The fact that genes encoding ribosomal proteins may not be highly efficient at translation elongation was shown on several Mycoplasma species (C. M. haemolamae, M. haemocanis, M. wenyonii, M. haemofelis, M. pneumonia, C. M. haemominutum, and M. suis). These species demonstrate decreased M values and a reduced number of perfect local inverted repeats (potential hairpins) in mRNA of both ribosomal and non-ribosomal genes. It makes translation elongation efficiency of non-ribosomal genes similar to ribosomal ones [72]. Thus, there are various situations where either an organism possesses a quite compact and evenly optimized genome or translation elongation efficiency does not appear to be a limiting stage in determining protein level. However, we have also demonstrated that, in general, the initial approach used by the EloE that relies on assessing the ranks of ribosomal proteins in the gene list sorted by the base EEI values is adequate to the experimental data of the organisms under study, especially for the organisms with a high number of ribosomal protein genes and low GC content. Therefore, it can be used in further development of the algorithms that would take into account not only translation elongation, but also other stages that affect the level of gene expression.
One of the difficulties in studying the relationships between elongation efficiency characteristics and protein abundance at the organism level is the lack of the genome-wide protein abundance profiles to assess the actual correlation between protein abundance and elongation efficiency indices based on representative datasets, which would include protein-encoding genes with various expression levels for taxonomically divergent organisms, including non-model ones. However, as more proteomic studies generating a full protein profile of an organism under study are published, the whole picture of how the particular aspects of optimization of translation elongation efficiency affect the protein abundance in various microorganisms will become more clear and detailed. We believe that a thorough bioinformatic estimation of factors contributing to protein abundance, such as elongation efficiency, paying attention to the actual biodiversity of prokaryotic species, is an important step towards in silico prediction of protein abundance levels.

4. Materials and Methods

4.1. Proteomic and Genomic Data

Gene expression at the protein level data was taken from the PaxDb database [73]. This database stores proteomic data obtained by MS-MS spectroscopy, which provides quantitative protein abundance information. Proteomes were obtained from 26 prokaryotic organisms, including 24 bacteria and 2 archaea. Since for many organisms, numerous experimental data are present, in the further calculation we used a median abundance of each protein per organism. The genomes of these strains with the corresponding loci identifiers were obtained from the NCBI Assembly database [74], the list of species presented in Table 2.
Neisseria meningitidis has been excluded from the subsequent analysis due to its insignificant correlation between protein abundance and base EEI values.

4.2. EloE Elongation Efficiency Indices (EEI)

In this article, we have analyzed elongation efficiency indices calculated by EloE (Elongation Efficiency) tool [7] (developed by Sokolov V.S, Novosibirsk, Russia). The executable file of the tool as well as the user’s manual, input and output data are deposited in the Supplementary files. It requires a Java Runtime Environment (JRE) SE 7 or higher to run the program. The EloE algorithm calculates elongation efficiency indices (EEI) for each organism’s protein-coding gene in five ways (Table 3) [9,10] using annotated genome sequence. The elongation efficiency indices are calculated according to Formula (2):
E E I ( i ) = K / ( w 1 Ta ( i ) + w 2 Te ( i ) )
where K—normalization constant, which is used to assure the range of EEI(i) within [0, 10]; w1 and w2—weight coefficients (equals 0 if a parameter is excluded and 1 if it is considered); Ta ( i ) estimates the average time required for the fixation of isoacceptor aminoacyl-tRNA in the A site of the ribosome; and Te ( i ) estimates the average time demanded by the ribosome for the translocation stage. There are two options for Te ( i ) function: L C I L ( i ) , which calculates local complementarity based on number of potential secondary structures in mRNA, and L C I E ( i ) , which calculates local complementarity taking into account the energy of potential secondary structures in mRNA. The formulae can be found in Appendix A.
For each elongation efficiency index (EEI), protein-coding genes are sorted in descending order according to the corresponding EEI values. To determine the index type that properly describes the efficiency of elongation translation in the particular organism under study, mean (M) and standard deviation (R) of the ranks of ribosomal protein genes are calculated. The ribosomal protein genes are known to be intensely expressed along a wide range of organisms and assumed to be, therefore, optimized in the efficiency of translation elongation. However, a list of highly expressed genes can be manually set by a user.
M and R values are calculated for each of the five EEI types [10].
M r a n k = 1 N r i b i = 1 N r i b x i ;
R r a n k = 1 N r i b i = 1 N r i b ( M r a n k x i ) 2 ;
where   M r a n k the mean rank of ribosomal protein genes, R r a n k the standard deviation of ribosomal protein genes’ ranks, N r i b the number of ribosomal protein-coding genes, and x i is the rank of ribosomal protein-coding gene in the gene set arranged in order of increasing EEI values.
M = 100 ( 2 ( M r a n k 1 ) N t o t 1 1 ) ,
R = 100 2 ( R r a n k 1 ) N t o t 1 ,
N t o t is the total number of protein-coding genes, M is the normalized mean rank of ribosomal genes, and R is the normalized standard deviation for ranks of ribosomal genes calculated for the EEI type.
We regard the type that has the maximum M parameter as the base organism type. If there are several types sharing the maximum M value, a type with the minimum R is defined as the base type. Elongation efficiency indices for protein-coding genes of an organism are calculated by this type. In this study, we have analyzed EEIs of the base type for each organism from Table 2.

4.3. Statistical Analysis and Regression Model

To estimate the power of elongation efficiency indices as predictors of gene expression at the protein level, we calculated the correlation coefficient between experimentally measured protein abundance and base EEI for each organism; we will further refer to it as (corr(PA|EEI). As EEI indices have a rank-size distribution, we used Spearman’s rank correlation coefficient [75] with the p-value threshold for statistical significance of 0.05. Then we have also calculated the correlation between corr(PA|EEI) and other parameters (doubling time, M, R parameters). In this case, we have used Pearson’s correlation. Since the values do not correspond to the normal distribution, verification of statistical significance has been provided by the bootstrap method [76] using the “boot” package in R.
The linear regression model has been built for 25 samples (see Table 2). Predictor variable is represented by M parameter, the dependent variable is corr(PA|EEI). The linear regression model has been built using the Sklearn package in Python using the entire dataset. The quality of the model was assessed by R2, mean square error (MSE), and mean absolute error (MAE) using Monte Carlo cross validation from the cross_validate and ShuffleSplit functions from the Sklearn package with splitting the dataset 2000 times into 20% test and 80% training sets.

4.4. Minimal Doubling Time

The value of minimal doubling time for each organism has been obtained from the literature (see Table 4). If the doubling time differs depending on medium and temperature, the minimum value is selected. All values are turned into hours.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms231911996/s1.

Author Contributions

Conceptualization, A.I.K. and A.E.K.; investigation, A.E.K.; methodology, A.I.K. and Y.G.M.; supervision, A.I.K., Y.G.M. and S.A.L.; formal analysis, A.E.K.; visualization, A.E.K.; writing—original draft, A.E.K. and A.I.K.; writing—review and editing, A.E.K., A.I.K., S.A.L. and Y.G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Kurchatov Genomic Centre of the Institute of Cytology and Genetics, SB RAS (№ 075-15-2019-1662). The data analysis was performed using computational resources of the “Bioinformatics” Joint Computational Center supported by the Ministry of Science and Higher Education budget project № FWNR-2022-0020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated during this study are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Appendix A

Detailed description of Formula (2), which is used for the elongation efficiency indices calculation.

Appendix A.1. Codon Usage Calculation

In Formula (2), the effect of codon usage bias is calculated by the formula below:
T a ( i ) = j = 1 n i β δ ( i , j ) / n i ;
β δ ( i , j ) = m = 1 C a m a δ ( i ,   j ) ;
where 1 / β δ ( i , j ) reflects optimal relative concentration of aminoacyl-tRNA complementary to jth codon of the genetic code; αδ(i,j) and αm are the usage frequencies in the subset of genes for the δ(i,j) and m codons, respectively; ni—the number of codons in the gene i; C—the total number of codons.

Appendix A.2. Secondary Structures Calculation

In Formula (2), the effect of secondary structures is calculated by the formula below:
Te ( i ) = t m i n ( 1 p ( i ) ) + t m a x p ( i ) ,
tmin is the minimal time of translocation, tmax is the maximum time of translocation; p(i)—the probability of realizing the maximum translocation time for the gene i.
p ( i ) = 0 L C I ( i ) k n + 1 x n G ( n + 1 ) e k x d x ;
k = m σ 2 ;
n = ( m σ ) 2 ;
m and σ2 are expected value and variance of a positive random variable with density, respectively; G(n+1)—Gamma function, LCI(i)—local complementary index for the gene i.

Appendix A.3. Local Complementary Indices Calculation

Figure A1. The illustration of the analysis of local inverted repeats in a window m [ 1 ,   m i s m a x l m a x ] described in Formulas (A7) and (A8).
Figure A1. The illustration of the analysis of local inverted repeats in a window m [ 1 ,   m i s m a x l m a x ] described in Formulas (A7) and (A8).
Ijms 23 11996 g0a1
L C I 1 ( i ) = m = 1 m i s m a x l m a x { s = s m i n s m a x [ l = l m i n l m a x ζ ( c o n ( m ,   m + s 1 ) ,   c o n ( m + s + l , m + 2 s + l 2 ) ¯ ) ] } m i s m a x l m a x
L C I 2 ( i ) = m = 1 m i s m a x l m a x { s = s m i n s m a x [ l = l m i n l m a x ψ ( c o n ( m ,   m + s ) ,   c o n ( m + s + l 1 , m + 2 s + l 2 ) ¯ ) ] } m i s m a x l m a x
Calculation of local complementarity indexes (LCI1 and LCI2):
  • c o n ( k ,   j ) is a gene context from nucleotide k to nucleotide j, and c o n ( x ,   y ) ¯ is complementary gene context from nucleotide x to nucleotide y (x < y);
  • s is the length of the c o n ( k ,   j )   = c o n ( x ,   y ) ¯ , which is the number of nucleotides in the considered reverted repeat; it is not less than S m i n = 3 nucleotides and not higher than S m a x = 6 nucleotides;
  • l is the distance between such repeats, l m i n = 3, l m a x = 50 nucleotides;
  • m i is the length of gene i without the last three nucleotides (stop codon);
  • ζ(con1, con2) = 1 if the contexts are identical and ζ(con1, con2) = 0 in other cases;
  • ψ(con1, con2) is the energy of a hairpin formed by con1, con2, calculated conventionally.

Appendix B

Table A1. The table represents Pearson’s correlations between measured protein abundance and elongation efficiency indices calculated by EloE and corresponding p-values. These correlation coefficients have been evaluated using several EEI values: (a) EEI values which have been calculated for the base EEI type of the studied organism are presented in the “r base” column with the p-value in the “Pval base” column, and the base EEI type is presented in the “EEI type” column and represents which of the five EEI types is determined as the type of the organism under study; (b) EEI values which were calculated for the each EEI type (1–5), the correlation coefficients and the p-values presented in the “r” and “Pval” columns for each type, respectively.
Table A1. The table represents Pearson’s correlations between measured protein abundance and elongation efficiency indices calculated by EloE and corresponding p-values. These correlation coefficients have been evaluated using several EEI values: (a) EEI values which have been calculated for the base EEI type of the studied organism are presented in the “r base” column with the p-value in the “Pval base” column, and the base EEI type is presented in the “EEI type” column and represents which of the five EEI types is determined as the type of the organism under study; (b) EEI values which were calculated for the each EEI type (1–5), the correlation coefficients and the p-values presented in the “r” and “Pval” columns for each type, respectively.
OrganismEEITyper BasePval Baser1Pval 1r2Pval 2r3Pval 3r4Pval 4r5Pval 5
Bacteroides thetaiotaomicron10.572.52 × 10−670.572.52 × 10−6709.07 × 10−10.082.95 × 10−20.452.49 × 10−390.212.58 × 10−9
Deinococcus deserti10.386.58 × 10−480.386.58 × 10−480.195.05 × 10−120.164.75 × 10−90.422.77 × 10−590.371.88 × 10−43
Escherichia coli10.5700.57007.71 × 10−1−0.044.66 × 10−30.51.37 × 10−2520.386.89 × 10−142
Lactococcus lactis10.62.58 × 10−1280.62.58 × 10−1280.291.37 × 10−26−0.166.93 × 10−90.64.20 × 10−1270.041.55 × 10−1
Listeria monocytogenes10.571.46 × 10−410.571.46 × 10−410.13.24 × 10−2−0.062.02 × 10−10.533.84 × 10−350.236.85 × 10−7
Salmonella typhimurium10.451.79 × 10−1260.451.79 × 10−1260.162.70 × 10−150.171.10 × 10−180.468.48 × 10−1330.444.92 × 10−123
Shigella flexneri10.654.01 × 10−2020.654.01 × 10−2020.054.94 × 10−20.245.17 × 10−240.628.03 × 10−1780.61.71 × 10−162
Staphylococcus aureus10.668.46 × 10−2110.668.46 × 10−2110.227.89 × 10−20−0.164.63 × 10−110.592.74 × 10−1590.053.43 × 10−2
Streptococcus pyogenes10.633.60 × 10−1410.633.60 × 10−1410.086.81 × 10−3−0.122.22 × 10−050.593.28 × 10−1210.22.36 × 10−13
Synechocystis sp.10.48.01 × 10−480.48.01 × 10−480.091.02 × 10−3−0.032.71 × 10−10.32.11 × 10−260.112.02 × 10−4
Thermococcus gammatolerans10.443.71 × 10−660.443.71 × 10−66−0.024.18 × 10−1−0.163.25 × 10−90.376.11 × 10−440.116.38 × 10−5
Yersinia pestis10.44.86 × 10−470.44.86 × 10−470.017.23 × 10−10.064.22 × 10−20.366.63 × 10−390.312.53 × 10−27
Acidithiobacillusferrooxidans20.121.73 × 10−50.226.07 × 10−160.121.73 × 10−50.157.05 × 10−80.357.86 × 10−390.342.38 × 10−36
Campylobacter jejuni20.466.48 × 10−42−0.153.25 × 10−50.466.48 × 10−42−0.218.67 × 10−90.345.60 × 10−22−0.241.93 × 10−11
Helicobacter pylori20.281.22 × 10−290.075.65 × 10−30.281.22 × 10−29−0.163.60 × 10−100.391.06 × 10−57−0.124.31 × 10−6
Leptospira interrogans20.356.06 × 10−66−0.282.65 × 10−400.356.06 × 10−66−0.166.61 × 10−150.468.46 × 10−117−0.252.45 × 10−34
Mycoplasma pneumoniae20.241.60 × 10−60.062.14 × 10−10.241.60 × 10−60.018.88 × 10−10.263.25 × 10−70.036.04 × 10−1
Pseudomonas aeruginosa30.271.37 × 10−420.091.76 × 10−50.252.09 × 10−360.271.37 × 10−420.281.24 × 10−440.36.09 × 10−53
Bacillus anthracis40.523.42 × 10−1020.533.42 × 10−1020.11.86 × 10−4−0.171.57 × 10−100.521.01 × 10−94−0.11.39 × 10−4
Bartonella henselae40.351.21 × 10−410.371.21 × 10−410.091.91 × 10−3−0.064.79 × 10−20.351.15 × 10−380.132.43 × 10−6
Desulfovibrio vulgaris40.393.87 × 10−370.43.87 × 10−370.135.22 × 10−50.183.23 × 10−80.395.07 × 10−360.368.71 × 10−31
Halobacterium salinarum40.332.00 × 10−20.082.00 × 10−20.171.00 × 10−50.151.00 × 10−70.331.00 × 10−90.281.00 × 10−22
Legionella pneumophila40.422.50 × 10−50.152.50 × 10−50.174.34 × 10−6−0.033.57 × 10−10.422.72 × 10−3209.27 × 10−1
Microcystis aeruginosa40.141.14 × 10−83−0.271.14 × 10−830.175.59 × 10−35−0.173.95 × 10−320.141.43 × 10−22−0.291.43 × 10−96
Mycobacterium tuberculosis40.263.44 × 10−280.193.44 × 10−280.051.48 × 10−30.054.60 × 10−30.261.45 × 10−520.213.94 × 10−36
Neisseria meningitidis50.097.73 × 10−20.088.93 × 10−20.071.41 × 10−10.035.30 × 10−10.043.88 × 10−10.097.73 × 10−2
Table A2. The table represents M (the normalized mean rank of ribosomal genes) and R (the normalized standard deviation for ranks of ribosomal genes) calculated for each EEI type (1–5), the EEI type, and M and R values for the EEI type (the columns M_main, R_main).
Table A2. The table represents M (the normalized mean rank of ribosomal genes) and R (the normalized standard deviation for ranks of ribosomal genes) calculated for each EEI type (1–5), the EEI type, and M and R values for the EEI type (the columns M_main, R_main).
OrganismEEI TypeM1R1M2R2M3R3M4R4M5R5M_MainR_Main
Acidithiobacillus ferrooxidans225442473648345142474247
Bacillus anthracis476483846−43557746−48517746
Bacteroides thetaiotaomicron19120−14581760713850419120
Bartonella henselae454423253−557614116606141
Campylobacter jejuni2−43546737−25573467−49426737
Deinococcus deserti1853130531956842264498728
Desulfovibrio vulgaris4613442403847791970387918
Escherichia coli1873013631453823475368730
Halobacterium salinarum4−6361441542364129433641
Helicobacter pylori2−33555144−10633252−24595144
Lactococcus lactis176494650−37627552−27687649
Legionella pneumophila416612761−11596643−10576643
Leptospira interrogans2−61455942−38534952−61345942
Listeria monocytogenes179362859−2362773813617936
Microcystis aeruginosa4−42394448−46455546−53335546
Mycobacterium tuberculosis4−7601856861366329693663
Mycoplasma pneumoniae2−11553454−28582960−30523454
Pseudomonas aeruginosa3−752783188317254344348317
Salmonella typhimurium1844024633047824282388539
Shigella flexneri1956664753871979359412
Staphylococcus aureus183254847−32628129−18638325
Streptococcus pyogenes19126−462−2559882726719126
Synechocystis sp.153511960−30584051−13655351
Thermococcus gammatolerans17732065−375463473677732
Yersinia pestis19126−165−459823460569126

References

  1. Hui, S.; Silverman, J.M.; Chen, S.S.; Erickson, D.W.; Basan, M.; Wang, J.; Hwa, T.; Williamson, J.R. Quantitative proteomic analysis reveals a simple strategy of global resource allocation in bacteria. Mol. Syst. Biol. 2015, 11, 784. [Google Scholar] [CrossRef]
  2. Mehdi, A.M.; Patrick, R.; Bailey, T.L.; Boden, M. Predicting the dynamics of protein abundance. Mol. Cell. Proteom. 2014, 13, 1330–1340. [Google Scholar] [CrossRef] [Green Version]
  3. Magnusson, R.; Rundquist, O.; Kim, M.J.; Hellberg, S.; Na, C.H.; Benson, M.; Gomez-Cabrero, D.; Kockum, I.; Tegnér, J.; Piehl, F.; et al. On the prediction of protein abundance from RNA. bioRxiv 2019. [Google Scholar] [CrossRef] [Green Version]
  4. Ferreira, M.; Ventorim, R.; Almeida, E.; Silveira, S.; Silveira, W. Protein Abundance Prediction Through Machine Learning Methods. J. Mol. Biol. 2021, 433, 167267. [Google Scholar] [CrossRef]
  5. Guimaraes, J.C.; Rocha, M.; Arkin, A.P. Transcript level and sequence determinants of protein abundance and noise in Escherichia coli. Nucleic Acids Res. 2014, 42, 4791–4799. [Google Scholar] [CrossRef] [Green Version]
  6. Kudla, G.; Murray, A.W.; Tollervey, D.; Plotkin, J.B. Coding-sequence determinants of gene expression in Escherichia coli. Science 2009, 324, 255–258. [Google Scholar] [CrossRef] [Green Version]
  7. Sokolov, V.; Zuraev, B.; Lashin, S.; Matushkin, Y. Web application for automatic prediction of gene translation elongation efficiency. J. Integr. Bioinform. 2015, 12, 256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Likhoshvai, A.; Matushkin, Y.G. Nucleotide composition-based prediction of gene expression efficacy. Mol. Biol. 2000, 34, 397–405. [Google Scholar] [CrossRef]
  9. Likhoshvai, V.A.; Matushkin, Y.G. Differentiation of single-cell organisms according to elongation stages crucial for gene expression efficacy. FEBS Lett. 2002, 516, 87–92. [Google Scholar] [CrossRef] [Green Version]
  10. Vladimirov, N.V.; Likhoshvai, V.A.; Matushkin, Y.G. Correlation of codon biases and potential secondary structures with mRNA translation efficiency in unicellular organisms. Mol. Biol. 2007, 41, 843–850. [Google Scholar] [CrossRef]
  11. Sharp, P.M.; Tuohy, T.M.F.; Mosurski, K.R. Codon usage in yeast: Cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res. 1986, 14, 5125–5143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Sørensen, M.A.; Kurland, C.G.; Pedersen, S. Codon usage determines translation rate in Escherichia coli. J. Mol. Biol. 1989, 207, 365–377. [Google Scholar] [CrossRef]
  13. Wei, Y.; Silke, J.R.; Xia, X. An improved estimation of tRNA expression to better elucidate the coevolution between tRNA abundance and codon usage in bacteria. Sci. Rep. 2019, 9, 3184. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, S.E.; Brooks, A.E.S.; Poole, A.M.; Simoes-Barbosa, A. Determinants of translation efficiency in the evolutionarily-divergent protist Trichomonas vaginalis. BMC Mol. Cell Biol. 2020, 21, 54. [Google Scholar] [CrossRef]
  15. Cambray, G.; Guimaraes, J.C.; Arkin, A.P. Evaluation of 244,000 synthetic sequences reveals design principles to optimize translation in escherichia coli. Nat. Biotechnol. 2018, 36, 1005–1015. [Google Scholar] [CrossRef]
  16. Mohammad, F.; Green, R.; Buskirk, A.R. A systematically-revised ribosome profiling method for bacteria reveals pauses at single-codon resolution. eLife 2019, 8, 1–25. [Google Scholar] [CrossRef] [PubMed]
  17. Hia, F.; Takeuchi, O. The effects of codon bias and optimality on mRNA and protein regulation. Cell. Mol. Life Sci. 2021, 78, 1909–1928. [Google Scholar] [CrossRef]
  18. Frumkin, I.; Lajoie, M.J.; Gregg, C.J.; Hornung, G.; Church, G.M.; Pilpel, Y. Codon usage of highly expressed genes affects proteome-wide translation efficiency. Proc. Natl. Acad. Sci. USA 2018, 115, E4940–E4949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Jeacock, L.; Faria, J.; Horn, D. Codon usage bias controls mRNA and protein abundance in trypanosomatids. eLife 2018, 7, 1–20. [Google Scholar] [CrossRef] [PubMed]
  20. Boël, G.; Letso, R.; Neely, H.; Price, W.N.; Su, M.; Luff, J.; Valecha, M.; Everett, J.K.; Acton, T.B.; Xiao, R.; et al. Codon influence on protein expression in E.coli. Nature 2016, 529, 358–363. [Google Scholar] [CrossRef]
  21. Plotkin, J.B.; Kudla, G. Synonymous but not the same: The causes and consequences of codon bias. Nat. Rev. Genet. 2011, 12, 32–42. [Google Scholar] [CrossRef] [Green Version]
  22. Iriarte, A.; Lamolle, G.; Musto, H. Codon Usage Bias: An Endless Tale. J. Mol. Evol. 2021, 89, 589–593. [Google Scholar] [CrossRef] [PubMed]
  23. Parvathy, S.T.; Udayasuriyan, V.; Bhadana, V. Codon usage bias. Mol. Biol. Rep. 2022, 49, 539–565. [Google Scholar] [CrossRef] [PubMed]
  24. Andersson, S.G.E.; Sharp, P.M. Codon usage and base composition in Rickettsia prowazekii. J. Mol. Evol. 1996, 42, 525–536. [Google Scholar] [CrossRef]
  25. Lafay, B.; Lloyd, A.T.; McLean, M.J.; Devine, K.M.; Sharp, P.M.; Wolfe, K.H. Proteome composition and codon usage in spirochaetes: Species-specific and DNA strand-specific mutational biases. Nucleic Acids Res. 1999, 27, 1642–1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Frutos, R.; Viari, A.; Ferraz, C.; Bensaid, A.; Morgat, A.; Boyer, F.; Coissac, E.; Vachiéry, N.; Demaille, J.; Martinez, D. Comparative genomics of three strains of Ehrlichia ruminantium: A review. Ann. N. Y. Acad. Sci. 2006, 1081, 417–433. [Google Scholar] [CrossRef] [PubMed]
  27. Fuglsang, A. Intragenic codon usage in proteobacteria: Translational selection, IS expansion and genomic shrinkage. Gene 2021, 809, 146015. [Google Scholar] [CrossRef] [PubMed]
  28. Lafay, B.; Atherton, J.C.; Sharp, P.M. Absence of translationally selected synonymous codon usage bias in Helicobacter pylori. Microbiology 2000, 146 Pt 4, 851–860. [Google Scholar] [CrossRef] [Green Version]
  29. Rispe, C.; Delmotte, F.; van Ham, R.C.H.J.; Moya, A. Mutational and Selective Pressures on Codon and Amino Acid Usage in Buchnera, Endosymbiotic Bacteria of Aphids. Genome Res. 2004, 14, 44–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Sharp, P.M.; Bailes, E.; Grocock, R.J.; Peden, J.F.; Sockett, E. Variation in the strength of selected codon usage bias among bacteria. Nucleic Acids Res. 2005, 33, 1141–1153. [Google Scholar] [CrossRef] [PubMed]
  31. Sokolov, V.S.; Likhoshvai, V.A.; Matushkin, Y.G. Gene expression and secondary mRNA structures in different Mycoplasma species. Russ. J. Genet. Appl. Res. 2014, 4, 208–217. [Google Scholar] [CrossRef]
  32. Sharp, P.M.; Li, W.-H. The codon adaptation index-a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 1987, 15, 1281–1295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Lee, B.D. Python Implementation of Codon Adaptation Index. J. Open Source Softw. 2018, 3, 905. [Google Scholar] [CrossRef] [Green Version]
  34. Anwar, A. BCAWT: Automated tool for codon usage bias analysis for molecular evolution. J. Open Source Softw. 2019, 4, 1500. [Google Scholar] [CrossRef] [Green Version]
  35. Carbone, A.; Zinovyev, A.; Kepes, F. Codon adaptation index as a measure of dominating codon bias. Bioinformatics 2003, 19, 2005–2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Hanson, G.; Coller, J. Codon optimality, bias and usage in translation and mRNA decay. Nat. Rev. Mol. Cell Biol. 2018, 19, 20–30. [Google Scholar] [CrossRef] [PubMed]
  37. Pop, C.; Rouskin, S.; Ingolia, N.T.; Han, L.; Phizicky, E.M.; Weissman, J.S.; Koller, D. Causal signals between codon bias, mRNA structure, and the efficiency of translation and elongation. Mol. Syst. Biol. 2014, 10, 770. [Google Scholar] [CrossRef]
  38. Shah, P.; Ding, Y.; Niemczyk, M.; Kudla, G.; Plotkin, J.B. XRate-limiting steps in yeast protein translation. Cell 2013, 153, 1589. [Google Scholar] [CrossRef] [Green Version]
  39. Quax, T.E.F.; Claassens, N.J.; Söll, D.; van der Oost, J. Codon Bias as a Means to Fine-Tune Gene Expression. Mol. Cell. 2015, 59, 149–161. [Google Scholar] [CrossRef] [Green Version]
  40. Chiaruttini, C.; Guillier, M. On the role of mRNA secondary structure in bacterial translation. Wiley Interdiscip. Rev. RNA 2019, 11, 1–21. [Google Scholar] [CrossRef]
  41. Thanaraj, T.A.; Argos, P. Ribosome-mediated translational pause and protein domain organization. Protein Sci. 1996, 5, 1594–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Wen, J.-D.; Lancaster, L.; Hodges, C.; Zeri, A.-C.; Yoshimura, S.H.; Noller, H.F.; Bustamante, C.; Tinoco, I. Following translation by single ribosomes one codon at a time. Nature 2008, 452, 598–603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Qu, X.; Wen, J.-D.; Lancaster, L.; Noller, H.F.; Bustamante, C.; Tinoco, I.J. The ribosome uses two active mechanisms to unwind messenger RNA during translation. Nature 2011, 475, 118–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Xie, P. Model of ribosome translation and mRNA unwinding. Eur. Biophys. J. 2013, 42, 347–354. [Google Scholar] [CrossRef]
  45. Xie, P.; Chen, H. Mechanism of ribosome translation through mRNA secondary structures. Int. J. Biol. Sci. 2017, 13, 712–722. [Google Scholar] [CrossRef]
  46. Takyar, S.; Hickerson, R.P.; Noller, H.F. mRNA Helicase Activity of the Ribosome. Cell 2005, 120, 49–58. [Google Scholar] [CrossRef] [Green Version]
  47. Burkhardt, D.H.; Rouskin, S.; Zhang, Y.; Li, G.W.; Weissman, J.S.; Gross, C.A. Operon mRNAs are organized into ORF-centric structures that predict translation efficiency. eLife 2017, 6, 1–23. [Google Scholar] [CrossRef]
  48. Kertesz, M.; Wan, Y.; Mazor, E.; Rinn, J.L.; Nutter, R.C.; Chang, H.Y.; Segal, E. Genome-wide Measurement of RNA Secondary Structure in Yeast. Nature 2010, 467, 103–107. [Google Scholar] [CrossRef] [Green Version]
  49. Faure, G.; Ogurtsov, A.Y.; Shabalina, S.A.; Koonin, E.V. Role of mRNA structure in the control of protein folding. Nucleic Acids Res. 2016, 44, 10898–10911. [Google Scholar] [CrossRef] [Green Version]
  50. Mauger, D.M.; Joseph Cabral, B.; Presnyak, V.; Su, S.V.; Reid, D.W.; Goodman, B.; Link, K.; Khatwani, N.; Reynders, J.; Moore, M.J.; et al. mRNA structure regulates protein expression through changes in functional half-life. Proc. Natl. Acad. Sci. USA 2019, 116, 24075–24083. [Google Scholar] [CrossRef]
  51. Zhang, Q.; Ma, D.; Wu, F.; Standage-Beier, K.; Chen, X.; Wu, K.; Green, A.A.; Wang, X. Predictable control of RNA lifetime using engineered degradation-tuning RNAs. Nat. Chem. Biol. 2021, 17, 828–836. [Google Scholar] [CrossRef] [PubMed]
  52. Mohanty, B.K.; Kushner, S.R. Regulation of mRNA Decay in Bacteria. Annu. Rev. Microbiol. 2016, 70, 25–44. [Google Scholar] [CrossRef]
  53. Jacks, T.; Madhani, H.D.; Masiarz, F.R.; Varmus, H.E. Signals for ribosomal frameshifting in the rous sarcoma virus gag-pol region. Cell 1988, 55, 447–458. [Google Scholar] [CrossRef]
  54. Lopinski, J.D.; Dinman, J.D.; Bruenn, J.A. Kinetics of Ribosomal Pausing during Programmed −1 Translational Frameshifting. Mol. Cell. Biol. 2000, 20, 1095–1103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Chemla, Y.; Peeri, M.; Heltberg, M.L.; Eichler, J.; Jensen, M.H.; Tuller, T.; Alfonta, L. A possible universal role for mRNA secondary structure in bacterial translation revealed using a synthetic operon. Nat. Commun. 2020, 11, 4827. [Google Scholar] [CrossRef] [PubMed]
  56. Gorochowski, T.E.; Ignatova, Z.; Bovenberg, R.A.L.; Roubos, J.A. Trade-offs between tRNA abundance and mRNA secondary structure support smoothing of translation elongation rate. Nucleic Acids Res. 2015, 43, 3022–3032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Gebert, D.; Jehn, J.; Rosenkranz, D. Widespread selection for extremely high and low levels of secondary structure in coding sequences across all domains of life. Open Biol. 2019, 9, 190020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Victor, M.P.; Acharya, D.; Begum, T.; Ghosh, T.C. The optimization of mRNA expression level by its intrinsic properties—Insights from codon usage pattern and structural stability of mRNA. Genomics 2019, 111, 1292–1297. [Google Scholar] [CrossRef] [PubMed]
  59. Matushkin, Y.; Institute of Cytology and Genetics, Siberian Branch of the Russian Academy of Science, Russia. Personal communication, 2022.
  60. Letunic, I.; Bork, P. Interactive tree of life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef] [PubMed]
  61. Gibson, B.; Wilson, D.J.; Feil, E.; Eyre-Walker, A. The distribution of bacterial doubling times in the wild. Proc. R. Soc. B Biol. Sci. 2018, 285. [Google Scholar] [CrossRef] [PubMed]
  62. Neidhardt, F.C.; Magasanik, B. Studies on the role of ribonucleic acid in the growth of bacteria. Biochim. Biophys. Acta 1960, 42, 99–116. [Google Scholar] [CrossRef]
  63. Dai, X.; Zhu, M.; Warren, M.; Balakrishnan, R.; Patsalo, V.; Okano, H.; Williamson, J.R.; Fredrick, K.; Wang, Y.P.; Hwa, T. Reduction of translating ribosomes enables Escherichia coli to maintain elongation rates during slow growth. Nat. Microbiol. 2016, 2, 1–9. [Google Scholar] [CrossRef] [Green Version]
  64. Higgs, P.G.; Ran, W. Coevolution of codon usage and tRNA genes leads to alternative stable states of biased codon usage. Mol. Biol. Evol. 2008, 25, 2279–2291. [Google Scholar] [CrossRef] [Green Version]
  65. Vieira-Silva, S.; Rocha, E.P.C. The systemic imprint of growth and its uses in ecological (meta)genomics. PLoS Genet. 2010, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Weissman, J.L.; Hou, S.; Fuhrman, J.A. Estimating maximal microbial growth rates from cultures, metagenomes, and single cells via codon usage patterns. Proc. Natl. Acad. Sci. USA 2021, 118, 1–10. [Google Scholar] [CrossRef]
  67. Riba, A.; Nanni, N. Di, Mittal, N.; Arhné, E.; Schmidt, A.; Zavolan, M. Protein synthesis rates and ribosome occupancies reveal determinants of translation elongation rates. Proc. Natl. Acad. Sci. USA 2019, 116, 15023–15032. [Google Scholar] [CrossRef] [Green Version]
  68. Ishihama, Y.; Schmidt, T.; Rappsilber, J.; Mann, M.; Harlt, F.U.; Kerner, M.J.; Frishman, D. Protein abundance profiling of the Escherichia coli cytosol. BMC Genom. 2008, 9, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Maurizi, M.R. Proteases and protein degradation in Escherichia coli. Experientia 1992, 48, 178–201. [Google Scholar] [CrossRef] [PubMed]
  70. Goodman, D.B.; Church, G.M.; Kosuri, S. Causes and effects of N-terminal codon bias in bacterial genes. Science 2013, 342, 475–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Samatova, E.; Daberger, J.; Liutkute, M.; Rodnina, M.V. Translational Control by Ribosome Pausing in Bacteria: How a Non-uniform Pace of Translation Affects Protein Production and Folding. Front. Microbiol. 2021, 11. [Google Scholar] [CrossRef] [PubMed]
  72. Sokolov, V.S.; Likhoshvai, V.A.; Matushkun, Y.G. Gene expression and mRNA secondary structures in different Mycoplasma species. Vavilov J. Genet. Breed. 2013, 17, 639–650. [Google Scholar] [CrossRef]
  73. Wang, M.; Herrmann, C.J.; Simonovic, M.; Szklarczyk, D.; von Mering, C. Version 4.0 of PaxDb: Protein abundance data, integrated across model organisms, tissues, and cell-lines. Proteomics 2015, 15, 3163–3168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kitts, P.A.; Church, D.M.; Thibaud-Nissen, F.; Choi, J.; Hem, V.; Sapojnikov, V.; Smith, R.G.; Tatusova, T.; Xiang, C.; Zherikov, A.; et al. Assembly: A resource for assembled genomes at NCBI. Nucleic Acids Res. 2016, 44, D73–D80. [Google Scholar] [CrossRef]
  75. Wissler, C. The Spearman correlation formula. Science 1905, 22, 309–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Efron, B.; Tibshirani, R.J. An Introduction to the Bootstrap; Chapman & Hall/CRC: Philadelphia, PA, USA, 1994. [Google Scholar]
  77. Kai, T.; Nagano, T.; Fukumoto, T.; Nakajima, M.; Takahashi, T. Autotrophic growth of Acidithiobacillus ferrooxidans by oxidation of molecular hydrogen using a gas-liquid contactor. Bioresour. Technol. 2007, 98, 460–464. [Google Scholar] [CrossRef]
  78. Bugrysheva, J.V.; Lascols, C.; Sue, D.; Weigel, L.M. Rapid antimicrobial susceptibility testing of bacillus anthracis, yersinia pestis, and burkholderia pseudomallei by use of laser light scattering technology. J. Clin. Microbiol. 2016, 54, 1462–1471. [Google Scholar] [CrossRef] [Green Version]
  79. Sonnenburg, E.D.; Zheng, H.; Joglekar, P.; Higginbottom, S.K.; Firbank, S.J.; Bolam, D.N.; Sonnenburg, J.L. Specificity of polysaccharide use in intestinal bacteroides species determines diet-induced microbiota alterations. Cell 2010, 141, 1241–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Chenoweth, M.R.; Somerville, G.A.; Krause, D.C.; O’Reilly, K.L.; Gherardini, F.C. Growth Characteristics of Bartonella henselae in a Novel Liquid Medium: Primary Isolation, Growth-Phase-Dependent Phage Induction, and Metabolic Studies. Appl. Environ. Microbiol. 2004, 70, 656–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Ducati, R.G.; Harijan, R.K.; Cameron, S.A.; Tyler, P.C.; Evans, G.B.; Schramm, V.L. Transition-State Analogues of Campylobacter jejuni 5′-Methylthioadenosine Nucleosidase. ACS Chem. Biol. 2018, 13, 3173–3183. [Google Scholar] [CrossRef]
  82. Bornot, J.; Molina-Jouve, C.; Uribelarrea, J.L.; Gorret, N. Quantitative characterization of the growth of deinococcus geothermalis dsm-11302: Effect of inoculum size, growth medium and culture conditions. Microorganisms 2015, 3, 441–463. [Google Scholar] [CrossRef]
  83. Fievet, A.; Ducret, A.; Mignot, T.; Valette, O.; Robert, L.; Pardoux, R.; Dolla, A.R.; Aubert, C. Single-cell analysis of growth and cell division of the anaerobe Desulfovibrio vulgaris hildenborough. Front. Microbiol. 2015, 6, 1378. [Google Scholar] [CrossRef] [Green Version]
  84. Gonzalez, O.; Gronau, S.; Pfeiffer, F.; Mendoza, E.; Zimmer, R.; Oesterhelt, D. Systems Analysis of Bioenergetics and Growth of the Extreme Halophile Halobacterium salinarum. PLoS Comput. Biol. 2009, 5. [Google Scholar] [CrossRef] [Green Version]
  85. Andersen, A.P.; Elliott, D.A.; Lawson, M.; Barland, P.; Hatcher, V.B.; Puszkin, E.G. Growth and morphological transformations of Helicobacter pylori in broth media. J. Clin. Microbiol. 1997, 35, 2918–2922. [Google Scholar] [CrossRef] [Green Version]
  86. Holubová, J.; Josephsen, J. Potential of AbiS as defence mechanism determined by conductivity measurement. J. Appl. Microbiol. 2007, 103, 2382–2391. [Google Scholar] [CrossRef]
  87. O’Connor, T.J.; Adepoju, Y.; Boyd, D.; Isberg, R.R. Minimization of the Legionella pneumophila genome reveals chromosomal regions involved in host range expansion. Proc. Natl. Acad. Sci. USA 2011, 108, 14733–14740. [Google Scholar] [CrossRef] [Green Version]
  88. Ratet, G.; Veyrier, F.J.; Fanton d’Andon, M.; Kammerscheit, X.; Nicola, M.A.; Picardeau, M.; Boneca, I.G.; Werts, C. Live Imaging of Bioluminescent Leptospira interrogans in Mice Reveals Renal Colonization as a Stealth Escape from the Blood Defenses and Antibiotics. PLoS Negl. Trop. Dis. 2014, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Glomski, I.J.; Decatur, A.L.; Portnoy, D.A. Listeria monocytogenes Mutants That Fail to Compartmentalize Listerolysin O Activity Are Cytotoxic, Avirulent, and Unable to Evade Host Extracellular Defenses. Infect. Immun. 2003, 71, 6754–6765. [Google Scholar] [CrossRef] [Green Version]
  90. Pereira, V. Isolation, Culture and Morphological Characterization of Microcystis Sp Toxic Strain From the Tacuary Reservoir. Int. J. Adv. Res. 2018, 6, 387–393. [Google Scholar] [CrossRef] [Green Version]
  91. James, B.W.; Williams, A.; Marsh, P.D. The physiology and pathogenicity of Mycobacterium tuberculosis grown under controlled conditions in a defined medium. J. Appl. Microbiol. 2000, 88, 669–677. [Google Scholar] [CrossRef] [PubMed]
  92. Gaspari, E.; Malachowski, A.; Garcia-Morales, L.; Burgos, R.; Serrano, L.; Martins dos Santos, V.A.P.; Suarez-Diez, M. Model-driven design allows growth of Mycoplasma pneumoniae on serum-free media. npj Syst. Biol. Appl. 2020, 6. [Google Scholar] [CrossRef]
  93. McBirney, S.E.; Trinh, K.; Wong-Beringer, A.; Armani, A.M. Wavelength-normalized spectroscopic analysis of Staphylococcus aureus and Pseudomonas aeruginosa growth rates. Biomed. Opt. Express 2016, 7, 4034–4042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Abshire, K.Z.; Neidhardt, F.C. Growth rate paradox of Salmonella typhimurium within host macrophages. J. Bacteriol. 1993, 175, 3744–3748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lucchini, S.; Liu, H.; Jin, Q.; Hinton, J.C.D.; Yu, J. Transcriptional adaptation of Shigella flexneri during infection of macrophages and epithelial cells: Insights into the strategies of a cytosolic bacterial pathogen. Infect. Immun. 2005, 73, 88–102. [Google Scholar] [CrossRef] [Green Version]
  96. Gera, K.; McIver, K.S. Laboratory growth and maintenance of streptococcus pyogenes (The Group A Streptococcus, GAS). Curr. Protoc. Microbiol. 2013, 30, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Touloupakis, E.; Cicchi, B.; Torzillo, G. A bioenergetic assessment of photosynthetic growth of Synechocystis sp. PCC 6803 in continuous cultures. Biotechnol. Biofuels 2015, 8, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Lagorce, A.; Fourçans, A.; Dutertre, M.; Bouyssiere, B.; Zivanovic, Y.; Confalonieri, F. Genome-wide transcriptional response of the Archaeon Thermococcus gammatolerans to Cadmium. PLoS ONE 2012, 7. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Spearman’s correlation coefficients between protein abundance and EEI for 25 prokaryotes.
Figure 1. Spearman’s correlation coefficients between protein abundance and EEI for 25 prokaryotes.
Ijms 23 11996 g001
Figure 2. Spearman’s correlation coefficients between protein abundance and EEI distribution among 5 EEI types for 25 prokaryotes (Neisseria meningitidis, the organism with p-value > 0.05, is excluded).
Figure 2. Spearman’s correlation coefficients between protein abundance and EEI distribution among 5 EEI types for 25 prokaryotes (Neisseria meningitidis, the organism with p-value > 0.05, is excluded).
Ijms 23 11996 g002
Figure 3. The distributions of the Spearman’s correlation coefficient values between protein abundance and EEI for the organisms belonged to the different elongation efficiency optimization types that take into account: number of secondary structures (EEI2, panel (a)), energy of secondary structures (EEI3, panel (b)), codon bias and number of secondary structures (EEI4, panel (c)). All these distributions are compared with the correlation between protein abundance and indices for codon bias-based index (EEI1) calculated for the same organisms.
Figure 3. The distributions of the Spearman’s correlation coefficient values between protein abundance and EEI for the organisms belonged to the different elongation efficiency optimization types that take into account: number of secondary structures (EEI2, panel (a)), energy of secondary structures (EEI3, panel (b)), codon bias and number of secondary structures (EEI4, panel (c)). All these distributions are compared with the correlation between protein abundance and indices for codon bias-based index (EEI1) calculated for the same organisms.
Ijms 23 11996 g003
Figure 4. The distribution of the analyzed 25 organisms by taxonomical categories. Arrows point to corresponding higher-order taxa of the analyzed species. A number near a species name corresponds to the base EEI type and a colored square shows a Spearman’s correlation (corr(PA|EEI)) coefficient value (see the legend).
Figure 4. The distribution of the analyzed 25 organisms by taxonomical categories. Arrows point to corresponding higher-order taxa of the analyzed species. A number near a species name corresponds to the base EEI type and a colored square shows a Spearman’s correlation (corr(PA|EEI)) coefficient value (see the legend).
Ijms 23 11996 g004
Figure 5. (a) Dependence of corr(PA|EEI) coefficient from the logarithm of minimal doubling time for 25 organisms. The color describes EEI type (see the legend). The trend line is colored purple; (b) distribution of bacteria with diverse growth rates (with the minimal doubling time <2 h, ≥2 h and <5 h, and ≥5 h for the fast, medium, and slow growing bacteria, respectively) by the M parameter.
Figure 5. (a) Dependence of corr(PA|EEI) coefficient from the logarithm of minimal doubling time for 25 organisms. The color describes EEI type (see the legend). The trend line is colored purple; (b) distribution of bacteria with diverse growth rates (with the minimal doubling time <2 h, ≥2 h and <5 h, and ≥5 h for the fast, medium, and slow growing bacteria, respectively) by the M parameter.
Ijms 23 11996 g005
Figure 6. Dependence between corr(PA|EEI) and M (mean ribosome protein-coding gene rank) and R (standard deviation of ribosome protein-coding genes’ rank) parameters.
Figure 6. Dependence between corr(PA|EEI) and M (mean ribosome protein-coding gene rank) and R (standard deviation of ribosome protein-coding genes’ rank) parameters.
Ijms 23 11996 g006
Table 1. Values of the analyzed parameters for the studied organisms: elongation efficiency type (EEI type), which was obtained by EloE (see the details in Materials and Methods section); coverage of proteomic data; Spearman correlation coefficients between protein abundance and base EEI index; corresponding p-value; minimal doubling time (see the references in Table 4); mean (M) and standard deviation (R) values of ranks of ribosomal protein genes measured on the base EEI scale.
Table 1. Values of the analyzed parameters for the studied organisms: elongation efficiency type (EEI type), which was obtained by EloE (see the details in Materials and Methods section); coverage of proteomic data; Spearman correlation coefficients between protein abundance and base EEI index; corresponding p-value; minimal doubling time (see the references in Table 4); mean (M) and standard deviation (R) values of ranks of ribosomal protein genes measured on the base EEI scale.
OrganismEEI TypeCoverageCorrelation Coefficientp-ValueDoubling Time (h)M_MainR_Main
Staphylococcus aureus162.50.668.46 × 10−2110.48325
Shigella flexneri139.40.654.01 × 10−2020.59412
Streptococcus pyogenes175.90.633.60 × 10−1410.66679126
Lactococcus lactis157.10.602.58 × 10−1280.57649
Bacteroides thetaiotaomicron115.90.572.52 × 10−672.79120
Listeria monocytogenes116.40.571.46 × 10−410.51677936
Escherichia coli197.400.5700.33338730
Bacillus anthracis426.200.523.42 × 10−1020.57746
Campylobacter jejuni247.400.466.48 × 10−422.46676737
Salmonella typhimurium156.30.451.79 × 10−1260.58539
Thermococcus gammatolerans162.20.443.71 × 10−664.57732
Legionella pneumophila425.20.422.50 × 10−053.36643
Synechocystis sp.137.80.408.01 × 10−485.85351
Yersinia pestis129.60.404.86 × 10−4719126
Desulfovibrio vulgaris427.10.393.87 × 10−372.487918
Deinococcus deserti138.50.386.58 × 10−482.68728
Bartonella henselae485.70.351.21 × 10−4136141
Leptospira interrogans266.20.356.06 × 10−668.25942
Halobacterium salinarum454.20.332.00 × 10−02113641
Helicobacter pylori298.80.281.22 × 10−290.83335144
Pseudomonas aeruginosa343.60.271.37 × 10−420.58317
Mycobacterium tuberculosis4840.263.44 × 10−2814.73663
Microcystis aeruginosa479.000.241.60 × 10−06465546
Mycoplasma pneumoniae260.90.141.14 × 10−8383454
Acidithiobacillus ferrooxidans241.90.121.73 × 10−0554247
Table 2. The list of species under study (species for which proteomic data were collected) and corresponding assembly accessions.
Table 2. The list of species under study (species for which proteomic data were collected) and corresponding assembly accessions.
SpeciesAssembly Accession
1Acidithiobacillus ferrooxidans ATCC23270GCF_000021485.1
2Bacillus anthracis str. SterneGCF_000008165.1
3Bacteroides thetaiotaomicron VPI-5482GCF_000011065.1
4Bartonella henselae str. Houston-1GCF_000046705.1
5Campylobacter jejuni NCTC11168GCF_000009085.1
6Deinococcus deserti VCD115GCF_000020685.1
7Desulfovibrio vulgaris str. HildenboroughGCF_000195755.1
8Escherichia coli K12 MG1655GCF_000005845.2
9Halobacterium salinarum NRC-1GCF_000006805.1
10Helicobacter pylori 26695GCF_000008525.1
11Lactococcus lactis subsp. lactis Il1403GCF_000006865.1
12Legionella pneumophila subsp. pneumophila str. Philadelphia 1GCF_000008485.1
13Leptospira interrogans serovar Lai str. 56601GCF_000007685.1
14Listeria monocytogenes EGD-eGCF_000196035.1
15Microcystis aeruginosa NIES-843GCF_000010625.1
16Mycobacterium tuberculosis H37RvGCF_000195955.2
17Mycoplasma pneumoniae FHGCF_001272835.1
18Neisseria meningitidis MC58GCF_000008805.1
19Pseudomonas aeruginosa PAO1GCF_000006765.1
20Salmonella enterica subsp. enterica serovar Typhimurium str. LT2GCF_000006945.2
21Shigella flexneri 2a str. 301GCF_000006925.2
22Staphylococcus aureusGCF_000009665.1
23Streptococcus pyogenesGCF_000006785.2
24Synechocystis sp. PCC 6803GCF_000009725.1
25Thermococcus gammatolerans EJ3GCF_000022365.1
26Yersinia pestis CO92 (enterobacteria)GCF_000009065.1
Table 3. The description of elongation efficiency types (EEI types) calculation.
Table 3. The description of elongation efficiency types (EEI types) calculation.
TypeCodon Usage
( T a ( i ) )
Local Complementarity Level (Potential mRNA Secondary Structures,
T e ( i )   D e p e n d i n g   o n   L C I L ( i ) )
Local Complementarity Level with the Energy of Potential mRNA Secondary Structures
( T e ( i )   D e p e n d i n g   o n   L C I E ( i ) )
EEI1+
EEI2+
EEI3+
EEI4++
EEI5++
Table 4. The table reflects minimal doubling time for each species in hours and in a logarithmic form. The article from which the data were taken is also presented for each of the organisms (column DT_source).
Table 4. The table reflects minimal doubling time for each species in hours and in a logarithmic form. The article from which the data were taken is also presented for each of the organisms (column DT_source).
OrganismDoubling_Time (DT), hLog (DT)DT_Source
Acidithiobacillus ferrooxidans50.69897[77]
Bacillus anthracis0.5−0.30103[78]
Bacteroides thetaiotaomicron2.70.431364[79]
Bartonella henselae30.477121[80]
Campylobacter jejuni2.4666670.39211[81]
Deinococcus deserti2.60.414973[82]
Desulfovibrio vulgaris2.480.394452[83]
Escherichia coli0.333333−0.47712[61]
Halobacterium salinarum111.041393[84]
Helicobacter pylori0.833333−0.07918[85]
Lactococcus lactis0.5−0.30103[86]
Legionella pneumophila3.30.518514[87]
Leptospira interrogans8.20.913814[88]
Listeria monocytogenes0.516667−0.28679[89]
Microcystis aeruginosa461.662758[90]
Mycobacterium tuberculosis14.71.167317[91]
Mycoplasma pneumoniae80.90309[92]
Pseudomonas aeruginosa0.5−0.30103[93]
Salmonella typhimurium0.5−0.30103[94]
Shigella flexneri0.5−0.30103[95]
Staphylococcus aureus0.4−0.39794[93]
Streptococcus pyogenes0.666667−0.17609[96]
Synechocystis sp.5.80.763428[97]
Thermococcus gammatolerans4.50.653213[98]
Yersinia pestis10[78]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Korenskaia, A.E.; Matushkin, Y.G.; Lashin, S.A.; Klimenko, A.I. Bioinformatic Assessment of Factors Affecting the Correlation between Protein Abundance and Elongation Efficiency in Prokaryotes. Int. J. Mol. Sci. 2022, 23, 11996. https://doi.org/10.3390/ijms231911996

AMA Style

Korenskaia AE, Matushkin YG, Lashin SA, Klimenko AI. Bioinformatic Assessment of Factors Affecting the Correlation between Protein Abundance and Elongation Efficiency in Prokaryotes. International Journal of Molecular Sciences. 2022; 23(19):11996. https://doi.org/10.3390/ijms231911996

Chicago/Turabian Style

Korenskaia, Aleksandra E., Yury G. Matushkin, Sergey A. Lashin, and Alexandra I. Klimenko. 2022. "Bioinformatic Assessment of Factors Affecting the Correlation between Protein Abundance and Elongation Efficiency in Prokaryotes" International Journal of Molecular Sciences 23, no. 19: 11996. https://doi.org/10.3390/ijms231911996

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop