Skip to main content

A genome-scale metabolic flux model of Escherichia coli K–12 derived from the EcoCyc database

Abstract

Background

Constraint-based models of Escherichia coli metabolic flux have played a key role in computational studies of cellular metabolism at the genome scale. We sought to develop a next-generation constraint-based E. coli model that achieved improved phenotypic prediction accuracy while being frequently updated and easy to use. We also sought to compare model predictions with experimental data to highlight open questions in E. coli biology.

Results

We present EcoCyc–18.0–GEM, a genome-scale model of the E. coli K–12 MG1655 metabolic network. The model is automatically generated from the current state of EcoCyc using the MetaFlux software, enabling the release of multiple model updates per year. EcoCyc–18.0–GEM encompasses 1445 genes, 2286 unique metabolic reactions, and 1453 unique metabolites. We demonstrate a three-part validation of the model that breaks new ground in breadth and accuracy: (i) Comparison of simulated growth in aerobic and anaerobic glucose culture with experimental results from chemostat culture and simulation results from the E. coli modeling literature. (ii) Essentiality prediction for the 1445 genes represented in the model, in which EcoCyc–18.0–GEM achieves an improved accuracy of 95.2% in predicting the growth phenotype of experimental gene knockouts. (iii) Nutrient utilization predictions under 431 different media conditions, for which the model achieves an overall accuracy of 80.7%. The model’s derivation from EcoCyc enables query and visualization via the EcoCyc website, facilitating model reuse and validation by inspection. We present an extensive investigation of disagreements between EcoCyc–18.0–GEM predictions and experimental data to highlight areas of interest to E. coli modelers and experimentalists, including 70 incorrect predictions of gene essentiality on glucose, 80 incorrect predictions of gene essentiality on glycerol, and 83 incorrect predictions of nutrient utilization.

Conclusion

Significant advantages can be derived from the combination of model organism databases and flux balance modeling represented by MetaFlux. Interpretation of the EcoCyc database as a flux balance model results in a highly accurate metabolic model and provides a rigorous consistency check for information stored in the database.

Background

Constraint-based modeling techniques such as flux balance analysis (FBA) have become central to systems biology [1, 2], enabling a wealth of informative simulations of cellular metabolism. Many constraint-based modeling techniques have been first demonstrated for the Escherichia coli K–12 MG1655 metabolic network. A series of E. coli constraint-based models have been published by the group of B. Palsson [36], extending work on stoichiometric constraint-based modeling of E. coli dating back more than twenty years [710]. These models constitute a gold standard for E. coli modeling, and have seen a range of applications [1113] including metabolic engineering, model-driven discovery, cellular-phenotype prediction, analysis of metabolic network properties, studies of evolutionary processes, and modeling of interspecies interactions.

Motivated by the widespread use of E. coli metabolic models, we aimed to illustrate the benefits of integrating metabolic modeling into model organism databases by developing an E. coli model derived directly from the EcoCyc bioinformatics database [14]. First, we aimed to use the extensive biochemical literature referenced in EcoCyc to develop a model with improved accuracy for phenotypic prediction, specifically for predicting the phenotypes of gene knock-outs, and for predicting growth or lack thereof under different nutrient conditions.

Second, we sought to make the model easy to understand and operate. Our goal was a high level of model accessibility and readability through (a) tight web-based integration of the model with extensive model query and visualization tools, and (b) a model representation that captures extensive information that enriches the model and aids its understanding, such as metabolic pathways, chemical structures, and genetic regulatory information. Metabolic models are not just mathematical entities that output predictions; they are also artifacts that scientists interact with in multiple ways. If a model can be quickly and easily understood, scientists are more likely to trust its predictions, and the model is easier to reuse, to modify and extend, to learn from, and to validate through inspection. These aspects of a metabolic model depend strongly on how the model is represented, on the software tools available to interactively inspect the model, and on how tightly integrated the model is with those software tools.

Third, we sought to produce a model that is frequently updated to integrate new knowledge of E. coli metabolism.

Fourth, we sought to use the EcoCyc-derived E. coli metabolic model to identify errors in EcoCyc, and open problems in E. coli biology, by performing in-depth investigations of the disagreements between the phenotypic predictions of the model and experimental results.

We present EcoCyc–18.0–GEM, a constraint-based genome-scale metabolic model for E. coli K–12 MG1655 that is directly derived from the EcoCyc model organism database ( http://EcoCyc.org) built on the genome sequence of E. coli K–12 MG1655. The model is implemented using the MetaFlux [15] component of the Pathway Tools software [16].

Results and discussion

The EcoCyc–18.0–GEM generated from EcoCyc 18.0 encompasses 1445 genes, 2286 unique cytosolic and periplasmic reactions, and 1453 unique metabolites. Table 1 compares the statistics of EcoCyc–18.0–GEM with previous E. coli metabolic models. EcoCyc–18.0–GEM is an advance over previous stoichiometric models of E. coli metabolism in four respects: in its size; in its accuracy; in its form, readability and accessibility; and in its update frequency. Here we summarize these results; these points will be expanded in subsequent subsections.

Table 1 Survey of recent E. coli genome-scale model statistics

The MetaFlux component of Pathway Tools translates Pathway/Genome Database (PGDB) reactions and compounds into constraint-based metabolic models. Our methodology of fusing systems-biology models and bioinformatics databases has several advantages because of strong synergies between these approaches. Databases and models both require extensive literature-based curation and refinement. It is more efficient to perform that curation once in a manner that benefits a database and a model, than to duplicate curation efforts for a database project and a modeling project. Furthermore, the modeling process identifies errors, omissions, and inconsistencies in the description of a metabolic model, and therefore drives correction and further curation of the database if the two efforts are coupled. We made more than 80 EcoCyc updates as a result of comparing model predictions with experimental data and literature for this work. In addition, bioinformatics database curation methods such as the use of evidence codes and citations to provide data provenance, and the incorporation of mini-review summaries that describe enzymes and pathways, benefit systems-biology models, which typically lack data provenance and explanations.

Advances in model size

Compared with iJO1366, EcoCyc–18.0–GEM represents a 6% increase in the number of genes, a 23% increase in the number of unique cytosolic and periplasmic reactions, and a 28% increase in the number of unique metabolites. The size of EcoCyc–18.0–GEM is currently exceeded only by the more mathematically complex ME-model of O’Brien et al.[17], which includes simulation of gene expression, transcriptional regulation, and protein synthesis.

Advances in model accuracy

We conducted a threephase validation of EcoCyc–18.0–GEM to assess its accuracy (see Table 1). In phase I we compared simulated EcoCyc–18.0–GEM rates of nutrient uptake and product secretion in aerobic and anaerobic glucose culture with experimental rates derived from chemostat culture; the performance of EcoCyc–18.0–GEM was equivalent to previous models. In phase II we compared essentiality prediction for all 1445 genes involved in the model with experimental gene essentiality datasets; its error rate in predicting gene-knockout phenotypes decreased by 46% over the best previous model. In phase III we compared nutrient utilization predictions of EcoCyc–18.0–GEM with 431 experimental nutrient utilization tests; its accuracy in predicting growth and respiration under different nutrient conditions increased by 4.8% over previous models as the number of nutrient conditions expanded 2.5-fold. We investigated conflicts between experimental results and predictions of EcoCyc–18.0–GEM in detail, and provide an extensive discussion of these conflicts within the context of EcoCyc and the literature.

Subjects of particular interest include alternative catalytic routes capable of replacing genes thought to be essential; compounds with unclear routes of catabolism which are capable of supporting growth and/or cellular respiration; regulatory and environmental perturbations of the stoichiometric network model; and investigations of what, exactly, constitutes gene essentiality.

Advances in model form, readability, and accessibility

Another benefit of coupling systems-biology models with databases, and a corresponding advance of our model, is that generating a constraint-based model from a database that has associated web-based visualization tools leads to a literate model (by analogy to Knuth’s notion of literate programming [18]). A literate model is easy to read, and is highly accessible to and understandable by scientists.

Advances in model update frequency

Because the MetaFlux component of Pathway Tools generates constraint-based models directly from the EcoCyc PGDBs, as the database is refined through new curation, those refinements are automatically incorporated into newly generated versions of the model. We release new versions of the EcoCyc-based model three times per year; previous models were updated every four years [46]. Although there are reasons to limit the frequency of releases in order to tie them to a well-defined version of the database and throughly test the accuracy of new versions, we believe that more frequent model updates are useful for an organism as important as E. coli.

Validation of biomass metabolites, nutrients, and secretions

Refinement of EcoCyc–18.0–GEM began with the validation of the biomass, nutrient, and secretion metabolite sets, which are detailed at length in Additional file 1 and Additional file 2: Table S1. The biomass metabolite set establishes requirements for growth and determines the growth rate of the simulation. The biomass metabolite set for EcoCyc–18.0–GEM was based on the iJO1366 wild-type and core biomass reaction sets published by Orth et al., with several revisions stemming from differences in content and functionality between EcoCyc and the iJO1366 model. Gene essentiality in constraint-based models is principally determined by the biomass demands of the cell. Inclusion of a metabolite in the biomass metabolite set forces the genes required for manufacture of that metabolite to become essential in the simulation.

A wild-type biomass metabolite set, which is derived from measurement of the biomolecular composition of healthy, growing cells, is not representative of the minimal set of biomass metabolites required for cell survival. Because biomass metabolites not truly required for cell survival will generate false simulation predictions of essentiality in their biosynthetic pathways, the concept of a core biomass metabolite set was developed by Feist et al. The core biomass metabolite set is a biomass metabolite set that is defined with the aim of maintaining quantitative accuracy with regards to cell performance while predicting the observed experimental essentiality data as accurately as possible. Because much of this work focuses on testing the minimum requirements for cell growth, we frequently employed the core biomass metabolite set in our simulations. We use the term “expanded biomass set” to refer to our version of the wild-type biomass metabolite set described in Orth et al., because we do not wish to imply that the simulated cells always represented the wild-type state.

The biomass metabolite sets described here underwent several revisions reflecting differences in scope between EcoCyc–18.0–GEM and iJO1366. Whereas iJO1366 is a purpose-built model developed using the COBRA Toolbox with input from KEGG, EcoCyc, and other databases, EcoCyc is a database with its own schema whose entries are programmatically transformed into an FBA model. The specific metabolites present in iJO1366 therefore cannot always be matched with the specific metabolites generated from EcoCyc by MetaFlux on a one-to-one basis. Several biomass metabolites represented as distinct within iJO1366, such as phosphatidylethanolamines with different chain lengths and saturations, are summed under the heading of a single representative metabolite in the EcoCyc–18.0–GEM biomass set. Additionally, not all processes covered in EcoCyc–18.0–GEM are covered in iJO1366, and the reverse is also true. As a result, the biomass metabolite sets differ slightly. Additional file 2: Table S2 contains a complete side-by-side comparison of the EcoCyc–18.0–GEM and iJO1366 biomass metabolite sets, and lists the differences between them.

We constructed standard nutrient sets based on culture conditions reflecting experiments in glucose or glycerol minimal media and on the model’s capability to use substrates. Those substrates include glucose or glycerol as appropriate, O 2, N H 4 + , phosphate, sulfate, ferrous iron, water, CO 2, minerals appropriate to the biomass objective function, and MOPS buffer (usable as a sulfur source) where appropriate. Because of passive diffusion at the high concentration of ammonium used in experimental culture, N H 4 + is supplied directly in the cytosol (avoiding false negative essentiality predictions for the high-affinity nitrogen transporter amtB), whereas all other nutrients are supplied in the periplasmic space.

Finally, we developed a large set of secreted compounds that could be supplied across all growth conditions explored with our model. It contains both plausible products of E. coli metabolism and dead-end metabolites [19] within the model. The presence or absence of metabolites in this set should not be construed to indicate their presence or absence in E. coli culture media.

We verified the metabolic reachability of each component within the EcoCyc biomass metabolite set by supplying nutrients representing an aerobic glucose minimal medium and setting the production of each individual metabolite in turn as the optimization goal of MetaFlux, and repaired gaps by means of literature-based manual curation of EcoCyc and expansion of the relevant metabolite sets.

ATP maximization validation

We next confirmed that simulations of aerobic growth on glucose run with maximization of ATP production as their objective made appropriate use of the glycolytic and TCA cycle pathways and agreed with previous work on E. coli FBA.

The maximization of ATP production under aerobic conditions was studied by setting the ATP consumption reaction ATP + H 2O → ADP + Pi + H + as the objective function to be maximized by MetaFlux. The fluxes resulting from the maximization of ATP production on glucose under aerobic conditions were compared with fluxes from COBRA Toolbox [20] simulations of iJO1366 under the same conditions and were found to be largely identical (Table 2). Differences arose from variances in proton translocation stoichiometries between the EcoCyc–18.0–GEM version of the NADH:ubiquinone oxidoreductase I (NADH-DEHYDROG-A-RXN) (4 H + translocated per 2 e -, as proposed by Treberg et al.[21]) and the iJO1366 version of the NADH:ubiquinone oxidoreductase (NADH16pp) (3 H + translocated per 2 e -, as proposed by Wikstrom and Hummer [22]). The exact number of protons translocated by the NADH:ubiquinone oxidoreductase is an issue of open discussion in the scientific literature, and this uncertainty is described in the EcoCyc summary for the enzyme. If a consensus develops behind the 3 H + per 2 e - view of translocation stoichiometry, future versions of EcoCyc will be changed to reflect this fact.

Table 2 Flux comparison between EcoCyc–18.0–GEM and iJO1366: ATP maximization objective under aerobic conditions

Further numerical differences are due to a technical consideration: EcoCyc cytochrome bo oxidase reaction stoichiometry is written in terms of whole molecules of oxygen, while iJO1366 CYTBO3_4pp is written in terms of half-molecules (1 O 2 consumed vs. 0.5 O 2).

Comparison with iJO1366 simulation and chemostat data

After completing our basic validation of biomass production and energy generation, we maximized the rate of EcoCyc–18.0–GEM biomass metabolite set production under several minimal media conditions and ensured that we obtained results comparable to the iJO1366 results for the same conditions obtained using the COBRA Toolbox. Divergences were addressed by literature-based manual curation of EcoCyc and modification of MetaFlux reaction sets. We further compared the extracellular flux distributions resulting from these simulations with the experimental data obtained in carbon-limited chemostat environments under both aerobic and anaerobic conditions.

Tables 3 and 4 compare extracellular metabolite flux results derived from EcoCyc–18.0–GEM simulation, iJO1366 simulation, and experimental data [23, 24] for the canonical cases of aerobic and anaerobic growth on glucose-limited chemostat culture. In all simulations, the experimental rate of glucose supply was the only fixed constraint; all other nutrients and secretions were left unconstrained.

Table 3 Comparison of experimental aerobic glucose-limited chemostat growth data with EcoCyc–18.0–GEM and iJO1366 constraint-based model predictions
Table 4 Comparison of experimental anaerobic glucose-limited chemostat growth data with EcoCyc–18.0–GEM and iJO1366 constraint-based model predictions

The behavior of MetaFlux/EcoCyc–18.0–GEM simulations was very similar in most regards to the behavior of COBRA/iJO1366 simulations. Respiration and fermentation rates scaled with nutrient uptake at comparable rates. The generally higher rates of O 2 uptake observed experimentally lend support to a lower practical efficiency of proton translation stoichiometry in vivo, perhaps augmented by respiratory inefficiencies such as futile cycling and generation of reactive oxygen species. Both models secrete the expected 1:2:1 mix of acetate, formate, and ethanol during anaerobic growth on glucose that Varma et al.[25] originally identified as stoichiometrically optimal. During the transition between purely anaerobic and aerobic domains, the competing demands of energy metabolism and redox elimination cause a characteristic pattern of mixed acid fermentation described by Varma et al., in which ethanol, then formate, and finally acetate production are eliminated as the cell’s oxygen supply becomes completely sufficient to support aerobic respiration. Figures 1 and 2 use the Cellular Overview and Omics Popup visualization functionalities of Pathway Tools to illustrate this behavior in EcoCyc–18.0–GEM during a transition from anaerobicity to aerobicity.

Figure 1
figure 1

Pathway Tools visualization of EcoCyc–18.0–GEM flux during aerobic transition. Example visualization of EcoCyc–18.0–GEM flux during a transition from anaerobic to aerobic growth, created within the interactive Cellular Overview diagram in Pathway Tools. The upper bound of glucose uptake is set to 10 mmol/gCDW/hr, while the upper bound of oxygen uptake is increased from 0 to 20 mmol/gCDW/hr in 2.5 mmol/gCDW/hr steps. Omics Popups are used to illustrate flux through acetaldehyde dehydrogenase, pyruvate-formate lyase, phosphoglucose isomerase, glyceraldehyde 3-phosphate dehydrogenase, cis-aconitate hydratase, and valine biosynthesis.

Figure 2
figure 2

Pathway Tools visualization of mixed-acid fermentation flux during aerobic transition. Visualization of EcoCyc–18.0–GEM flux in mixed-acid fermentation during a transition from anaerobic to aerobic growth, created within the EcoCyc mixed-acid fermentation pathway page in Pathway Tools. The upper bound of glucose uptake is set to 10 mmol/gCDW/hr, while the upper bound of oxygen uptake is increased from 0 to 20 mmol/gCDW/hr in 2.5 mmol/gCDW/hr steps. Omics Popups are used to illustrate changes in flux to the mixed-acid fermentation products formate, acetate, and ethanol as the cellular energy and redox balance evolves during the aerobic transition.

Comparisons between FBA-predicted extracellular fluxes and experimental fluxes show that EcoCyc–18.0–GEM and iJO1366 FBA predictions agree more closely with each other than with experimental flux results, although the correspondence between simulation and experiment was quite close for the experimental fluxes under consideration. This result was expected given the adaptation of the iJO1366 biomass function for use in EcoCyc–18.0–GEM, the use of iJO1366 and preceding reconstructions as benchmarks in the development of EcoCyc–18.0–GEM, and the use of EcoCyc as a reference in the construction of iJO1366 and its predecessors. The experimental measurements generally demonstrate higher fluxes of the respiratory gases O 2 and CO 2 than the simulated fluxes, suggesting a degree of respiratory inefficiency not properly modeled by FBA. Similarly, small quantities of succinate and lactate were produced by experimental fermentation, indicating a degree of divergence from metabolic optimality in vivo. Broader cellular constraints such as regulation, protein crowding, pathway enzyme synthesis requirements, and pathway-throughput limits underlie these differences [17, 26, 27]. Successive generations of evolution under constant growth conditions might bring the experimental result closer to theory, as described in Ibarra et al.[28].

Gene essentiality analysis

One of the most exciting aspects of genome-scale flux modeling is the ability to rapidly test computational gene knockouts (KOs) for their effects on metabolic function. Gene KO simulation is useful both for prediction and for validation: in silico FBA screens of gene KOs have been applied in a variety of metabolic engineering efforts [2931], and E. coli KO library collections with well-characterized growth behavior provide an important tool for flux model validation.

FBA gene KO essentiality prediction depends on two types of database associations between genes and chemical reactions: genes whose products catalyze reactions, and genes whose products are reaction substrates (e.g., acyl-carrier protein). Simulation gene KOs are carried out by identifying all reactions involving the gene, and then identifying all other genes capable of catalyzing the reactions or supplying the substrates thus identified. Reactions for which no isozymes or alternative substrate supplies are found are removed from the FBA stoichiometric network. An FBA solution is then calculated for the new model. If the simulated gene KO has caused the deletion of one or more reactions required for the synthesis of a biomass metabolite, generation of the full biomass metabolite set will be blocked and the FBA simulation returns a no-growth result. Such a result represents a prediction of gene essentiality. If the complete biomass metabolite set can still be produced in spite of the simulated gene knockout, the FBA simulation returns a growth result, indicating a prediction of gene non-essentiality.

The experimental essentiality data used in our tests consisted of two major datasets. The first, used to study gene essentiality on rich and glucose minimal media, was the deletion study of Baba et al.[32] as updated by Yamamoto et al.[33], which tested the Keio collection library of 4288 E. coli gene deletion strains for growth on LB rich media and MOPS minimal media with 0.4% glucose. We conducted our glycerol minimal media tests using the gene knockout essentiality data of Joyce et al.[34], an expansion of the study of the Keio collection essentiality to include growth on M9 minimal medium with 1% glycerol.

Several E. coli gene deletions strongly affect growth on various types of minimal media, but are nonessential to growth on rich media. Because the FBA simulation result is treated as a binary test (growth or no-growth), gene deletions that strongly affect growth on minimal media without producing a completely lethal phenotype must be defined either as experimentally essential or as experimentally nonessential.

Two representative perspectives on this definition are the narrow essentiality criteria of no observable growth in minimal media and the broad essentiality criteria used by Orth et al. The narrow glucose essentiality criteria treat as essential those Baba et al. and Yamamoto et al. gene deletion mutants with OD600 ≤ 0.005 after 24 and 48 hr growth on glucose minimal media. This requires no perceptible growth over a long period. The broad essentiality criteria was originally defined in relative terms by Joyce et al., as the slowest-growing ninth of all Keio collection deletion mutants.

In absolute terms, that approach treats as essential those deletion mutants measured by Baba et al. to have OD ≤ 0.091 after 24 hr growth on glucose minimal media, which indicates impaired growth over a shorter period. The practical difference between these two perspectives is illustrated in Figure 3, which displays the distribution of OD600 data for all rich media-viable Keio collection mutants after 24 hr of growth on MOPS media containing 0.4% glucose, as originally published in Supplementary Table three of Baba et al. As the figure illustrates, the broad essentiality criteria include a population of cells with severe growth defects that is not contained in the narrow essentiality data. The comparison between narrow and broad essentiality criteria can be expanded to glycerol minimal media by comparing narrow glycerol essentiality criteria of no observed growth on rich media with the glycerol essentiality criteria of Orth et al., again derived from the criteria of Joyce et al. involving successive division into thirds.

Figure 3
figure 3

Essentiality criteria basis in high-throughput KO data. Histogram of OD600 measurements for all rich media-viable Baba et al. deletion mutants after 24 hr of growth on MOPS media containing 0.4% glucose. Data from Supplementary Table three of Baba et al.

In order to examine criteria for experimental gene essentiality more deeply and to illustrate the effect of defining a core biomass metabolite set, we conducted essentiality testing using both the expanded and core biomass metabolite sets proposed by Orth et al. Differences in essentiality predictions between the two data sets illustrated the differences between standard cell composition under nominal conditions and the minimal composition required for cell growth.

We simulated single gene KOs on glucose and glycerol minimal media for the 1445 genes in EcoCyc–18.0–GEM to test whether the resulting EcoCyc–18.0–GEM gene deletion mutants were capable of generating core and expanded biomass metabolite sets from sets of nutrients based on the experimental culture media of Baba et al. and Joyce et al. Gene KO simulations capable of generating any growth at all were scored as nonessential, whereas gene KOs blocking generation of the biomass metabolite set were scored as essential.

We compared the results of this simulated essentiality screen with experimental gene essentiality results based on both narrow and broad gene essentiality criteria. Incorrect essentiality predictions were addressed by literature-based manual curation of EcoCyc and modification of MetaFlux metabolite sets. Final essentiality prediction results after curation are summarized in Tables 5 for glucose and 6 for glycerol. The overall accuracy of prediction for growth on glucose with the core biomass metabolite set and broad essentiality criteria was 1375/1445 (95.2% accuracy, 99.0% sensitivity, 77.5% specificity). For prediction of growth on glycerol under the same simulation conditions, the overall accuracy of prediction was 1365/1445 (94.5% accuracy, 98.1% sensitivity, 77.5% specificity). Sensitivity here refers to the percentage of gene deletions resulting in growth that are correctly identified by simulation, while specificity refers to the percentage of gene deletions resulting in no growth that are correctly identified by simulation.

Table 5 Comparison of experimental gene essentiality results with computational EcoCyc–18.0–GEM results for aerobic growth on MOPS medium with 0.4% glucose
Table 6 Comparison of experimental gene essentiality results with computational EcoCyc–18.0–GEM results for aerobic growth on MOPS medium with 1% glycerol

Tables 5 and 6 illustrate that the gene essentiality predictions in EcoCyc–18.0–GEM differed in a number of cases from the gene essentiality conclusions generated by high-throughput gene KO screening. Because these are situations of considerable interest to the development of EcoCyc as a reference, we examined them in greater detail for the case of growth on glucose, with reference to the E. coli literature. Our examination covered two types of incorrect gene deletion growth predictions. The first type was a false positive growth prediction. These genes, which are experimentally essential under the conditions tested by Baba et al., were predicted to be nonessential by EcoCyc–18.0–GEM. The second type was a false negative growth prediction. These genes, which are not experimentally essential under the conditions tested by Baba et al., were predicted to be essential by EcoCyc–18.0–GEM.

Tables 7, 8, 9, 10, 11 and 12 present five broad categories of incorrect gene deletion predictions from EcoCyc–18.0–GEM. Table 7 cover false predictions involving open questions of E. coli biology, false predictions resulting from interesting facets of experimental or simulation methods, and other situations of special relevance. Table 8 covers false predictions in core glycolytic, pentose phosphate, Entner-Doudoroff, and TCA cycle metabolism. This highly interconnected region of E. coli metabolism contains several isozymes and opportunities for reversibility, and presents a challenge to FBA essentiality predictions in the absence of complete regulatory modeling. Table 9 cover false predictions that are the result of unmodeled regulation of gene expression or enzyme activity. Genes repressed under Baba et al. experimental growth conditions, insufficiently expressed isozymes, and cases of enzyme inhibition all fall into this category. Table 10 covers situations in which the essentiality conclusions of the high-throughput essentiality screen differed significantly from the essentiality conclusions made by the E. coli K–12 literature. Table 11 covers false gene essentiality predictions relating to systems beyond the scope of EcoCyc–18.0–GEM’s biomass objective function. Finally, Table 12 covers false gene essentiality predictions made as a result of MetaFlux and EcoCyc technical problems discovered in the course of this study.

Table 7 False gene essentiality predictions resulting from open questions in E. coli biology and gene essentiality
Table 8 False gene essentiality predictions within glycolytic and TCA cycle metabolism
Table 9 False gene essentiality predictions resulting from isozymes or pathways not operational under the experimental conditions of Baba et al.
Table 10 Genes for which EcoCyc–18.0–GEM predictions identified likely errors in high-throughput essentiality screening, and the EcoCyc–18.0–GEM predictions were confirmed by conventional essentiality experiments
Table 11 False gene essentiality predictions for genes representing systems beyond the scope of the EcoCyc–18.0–GEM biomass function
Table 12 False gene essentiality predictions caused by technical issues in MetaFlux and EcoCyc

Several of the false gene essentiality predictions described within these tables were discussed in the work of Kim and Copley, who examined the essentiality conclusions of Baba et al. in E. coli core metabolism with reference to the then-current state of EcoCyc. Constraint-based model improvement and gap-filling based on gene essentiality predictions derived from the work of Baba et al. have been examined for the COBRA family of constraint-based models of E. coli metabolism by Reed et al.[140], Kumar et al.[141], Kumar and Maranas [142], Barua et al.[143], Orth and Palsson [73, 144], and Tervo and Reed [145]. Our revisions of EcoCyc–18.0–GEM included manual application of a subset of GrowMatch [142] gap-filling methods, specifically resolution of false positive gene essentiality predictions associated with blocked genes and false negative results associated with secretion of metabolites. The essentiality prediction accuracy resulting from our manual curation process is similar to the accuracy resulting from applying the full GrowMatch algorithm to the iAF1260 model.

Additional file 2: Table S3 provides detailed listings of essentiality status and model predictions, including a breakdown of gene essentiality prediction status by criteria used.

Nutrient utilization analysis

Observation of culture growth on various nutrient sources is a foundation of microbiology [146]. EcoCyc 18.0 contains information on E. coli respiration for 428 types of media, including 22 conventional types of minimal growth media and 383 Biolog Phenotype Microarray (PM) wells. The 383 Biolog PM media conditions represent a high-throughput method of evaluating metabolic phenotypes in culture based on a tetrazolium dye assay of cellular respiration. Each well in a Biolog 96-well PM plate contains a standard minimal media composition plus a nutrient source that is varied across the PM plate, with the element supplied by the varying nutrient source dependent on the type of Biolog PM plate in use [147149].

We evaluated the performance of EcoCyc–18.0–GEM in predicting growth for the following available datasets: (1) aerobic E. coli growth on the 22 common conventional minimal growth media; (2) consensus estimates of respiration based on four different experimentalists’ measurements of aerobic Biolog 96-well plates PM1–4, representing 313 conflict-free growth observations; and (3) an anaerobic Biolog PM1 plate assay surveying carbon source utilization in the absence of oxygen, representing 96 anaerobic growth observations. Biolog PM data stored in EcoCyc measures utilization of nutrients as sources of carbon (PM1–2), nitrogen (PM3), sulfur (PM4), and phosphorus (PM4).

Conventional media compositions and growth results were drawn from the literature. Aerobic Biolog PM nutrient utilization assay results were compiled from four different datasets captured in EcoCyc: (1) from our own experiments; (2) from a dataset obtained from B. Bochner; and from the recent publications of (3) AbuOun et al.[150] and (4) Yoon et al.[151]. Anaerobic Biolog PM nutrient utilization assay results were obtained from B. Bochner. We did not include the data of Baumler et al. in our analysis of Biolog PM results because of variation in culture conditions and a high degree of conflict with other datasets under both aerobic and anaerobic conditions [152]. See the Methods section for additional details.

Growth on a given type of media was tested by constructing simulated MetaFlux nutrient sets corresponding to the contents of the media in question and comparing EcoCyc–18.0–GEM growth predictions with experimental growth results. Due to the absence of enterobactin iron uptake modeling in EcoCyc, Fe 3+ in the medium was replaced with Fe 2+. Anaerobic simulations were prepared identically to those performed for aerobic growth, except for the removal of oxygen from the nutrient set, inclusion of the formate-hydrogen lyase reaction, and the removal of the protoheme and pyridoxal 5’-phosphate synthesis requirement from the biomass.

Literature-based EcoCyc curation and appropriate modifications of MetaFlux metabolite sets were used to address incorrect nutrient utilization predictions. The final results for PM array validation after curation are listed in Table 13. Overall accuracy of growth prediction for aerobic Biolog PM assays was 252/313 (80.5%), with 70 assays not evaluated because of experimental conflicts (see the Methods section). Anaerobic Biolog PM assay predictions had an overall accuracy of 74/96 (77.1%). Aerobic growth tests on conventional minimal media contained in EcoCyc had an overall accuracy of 22/22 (100.0%).

Table 13 EcoCyc–18.0–GEM nutrient utilization prediction results

The overall accuracy of nutrient utilization prediction across all aerobic and anaerobic PM and conventional growth media is 348/431 (80.7%). Tables 14, 15, 16 and 17 provide detailed discussions of false negatives and false positives for aerobic PM assays. Tables of results for anaerobic PM assays and conventional growth media are available in Additional file 2: Tables S8 and S9, respectively.

Table 14 Conflicts between EcoCyc–18.0–GEM growth predictions and experimental carbon source utilization data for aerobic growth on Biolog PM plates at 37°C
Table 15 Conflicts between EcoCyc–18.0–GEM growth predictions and experimental nitrogen source utilization data for aerobic growth on Biolog PM plates at 37°C
Table 16 Conflicts between EcoCyc–18.0–GEM growth predictions and experimental sulfur source utilization data for aerobic growth on Biolog PM plates at 37°C
Table 17 Conflicts between EcoCyc–18.0–GEM growth predictions and experimental sulfur source utilization data for aerobic growth on Biolog PM plates at 37°C

Model readability and accessibility

Scientists naturally need to ask many questions of a metabolic model, such as “What are the chemical structures of all substrates in reaction X, and is X chemically balanced?” “What metabolic pathway(s) is reaction X a member of, and what are the adjacent reactions?” “Which E. coli enzymes are inhibited by ADP?” “What transcriptional regulators affect the expression of the enzymes for reaction X?” Their ability to answer these questions rapidly and accurately is strongly dependent on the model representation, the software tools available for querying and visualizing that representation, the tightness with which those tools are integrated with the model, and the presence of additional enriching information for the model.

Existing E. coli models are represented as spreadsheet files and as SBML files, making it tedious or impossible for non-programmers to answer the preceding questions directly from those files. Although SBML files can be imported into software tools such as the RAVEN Toolbox [181] and rbionet [182], in practice that approach is limited because of variations in SBML encodings, the effort required to install and integrate multiple software tools with disparate capabilities, and the limited visualization capabilities of those tools. More fundamentally, previous E. coli models do not capture (nor can SBML capture) additional enriching information that, while not required for the mathematical operation of a model, greatly enhances our ability to validate and understand a model, and to answer the preceding questions. Examples of such enriching information present in EcoCyc–18.0–GEM are metabolite chemical structures, arrangements of reactions within metabolic pathways, and gene regulatory information. Note that introducing ad-hoc definitions of these data (e.g., pathways) in the SBML “Notes” field, or introducing SBML links to external databases, would be considered out of bounds: since pathways are not captured formally in the SBML specification today, there is no guarantee regarding interoperability of software tools with such ad-hoc data.

EcoCyc–18.0–GEM is highly understandable because it can be interactively queried and visualized through the EcoCyc web site and desktop Pathway Tools software, which supports visualization of metabolic pathways and reaction diagrams; metabolite pages that depict metabolite structures and all reactions a metabolite is involved in; depiction of gene/reaction connections and of genome organization via a genome browser; navigation through the E. coli gene regulatory network; constructing structured queries such as: find all reactions of a given metabolite; find all enzymes utilizing a given cofactor; and presentation of text summaries and citations that explain and support aspects of the model. In general, other tools for metabolic model visualization tend to be less comprehensive, and to be less closely coupled to the model; see [183186] for recent reviews.

The reaction fluxes computed from EcoCyc–18.0–GEM are more understandable than those from previous E. coli models because EcoCyc–18.0–GEM fluxes can be immediately painted onto the EcoCyc Cellular Overview, a zoomable diagram of the complete metabolic map of E. coli that allows immediate visual inspection of flux patterns. Although other software tools exist for visualizing flux patterns on metabolic networks, e.g., the RAVEN Toolbox, they are unlikely to be easily usable with previous E. coli models. For example, RAVEN Toolbox requires that the user manually construct the metabolic network diagram, which could take days or weeks of effort. In contrast, Pathway Tools generates metabolic map diagrams algorithmically from a PGDB.

Conclusions

EcoCyc–18.0–GEM demonstrates the advantages of literate modeling based on comprehensive organism databases. It provides comprehensive genome-scale coverage of the E. coli metabolic network, representing gene function with an unprecedented degree of accuracy.

Integration of EcoCyc–18.0–GEM into the EcoCyc database gives investigators working with the model access to the full Pathway Tools bioinformatics and data visualization suite. This allows construction of complex database queries involving the full range of biochemical entities within E. coli, and visualization of pathways and reactions within the model as they change throughout the course of construction. As part of EcoCyc, EcoCyc–18.0–GEM will receive frequent updates to remain abreast of recent research developments.

The process of EcoCyc–18.0–GEM construction and validation resulted in more than 80 updates to EcoCyc. These included expansion and revision of periplasmic phosphatase activities many updates to sugar transport and phosphotransferase system modeling; correction of incorrect compartment assignments; fixes for L-lactate dehydrogenase action; revisions to glutathione hydrolysis; new transport reactions for compounds identified as nutrient sources during Biolog PM testing; addition of MOPS catabolism via the alkanesulfonate pathway; removal of several incorrect reactions and gene-protein relationships; numerous fixes to reaction reversibility and directionality; several compound class reassignments to correct issues with reaction instantiation; mass rebalancing for several reactions; and revisions to ATP synthase proton stoichiometry. These updates are outlined in Additional file 2: Table S11.

The MetaFlux software has also been improved as a result of the FBA validation of EcoCyc. These improvements include upgrades to compartmentalization handling, gene deletion code, and electron transfer reaction handling. MetaFlux solution and log files have been updated to contain additional statistical information and provide a more detailed explanation of the metabolic network construction process to the user. Numerous updates and revisions to the MetaFlux model of E. coli have been introduced as part of this effort. Biomass metabolite sets have been revised to reflect the work of Orth et al., and additional updates have been made on the basis of the validation process in order to create the most accurate final product possible.

Many questions of interest to E. coli modelers and experimentalists were raised in the course of EcoCyc–18.0–GEM development. By highlighting these questions and presenting them within the context of EcoCyc as a reference database, we address the interests of the general metabolic modeling audience and of E. coli experimentalists interested in using models to explore their results and generate new leads for research. We summarize these questions here.

Experimental measurements of respiratory fluxes in glucose-fed aerobic chemostat culture are higher than those predicted by simulation, and small quantities of succinate and lactate are generated in experimental anaerobic fermentations, suggesting interesting in vivo deviations from theoretical in silico optimality.

A small number of metabolic byproducts must be removed directly from the cytosol in the secretion set because of a lack of known salvage or excretion pathways. The fate of these metabolites is of interest.

Several incorrect essentiality predictions are associated with unclear cellular biomass requirements, pathways with potential alternate routes of catalysis, uncertain determinations of essentiality on glucose minimal media, and ambiguous or missing gene function. Similarly, nutrient utilization predictions have identified a number of compounds that lack clear pathways of entry into metabolism but are capable of supporting respiration and/or growth. Resolution of these uncertainties would improve our understanding of E. coli function in varying environments.

Methods

The data used for this project were obtained from EcoCyc version 18.0, and the bioinformatics and flux analysis procedures documented here were performed in either the Web or desktop environment of the Pathway Tools 18.0 software. Pathway Tools can be downloaded, along with documentation and example files, at http://brg.ai.sri.com/ptools/. The simulation tests were constructed by using Lisp scripting and the Pathway Tools Lisp API, documented at http://brg.ai.sri.com/ptools/api/. Further details on the construction of EcoCyc–18.0–GEM can be found within Additional file 1.

EcoCyc–18.0–GEM development employed the MetaFlux component of Pathway Tools, documented in the Pathway Tools User’s Guide and in [15]. All simulations were run in solving mode on a 2.7 GHz i7 MacBook Pro with 16 GB RAM. The “minimize-fluxes: yes” option was used for taxicab norm minimization of fluxes. For additional information, see the Pathway Tools User’s Guide. A MetaFlux.fba file demonstrating simulation of aerobic growth of E. coli BW25113 on glucose is included as Additional file 3.

We note that the choice of stoichiometric representation of equations can affect flux balance solutions when minimization of summed flux is used as part of the objective function; see [187] for further details. Stoichiometric coefficients within EcoCyc–18.0–GEM are scaled so as to provide minimum whole-integer stoichiometry. Application of different scaling within the flux network may lead to altered flux solutions.

Briefly, MetaFlux creates a stoichiometric metabolic flux network at run time from the metabolites and reactions contained in a Pathway Tools PGDB. During network construction, MetaFlux removes PGDB reactions that are: ambiguously instantiated (see below); unbalanced or having an undetermined balance state; disconnected from the network; marked as physiologically irrelevant; involved in polymerization; involved in polymer segment or protein modification; lacking substrates on one side; containing substrate entities that are described only by strings; possessed of variable stoichiometry; or have more than 10,000 permutations that must be checked during instantiation. MetaFlux then instantiates reactions containing compound classes, replacing the class reactions with mass-balanced reactions containing instances of the relevant compound classes. The resulting set of metabolites and reactions constitutes the metabolic flux network operated on by MetaFlux. Enzymatic reaction and gene-protein relationship data encoded in the PGDB are used to associate the reactions of the network with enzymes and genes as appropriate.

COBRA [20, 188] simulations of the iJO1366 E. coli genome-scale reconstruction were employed in order to validate the EcoCyc FBA simulations and to provide a point of comparison to existing reference models. The iJO1366 simulations were performed using the COBRA Toolbox 5.0.0 within MATLAB R2010a and cobrapy 0.2 within Python 2.7.6. All iJO1366 simulations used taxicab norm minimization of fluxes as described in the documentation for the optimizeCbModel function.

Biomass metabolite sets were constructed using the wild-type and core biomass sets of Orth et al. and modified according to our research findings and experimental data on gene essentiality. The ATP turnover requirement for growth-associated maintenance costs (GAM) was set to 53.95 mmol ATP/gCDW, while the ATP turnover requirement for non-growth-associated maintenance costs (NGAM) was set to 3.15 mmol ATP/gCDW, per Orth et al.

All coefficients of the biomass metabolite set represent millimoles (mmol) of each metabolite required per gram of cell dry weight (gCDW). Coefficients of the nutrient and metabolite sets represent specific uptake and output fluxes, in millimoles of each metabolite supplied per gram of cell dry weight per hour (mmol/gCDW/hr). An overall biomass flux of 1.0 thus represents a specific growth rate μ of 1.0 hr -1 (one new gram of cell dry weight per gram of cell dry weight per hour).

Aerobic glucose chemostat data for experimental comparisons were obtained from [23]. Anaerobic glucose chemostat data were obtained from [24] via [25].

The protocols followed in our PM experiments are as follows: PM plates 1–4 containing 190 sole carbon sources, 95 sole nitrogen sources, 59 sole phosphate sources and 35 sole sulfur sources were used in this analysis. E. coli MG1655 was obtained from the Yale Coli Genetic Stock Center, pre-grown on nutrient agar and used to inoculate the plates following Biolog instructions. The data were collected and analyzed using the OmniLogH PM system, which records the color change every 15 min for each well in the 96 well assay plates. All incubations were performed at 37°C over 48 hr. For the complete details of all PM assay conditions, please refer to the original publications.

The PM nutrient-source assay is an assay of respiration based on the generation of NADH by carbon metabolism and the subsequent reduction of a tetrazolium redox dye by NADH. As such, it does not directly measure either the cell growth simulated by FBA biomass objectives or the uptake of noncarbon sources. However, checkpoint linkage of carbon-source catabolism to nitrogen, phosphorus, and sulfur source catabolism enables the tetrazolium redox dye assay to probe the metabolism of non-carbon sources [189]. Bochner describes the phenomenon of checkpoint linkage as starvation for elemental nutrient source leading to arrest of cellular respiration, due to redox imbalance or alarmone synthesis. We therefore compared PM respiration results directly with FBA growth simulation.

Biolog PM results are stored within the EcoCyc database and are accessible via the Pathway Tools API and EcoCyc website. Individual Biolog PM assay scores for each well were compared across experimental datasets to establish a consensus for comparison with EcoCyc–18.0–GEM simulation. Because simulations of nutrient utilization were scored according to growth or no growth, experimental Biolog PM results indicating ‘normal’ and ‘low’ respiration were combined into a ‘positive’ result. The majority of Biolog PM tests (313/383) displayed a consensus of either respiration or no respiration across all four experimental datasets used. In 70 of 383 cases, no clear consensus could be reached. These divergent cases were omitted from the nutrient-utilization assay validation because no reliable conclusion could be reached regarding the results. See Additional file 2: Table S10 for a list of omitted PM data.

References

  1. Orth JD, Thiele I, Palsson BØ: What is flux balance analysis?. Nat Biotechnol. 2010, 28 (3): 245-248.

    PubMed Central  PubMed  Google Scholar 

  2. Thiele I, Palsson BØ: A protocol for generating a high-quality genome-scale metabolic reconstruction. Nat Protoc. 2010, 5 (1): 93-121.

    PubMed Central  PubMed  Google Scholar 

  3. Edwards J, Palsson BØ: The Escherichia coli MG1655in silico metabolic genotype: its definition, characteristics, and capabilities. Proc Natl Acad Sci U S A. 2000, 97: 5528-5533.

    PubMed Central  PubMed  Google Scholar 

  4. Reed JL, Vo TD, Schilling CH, Palsson BØ: An expanded genome-scale model of Escherichia coliK-12 (iJR904 GSM/GPR). Genome Biol. 2003, 4 (9): 54-

    Google Scholar 

  5. Feist AM, Henry CS, Reed JL, Krummenacker M, Joyce AR, Karp PD, Broadbelt LJ, Hatzimanikatis V, Palsson BØ: A genome-scale metabolic reconstruction for Escherichia coliK–12 MG1655 that accounts for 1260 ORFs and thermodynamic information. Mol Syst Biol. 2007, 3: 121-

    PubMed Central  PubMed  Google Scholar 

  6. Orth JD, Conrad TM, Na J, Lerman JA, Nam H, Feist AM, Palsson BØ: A comprehensive genome-scale reconstruction of Escherichia colimetabolism — 2011. Mol Syst Biol. 2011, 7: 535-

    PubMed Central  PubMed  Google Scholar 

  7. Majewski RA, Domach MM: Simple constrained-optimization view of acetate overflow in E. coli. Biotechnol Bioeng. 1990, 35 (7): 732-738.

    PubMed  Google Scholar 

  8. Varma A, Boesch BW, Palsson BØ: Biochemical production capabilities ofEscherichia coli. Biotechnol Bioeng. 1993, 42 (1): 59-73.

    PubMed  Google Scholar 

  9. Varma A, Palsson BØ: Metabolic capabilities of Escherichia coli: I. synthesis of biosynthetic precursors and cofactors. J Theor Biol. 1993, 165 (4): 477-502.

    PubMed  Google Scholar 

  10. Pramanik J, Keasling JD: Stoichiometric model of Escherichia colimetabolism: incorporation of growth-rate dependent biomass composition and mechanistic energy requirements. Biotechnol Bioeng. 1997, 56 (4): 398-421.

    PubMed  Google Scholar 

  11. Oberhardt MA, Palsson BØ: Applications of genome-scale metabolic reconstructions. Mol Syst Biol. 2009, 5: 320-

    PubMed Central  PubMed  Google Scholar 

  12. Lewis NE, Nagarajan H, Palsson BØ: Constraining the metabolic genotype-phenotype relationship using a phylogeny of in silico methods. Nat Rev Microbiol. 2012, 10 (4): 291-305.

    PubMed Central  PubMed  Google Scholar 

  13. McCloskey D, Palsson BØ: Basic and applied uses of genome-scale metabolic network reconstructions ofEscherichia coli. Mol Syst Biol. 2013, 9: 661-

    PubMed Central  PubMed  Google Scholar 

  14. Keseler IM, Mackie A, Peralta-Gil M, Santos-Zavaleta A, Gama-Castro S, Bonavides-Martinez C, Fulcher C, Huerta AM, Kothari A, Krummenacker M, Latendresse M, Muniz-Rascado L, Ong Q, Paley S, Schroder I, Shearer AG, Subhraveti P, Travers M, Weerasinghe D, Weiss V, Collado-Vides J, Gunsalus RP, Paulsen I, Karp PD: EcoCyc: fusing model organism databases with systems biology. Nucleic Acids Res. 2013, 41 (Database issue): 605-612.

    Google Scholar 

  15. Latendresse M, Krummenacker M, Trupp M, Karp PD: Construction and completion of flux balance models from pathway databases. Bioinformatics. 2012, 28: 388-396.

    PubMed Central  PubMed  Google Scholar 

  16. Karp PD, Paley SM, Krummenacker M, Latendresse M, Dale JM, Lee T, Kaipa P, Gilham F, Spaulding A, Popescu L, Altman T, Paulsen I, Keseler IM, Caspi R: Pathway Tools version 13.0: integrated software for pathway/genome informatics and systems biology. Brief Bioinform. 2010, 11: 40-79. [doi:10.1093/bib/bbp043],

    PubMed Central  PubMed  Google Scholar 

  17. O’Brien EJ, Lerman JA, Chang RL, Hyduke DR, Palsson BØ: Genome-scale models of metabolism and gene expression extend and refine growth phenotype prediction. Mol Syst Biol. 2013, 9: 693-

    PubMed Central  PubMed  Google Scholar 

  18. Knuth DE: Literate programming. Comput J. 1984, 27 (2): 97-111.

    Google Scholar 

  19. Mackie A, Keseler IM, Nolan L, Karp PD, Paulsen IT: Dead end metabolites — defining the known unknowns of the E. colimetabolic network. PLoS One. 2013, 8 (9): 75210-

    Google Scholar 

  20. Schellenberger J, Que R, Fleming RM, Thiele I, Orth JD, Feist AM, Zielinski DC, Bordbar A, Lewis NE, Rahmanian S, Kang J, Hyduke DR, Palsson BØ: Quantitative prediction of cellular metabolism with constraint-based models: the COBRA toolbox v2.0. Nat Protoc. 2011, 6: 1290-1307.

    PubMed Central  PubMed  Google Scholar 

  21. Treberg JR, Brand MD: A model of the proton translocation mechanism of complex I. JBC. 2011, 286 (20): 17579-17584.

    Google Scholar 

  22. Wikstrom M, Hummer G: Stoichiometry of proton translocation by respiratory complex I and its mechanistic implications. PNAS. 2012, 109 (12): 4431-4436.

    PubMed Central  PubMed  Google Scholar 

  23. Kayser A, Weber J, Hecht V, Rinas U: Metabolic flux analysis of Escherichia coliin glucose-limited continuous culture. i. growth-rate-dependent metabolic efficiency at steady state. Microbiology. 2005, 151: 693-706.

    PubMed  Google Scholar 

  24. Belaich A, Belaich JP: Microcalorimetric study of the anaerobic growth of Escherichia coli: growth thermograms in a synthetic medium. J Bacteriol. 1976, 72: 497-499.

    Google Scholar 

  25. Varma A, Boesch BW, Palsson BØ: Stoichiometric interpretation of Escherichia coli glucose catabolism under various oxygenation rates. Appl Environ Microbiol. 1993, 59 (8): 2465-2473.

    PubMed Central  PubMed  Google Scholar 

  26. Zhuang K, Vemuri GN, Mahadevan R: Economics of membrane occupancy and respiro-fermentation. Mol Syst Biol. 2011, 7: 500-

    PubMed Central  PubMed  Google Scholar 

  27. van Hoek MJA, Merks RMH: Redox balance is key to explaining full vs. partial switching to low-yield metabolism. BMC Syst Biol. 2012, 6: 22-

    PubMed Central  PubMed  Google Scholar 

  28. Ibarra RU, Edwards JS, Palsson BØ: Escherichia coliK-12 undergoes adaptive evolution to achieve in silicopredicted optimal growth. Nature. 2002, 420: 186-189.

    PubMed  Google Scholar 

  29. Milne CB, Kim PJ, Eddy JA, Price ND: Accomplishments in genome-scale in silico modeling for industrial and medical biotechnology. Biotechnol J. 2009, 4 (12): 1653-16570.

    PubMed Central  PubMed  Google Scholar 

  30. Gianchandani EP, Chavali AK, Papin JA: The application of flux balance analysis in systems biology. WIREs Syst Biol Med. 2010, 2 (3): 372-382.

    Google Scholar 

  31. Lee JW, Na D, Park JM, Lee J, Choi S, Lee SY: Systems metabolic engineering of microorganisms for natural and non-natural chemicals. Nat Chem Biol. 2012, 8: 536-546.

    PubMed  Google Scholar 

  32. Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, Datsenko KA, Tomita M, Wanner BL: Construction of Escherichia coli K–12in-frame, single-gene knockout mutants: The Keio collection. Mol Syst Biol. 2006, 2 (2006): 0008-

    PubMed  Google Scholar 

  33. Yamamoto N, Nakahigashi K, Nakamichi T, Yoshino M, Takai Y, Touda Y, Furubayashi A, Kinjyo S, Dose H, Hasegawa M, Datsenko KA, Nakayashiki T, Tomita M, Wanner BL, Mori H: Update on the collection of Escherichia colisingle-gene deletion mutants. Mol Syst Biol. 2009, 5: 335-

    PubMed Central  PubMed  Google Scholar 

  34. Joyce AR, Reed JL, White A, Edwards R, Osterman A, Baba T, Mori H, Lesely SA, Palsson BØ: Experimental and computational assessment of conditionally essential genes in Escherichia coli. J Bacteriol. 2006, 188 (23): 8259-8271.

    PubMed Central  PubMed  Google Scholar 

  35. Cox RJ, Wang PSH: Is N-acetylornithine aminotransferase the real N-succinyl-LL-diaminopimelate aminotransferase in Escherichia coliandMycobacterium smegmatis?. J Chem Soc, Perkin Trans. 2001, 1: 2006-2008.

    Google Scholar 

  36. Kim J, Copley S: Why metabolic enzymes are essential or nonessential for growth of Escherichia coliK–12 on glucose. Biochemistry. 2007, 46 (44): 12501-12511.

    PubMed  Google Scholar 

  37. van der Ploeg JR, Eichhorn E, Leisinger T: Sulfonate-sulfur metabolism and its regulation in Escherichia coli. Arch Microbiol. 2001, 176: 1-8.

    PubMed  Google Scholar 

  38. Bykowski T, van der Ploeg JR, Iwanicka-Nowicka R, Hryniewicz MM: The switch from inorganic to organic sulphur assimilation in Escherichia coli: adenosine 5’-phosphosulphate (APS) as a signalling molecule for sulphate excess. Mol Microbiol. 2002, 43 (5): 1347-1358.

    PubMed  Google Scholar 

  39. Richaud C, Higgins W, Mengin-Lecruelx D, Stragier P: Molecular cloning, characterization, and chromosomal localization of dapF, the Escherichia coligene for diaminopimelate epimerase. J Bacteriol. 1987, 169 (4): 1454-1459.

    PubMed Central  PubMed  Google Scholar 

  40. Mengin-Lecruelx D, Michaud C, Richaud C, Blanot D, van Heijenoort J: Incorporation of LL-diaminopimelic acid into peptidoglycan of Escherichia colimutants lacking diaminopimelate epimerase encoded by dapF. J Bacteriol. 1988, 170 (5): 2031-2039.

    Google Scholar 

  41. el-Hajj HH, Zhang H, Weiss B: Lethality of a dut (deoxyuridine triphosphatase) mutation in Escherichia coli. J Bacteriol. 1988, 170 (3): 1069-1075.

    PubMed Central  PubMed  Google Scholar 

  42. Gross M, Marianovsky I, Glaser G: MazG - a regulator of programmed cell death in Escherichia coli. Mol Microbiol. 2006, 59 (2): 590-601.

    PubMed  Google Scholar 

  43. Haussmann C, Rohdich F, Schmidt E, Bacher A, Richter G: Biosynthesis of pteridines in Escherichia coli. Structural and mechanistic similarity of dihydroneopterin-triphosphate epimerase and dihydroneopterin aldolase. J Biol Chem. 1998, 273 (28): 17418-17424.

    PubMed  Google Scholar 

  44. Kato J, Hashimoto M: Construction of consecutive deletions of the Escherichia colichromosome. Mol Syst Biol. 2007, 3: 132-

    PubMed Central  PubMed  Google Scholar 

  45. Fermer C, Swedberg G: Adaptation to sulfonamide resistance in Neisseria meningitidismay have required compensatory changes to retain enzyme function: kinetic analysis of dihydropteroate synthases from N. meningitidisexpressed in a knockout mutant of Escherichia coli. J Bacteriol. 1997, 179 (3): 831-837.

    PubMed Central  PubMed  Google Scholar 

  46. Inoue A, Murata Y, Takahashi H, Tsuji N, Fujisaki S, Kato J: Involvement of an essential gene,mviN, in murein synthesis inEscherichia coli. J Bacteriol. 2008, 190 (21): 7298-7301.

    PubMed Central  PubMed  Google Scholar 

  47. Ruiz N: Bioinformatics identification of MurJ (MviN) as the peptidoglycan lipid II flippase in Escherichia coli. PNAS. 2008, 105 (21): 15553-15557.

    PubMed Central  PubMed  Google Scholar 

  48. Mohammadi T, van Dam V, Sijbrandi R, Vernet T, Zapun A, Bouhss A, MD-d Bruin, Nguyen-Disteche M, de Kruijff B: Identification of FtsW as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J. 2011, 30 (8): 1425-1432.

    PubMed Central  PubMed  Google Scholar 

  49. Jermy A: Bacterial physiology: flipping lipids. Nat Rev Microbiol. 2011, 9 (5): 314-

    Google Scholar 

  50. Butler EK, Davis RM, Bari V, Nicholson PA, Ruiz N: Structure-function analysis of MurJ reveals a solvent-exposed cavity containing residues essential for peptidoglycan biogenesis inEscherichia coli. J Bacteriol. 2013, 195 (20): 4639-4649.

    PubMed Central  PubMed  Google Scholar 

  51. Sperandeo P, Pozzi C, Deho G, Polissi A: Non-essential kdo biosynthesis and new essential cell envelope biogenesis genes in the Escherichia coli yrbG-yhbGlocus. Res Microbiol. 2006, 157 (6): 547-558.

    PubMed  Google Scholar 

  52. Mamat U, Meredith TC, Aggarwal P, Kuhl A, Kirchoff P, Lindner B, Hanuszkiewicz A, Sun J, Holst O, Woodard RW: Single amino acid substitutions in either YhjD or MsbA confer viability to 3-deoxy-d-manno-oct-2-ulosonic acid-depleted Escherichia coli. Mol Microbiol. 2008, 67 (3): 633-648.

    PubMed  Google Scholar 

  53. Klein G, Lindner B, Brabetz W, Brade H, Raina S: Escherichia coli, K–12 suppressor-free mutants lacking early glycosyltransferases and late acyltransferases: minimal lipopolysaccharide structure and induction of envelope stress response. J Biol Chem. 2008, 284 (23): 15369-151389.

    Google Scholar 

  54. Green JM, Merkel WK, Nichols BP: Characterization and sequence of Escherichia coli pabC, the gene encoding aminodeoxychorismate lyase, a pyridoxal phosphate-containing enzyme. J Bacteriol. 1992, 174 (16): 5317-5323.

    PubMed Central  PubMed  Google Scholar 

  55. Langley D, Guest JR: Biochemical genetics of the α-keto acid dehydrogenase complexes of Escherichia coliK–12: Isolation and biochemical properties of deletion mutants. J Gen Microbiol. 1977, 99: 263-276.

    PubMed  Google Scholar 

  56. Chang YY, Cronan JE: Mapping nonselectable genes of Escherichia coliby using transposon Tn10: location of a gene affecting pyruvate oxidase. J Bacteriol. 1982, 151 (3): 1279-1289.

    PubMed Central  PubMed  Google Scholar 

  57. Abdel-Hamid AM, Attwood MM, Guest JR: Pyruvate oxidase contributes to the aerobic growth efficiency of Escherichia coli. Microbiology. 2001, 147: 1483-1498.

    PubMed  Google Scholar 

  58. Hillman JD, Fraenkel DG: Glyceraldehyde 3-phosphate dehydrogenase mutants of Escherichia coli. J Bacteriol. 1975, 122 (3): 1175-1179.

    PubMed Central  PubMed  Google Scholar 

  59. Irani M, Maitra PK: Properties of Escherichia colimutants deficient in enzymes of glycolysis. J Bacteriol. 1977, 132 (2): 398-410.

    PubMed Central  PubMed  Google Scholar 

  60. Reeves HC, Ajl SJ: Alpha-hydroxyglutaric acid synthetase. J Bacteriol. 1962, 84: 186-187.

    PubMed Central  PubMed  Google Scholar 

  61. Wegener WS, Reeves HC, Ajl SJ: Heterogeneity of the glyoxylate-condensing enzymes. J Bacteriol. 1965, 90 (3): 594-598.

    PubMed Central  PubMed  Google Scholar 

  62. Helling RB, Kukora JS: Nalidixic acd-resistant mutants of Escherichia colideficient in isocitrate dehydrogenase. J Bacteriol. 1971, 105 (3): 1224-1226.

    PubMed Central  PubMed  Google Scholar 

  63. Lakshmi TM, Helling RB: Selection for citrate synthase deficiency in icdmutants of Escherichia coli. J Bacteriol. 1976, 127 (1): 76-83.

    PubMed Central  PubMed  Google Scholar 

  64. Kabir MM, Shimizu K: Metabolic regulation analysis of icd-gene knockout Escherichia colibased on 2d electrophoresis with maldi-tof mass spectrometry and enzyme activity measurements. Appl Microbiol Biotechnol. 2004, 65 (1): 84-96.

    PubMed  Google Scholar 

  65. Lin H, Bennett GN, San KY: Genetic reconstruction of the aerobic central metabolism in Escherichia colifor the absolute aerobic production of succinate. Biotechnol Bioeng. 2005, 89 (2): 148-156.

    PubMed  Google Scholar 

  66. Kalliri E, Mulrooney SB, Hausinger RP: Identification of Escherichia coli ygaFas an L-2-hydroxyglutarate oxidase. J Bacteriol. 2008, 190 (11): 3793-3798.

    PubMed Central  PubMed  Google Scholar 

  67. Reinoso CA, Appanna VD, Vasquez CC: α-ketoglutarate accumulation is not dependent on isocitrate dehydrogenase activity during tellurite detoxification in Escherichia coli. Biomed Res Int. 2013, 2013: 784190-

    PubMed Central  PubMed  Google Scholar 

  68. Stribling D, Perham RN: Purification and characterization of two fructose diphosphate aldolases from Escherichia coli(Crookes’ strain). Biochem J. 1973, 131 (4): 833-841.

    PubMed Central  PubMed  Google Scholar 

  69. Scamuffa MD, Caprioli RM: Comparison of the mechanisms of two distinct aldolases from Escherichia coligrown on gluconeogenic substrates. Biochim Biophys Acta. 1980, 614 (2): 583-590.

    PubMed  Google Scholar 

  70. Schurmann M, Sprenger GA: Fructose-6-phosphate aldolase is a novel class I aldolase from Escherichia coliand is related to a novel group of bacterial transaldolases. JBC. 2001, 276: 11055-11061.

    Google Scholar 

  71. Baechler C, Schneider P, Baehler P, Lustig A, Erni B: Escherichia colidihydroxyacetone kinase controls gene expression by binding to transcription factor dhaR. EMBO J. 2005, 24 (2): 283-293.

    Google Scholar 

  72. Gerike U, Hough DW, Russell NJ, Dyall-Smith ML, Danson MJ: Citrate synthase and 2-methylcitrate synthase: structural, functional and evolutionary relationships. Microbiology. 1998, 144 (4): 929-935.

    PubMed  Google Scholar 

  73. Orth JD, Palsson B: Gap-filling analysis of the iJO1366 Escherichia colimetabolic network reconstruction for discovery of metabolic functions. BMC Syst Biol. 2012, 6: 30-

    PubMed Central  PubMed  Google Scholar 

  74. Kotlarz D, Garreau H, Buc H: Regulation of the amount and of the activity of phosphofructokinases and pyruvate kinases in Escherichia coli. Biochim Biophys Acta. 1975, 381 (2): 257-268.

    PubMed  Google Scholar 

  75. Lovingshimer MR, Siegele D, Reinhart GD: Construction of an inducible, pfkAand pfkBdeficient strain of Escherichia colifor the expression and purification of phosphofructokinase from bacterial sources. Protein Expr Purif. 2006, 46: 475-482.

    PubMed  Google Scholar 

  76. Peng L, Arauzo-Bravo MJ, Shimizu K: Metabolic flux analysis for a ppcmutant Escherichia colibased on 13C-labelling experiments together with enzyme activity assays and intracellular metabolite measurements. FEMS Microbiol Lett. 2004, 235 (1): 17-23.

    PubMed  Google Scholar 

  77. Patrick WM, Quandt EM, Swartzlander DB, Matsumura I: Multicopy suppression underpins metabolic evolvability. Mol Biol Evol. 2007, 24 (12): 2716-2722.

    PubMed Central  PubMed  Google Scholar 

  78. Ferguson GP, Totemeyer S, MacLean MJ, Booth IR: Methylglyoxal production in bacteria: suicide or survival?. Arch Microbiol. 1998, 170 (4): 209-219.

    PubMed  Google Scholar 

  79. Fong SS, Nanchen A, Palsson BØ: Latent pathway activation and increased pathway capacity enable Escherichia coliadaptation to loss of key metabolic enzymes. J Biol Chem. 2006, 281 (12): 8024-8033.

    PubMed  Google Scholar 

  80. Michel G, Roszak AW, Sauve V, Maclean J, Matte A, Coggins JR, Cygler M, Lapthorn AJ: Structures of shikimate dehydrogenase AroE and its paralog YdiB. a common structural framework for different activities. J Biol Chem. 2003, 278 (21): 19463-19472.

    PubMed  Google Scholar 

  81. Johansson L, Liden G: Transcriptome analysis of a shikimic acid producing strain of Escherichia coliW3110 grown under carbon- and phosphate-limited conditions. J Biotechnol. 2006, 126 (4): 528-545.

    PubMed  Google Scholar 

  82. Guilloton MB, Lamblin AF, Kozliak EI, Gerami-Nejad M, Tu C, Silverman D, Anderson PM, Fuchs JA: A physiological role for cyanate-induced carbonic anhydrase in Escherichia coli. J Bacteriol. 1993, 175 (5): 1443-1451.

    PubMed Central  PubMed  Google Scholar 

  83. Howell EE, Foster PG, Foster LM: Construction of a dihydrofolate reductase-deficient mutant of Escherichia coliby gene replacement. J Bacteriol. 1988, 170 (7): 3040-3045.

    PubMed Central  PubMed  Google Scholar 

  84. Pribat A, Blaby IK, Lara-Núñez A, Gregory JF3rd, de Crécy-Lagard V: FolX and FolM are essential for tetrahydromonapterin synthesis in Escherichia coliand Pseudomonas aeruginosa. J Bacteriol. 2010, 192 (2): 475-482.

    PubMed Central  PubMed  Google Scholar 

  85. Shim JH, Benkovic SJ: Evaluation of the kinetic mechanism of Escherichia coliglycinamide ribonucleotide transformylase. Biochemistry. 1998, 37 (24): 8776-8782.

    PubMed  Google Scholar 

  86. Newman EB, Kapoor V, Potter R: Role of L-threonine dehydrogenase in the catabolism of threonine and synthesis of glycine by Escherichia coli. J Bacteriol. 1976, 126 (3): 1245-1249.

    PubMed Central  PubMed  Google Scholar 

  87. Liu JQ, Dairi T, Itoh N, Kataoka M, Shimizu S, Yamada H: Gene cloning, biochemical characterization and physiological role of a thermostable low-specificity L-threonine aldolase from Escherichia coli. Eur J Biochem. 1998, 255 (1): 220-226.

    PubMed  Google Scholar 

  88. Vales LD, Chase JW, Murphy JB: Orientation of the guanine operon of Escherichia coliK–12 by utilizing strains containing guaB-xseand guaB-uppdeletions. J Bacteriol. 1979, 139 (1): 320-322.

    PubMed Central  PubMed  Google Scholar 

  89. Umbarger HE, Brown B: Threonine deamination in Escherichia coli. II. Evidence for two L-threonine deaminases. J Bacteriol. 1957, 73 (1): 105-112.

    PubMed Central  PubMed  Google Scholar 

  90. Zhao X, Miller JR, Jiang Y, Marletta MA, Cronan JE: Assembly of the covalent linkage between lipoic acid and its cognate enzymes. Chem Biol. 2003, 10 (12): 1293-1302.

    PubMed  Google Scholar 

  91. Reidl J, Boos W: The malX malYoperon of Escherichia coliencodes a novel enzyme II of the phosphotransferase system recognizing glucose and maltose and an enzyme abolishing the endogenous induction of the maltose system. J Bacteriol. 1991, 173 (15): 4862-4876.

    PubMed Central  PubMed  Google Scholar 

  92. Zdych E, Peist R, Reidl J, Boos W: MalY of Escherichia coliis an enzyme with the activity of a beta C-S lyase (cystathionase). J Bacteriol. 1995, 177 (17): 5035-5039.

    PubMed Central  PubMed  Google Scholar 

  93. Boos W, Shuman H: Maltose/maltodextrin system of Escherichia coli: transport, metabolism, and regulation. Microbiol Mol Biol Rev. 1998, 62 (1): 204-229.

    PubMed Central  PubMed  Google Scholar 

  94. Cohen GN, Stanier RY, Bras GL: Regulation of the biosynthesis of amino acids of the aspartate family in coliform bacteria and pseudomonads. J Bacteriol. 1969, 99 (3): 791-801.

    PubMed Central  PubMed  Google Scholar 

  95. Garriga X, Eliasson R, Torrents E, Jordan A, Barbe J, Gibert I, Reichard P: nrdDand nrdGgenes are essential for strict anaerobic growth of Escherichia coli. Biochem Biophys Res Commun. 1996, 229 (1): 189-192.

    PubMed  Google Scholar 

  96. Lam HM, Winkler ME: Metabolic relationships between pyridoxine (vitamin B6) and serine biosynthesis in Escherichia coliK–12. J Bacteriol. 1990, 172 (11): 6518-6528.

    PubMed Central  PubMed  Google Scholar 

  97. Kim J, Kershner JP, Novikov Y, Shoemaker RK, Copley SD: Three serendipitous pathways in E. colican bypass a block in pyridoxal-5’-phosphate synthesis. Mol Syst Biol. 2010, 6: 436-

    PubMed Central  PubMed  Google Scholar 

  98. Hove-Jensen B, Rosenkrantz TJ, Zechel DL, Willemoes M: Accumulation of intermediates of the carbon-phosphorus lyase pathway for phosphonate degradation in phn mutants of Escherichia coli. J Bacteriol. 2010, 192 (1): 370-374.

    PubMed Central  PubMed  Google Scholar 

  99. Zhang Q, van der Donk WA: Answers to the carbon-phosphorus lyase conundrum. Chembiochem. 2012, 13 (5): 627-629.

    PubMed Central  PubMed  Google Scholar 

  100. Wanner BL: Gene regulation by phosphate in enteric bacteria. J Cell Biochem. 1993, 51 (1): 47-54.

    PubMed  Google Scholar 

  101. Wanner BL, Boline JA: Mapping and molecular cloning of the phn (psiD) locus for phosphonate utilization in Escherichia coli. J Bacteriol. 1990, 172 (3): 1186-1196.

    PubMed Central  PubMed  Google Scholar 

  102. Ravnikar PD, Somerville RL: Genetic characterization of a highly efficient alternate pathway of serine biosynthesis in Escherichia coli. J Bacteriol. 1987, 169 (6): 2611-2617.

    PubMed Central  PubMed  Google Scholar 

  103. Kim C, Song S, Park C: The D-allose operon of Escherichia coliK–12. J Bacteriol. 1997, 179 (24): 7631-7637.

    PubMed Central  PubMed  Google Scholar 

  104. Poulsen TS, Chang YY, Hove-Jensen B: D-allose catabolism of Escherichia coli: involvement of alsIand regulation of alsregulon expression by allose and ribose. J Bacteriol. 1999, 181 (22): 7126-7130.

    PubMed Central  PubMed  Google Scholar 

  105. Pittard J, Wallace BJ: Distribution and function of genes concerned with aromatic biosynthesis inEscherichia coli. J Bacteriol. 1966, 91 (4): 1494-1508.

    PubMed Central  PubMed  Google Scholar 

  106. Bottomley JR, Clayton CL, Chalk PA, Kleanthous C: Cloning, sequencing, expression, purification, and preliminary characterization of a type II dehydroquinase from Helicobacter pylori. Biochem J. 1996, 319 (2): 559-565.

    PubMed Central  PubMed  Google Scholar 

  107. von Meyenburg K, Jorgensen BB, Nielsen J, Hansen FG: Promoters of the atpoperon coding for the membrane-bound ATP synthase of Escherichia colimapped by Tn10 insertion mutations. Mol Gen Genet. 1982, 188 (2): 240-248.

    PubMed  Google Scholar 

  108. Jensen PR, Michelsen O: Carbon and energy metabolism of atpmutants of Escherichia coli. J Bacteriol. 1992, 174 (23): 7635-7641.

    PubMed Central  PubMed  Google Scholar 

  109. Green GN, Gennis RB: Isolation and characterization of an Escherichia colimutant lacking cytochrome d terminal oxidase. J Bacteriol. 1983, 154 (3): 1269-1275.

    PubMed Central  PubMed  Google Scholar 

  110. Georgiou CD, Fang H, Gennis RB: Identification of the cydClocus required for expression of the functional form of the cytochrome d terminal oxidase complex in Escherichia coli. J Bacteriol. 1987, 169 (5): 2107-2112.

    PubMed Central  PubMed  Google Scholar 

  111. Puustinen A, Finel M, Haltia T, Gennis RB, Wikstrom M: Properties of the two terminal oxidases of Escherichia coli. Biochemistry. 1991, 30 (16): 3936-3942.

    PubMed  Google Scholar 

  112. Pittman MS, Robinson HC, Poole RK: A bacterial glutathione transporter (Escherichia coliCydDC) exports reductant to the periplasm. J Biol Chem. 2005, 280 (37): 32254-32261.

    PubMed  Google Scholar 

  113. Portnoy VA, Herrgard MJ, Palsson BØ: Aerobic fermentation of d-glucose by an evolved cytochrome oxidase-deficient Escherichia colistrain. Appl Environ Microbiol. 2008, 74 (24): 7561-7569.

    PubMed Central  PubMed  Google Scholar 

  114. Borisov VB, Gennis RB, Hemp J, Verkhovsky MI: The cytochrome bd respiratory oxygen reductases. Biochim Biophys Acta. 2011, 1807 (11): 1398-1413.

    PubMed Central  PubMed  Google Scholar 

  115. Chou CH, Bennett GN, San KY: Effect of modified glucose uptake using genetic engineering techniques on high-level recombinant protein production in Escherichia colidense cultures. Biotechnol Bioeng. 1994, 44 (8): 952-960.

    PubMed  Google Scholar 

  116. Flores N, Xiao J, Berry A, Bolivar F, Valle F: Pathway engineering for the production of aromatic compounds in Escherichia coli. Nat Biotechnol. 1996, 14 (5): 620-623.

    PubMed  Google Scholar 

  117. Chen R, Yap WM, Postma PW, Bailey JE: Comparative studies of Escherichia colistrains using different glucose uptake systems: metabolism and energetics. Biotechnol Bioeng. 1997, 56 (5): 583-590.

    PubMed  Google Scholar 

  118. Zeppenfeld T, Larisch C, Lengeler JW, Jahreis K: Glucose transporter mutants of Escherichia coliK-12 with changes in substrate recognition of IICB(Glc) and induction behavior of the ptsGgene. J Bacteriol. 2000, 182 (16): 4443-4452.

    PubMed Central  PubMed  Google Scholar 

  119. Flores N, Flores S, Escalante A, de Anda R: Adaptation for fast growth on glucose by differential expression of central carbon metabolism and galregulon genes in an Escherichia colistrain lacking the phosphoenolpyruvate:carbohydrate phosphotransferase system. Metab Eng. 2005, 7 (2): 70-87.

    PubMed  Google Scholar 

  120. Steinsiek S, Bettenbrock K: Glucose transport in Escherichia colimutant strains with defects in sugar transport systems. J Bacteriol. 2012, 194 (21): 5897-5908.

    PubMed Central  PubMed  Google Scholar 

  121. Escalante A, Cervantes AS, Gosset G, Bolivar F: Current knowledge of theEscherichia coliphosphoenolpyruvate-carbohydrate phosphotransferase system: peculiarities of regulation and impact on growth and product formation. Appl Microbiol Biotechnol. 2012, 94 (6): 1483-1494.

    PubMed  Google Scholar 

  122. Sarubbi E, Rudd KE, Cashel M: Basal ppgpp level adjustment shown by new spoTmutants affect steady state growth rates and rrnAribosomal promoter regulation in Escherichia coli. Mol Gen Genet. 1988, 213 (2–3): 214-222.

    PubMed  Google Scholar 

  123. Sarubbi E, Rudd KE, Xiao H, Ikehara K, Kalman M, Cashel M: Characterization of the spoTgene of Escherichia coli. J Biol Chem. 1989, 264 (25): 15074-15082.

    PubMed  Google Scholar 

  124. Xiao H, Kalman M, Ikehara K, Zemel S, Glaser G, Cashel M: Residual guanosine 3’,5’-bispyrophosphate synthetic activity of relAnull mutants can be eliminated by spoTnull mutations. J Biol Chem. 1991, 266 (9): 5980-5990.

    PubMed  Google Scholar 

  125. Gentry DR, Cashel M: Mutational analysis of the Escherichia coli spoTgene identifies distinct but overlapping regions involved in ppgpp synthesis and degradation. Mol Microbiol. 1996, 19 (6): 1373-1384.

    PubMed  Google Scholar 

  126. Jin DJ, Cagliero C, Zhou YN: Growth rate regulation in Escherichia coli. FEMS Microbiol Rev. 2012, 36 (2): 269-287.

    PubMed Central  PubMed  Google Scholar 

  127. Cox GB, Gibson F, Pittard J: Mutant strains of Escherichia coliK–12 unable to form ubiquinone. J Bacteriol. 1968, 95 (5): 1591-1598.

    PubMed Central  PubMed  Google Scholar 

  128. Wu G, Williams HD, Gibson F, Poole RK: Mutants of Escherichia coliaffected in respiration: the cloning and nucleotide sequence of ubiA, encoding the membrane-bound p-hydroxybenzoate:octaprenyltransferase. J Gen Microbiol. 1993, 139 (8): 1795-1805.

    PubMed  Google Scholar 

  129. Klena JD, Ashford RS, Schnaitman CA: Role of Escherichia coliK-12 rfa genes and the rfp gene of Shigella dysenteriae1 in generation of lipopolysaccharide core heterogeneity and attachment of o antigen. J Bacteriol. 1992, 174 (22): 7297-7307.

    PubMed Central  PubMed  Google Scholar 

  130. Roncero C, Casadaban MJ: Genetic analysis of the genes involved in synthesis of the lipopolysaccharide core in Escherichia coliK–12: three operons in the rfalocus. J Bacteriol. 1992, 174 (10): 3250-3260.

    PubMed Central  PubMed  Google Scholar 

  131. Hwang J, Inouye M: The tandem GTPase, der, is essential for the biogenesis of 50S ribosomal subunits in Escherichia coli. Mol Microbiol. 2006, 61 (6): 1660-1672.

    PubMed  Google Scholar 

  132. Bharat A, Jiang M, Sullivan SM, Maddock JR, Brown ED: Cooperative and critical roles for both G domains in the GTPase activity and cellular function of ribosome-associated Escherichia coliEngA. J Bacteriol. 2006, 188 (22): 7992-7996.

    PubMed Central  PubMed  Google Scholar 

  133. Tomar SK, Dhimole N, Chatterjee M, Prakash B: Distinct GDP/GTP bound states of the tandem G-domains of EngA regulate ribosome binding. Nucleic Acids Res. 2009, 37 (7): 2359-2370.

    PubMed Central  PubMed  Google Scholar 

  134. Spratt BG: Distinct penicillin binding proteins involved in the division, elongation, and shape of Escherichia coliK–12. PNAS. 2009, 72 (8): 2999-3003.

    Google Scholar 

  135. Ogura T, Bouloc P, Niki H, D’Ari R, Hiraga S, Jaffe A: Penicillin-binding protein 2 is essential in wild-type Escherichia colibut not in lovor cyamutants. J Bacteriol. 1989, 171 (6): 3025-3030.

    PubMed Central  PubMed  Google Scholar 

  136. Wang Y, Stieglitz KA, Bubunenko M, Court DL, Stec B, Roberts MF: The structure of the R184A mutant of the inositol monophosphatase encoded by suhBand implications for its functional interactions in Escherichia coli. J Biol Chem. 2007, 282 (37): 26989-26996.

    PubMed  Google Scholar 

  137. Puan KJ, Wang H, Dairi T, Kuzuyama T, Morita CT: fldAis an essential gene required in the 2-C-methyl-D-erythritol 4-phosphate pathway for isoprenoid biosynthesis. FEBS Lett. 2005, 579 (17): 3802-3806.

    PubMed  Google Scholar 

  138. Lu Q, Inouye M: Adenylate kinase complements nucleoside diphosphate kinase deficiency in nucleotide metabolism. PNAS. 1996, 93 (12): 5720-5725.

    PubMed Central  PubMed  Google Scholar 

  139. Gerhart JC, Schachman HK: Distinct subunits for the regulation and catalytic activity of aspartate transcarbamylase. Biochemistry. 1965, 4 (6): 1054-1062.

    PubMed  Google Scholar 

  140. Reed JL, Patel TR, Chen KH, Joyce AR, Applebee MK, Herring CD, Bui OT, Knight EM, Fong SS, Palsson BØ: Systems approach to refining genome annotation. Proc Natl Acad Sci U S A. 2006, 103 (46): 17480-17484.

    PubMed Central  PubMed  Google Scholar 

  141. Kumar VS, Dasika MS, Maranas CD: Optimization based automated curation of metabolic reconstructions. BMC Bioinformatics. 2007, 8: 212-

    Google Scholar 

  142. Kumar VS, Maranas CD: GrowMatch: An automated method for reconciling in silico/in vivo growth predictions. PLoS Comput Biol. 2009, 5 (3): 1000308-

    Google Scholar 

  143. Barua D, Kim J, Reed JL: An automated phenotype-driven approach (geneforce) for refining metabolic and regulatory models. PLoS Comput Biol. 2010, 6 (10): 1000970-

    Google Scholar 

  144. Orth JD, Palsson BØ: Systematizing the generation of missing metabolic knowledge. Biotechnol Bioeng. 2010, 107 (3): 403-412.

    PubMed Central  PubMed  Google Scholar 

  145. Tervo CJ, Reed JL: BioMog: a computational framework for the de novo generation or modification of essential biomass components. PLoS One. 2013, 8 (12): 81322-

    Google Scholar 

  146. Monod J: The growth of bacterial cultures. Annu Rev Microbiol. 1949, 3: 371-394.

    Google Scholar 

  147. Bochner BR, Gadzinski P, Panomitros E: Phenotype microarrays for high-throughput phenotypic testing and assay of gene function. Genome Res. 2001, 11 (7): 1246-1255.

    PubMed Central  PubMed  Google Scholar 

  148. Shea A, Wolcott M, Daefler S, Rozak DA: Biolog phenotype microarrays. Microbial Systems Biology: Methods and Protocols. 2012, New York: Humana Press, 331-73. Chap. 12,

    Google Scholar 

  149. Mackie AM, Hassan KA, Paulsen IT, Tetu SG: Biolog phenotype microarrays for phenotypic characterization of microbial cells. Methods Mol Biol. 1096, 2014: 123-301.

    Google Scholar 

  150. Oun MA, Suthers PF, Jones GI, Carter BR, Saunders MP, Maranas CD, Woodward MJ, Anjum MF: Genome scale reconstruction of a salmonella metabolic model: comparison of similarity and differences with a commensal Escherichia colistrain. J Biol Chem. 2009, 284 (43): 29480-29488.

    Google Scholar 

  151. Yoon SH, Han MJ, Jeong H, Lee CH, Xia XX, Lee DH, Shim JH, Lee SY, Oh TK, Kim JF: Comparative multi-omics systems analysis of Escherichia colistrains B and K–12. Genome Biol. 2012, 13 (5): 37-

    Google Scholar 

  152. Mackie A, Paley S, Keseler IM, Shearer A, Paulsen IT, Karp PD: Addition of Escherichia coli K–12 growth-observation and gene essentiality data to the EcoCyc database. J Bacteriol. 2014, 196 (5): 982-988.

    PubMed Central  PubMed  Google Scholar 

  153. Wilson DM, Wilson TH: Cation specificity for sugar substrates of the melibiose carrier in Escherichia coli. Biochim Biophys Acta. 1987, 904 (2): 191-200.

    PubMed  Google Scholar 

  154. Sandermann H Jr: β-D-galactoside transport in Escherichia coli: substrate recognition. Eur J Biochem. 1977, 80 (2): 507-515.

    PubMed  Google Scholar 

  155. Sahota SS, Bramley PM, Menzies IS: The fermentation of lactulose by colonic bacteria. Microbiology. 1981, 128 (2): 319-325.

    Google Scholar 

  156. Wallenfels K, Weil R: β-galactosidase. The Enzymes, 3rd Ed. 1972, New York: Academic Press, 617-663.

    Google Scholar 

  157. Roderick SL: The lac operon galactoside acetyltransferase. CR Biologies. 2005, 328: 568-575.

    Google Scholar 

  158. Lang VJ, Leystra-Lantz C, Cook RA: Characterization of the specific pyruvate transport system in Escherichia coliK–12. J Bacteriol. 1987, 169 (1): 380-385.

    PubMed Central  PubMed  Google Scholar 

  159. Kim OB, Reimann J, Lukas H, Schumacher U, Grimpo J, Dunnwald P, Unden G: Regulation of tartrate metabolism by TtdR and relation to the DcuS/DcuR-regulated C4-dicarboxylate metabolism of Escherichia coli. Microbiol. 2009, 155 (11): 3632-3640.

    Google Scholar 

  160. Lukas H, Reimann J, Kim OB, Grimpo J, Unden G: Regulation of aerobic and anaerobic d-malate metabolism of Escherichia coliby the LysR-type regulator DmlR (YeaT). J Bacteriol. 2010, 192 (10): 2503-2511.

    PubMed Central  PubMed  Google Scholar 

  161. Lauritzen AM, Lipscomb WN: Modification of three active site lysine residues in the catalytic subunit of aspartate transcarbamylase by D- and L-bromosuccinate. J Biol Chem. 1982, 257: 1312-1319.

    PubMed  Google Scholar 

  162. Hall BG: Chromosomal mutation for citrate utilization by Escherichia coliK–12. J Bacteriol. 1982, 151 (1): 269-273.

    PubMed Central  PubMed  Google Scholar 

  163. Lutgens M, Gottschalk G: Why a co-substrate is required for anaerobic growth of Escherichia colion citrate. J Gen Microbiol. 1980, 119 (1): 63-70.

    PubMed  Google Scholar 

  164. Schneider BL, Ruback S, Kiupakis AK, Kasbarian H, Pybus C, Reitzer L: The Escherichia coli gabDTPCoperon: specific gamma-aminobutyrate catabolism and nonspecific induction. J Bacteriol. 2002, 184 (24): 6976-6986.

    PubMed Central  PubMed  Google Scholar 

  165. Cunin R, Glansdorff N, Pierard A, Stalon V: Biosynthesis and metabolism of arginine in bacteria. Microbiol Rev. 1986, 50 (3): 314-352.

    PubMed Central  PubMed  Google Scholar 

  166. Schneider BL, Kiupakis AK, Reitzer LJ: Arginine catabolism and the arginine succinyltransferase pathway in Escherichia coli. J Bacteriol. 1998, 180 (16): 4278-4286.

    PubMed Central  PubMed  Google Scholar 

  167. Plumbridge J, Pellegrini O: Expression of the chitobiose operon of Escherichia coliis regulated by three transcription factors: NagC, ChbR and CAP. Mol Microbiol. 2004, 52 (2): 437-449.

    PubMed  Google Scholar 

  168. Kachroo AH, Kancherla AK, Singh NS, Varshney U, Mahadevan S: Mutations that alter the regulation of the chb operon of Escherichia coliallow utilization of cellobiose. Mol Microbiol. 2007, 66 (2): 1382-1395.

    PubMed  Google Scholar 

  169. Chang GW, Chang JT: Evidence for the B12-dependent enzyme ethanolamine deaminase inSalmonella. Nature. 1975, 254 (5496): 150-151.

    PubMed  Google Scholar 

  170. Jones PW, Turner JM: Interrelationships between the enzymes of ethanolamine metabolism in Escherichia coli. J Gen Microbiol. 1984, 130 (2): 299-308.

    PubMed  Google Scholar 

  171. Blackwell CM, Turner JM: Microbial metabolism of amino alcohols. formation of coenzyme B12-dependent ethanolamine ammonia-lyase and its concerted induction in Escherichia coli. Biochem J. 1978, 176 (3): 751-757.

    PubMed Central  PubMed  Google Scholar 

  172. Akita K, Hieda N, Baba N, Kawaguchi S, Sakamoto H, Nakanishi Y, Yamanishi M, Mori K, Toraya T: Purification and some properties of wild-type and N-terminal-truncated ethanolamine ammonia-lyase of Escherichia coli. J Biochem. 2010, 147 (1): 83-93.

    PubMed  Google Scholar 

  173. Xi H, Schneider BL, Reitzer L: Purine catabolism in Escherichia coliand function of xanthine dehydrogenase in purine salvage. J Bacteriol. 2000, 182 (19): 5332-5341.

    PubMed Central  PubMed  Google Scholar 

  174. Reitzer L: Nitrogen assimilation and global regulation in Escherichia coli. Annu Rev Microbiol. 2003, 57: 155-176.

    PubMed  Google Scholar 

  175. Cusa E, Obradors N, Baldoma L, Badia J, Aguilar J: Genetic analysis of a chromosomal region containing genes required for assimilation of allantoin nitrogen and linked glyoxylate metabolism in Escherichia coli. J Bacteriol. 1999, 181 (24): 7479-7484.

    PubMed Central  PubMed  Google Scholar 

  176. Harborne NR, Griffiths L, Busby SJ, Cole JA: Transcriptional control, translation and function of the products of the five open reading frames of the Escherichia coli niroperon. Mol Microbiol. 1992, 6 (19): 2805-2813.

    PubMed  Google Scholar 

  177. Wang H, Tseng CP, Gunsalus RP: The napFand narGnitrate reductase operons in Escherichia coliare differentially expressed in response to submicromolar concentrations of nitrate but not nitrite. J Bacteriol. 1999, 181 (17): 5303-5308.

    PubMed Central  PubMed  Google Scholar 

  178. Bertero MG, Rothery RA, Palak M, Hou C, Lim D, Blasco F, Weiner JH, Strynadka NC: Insights into the respiratory electron transfer pathway from the structure of nitrate reductase A. Nat Struct Biol. 2003, 10 (9): 681-687.

    PubMed  Google Scholar 

  179. Seiflein TA, Lawrence JG: Methionine-to-cysteine recycling in Klebsiella aerogenes. J Bacteriol. 2001, 183 (1): 336-346.

    PubMed Central  PubMed  Google Scholar 

  180. Delavier-Klutchko C, Flavin M: Enzymatic synthesis and cleavage of cystathionine in fungi and bacteria. J Biol Chem. 1965, 240: 2537-2549.

    PubMed  Google Scholar 

  181. Agren R, Liu L, Shoaie S, Vongsangnak W, Nookaew I, Nielsen J: The RAVEN toolbox and its use for generating a genome-scale metabolic model for Penicillium chrysogenum. PLoS Comput Biol. 2013, 9 (3): 1002980-

    Google Scholar 

  182. Thorleifsson SG, Thiele I: rBioNet: a COBRA toolbox extension for reconstructing high-quality biochemical networks. Bioinformatics. 2011, 27 (14): 2009-2010.

    PubMed  Google Scholar 

  183. Dandekar T, Fieselmann A, Majeed S, Ahmed Z: Software applications toward quantitative metabolic flux analysis and modeling. Brief Bioinform. 2014, 15 (1): 91-107.

    PubMed  Google Scholar 

  184. Lakshmanan M, Koh G, Chung BK, Lee DY: Software applications for flux balance analysis. Brief Bioinform. 2014, 15 (1): 108-122.

    PubMed  Google Scholar 

  185. Tomar N, De RK: Comparing methods for metabolic network analysis and an application to metabolic engineering. Gene. 2013, 521 (1): 1-14.

    PubMed  Google Scholar 

  186. Hamilton JJ, Reed JL: Software platforms to facilitate reconstructing genome-scale metabolic networks. Environ Microbiol. 2014, 16 (1): 49-59.

    PubMed  Google Scholar 

  187. Brochado AR, Andrejev S, Maranas CD, Patil KR: Impact of stoichiometry representation on simulation of genotype-phenotype relationships in metabolic networks. PLoS Comput Biol. 2012, 8 (11): 1002758-

    Google Scholar 

  188. Ebrahim A, Lerman JA, Palsson BØ: COBRApy: COnstraints-Based Reconstruction and Analysis for Python. BMC Syst Biol. 2013, 7: 74-

    PubMed Central  PubMed  Google Scholar 

  189. Bochner BR: Global phenotypic characterization of bacteria. FEMS Microbiol Rev. 2009, 33 (1): 191-205.

    PubMed Central  PubMed  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Mario Latendresse and Markus Krummenacker for MetaFlux development and support, and Larry Reitzer for calling our attention to the problem of dapC essentiality. This work was supported by award number U24GM077678 from the National Institute of General Medical Sciences. The content of this article is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of General Medical Sciences or the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel S Weaver.

Additional information

Competing interests

DSW, IMK, and PDK receive fees from SRI International commercial licensing of Pathway Tools.

Authors’ contributions

PDK and DSW conceived and designed the experiments. DSW constructed the model and conducted the simulations. DSW analyzed the data. IMK, AM, and DSW curated the database. DSW and PDK wrote the manuscript. AM and ITP performed phenotype microarray experiments. All authors read and approved the final manuscript.

Electronic supplementary material

Additional file 1: Supplementary information text and table descriptions.(PDF 164 KB)

Additional file 2: Supplementary tables (11 tables).(XLSX 246 KB)

12918_2014_1341_MOESM3_ESM.zip

Additional file 3: MetaFlux .fba file demonstrating simulation of aerobic growth of E. coli BW25113 on glucose.(ZIP 3 KB)

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Authors’ original file for figure 2

Authors’ original file for figure 3

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly credited. 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Weaver, D.S., Keseler, I.M., Mackie, A. et al. A genome-scale metabolic flux model of Escherichia coli K–12 derived from the EcoCyc database. BMC Syst Biol 8, 79 (2014). https://doi.org/10.1186/1752-0509-8-79

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1752-0509-8-79

Keywords