Skip to main content
Log in

On the Question of the Metabolic Costs of the Main Metabolic Precursors in Escherichia coli

  • SYSTEM ANALYSIS, MATHEMATICAL MODELING, AND INFORMATION SYSTEMS
  • Published:
Applied Biochemistry and Microbiology Aims and scope Submit manuscript

Abstract

The term metabolic cost (MС) is often used for assessment of energy consumption in the biosynthesis of various substances under different growth conditions or by different cell types. The MC of the metabolite is calculated according to a specified algorithm in universal ~P units, multiples of ATP molecules hydrolyzed to ADP and inorganic phosphate. Our analysis of the published data showed that the interpretation of the MC concept and the algorithms for its calculation, proposed by different authors, differ significantly. Since the MС is often considered in connection with system-level tasks, such as the metabolic flux analysis and the natural selection mechanisms, it seems appropriate to characterize this concept in detail. In this work, the term MС was clearly defined and used to calculate the energy consumption for the synthesis of 13 metabolic precursors of Escherichia coli biomass based on the modern model of the central metabolism of this bacterium. It was found that the MC, expressed in units of stored or hydrolyzed ATP molecules (~P), depends on the characteristics of the metabolism of an individual organism, its culturing conditions, and the P/O ratio, which characterizes the number of ATP molecules formed during the transfer of one electron pair to one oxygen atom in oxidative phosphorylation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Notes

  1. In a strict sense, glycolysis refers to the full set of biochemical reactions in the organism that are involved in the glucose conversion to pyruvate. For instance, the E. coli cell typically has 3 glycolytic pathways: Embden–Meyerhof–Parnas (EMP), pentose phosphate (PP), and Entner–Doudoroff (ED) [8]. However, another glycolytic pathway, via gluconate, can also be activated in the presence of pyrroloquinoline quinone (PQQ), a glucose dehydrogenase cofactor [9].

REFERENCES

  1. Varma, A. and Palsson, B.O., Metabolic flux balancing: basic concepts, scientific and practical use, BioTechnology, 1994, vol. 12, pp. 994–998. https://doi.org/10.1038/nbt1094-994

    Article  CAS  Google Scholar 

  2. Varma, A. and Palsson, B.O., Metabolic capabilities of Escherichia coli: I. Synthesis of biosynthetic precursors and cofactors, J. Theor. Biol., 1993, vol. 165, pp. 477–502. https://doi.org/10.1006/jtbi.1993.1202

    Article  CAS  PubMed  Google Scholar 

  3. Stouthamer, A.H., A theoretical study on the amount of ATP required for synthesis of microbial cell material, Antonie Van Leeuwenhoek, 1973, vol. 39, pp. 545–565. https://doi.org/10.1007/BF02578899

    Article  CAS  PubMed  Google Scholar 

  4. Akashi, H. and Gojobori, T., Metabolic efficiency and amino acid composition in the proteomes of Escherichia coli and Bacillus subtilis, Proc. Natl. Acad. Sci. U. S. A., 2002, vol. 99, pp. 3695–3700. https://doi.org/10.1073/pnas.062526999

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Craig, C. and Weber, R., Selection costs of amino acid substitutions in ColE1 and ColIa gene clusters harbored by Escherichia coli, Mol. Biol. Evol., 1998, vol. 15, no. 6, pp. 774–776. https://doi.org/10.1093/oxfordjournals.molbev.a025981

    Article  CAS  PubMed  Google Scholar 

  6. Zamboni, N., Fendt, S.-M., Ruhl, M., and Sauer, U., 13C-based metabolic flux analysis, Nat. Protoc., 2009, vol. 4, pp. 878–892. https://doi.org/10.1038/nprot.2009.58

    Article  CAS  PubMed  Google Scholar 

  7. Long, C.P. and Antoniewicz, M.R., High-resolution 13C metabolic flux analysis, Nat. Protoc., 2019, vol. 14, pp. 2856–2877. https://doi.org/10.1038/s41596-019-0204-0

    Article  CAS  PubMed  Google Scholar 

  8. Stephanopoulos, G.N., Aristidou, A.A., and Nielsen, J.H., Metabolic Engineering: Principles and Methodologies, Stephanopoulos, G., Ed., San Diego, USA: Academic, 1998.

    Google Scholar 

  9. Andreeva, I.G., Golubeva, L.I., Kuvaeva, T.M., et al., Identification of Pantoea ananatis gene encoding membrane pyrroloquinoline quinone (PQQ)-dependent glucose dehydrogenase and pqqABCDEF operon essential for PQQ biosynthesis, FEMS Microbiol. Lett., 2011, vol. 318, pp. 55–60. https://doi.org/10.1111/j.1574-6968.2011.02240.x

    Article  CAS  PubMed  Google Scholar 

  10. Sovremennaya mikrobiologiya. Prokarioty (v 2 tomakh) (Modern Microbiology. Prokaryotes (in 2 vols.)), Lengeler, I., Drevs, G., and Shlegel, G., Eds., Moscow: Mir, 2005, vol. 1.

  11. Noor, E., Eden, E., Milo, R., and Alon, U., Central carbon metabolism as a minimal biochemical walk between precursors for biomass and energy, Mol. Cell, 2010, vol. 39, pp. 809–820. https://doi.org/10.1016/j.molcel.2010.08.031

    Article  CAS  PubMed  Google Scholar 

  12. Gopalakrishnan, S. and Maranas, C.D., 13C metabolic flux analysis at a genome-scale, Metab. Eng., 2015, vol. 32, pp. 12–22. https://doi.org/10.1016/j.ymben.2015.08.006

    Article  CAS  PubMed  Google Scholar 

  13. Neidhardt, F.C. and Umbarger, H.E., Chemical composition of Escherichia coli, in Escherichia coli and Salmonella: Cellular and Molecular Biology, Neidhardt, F.C., Ed., Washington, USA: ASM Press, 1996, 2nd ed., pp. 13–16.

    Google Scholar 

  14. Ferreira, G.C. and Gong, J., 5-Aminolevulinate synthase and the first step of heme biosynthesis, J. Bioenerg. Biomembr., 1995, vol. 27, no. 2, pp. 151–159. https://doi.org/10.1007/BF02110030

    Article  CAS  PubMed  Google Scholar 

  15. Reed, J.L., Vo, T.D., Schilling, C.H., and Palsson, B.O., An expanded genome-scale model of Escherichia coli K-12 (iJR904 GSM/GPR), Genome Biol., 2003, vol. 4, no. 9, p. R54. https://doi.org/10.1186/gb-2003-4-9-r54

    Article  PubMed  PubMed Central  Google Scholar 

  16. Seligmann, H., Cost-minimization of amino acid usage, J. Mol. Evol., 2003, vol. 56, pp. 151–161. https://doi.org/10.1007/s00239-002-2388-z

    Article  CAS  PubMed  Google Scholar 

  17. Heizer, E.M., Jr., Raiford, D.W., Raymer, M.L., et al., Amino acid cost and codon-usage biases in 6 prokaryotic genomes: a whole-genome analysis, Mol. Biol. Evol., 2006, vol. 23, pp. 1670–1680. https://doi.org/10.1093/molbev/msl029

    Article  CAS  PubMed  Google Scholar 

  18. Heizer, E.M., Jr., Raymer, M.L., and Krane, D.E., Amino acid biosynthetic cost and protein conservation, J. Mol. Evol., 2011, vol. 72, pp. 466–473. https://doi.org/10.1007/s00239-011-9445-4

    Article  CAS  PubMed  Google Scholar 

  19. Kaleta, C., Schauble, S., Rinas, U., and Schuster, S., Metabolic costs of amino acid and protein production in Escherichia coli, Biotechnol. J., 2013, vol. 8, pp. 1105–1114. https://doi.org/10.1002/biot.201200267

    Article  CAS  PubMed  Google Scholar 

  20. Zhang, H., Wang, Y., Li, J., et al., Biosynthetic energy cost for amino acids decreases in cancer evolution, Nat. Commun., 2018, vol. 9, p. 4124. https://doi.org/10.1038/s41467-018-06461-1

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Strier, L., Biokhimiya (v 3 tomakh) (Biochemistry (in 3 vols.)), Severin, S.E, Ed., Moscow: Mir, 1985, vol. 2.

  22. Wagner, A., Energy constraints on the evolution of gene expression, Mol. Biol. Evol., 2005, vol. 22, pp. 1365–1374. https://doi.org/10.1093/molbev/msi126

    Article  CAS  PubMed  Google Scholar 

  23. Taymaz-Nikerel, H., Borujeni, A.E., Verheijen, P.J.T., et al., Genome-derived minimal metabolic models for Escherichia coli MG1655 with estimated in vivo respiratory ATP stoichiometry, Biotechnol. Bioeng., 2010, vol. 107, pp. 369–381. https://doi.org/10.1002/bit.22802

    Article  CAS  PubMed  Google Scholar 

  24. Bekker, M., de Vries, S., Beek, A.T., et al., Respiration of Escherichia coli can be fully uncoupled via the nonelectrogenic terminal cytochrome bd-II oxidase, J. Bacteriol., 2009, vol. 191, pp. 5510–5517. . https://doi.org/10.1128/JB.00562-09

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Magalon, A. and Alberge, F., Distribution and dynamics of OXPHOS complexes in the bacterial cytoplasmic membrane, Biochim. Biophys. Acta, 2016, vol. 1857, pp. 198–213. https://doi.org/10.1016/j.bbabio.2015.10.015

    Article  CAS  PubMed  Google Scholar 

  26. Feist, A.M., Henry, C.S., Reed, J.L., et al., A genome-scale metabolic reconstruction for Escherichia coli K-12 MG1655 that accounts for 1260 ORFs and thermodynamic information, Mol. Syst. Biol., 2007, vol. 3, p. 121. https://doi.org/10.1038/msb4100155

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Tomashek, J.J. and Brusilow, W.S.A., Stoichiometry of energy coupling by proton-translocating ATPases: a history of variability, J. Bioenerg. Biomembr., 2000, vol. 32, pp. 493–500. https://doi.org/10.1023/A:1005617024904

    Article  CAS  PubMed  Google Scholar 

  28. Neidhardt, F.C., Ingraham, J.L., and Schaechter, M., Physiology of the Bacterial Cell: A Molecular Approach, Neidhardt, F.C., Ed., Sunderland, USA: Sinauer Associates, 1990.

    Google Scholar 

  29. Krick, T., Verstraete, N., Alonso, L.G., et al., Amino acid metabolism conflicts with protein diversity, Mol. Biol. Evol., 2014, vol. 31, pp. 2905–2912. https://doi.org/10.1093/molbev/msu228

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Stahlberg, H., Muller, D.J., Suda, K., et al., Bacterial Na+-ATP synthase has an undecameric rotor, EMBO Rep., 2001, vol. 2, pp. 229–233. https://doi.org/10.1093/embo-reports/kve047

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Steigmiller, S., Turina, P., and Graber, P., The thermodynamic H+/ATP ratios of the H+-ATP synthases from chloroplasts and Escherichia coli, Proc. Natl. Acad. Sci. U. S. A., 2008, vol. 105, pp. 3745–3750. https://doi.org/10.1073/pnas.0708356105

    Article  PubMed  PubMed Central  Google Scholar 

  32. Antoniewicz, M.R., Kraynie, D.F., Laffend, L.A., et al., Metabolic flux analysis in a nonstationary system: fedbatch fermentation of a high yielding strain of E. coli producing 1,3-propanediol, Metab. Eng., 2007, vol. 9, pp. 277–292. https://doi.org/10.1016/j.ymben.2007.01.003

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Leighty, R.W. and Antoniewicz, M.R., Parallel labeling experiments with [U-13C]glucose validate E. coli metabolic network model for 13C metabolic flux analysis, Metab. Eng., 2012, vol. 14, pp. 533–541. https://doi.org/10.1016/j.ymben.2012.06.003

    Article  CAS  PubMed  Google Scholar 

  34. Leighty, R.W. and Antoniewicz, M.R., COMPLETE-MFA: complementary parallel labeling experiments technique for metabolic flux analysis, Metab. Eng., 2013, vol. 20, pp. 49–55. https://doi.org/10.1016/j.ymben.2013.08.006

    Article  CAS  PubMed  Google Scholar 

  35. Long, C.P., Gonzalez, J.E., Feist, A.M., et al., Fast growth phenotype of E. coli K K-12 from adaptive laboratory evolution does not require intracellular flux rewiring, Metab. Eng., 2017, vol. 44, pp. 100–107. https://doi.org/10.1016/j.ymben.2017.09.012

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gonzalez, J.E., Long, C.P., and Antoniewicz, M.R., Comprehensive analysis of glucose and xylose metabolism in Escherichia coli under aerobic and anaerobic conditions by 13C metabolic flux analysis, Metab. Eng., 2017, vol. 39, pp. 9–18. https://doi.org/10.1016/j.ymben.2016.11.003

    Article  CAS  PubMed  Google Scholar 

  37. Orth, J.D., Conrad, T.M., Na, J., et al., A comprehensive genome-scale reconstruction of Escherichia coli metabolism–2011, Mol. System. Biol., 2011, vol. 7, p. 535. https://doi.org/10.1038/msb.2011.65

    Article  Google Scholar 

  38. Ravikirthi, P., Suthers, P.F., and Maranas, C.D., Construction of an E. coli genome-scale atom mapping model for MFA calculations, Biotechnol. Bioeng., 2011, vol. 108, pp. 1372–1382. https://doi.org/10.1002/bit.23070

    Article  CAS  PubMed  Google Scholar 

  39. Latendresse, M., Malerich, J.P., Travers, M., and Karp, P.D., Accurate atom-mapping computation for biochemical reactions, J. Chem. Inf. Model., 2012, vol. 52, pp. 2970–2982. https://doi.org/10.1021/ci3002217

    Article  CAS  PubMed  Google Scholar 

  40. Szyperski, T., Biosynthetically directed fractional 13C-labeling of proteinogenic amino acids: an efficient analytical tool to investigate intermediary metabolism, Eur. J. Biochem., 1995, vol. 232, pp. 433–448. https://doi.org/10.1111/j.1432-1033.1995.tb20829.x

    Article  CAS  PubMed  Google Scholar 

  41. Antoniewicz, M.R., Kelleher, J.K., and Stephanopoulos, G., Elementary metabolite units (EMU): a novel framework for modeling isotopic distributions, Metab. Eng., 2007, vol. 9, pp. 68–86. https://doi.org/10.1016/j.ymben.2006.09.001

    Article  CAS  PubMed  Google Scholar 

  42. Wiechert, W., Mollney, M., Petersen, S., and de Graaf, A., A universal framework for 13C metabolic flux analysis, Metab. Eng., 2001, vol. 3, pp. 265–283. https://doi.org/10.1006/mben.2001.0188

    Article  CAS  PubMed  Google Scholar 

  43. Shupletsov, M.S., Golubeva, L.I., Rubina, S.S., et al., OpenFLUX2: 13C-MFA modeling software package adjusted for the comprehensive analysis of single and parallel labeling experiments, Microb. Cell Fact., 2014, vol. 13, p. 152. https://doi.org/10.1186/s12934-014-0152-x

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Wittman, C. and Van De Graaf, A.A., Handbook of Corynebacterium glutamicum, Eggeling, L. and Bott, M., Boca Raton: CRC Press, 2005, 1st ed., ch. 12, pp. 277–304. https://doi.org/10.1201/9781420039696.ch12

  45. Becker, J. and Wittmann, C., Metabolic Flux Analysis: Methods and Protocols (Methods in Molecular Biology), Kromer, J.O., Nielsen, L., and Blank, L.M., Eds., Humana: Springer Nature, 2014, 1st ed., ch. 10, pp. 165–174. https://doi.org/10.1007/978-1-4939-1170-7_10

  46. Kawai, S., Mori, S., Mukai, T., et al., Molecular characterization of Escherichia coli NAD+ kinase, Eur. J. Biochem., 2001, vol. 268, pp. 4359–4365. https://doi.org/10.1046/j.1432-1327.2001.02358.x

    Article  CAS  PubMed  Google Scholar 

  47. Sakuraba, H., Kawakami, R., and Ohshima, T., First archaeal inorganic polyphosphate/ATP-dependent NAD+ kinase, from hyperthermophilic archaeon Pyrococcus horikoshii: cloning, expression, and characterization, Appl. Environ. Microbiol., 2005, vol. 71, pp. 4352–4358. https://doi.org/10.1128/AEM.71.8.4352-4358.2005

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Outten, C.E. and Culotta, V.C., A novel NADH kinase is the mitochondrial source of NADPH in Saccharomyces cerevisiae, EMBO J., 2003, vol. 22, pp. 2015–2024. https://doi.org/10.1093/emboj/cdg211

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Strand, M.K., Stuart, G.R., Longley, M.J., et al., POS5 gene of Saccharomyces cerevisiae encodes a mitochondrial NADH kinase required for stability of mitochondrial DNA, Eukaryot. Cell, 2003, vol. 2, pp. 809–820. https://doi.org/10.1128/EC.2.4.809-820.2003

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sauer, U., Canonaco, F., Heri, S., et al., The soluble and membrane-bound transhydrogenases UdhA and PntAB have divergent functions in NADPH metabolism of Escherichia coli, J. Biol. Chem., 2004, vol. 279, pp. 6613–6619. https://doi.org/10.1074/jbc.M311657200

    Article  CAS  PubMed  Google Scholar 

  51. Wiechert, W. and de Graaf, A.A., Modelling for improved bioreactor performance, in Modelling for Improved Bioreactor Performance, Bales, V., Bratislava: Male Centrum, 1993, pp. 19–24.

  52. Wiechert, W. and de Graaf, A.A., In vivo stationary flux analysis by 13C labeling experiments, Adv. Biochem. Eng. Biotechnol., 1996, vol. 54, pp. 109–154. https://doi.org/10.1007/BFb0102334

    Article  CAS  PubMed  Google Scholar 

  53. Quek, L.E., Wittmann, C., Nielsen, L.K., and Kromer, J.O., OpenFLUX: efficient modelling software for 13C-based metabolic flux analysis, Microb. Cell Fact., 2009, vol. 8, p. 25. https://doi.org/10.1186/1475-2859-8-25

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Quek, L.E. and Nielsen, L.K., Steady-state 13C fluxomics using OpenFLUX, Methods Mol. Biol., 2014, vol. 1191, pp. 209–224. https://doi.org/10.1007/978-1-4939-1170-7_13

    Article  CAS  PubMed  Google Scholar 

  55. van Winden, W., Verheijen, P., and Heijnen, S., Possible pitfalls of flux calculations based on 13C-labeling, Metab. Eng., 2001, vol. 3, pp. 151–162. https://doi.org/10.1006/mben.2000.0174

    Article  CAS  PubMed  Google Scholar 

  56. Kleijn, R.J., van Winden, W.A., van Gulik, W.M., and Heijen, J.J., Revisiting the 13C-label distribution of the non-oxidative branch of the pentose phosphate pathway based upon kinetic and genetic evidence, FEBS J., 2005, vol. 272, pp. 4970–4982. https://doi.org/10.1111/j.1742-4658.2005.04907.x

    Article  CAS  PubMed  Google Scholar 

  57. Melendez-Hevia, E. and Isidoro, A., The game of the pentose phosphate cycle, J. Theor. Biol., 1985, vol. 117, pp. 251–263. https://doi.org/10.1016/s0022-5193(85)80220-4

    Article  CAS  PubMed  Google Scholar 

  58. Shaikh, A.S., Tang, Y.J., Mukhopadhyay, A., et al., Study of stationary phase metabolism via isotopomer analysis of amino acids from an isolated protein, Biotechnol. Prog., 2010, vol. 6, pp. 52–56. https://doi.org/10.1002/btpr.325

    Article  CAS  Google Scholar 

  59. Wasylenko, T.M. and Stephanopoulos, G., Metabolomic and 13C-metabolic flux analysis of a xylose-consuming Saccharomyces cerevisiae strain expressing xylose isomerase, Biotechnol. Bioeng., 2015, vol. 112, pp. 470–483. https://doi.org/10.1002/bit.25447

    Article  CAS  PubMed  Google Scholar 

  60. Kochetov, G.A. and Solovjeva, O.N., Structure and functioning mechanism of transketolase, Biochim. Biophys. Acta, 2014, vol. 1844, pp. 1608–1618. https://doi.org/10.1016/j.bbapap.2014.06.003

    Article  CAS  PubMed  Google Scholar 

  61. van Winden, W.A., Gulik, W.M., SchiPer, D., et al. Metabolic flux and metabolic network analysis of Penicillium chrysogenum using 2D [13C, 1H] COSY NMR measurements and cumulative Bondomer simulation, Biotechnol. Bioeng., 2003, vol. 83, pp. 75–92. https://doi.org/10.1002/bit.10648

    Article  CAS  PubMed  Google Scholar 

  62. van Winden, W.A., van Dam, J.C., Ras, C., et al., Metabolic flux analysis of Saccharomyces cerevisiae CEN.PK113-7D based on mass isotopomer measurements of 13C-labeled primary metabolites, FEMS Yeast Res., 2005, vol. 5, pp. 559–568. https://doi.org/10.1016/j.femsyr.2004.10.007

    Article  CAS  PubMed  Google Scholar 

  63. Millard, P., Sokol, S., Letisse, F., and Portais, J.C., IsoDesign: a software for optimizing the design of 13C-metabolic flux analysis experiments, Biotechnol. Bioeng., 2015, vol. 111, pp. 202–208. https://doi.org/10.1002/bit.24997

    Article  CAS  Google Scholar 

  64. Martin, H.G., Kumar, V.S., Weaver, D., et al., A method to constrain genome-scale models with 13C labeling data, PLoS Comput. Biol., 2015, vol. 11, p. e1004363. https://doi.org/10.1371/journal.pcbi.1004363

    Article  CAS  PubMed  Google Scholar 

  65. Cordova, L.T., Lu, J., Cipolla, R.M., et al., Co-utilization of glucose and xylose by evolved Thermus thermophilus LC113 strain elucidated by 13C metabolic flux analysis and whole genome sequencing, Metab. Eng., 2016, vol. 37, pp. 63–71. https://doi.org/10.1016/j.ymben.2016.05.001

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Cordova, L.T., Long, C.P., Venkataramanan, K.P., and Antoniewicz, M.R., Complete genome sequence, metabolic model construction and phenotypic characterization of Geobacillus LC300, an extremely thermophilic, fast growing, xylose-utilizing bacterium, Metab. Eng., 2015, vol. 32, pp. 74–81. https://doi.org/10.1016/j.ymben.2015.09.009

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Au, J., Choi, J., Jones, S.W., et al., Parallel labeling experiments validate Clostridium acetobutylicum metabolic network model for 13C metabolic flux analysis, Metab. Eng., 2014, vol. 26, pp. 23–33. https://doi.org/10.1016/j.ymben.2014.08.002

    Article  CAS  PubMed  Google Scholar 

  68. Zelle, R.M., de Hulster, E., van Winden, W.A., et al., Malic acid production by Saccharomyces cerevisiae: engineering of pyruvate carboxylation, oxaloacetate reduction, and malate export, Appl. Environ. Microbiol., 2008, vol. 74, pp. 2766–2777. https://doi.org/10.1128/AEM.02591-07

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Shymansky, C.M., Wang, G., Baidoo, E.E.K., et al., Flux-enabled exploration of the role of Sip1 in galactose yeast metabolism, Front. Bioeng. Biotechnol., 2017, vol. 5, p. 31. https://doi.org/10.3389/fbioe.2017.00031

    Article  PubMed  PubMed Central  Google Scholar 

  70. Wasylenko, T.M., Ahn, W.S., and Stephanopoulos, G., The oxidative pentose phosphate pathway is the primary source of NADPH for lipid overproduction from glucose in Yarrowia lipolytica, Metab. Eng., 2015, vol. 30, pp. 27–39. https://doi.org/10.1016/j.ymben.2015.02.007

    Article  CAS  PubMed  Google Scholar 

  71. Kleijn, R.J., van Winden, W.A., Ras, C., et al., 13C-labeled gluconate tracing as a direct and accurate method for determining the pentose phosphate pathway split ratio in Penicillium chrysogenum, Appl. Environ. Microbiol., 2006, vol. 72, pp. 4743–4754. https://doi.org/10.1128/AEM.02955-05

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ahn, W.S. and Antoniewicz, M.R., Metabolic flux analysis of CHO cells at growth and non-growth phases using isotopic tracers and mass spectrometry, Metab. Eng., 2011, vol. 13, pp. 598–609. https://doi.org/10.1016/j.ymben.2011.07.002

    Article  CAS  PubMed  Google Scholar 

  73. Ahn, W.S., Crown, S.B., and Antoniewicz, M.R., Evidence for transketolase-like TKTL1 flux in CHO cells based on parallel labeling experiments and 13C-metabolic flux analysis, Metab. Eng., 2016, vol. 37, pp. 72–78. https://doi.org/10.1016/j.ymben.2016.05.005

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Long, C.P., Au, J., Gonzalez, J.E., and Antoniewicz, M.R., 13C metabolic flux analysis of microbial and mammalian systems is enhanced with GC-MS measurements of glycogen and RNA labeling, Metab. Eng., 2016, vol. 38, pp. 65–72. https://doi.org/10.1016/j.ymben.2016.06.007

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Jankowski, M.D., Henry, C.S., Broadbelt, L.J., and Hatzimanikatis, V., Group contribution method for thermodynamic analysis of complex metabolic networks, Biophys. J., 2008, vol. 95, pp. 1487–1499. https://doi.org/10.1529/biophysj.107.124784

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Silverstein, T., The mitochondrial phosphate-to-oxygen ratio is not an integer, Biochem. Mol. Biol. Edu., 2005, vol. 33, pp. 416–417. https://doi.org/10.1002/bmb.2005.49403306416

    Article  CAS  Google Scholar 

  77. Mahmoudabadi, G., Phillips, R., Lynch, M., and Milo, R., Defining the energetic costs of cellular structures, bioRxiv, 2019, vol. 66040. https://doi.org/10.1101/666040

  78. Du, B., Zhang, Z., Grubner, S., et al., Temperature-dependent estimation of Gibbs energies using an updated group-contribution method, Biophys. J., 2018, vol. 114, pp. 2691–2702. https://doi.org/10.1016/j.bpj.2018.04.030

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Mavrovouniotis, M.L., Estimation of standard Gibbs energy changes of biotransformations, J. Biol. Chem., 1991, vol. 266, pp. 14440–14445. https://doi.org/10.1016/S0021-9258(18)98705-3

    Article  CAS  PubMed  Google Scholar 

  80. Mavrovouniotis, M.L., Group contributions for estimating standard Gibbs energies of formation of biochemical compounds in aqueous-solution, Biotechnol. Bioeng., 1990, vol. 36, pp. 1070–1082. https://doi.org/10.1002/bit.260361013

    Article  CAS  PubMed  Google Scholar 

  81. Alberty, R.A., Calculation of standard transformed formation properties of biochemical reactants and standard apparent reduction potentials of half reactions, Arch. Biochem. Biophys., 1998, vol. 358, pp. 25–39. https://doi.org/10.1002/bit.260361013

    Article  CAS  PubMed  Google Scholar 

  82. Bennett, B.D., Kimball, E.H., Gao, M., et al., Absolute metabolite concentrations and implied enzyme active site occupancy in Escherichia coli, Nat. Chem. Biol., 2009, vol. 5, pp. 593–599. https://doi.org/10.1038/nchembio.186

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

This work was supported by ongoing institutional funding. No additional grants to carry out or direct this particular research were obtained.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. S. Kovaleva.

Ethics declarations

ETHICS APPROVAL AND CONSENT TO PARTICIPATE

This article does not contain any studies involving animals performed by any of the authors.

This article does not contain any studies involving human participants performed by any of the authors outside the scope of people’s normal professional activities.

CONFLICT OF INTEREST

The authors declare that they have no conflicts of interest.

Additional information

Translated by I. Gordon

Publisher’s Note.

Pleiades Publishing remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Abbreviations: aa, amino acid(s); BC, biosynthetic complexity; CM, central metabolism; 13С-MFA, 13С metabolic flux analysis; 13DPG, 1,3-diphosphoglycerate; EC, electron chain; EMP pathway, Embden–Meyerhof–Parnas pathway; ET chain, electron transport chain; [H], hydride ion; H, energy stored in reduced forms of NADH, NADPH and FADH2; LPS, lipopolysaccharides; MC, metabolic cost; OXPHOS complexes, oxidative phosphorylation complexes; MP, metabolic precursor; OXPHOS complex, oxidative phosphorylation complex; ∼Р, phosphoryl group; ∼Р, energy stored in ATP molecule; PEP, phosphoenolpyruvate; P/O ratio, number of ATP molecules generated during the transfer of 1 electron pair to one oxygen atom in oxidative phosphorylation; PP pathway, pentose phosphate pathway; SP/OP/ETP, substrate/oxidative/electron transport phosphorylation; TCA, tricarboxylic acid(s).

SUPPLEMENTARY MATERIALS

SUPPLEMENTARY MATERIALS

Calculation of the Gibbs energy for metabolic reactions

For a random chemical reaction:

$${{v}_{1}}{{S}_{1}} + {{v}_{2}}{{S}_{2}} \to {{\mu }_{1}}{{P}_{1}} + {{\mu }_{2}}{{P}_{2}},$$
(17)

where Si and Pj are the substrates and products of the reaction, respectively, vi and μj are their stoichiometric coefficients. The Gibbs energy of this reaction (ΔrG') can be calculated using the following formula:

$${{\Delta }_{{\text{r}}}}G{\kern 1pt} ' = {{\Delta }_{{\text{r}}}}G{\kern 1pt} {{'}^{{\text{o}}}} + RT{\text{ln}}{{\left( {{{{\left[ {{{P}_{1}}} \right]}}^{\mu }}^{1}{{{\left[ {{{P}_{2}}} \right]}}^{\mu }}^{2}} \right)} \mathord{\left/ {\vphantom {{\left( {{{{\left[ {{{P}_{1}}} \right]}}^{\mu }}^{1}{{{\left[ {{{P}_{2}}} \right]}}^{\mu }}^{2}} \right)} {\left( {{{{\left[ {{{S}_{1}}} \right]}}^{v}}^{1}{{{\left[ {{{S}_{2}}} \right]}}^{v}}^{2}} \right)}}} \right. \kern-0em} {\left( {{{{\left[ {{{S}_{1}}} \right]}}^{v}}^{1}{{{\left[ {{{S}_{2}}} \right]}}^{v}}^{2}} \right)}},$$
(18)

where [Si] and [Pj] are the molar concentration of substrates and products of the reaction, respectively; ΔrG'o are the standard Gibbs energy of the reaction under 1 atm, 25°С and 1 М concentration of each reagent; T is the absolute temperature, K; R is the universal gas constant (≈ 8.3144598(48) J mol–1 K–1).

The value of ΔrG'o for metabolic reactions can be extracted from specialized literature (for example, [2678]). The values of ΔrG'm for reactions under standard pressure and temperature are also given, but at 1 mM concentration of each reagents, as well as the values of standard Gibbs free energy of metabolite formation at 1 M (ΔfG'o) or 1 mM (ΔfG'm) concentrations. The available parameters of ΔfG'o are usually the result of theoretical evaluation using the group contribution method developed by M. Mavrovouniotis [7980] and supplemented by M. Jankowski et al. [75] and B. Du et al. [78] on the basis of calculated and experimental data obtained by R. Alberty [81].

If the values of ΔrG'o are known, we can determine the ΔrG ' value, for example, for the reaction at physiological reagent concentrations, i.e., specified for metabolites of E. coli cells exponentially growing under aerobic conditions on glucose. In these conditions, according to [82]: [ATP] = 9.6 × 10−3, [ADP] = 5.6 × 10−4, [Pi] = 2 × 10−2, [NAD+] = 2.6 × 10−3, [NADPH] = 1.2 × 10−4, [NADPH] = 8.3 × 10−5, [NADP+] = 2.1 × 10−6. These are the concentrations we used to calculate ΔrG'.

The energy released during the hydrolysis of ATP to ADP and Pi (ATP maintenance requirement) is called the phosphorylation potential, and its value is estimated for the reaction performed under standard conditions [82]:

$$\begin{gathered} {\text{ATP}} + {{{\text{H}}}_{{\text{2}}}}{\text{O}} \to {\text{ADP}} + {{{\text{P}}}_{{\text{i}}}} \\ + \,\,{\text{H}}({{\Delta }_{{\text{r}}}}G^{'{\text{o}}} = - 6.7{\text{ kcal/mol}}), \\ \end{gathered} $$
(19)

and at physiological metabolite concentrations the equation is represented as:

$$\begin{gathered} {{\Delta }_{{\text{r}}}}G{\kern 1pt} ' = {\text{ }}{{\Delta }_{{\text{r}}}}G^{'{\text{o}}} + RT{\text{ln}}{{\left( {[{\text{ADP}}][{{{\text{P}}}_{{\text{i}}}}]} \right)} \mathord{\left/ {\vphantom {{\left( {[{\text{ADP}}][{{{\text{P}}}_{{\text{i}}}}]} \right)} {\left[ {{\text{ATP}}} \right]}}} \right. \kern-0em} {\left[ {{\text{ATP}}} \right]}} \\ = \,\, - 10.9{\text{ kcal/mol}} = {\text{[}}{{{\text{P}}}_{{\text{i}}}}{\text{]}}. \\ \end{gathered} $$
(20)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Golubeva, L.I., Kovaleva, E.S. & Mashko, S.V. On the Question of the Metabolic Costs of the Main Metabolic Precursors in Escherichia coli. Appl Biochem Microbiol 59, 1201–1213 (2023). https://doi.org/10.1134/S0003683823090041

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0003683823090041

Keywords:

Navigation