Skip to main content
Log in

A mechanistic model of metabolic symbioses in microbes recapitulates experimental data and identifies a continuum of symbiotic interactions

  • Original Article
  • Published:
Theory in Biosciences Aims and scope Submit manuscript

Abstract

Microbial symbioses based on nutrient exchange and interdependence are ubiquitous in nature and biotechnologically promising; however, an in-depth mathematical description of their exact underlying dynamics from first principles is still missing. Hence, in this paper a novel mechanistic mathematical model of such a relationship in a continuous chemostat culture is derived. In contrast to preceding works on the topic, only parameters which can be directly measured and understood from biological first principles are used, allowing for a higher degree of mechanistic understanding of the underlying processes compared to previous approaches. The predictive power of the model is validated by demonstrating that it accurately recapitulates both the temporal dynamics as well as the final state of a previously published cross-feeding experiment. The model is then used to examine the influence of the biological traits of the involved organisms on the position and stability of the equilibrium states of the system using bifurcation analyses. It is additionally demonstrated how manipulating the external metabolite concentrations of the system can shift the species interaction on a continuous spectrum ranging from mutualism over commensalism to parasitism. This further reinforces the idea of a continuous spectrum of symbiotic interactions as opposed to static and discrete categories. Finally, the practical implications of the results for the biotechnological application of such microbial consortia are discussed.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Ames GF (1964) Uptake of amino acids by salmonella typhimurium. Arch Biochem Biophys 104(1):1–18

    CAS  PubMed  Google Scholar 

  • Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman J, Smith JA, Struhl K, Ferreira A, Neidhardt F, Ingraham J et al (1999) Short protocols in molecular biology: a compendium of methods from current protocols in molecular biology. Number 574.88 S559. Universidade de Sao Paulo, Piracicaba, SP (Brasil). Faculdade de Medicina

  • Bull JJ, Harcombe WR (2009) Population dynamics constrain the cooperative evolution of cross-feeding. PLoS ONE 4(1):e4115

    PubMed  PubMed Central  Google Scholar 

  • Bull JJ, Molineux IJ, Rice W (1991) Selection of benevolence in a host-parasite system. Evolution 45(4):875–882

    CAS  PubMed  Google Scholar 

  • Canfield DE, Kristensen E, Thamdrup B (2005) Aquatic geomicrobiology. Gulf Professional Publishing, Oxford

    Google Scholar 

  • de Bashan LE, Mayali X, Bebout BM, Weber PK, Detweiler AM, Hernandez J-P, Prufert-Bebout L, Bashan Y (2016) Establishment of stable synthetic mutualism without co-evolution between microalgae and bacteria demonstrated by mutual transfer of metabolites (nanosims isotopic imaging) and persistent physical association (fluorescent in situ hybridization). Algal Res 15:179–186

    Google Scholar 

  • Douglas A (1994) Symbiotic interactions: Oxford Science Publications. Oxford University Press, Oxford

    Google Scholar 

  • Drevon D, Fursa SR, Malcolm AL (2017) Intercoder reliability and validity of webplotdigitizer in extracting graphed data. Behav Modif 41(2):323–339

    PubMed  Google Scholar 

  • Filipe CD, Grady CL Jr (1998) Biological wastewater treatment, revised and expanded. CRC Press, Boca Raton

    Google Scholar 

  • Foster KR, Wenseleers T (2006) A general model for the evolution of mutualisms. J Evol Biol 19(4):1283–1293

    CAS  PubMed  Google Scholar 

  • Gardner TS, Cantor CR, Collins JJ (2000) Construction of a genetic toggle switch in escherichia coli. Nature 403(6767):339–342

    CAS  PubMed  Google Scholar 

  • Gerlee P, Lundh T (2016) Scientific models: red atoms, white lies and black boxes in a yellow book. Springer, New York

    Google Scholar 

  • Graves WG, Peckham B, Pastor J (2006) A bifurcation analysis of a differential equations model for mutualism. Bull Math Biol 68(8):1851–1872

    PubMed  Google Scholar 

  • Großkopf T, Soyer OS (2014) Synthetic microbial communities. Curr Opin Microbiol 18:72–77

    PubMed  PubMed Central  Google Scholar 

  • Helling RB, Vargas CN, Adams J (1987) Evolution of Escherichia coli during growth in a constant environment. Genetics 116(3):349–358

    CAS  PubMed  PubMed Central  Google Scholar 

  • Hoek TA, Axelrod K, Biancalani T, Yurtsev EA, Liu J, Gore J (2016) Resource availability modulates the cooperative and competitive nature of a microbial cross-feeding mutualism. PLoS Biol 14(8):e1002540

    PubMed  PubMed Central  Google Scholar 

  • Jeon KW (1972) Development of cellular dependence on infective organisms: micrurgical studies in amoebas. Science 176(4039):1122–1123

    CAS  PubMed  Google Scholar 

  • Jeon K, Jeon M (1976) Endosymbiosis in amoebae: recently established endosymbionts have become required cytoplasmic components. J Cell Physiol 89(2):337–344

    CAS  PubMed  Google Scholar 

  • Kerner A, Park J, Williams A, Lin XN (2012) A programmable Escherichia coli consortium via tunable symbiosis. PLoS ONE 7(3):e34032

    CAS  PubMed  PubMed Central  Google Scholar 

  • Kouzuma A, Kato S, Watanabe K (2015) Microbial interspecies interactions: recent findings in syntrophic consortia. Front Microbiol 6:477

    PubMed  PubMed Central  Google Scholar 

  • Mee MT, Collins JJ, Church GM, Wang HH (2014) Syntrophic exchange in synthetic microbial communities. Proc Nat Acad Sci 111(20):E2149–E2156

    CAS  PubMed  Google Scholar 

  • Megee R, Drake J, Fredrickson A, Tsuchiya H (1972) Studies in intermicrobial symbiosis. saccharomyces cerevisiae and lactobacillus casei. Can J Microbiol 18(11):1733–1742

    PubMed  Google Scholar 

  • Melnyk AH, Wong A, Kassen R (2015) The fitness costs of antibiotic resistance mutations. Evol Appl 8(3):273–283

    PubMed  Google Scholar 

  • Miller J (1972) Experiments in molecular genetics. Bacterial genetics—E. coli. Cold Spring Harbor Laboratory, Cold Spring Harbor

    Google Scholar 

  • Monod J (1949) The growth of bacterial cultures. Ann Rev Microbiol 3(1):371–394

    CAS  Google Scholar 

  • Neidhardt F, Ingraham J, Schaechter M (1990) Physiology of the bacterial cell. Sinauer Associates, Sunderland

    Google Scholar 

  • Overmann J, Schubert K (2002) Phototrophic consortia: model systems for symbiotic interrelations between prokaryotes. Arch Microbiol 177(3):201–208

    CAS  PubMed  Google Scholar 

  • Pande S, Kost C (2017) Bacterial unculturability and the formation of intercellular metabolic networks. Trends Microbiol 25(5):349–361

    CAS  PubMed  Google Scholar 

  • Prescott LM, Harley JP, Klein DA, Willey JM (2010) Microbiologie. De Boeck Supérieur, Louvain-la-Neuve

    Google Scholar 

  • Senn H, Lendenmann U, Snozzi M, Hamer G, Egli T (1994) The growth of escherichia coli in glucose-limited chemostat cultures: a re-examination of the kinetics. Biochim Biophys Acta (BBA)-Gen Subj 1201(3):424–436

    Google Scholar 

  • Sezonov G, Joseleau-Petit D, d’Ari R (2007) Escherichia coli physiology in luria-bertani broth. J Bacteriol 189(23):8746–8749

    CAS  PubMed  PubMed Central  Google Scholar 

  • Shapiro JW, Turner PE (2018) Evolution of mutualism from parasitism in experimental virus populations. Evolution 72(3):707–712

    PubMed  Google Scholar 

  • Shapiro JW, Williams ES, Turner PE (2016) Evolution of parasitism and mutualism between filamentous phage m13 and escherichia coli. PeerJ 4:e2060

    PubMed  PubMed Central  Google Scholar 

  • Smith HL, Waltman P (1995) The theory of the chemostat: dynamics of microbial competition, vol 13. Cambridge University Press, Cambridge

    Google Scholar 

  • Stewart EJ (2012) Growing unculturable bacteria. J Bacteriol 194(16):4151–4160

    CAS  PubMed  PubMed Central  Google Scholar 

  • Stump SM, Klausmeier CA (2016) Competition and coexistence between a syntrophic consortium and a metabolic generalist, and its effect on productivity. J Theor Biol 404:348–360

    PubMed  Google Scholar 

  • Sun Z, Koffel T, Stump SM, Grimaud GM, Klausmeier CA (2019) Microbial cross-feeding promotes multiple stable states and species coexistence, but also susceptibility to cheaters. J Theor Biol 465:63–77

    PubMed  Google Scholar 

  • Turner PE, Souza V, Lenski RE (1996) Tests of ecological mechanisms promoting the stable coexistence of two bacterial genotypes. Ecology 77(7):2119–2129

    Google Scholar 

  • Van Rossum G, Drake FL Jr (1995) Python tutorial. Centrum voor Wiskunde en Informatica, Amsterdam

    Google Scholar 

  • Vet S, de Buyl S, Faust K, Danckaert J, Gonze D, Gelens L (2018) Bistability in a system of two species interacting through mutualism as well as competition: chemostat vs. Lotka–Volterra equations. PloS ONE 13(6):e0197462

    PubMed  PubMed Central  Google Scholar 

  • Virtanen P, Gommers R, Oliphant TE, Haberland M, Reddy T, Cournapeau D, Burovski E, Peterson P, Weckesser W, Bright J, van der Walt SJ, Brett M, Wilson J, Jarrod Millman K, Mayorov N, Nelson ARJ, Jones E, Kern R, Larson E, Carey C, Polat İ, Feng Y, Moore EW, Vand erPlas J, Laxalde D, Perktold J, Cimrman R, Henriksen I, Quintero EA, Harris CR, Archibald AM, Ribeiro AH, Pedregosa F, van Mulbregt P, Contributors S (2020) SciPy 1.0: fundamental algorithms for scientific computing in python. Nat Methods 17:261–272

    CAS  PubMed  PubMed Central  Google Scholar 

  • Wu G, Yan Q, Jones JA, Tang YJ, Fong SS, Koffas MA (2016) Metabolic burden: cornerstones in synthetic biology and metabolic engineering applications. Trends Biotechnol 34(8):652–664

    CAS  PubMed  Google Scholar 

  • Yamamura N (1996) Evolution of mutualistic symbiosis: a differential equation model. Res Popul Ecol 38(2):211–218

    Google Scholar 

  • Yukalov VI, Yukalova E, Sornette D (2012) Modeling symbiosis by interactions through species carrying capacities. Phys D 241(15):1270–1289

    Google Scholar 

  • Zhang X, Reed JL (2014) Adaptive evolution of synthetic cooperating communities improves growth performance. PloS ONE 9(10):e108297

    PubMed  PubMed Central  Google Scholar 

  • Ziv N, Brandt NJ, Gresham D (2013) The use of chemostats in microbial systems biology. J Vis Exp JoVE 80:e50168

    Google Scholar 

Download references

Acknowledgements

I would like to thank Matthias Bild and Prof. Dr. R. Mutzel for helpful feedback and fruitful discussions during the derivation of the model. I am also indebted to Dr. C. v. Törne for his valuable support in the analytical examination of the model. Finally, I would like to thank three anonymous reviewers for their extensive critical, helpful and constructive feedback on an earlier version of this manuscript, which has greatly improved its quality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matthias M. Fischer.

Additional information

Publisher's Note

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

Appendices

Appendix A: Parametrisation of the model for the validation with experimental data

In order to validate the derived model using previously published experimental data, the model was parametrised using parameter values provided in the experimental paper by Zhang and Reed (2014) or, when such parameter values have not been provided by the authors, estimated from literature values.

Specifically, the initial glucose concentration of the medium was set to \(R(0) = 2 \; {\mathrm{{g/l}}}\) following the description of the culture medium given in the paper, while both initial metabolite concentrations were set to \(M_1(0) = M_2(0) = 0\), as neither of the two exchanged metabolites (leucine or lysine, respectively) was supplemented into the growth medium. The washout rate was set to \(\omega = 0\), as the coculture was not diluted over the course of the 75-h long experiment. Initial population densities were set to \(N_1(0) = N_2(0) = 5 \times 10^6\) cells per millilitre assuming a 50:50 mixture of the two strains at the initial OD of 0.01 described by the authors.

The intrinsic growth rates of the two auxotrophy mutants were set to the values provided by the authors, i.e. \(r_1 = 0.461 \; {\mathrm{{h}}}^{-1}, r_2 = 0.465 \; {\mathrm{{h}}}^{-1}\). The efflux rates of the two strains have been measured to be 0.027 mmol/(gDW \(\times\) h) leucine, and 0.02 mmol/(gDW \(\times\) h) lysine, respectively. Using a literature value of \(3 \times 10^{-13} \; \mathrm{{g}}\) dry weight (DW) per E.coli cell (Neidhardt et al. 1990), and molar masses of 131.17 g/mol for leucine and 146.19 g/mol for lysine, efflux rates of \(\epsilon _1 = 1.1 \; \mathrm{{fg/(cell}} \times {\mathrm{{h}}})\) and \(\epsilon _2 = 0.9 \; \mathrm{{fg/(cell}} \times {\mathrm{{h}}})\) were estimated. Metabolite requirements were measured by the authors as 0.35 mmol/(gDW) lysine and 0.473 mmol/(gDW) leucine, respectively. Using the same conversion factors as before, these amount to \(\gamma _1 = 15 \; {\mathrm{{fg/cell}}}\) lysine and \(\gamma _2 = 19 \; {\mathrm{{fg/cell}}}\) leucine.

The Monod constants of the two strains for the respective two amino acids has not been measured by the authors, so it was estimated from Kerner et al. (2012), who measured the Monod constants of E. coli for the amino acids tyrosine and trypsine, to be approximately \(K_1 = K_2 = 5 \; \upmu {\mathrm{{g/l}}}\). For the Monod constant of E. coli for glucose, the literature value of \(L_1 = L_2 = 0.1 \; {\mathrm{{g/l}}}\) from Senn et al. (1994) was used. As a side note, this striking difference of several orders of magnitude between \(K_1, K_2\) and \(L_1, L_2\) indicating a very high affinity of E. coli for the exchanged amino acids does not come as a surprise, as bacteria have evolved highly specialised permeases which are able to take up even extremely small concentrations of externally present aromatic amino acids, most likely in order to avoid the significantly more energetically expensive de novo synthesis of these amino acids which is switched off via feedback inhibition, whenever the respective amino acid is sufficiently present externally (Ames 1964). Finally, from Fig. 2b of the paper, the requirement of glucose per one bacterial cell was estimated to be \(\alpha _1 = \alpha _2 = 5000 \; {\mathrm{{fg/cell}}}\) by dividing the amount of utilised glucose by the final population density of the culture, assuming that bacterial death is neglectable for the short duration of the experiment of 75 h.

Appendix B: Proofs of Lemma 1 and 2

Proof

(Lemma 1—Invariance of the nonnegative orthant) Consider the behaviour of the five differential equations if the respective state variable equals zero. One gets:

$$\begin{aligned} \begin{aligned} \dot{N}_1(t) \arrowvert _{N_1(t)=0}&= 0 \\ \dot{N}_2(t) \arrowvert _{N_2(t)=0}&= 0 \\ \dot{R}(t) \arrowvert _{R(t)=0}&= \omega R_{{\mathrm{{in}}}}> 0 \\ \dot{M}_1(t) \arrowvert _{M_1(t)=0}&= \omega M_{1, {\mathrm{{in}}}} + \epsilon _1 N_1(t) - 0> 0\\ \dot{M}_2(t) \arrowvert _{M_2(t)=0}&= \omega M_{2, {\mathrm{{in}}}} + \epsilon _2 N_2(t) - 0 > 0\\ \end{aligned} \end{aligned}$$
(10)

Accordingly, the system will never be able to leave \({\mathbb {R}}_{\ge 0}^5\). \(\square\)

Proof

(Lemma 2—Boundedness) By inspection of the third differential equation, one can easily see the upper bound \(\forall t: R(t) < R_{{\mathrm{{in}}}}\).

Now introduce the quantity \(C(t) {:}{=}\alpha _1 N_1(t) + \alpha _2 N_2(t) + R(t)\), describing the complete concentration of resource in the culture vessel in form of free resource or microbial individuals. With some arithmetics, one may deduce that

$$\begin{aligned} \dot{C}(t) = \omega \cdot (R_{{\mathrm{{in}}}} - C(t)), \end{aligned}$$
(11)

from which directly follows that \(N_1(t), N_2(t)\) are bounded by the amount of resource influx below some values \(N_1^{\max }, N_2^{\max } \in {\mathbb {R}}^+\).

From this, one may obtain the following upper bounds of \(\dot{M}_1(t)\):

$$\begin{aligned} \forall t: \dot{M}_1(t) \le \omega M_{1, {\mathrm{{in}}}} + \epsilon _1 N_1^{\max } - \omega M_1(t) \end{aligned}$$
(12)

Thus, \(M_1(t)\) is bounded and similarly, \(M_2(t)\) is as well, which concludes the proof. \(\square\)

Appendix C: Jacobian of the system

The Jacobian of the system is given by

$$\begin{aligned} \mathbf{J } = \begin{bmatrix} \dfrac{M_2 R r_1}{(K_1+M_2)(L_1+R)} - \omega & 0 & \dfrac{L_1 M_2 N_1 r_1}{(K_1+M_2)(L_1+R)^2} & 0 & \dfrac{K_1 N_1 R r_1}{(K_1+M_2)^2 (L_1+R)} \\ 0 & \dfrac{M_1 R r_2}{(K_2+M_1)(L_2+R)} - \omega & \dfrac{L_2 M_1 N_2 r_2}{(K_2+M_1)(L_2+R)^2} & \dfrac{K_2 N_2 R r_2}{(K_2+M_1)^2 (L_2+R)} & 0 \\ - \dfrac{\alpha _1 M_2 R r_1}{(K_1 + M_2)(L_1+R)} & - \dfrac{\alpha _2 M_1 R r_2}{(K_2 + M_1)(L_2+R)} & - \dfrac{\alpha _1 L_1 M_2 N_1 r_1}{(K_1+M_2)(L_1+R)^2} - \dfrac{\alpha _2 L_2 M_1 N_2 r_2}{(K_2+M_1)(L_2+R)^2} - \omega & - \dfrac{\alpha _1 K_2 N_2 R r_2}{(K_2+M_1)^2 (L_2+R)} & - \dfrac{\alpha _2 K_1 N_1 R r_1}{(K_1+M_2)^2 (L_1+R)} \\ \epsilon _1 & - \dfrac{\gamma _2 M_1 R r_2}{(K_2 + M_1)(L_2+R)} & - \dfrac{\gamma _2 L_2 M_1 N_2 r_2}{(K_2+M_1)(L_2+R)^2} & - \dfrac{\gamma _2 K_2 N_2 R r_2}{(K_2+M_1)^2 (L_2+R)} - \omega & 0 \\ - \dfrac{\gamma _1 M_2 R r_1}{(K_1 + M_2)(L_1+R)} & \epsilon _2 & - \dfrac{\gamma _1 L_1 M_2 N_1 r_1}{(K_1+M_2)(L_1+R)^2} & 0 & - \dfrac{\gamma _1 K_1 N_1 R r_1}{(K_1+M_2)^2 (L_1+R)} - \omega \end{bmatrix}. \end{aligned}$$
(13)

As long as no metabolites are manually added to the culture, the Jacobian of the extinction state always reduces to

$$\begin{aligned} \mathbf{J } = \left[ \begin{matrix} - \omega & 0 & 0 & 0 & 0 \\ 0 & - \omega & 0 & 0 & 0 \\ 0 & 0 & - \omega & 0 & 0 \\ \epsilon _1 & 0 & 0 & - \omega & 0 \\ 0 & \epsilon _2 & 0 & 0 & - \omega \end{matrix} \right] . \end{aligned}$$
(14)

For the non-trivial case, where all system variables exceed zero, using Eq. (6) the Jacobian of the system simplifies to

$$\begin{aligned} \mathbf{J } = \begin{bmatrix} 0 & 0 & {\mathcal {L}}_1 & 0 & {\mathcal {K}}_1 \\ 0 & 0 & {\mathcal {L}}_2 & {\mathcal {K}}_2 & 0 \\ - \alpha _1 \omega & - \alpha _2 \omega & - \alpha _1 {\mathcal {L}}_1 - \alpha _2 {\mathcal {L}}_2 - \omega & - \alpha _1 {\mathcal {K}}_2& - \alpha _2 {\mathcal {K}}_1 \\ \epsilon _1 & - \gamma _2 \omega & - \gamma _2 {\mathcal {L}}_2 & - \gamma _2 {\mathcal {K}}_2 - \omega & 0 \\ - \gamma _1 \omega & \epsilon _2 & - \gamma _1 {\mathcal {L}}_1 & 0 & - \gamma _1 {\mathcal {K}}_1 - \omega \end{bmatrix}, \end{aligned}$$
(15)

where \({\mathcal {L}}_1 = \frac{L_1 N_1 \omega }{(L_1 + R) R}\), \({\mathcal {L}}_2 = \frac{L_2 N_2 \omega }{(L_2 + R) R}\), \({\mathcal {K}}_1 = \frac{K_1 N_1 \omega }{(K_1 + M_2) M_2}\), and \({\mathcal {K}}_2 = \frac{K_2 N_2 \omega }{(K_2 + M_1) M_1}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fischer, M.M. A mechanistic model of metabolic symbioses in microbes recapitulates experimental data and identifies a continuum of symbiotic interactions. Theory Biosci. 139, 265–278 (2020). https://doi.org/10.1007/s12064-020-00318-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12064-020-00318-2

Keywords

Navigation