Skip to main content
Log in

Cooperative metabolic resource allocation in spatially-structured systems

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Natural selection has shaped the evolution of cells and multi-cellular organisms such that social cooperation can often be preferred over an individualistic approach to metabolic regulation. This paper extends a framework for dynamic metabolic resource allocation based on the maximum entropy principle to spatiotemporal models of metabolism with cooperation. Much like the maximum entropy principle encapsulates ‘bet-hedging’ behaviour displayed by organisms dealing with future uncertainty in a fluctuating environment, its cooperative extension describes how individuals adapt their metabolic resource allocation strategy to further accommodate limited knowledge about the welfare of others within a community. The resulting theory explains why local regulation of metabolic cross-feeding can fulfil a community-wide metabolic objective if individuals take into consideration an ensemble measure of total population performance as the only form of global information. The latter is likely supplied by quorum sensing in microbial systems or signalling molecules such as hormones in multi-cellular eukaryotic organisms.

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

Similar content being viewed by others

References

  • Ackermann M (2015) A functional perspective on phenotypic heterogeneity in microorganisms. Nat Rev Microbiol 13:497–508

    Google Scholar 

  • An JH, Goo E, Kim H, Seo YS, Hwang I (2014) Bacterial quorum sensing and metabolic slowing in a cooperative population. Proc Natl Acad Sci USA 111:14912–14927

    Google Scholar 

  • Axelrod R, Hamilton WD (1981) The evolution of cooperation. Science 211:1390–1396

    MathSciNet  MATH  Google Scholar 

  • Barrat A, Barthélemy M, Vespignani A (2008) Dynamical processes on complex networks. Cambridge Univ Press, Cambridge

    MATH  Google Scholar 

  • Basan M, Hui S, Okano H, Zhang Z, Shen Y, Williamson JR, Hwa T (2015) Overflow metabolism in Escherichia coli results from efficient proteome allocation. Nature 528:99–104

    Google Scholar 

  • Bélanger M, Allaman I, Magistretti PJ (2011) Brain energy metabolism: focus on astrocyte-neuron metabolic cooperation. Cell Metab 14:724–738

    Google Scholar 

  • Biggs MB, Papin JA (2013) Novel multiscale modeling tool applied to Pseudomonas aeruginosa biofilm formation. PLoS ONE 8:e78011

    Google Scholar 

  • Boccaletti S, Bianconi G, Criado R, del Genio CI, Gómez-Gardeñes J, Romance M, deI Sendiña-Nadal I, Wang Z, Zanin M, (2014) The structure and dynamics of multilayer networks. Phys Rep 544:1–122

  • Boyle KE, Monaco H, van Ditmarsch D, Deforet M, Xavier JB (2015) Integration of metabolic and quorum sensing signals governing the decision to cooperate in a bacterial social trait. PLoS Comput Biol 11:e1004279

    Google Scholar 

  • Boyle KE, Monaco HT, Deforet M, Yan J, Wang Z, Rhee K, Xavier JB (2017) Metabolism and the evolution of social behavior. Mol Biol Evol 34:2367–2379

    Google Scholar 

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

    Google Scholar 

  • Čáp M, Štěpánek L, Harant K, Váchová L, Palková Z (2012) Cell differentiation within a yeast colony: metabolic and regulatory parallels with a tumor-affected organism. Mol Cell 46:436–448

    Google Scholar 

  • Chao L, Levin BR (1981) Structured habitats and the evolution of anticompetitor toxins in bacteria. Proc Natl Acad Sci USA 78:6324–6328

    Google Scholar 

  • Chen J, Gomez JA, Höffner K, Phalak P, Barton PI, Henson MA (2016) Spatiotemporal modeling of microbial metabolism. BMC Syst Biol 10:21

    Google Scholar 

  • Cole JA, Kohler L, Hedhli J, Luthey-Schulten Z (2015) Spatially-resolved metabolic cooperativity within dense bacterial colonies. BMC Syst Biol 9:15

    Google Scholar 

  • Davenport PW, Griffin JL, Welch M (2015) Quorum sensing is accompanied by global metabolic changes in the opportunistic human pathogen Pseudomonas aeruginosa. J Bacteriol 197:2072–2082

    Google Scholar 

  • De Martino D, Andersson AMC, Bergmiller B, Guet CC, Tkac̆ik G, (2017) Statistical mechanics for metabolic networks during steady state growth. Nat Commun 9:2988

  • D’Souza G, Shitut S, Preussger D, Yousif G, Waschina S, Kost C (2018) Ecology and evolution of metabolic cross-feeding interactions in bacteria. Nat Prod Rep 35:455–488

    Google Scholar 

  • Estrela S, Gudelj I (2010) Evolution of cooperative cross-feeding could be less challenging than originally thought. PLoS ONE 5:e14121

    Google Scholar 

  • Evans CR, Kempes CP, Price-Whelan A, Dietrich LEP (2020) Metabolic heterogeneity and cross-feeding in bacterial multicellular systems. Trends Microbiol. https://doi.org/10.1016/j.tim.2020.03.008 (in press)

    Article  Google Scholar 

  • Fernandez-de-Cossio-Diaz J, Mulet R (2019) Maximum entropy and population heterogeneity in continuous cell cultures. PLoS Comput Biol 15:e1006823

    Google Scholar 

  • Fernandez-de-Cossio-Diaz J, Mulet R (2020) Statistical mechanics of interacting metabolic networks. Phys Rev E 101:042401

    MathSciNet  Google Scholar 

  • Fleming RMT, Maes CM, Saunders MA, Ye Y, Palsson BØ (2012) A variational principle for computing non-equilibrium fluxes and potentials in genome-scale biochemical networks. J Theor Biol 292:71–77

    MATH  Google Scholar 

  • Françios J, Parrou JL (2001) Reserve carbohydrates metabolism in the yeast Saccharomyces cerevisiae. FEMS Microbiol Rev 25:125–145

    Google Scholar 

  • Gardner A, West SA, Griffin AS (2007) Is bacterial persistence a social trait? PLoS ONE 2:e752

    Google Scholar 

  • Germerodt S, Bohl K, Lück A, Pande S, Schröter A, Kaleta C, Schuster S, Kost C (2016) Pervasive selection for cooperative cross-feeding in bacterial communities. PLoS Comput Biol 12:e1004986

    Google Scholar 

  • Goo E, An JH, Kang Y, Hwang I (2015) Control of bacterial metabolism by quorum sensing. Trends Microbiol 23:567–576

    Google Scholar 

  • Granados AA, Crane MM, Montano-Gutierrez LF, Tanaka RJ, Voliotis M, Swain PS (2017) Distributing tasks via multiple input pathways increases cellular survival in stress. Elife 6:e21415

    Google Scholar 

  • Grote J, Krysciak D, Streit WR (2015) Phenotypic heterogeneity, a phenomenon that may explain why quorum sensing does not always result in truly homogenous cell behavior. Appl Environ Microbiol 81:5280–5289

    Google Scholar 

  • Hansen SK, Rainey PB, Haagensen JA, Molin S (2007) Evolution of species interactions in a biofilm community. Nature 445:533–536

    Google Scholar 

  • Harcombe W (2010) Novel cooperation experimentally evolved between species. Evolution 64:2166–2172

    Google Scholar 

  • Harcombe WR, Riehl WJ, Dukovski I, Granger BR, Betts A, Lang AH, Bonilla G, Kar A, Leiby N, Mehta P, Marx CJ, Segrè D (2014) Metabolic resource allocation in individual microbes determines ecosystem interactions and spatial dynamics. Cell Rep 7:1104–1115

    Google Scholar 

  • Harte J, Newman EA (2014) Maximum information entropy: a foundation for ecological theory. Trends Ecol Evol 29:384–389

    Google Scholar 

  • Hauert C, Doebeli M (2004) Spatial structure often inhibits the evolution of cooperation in the snowdrift game. Nature 428:643–646

    Google Scholar 

  • Hindmarsh AC, Brown PN, Grant KE, Lee SL, Serban R, Shumaker DE, Woodward CS (2005) SUNDIALS: suite of nonlinear and differential/algebraic equation solvers. ACM Trans Math Softw 31:363–396

    MathSciNet  MATH  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:e1002540

    Google Scholar 

  • Jaynes ET (1957) Information theory and statistical mechanics. Phys Rev 106:620–630

    MathSciNet  MATH  Google Scholar 

  • Jones KD, Kompala DS (1999) Cybernetic model of the growth dynamics of Saccharomyces cerevisiae in batch and continuous cultures. J Biotechnol 71:105–131

    Google Scholar 

  • Klamt S, Stelling J (2002) Combinatorial complexity of pathway analysis in metabolic networks. Mol Biol Rep 29:233–236

    Google Scholar 

  • Kussell E, Leibler S (2005) Phenotypic diversity, population growth, and information in fluctuating environments. Science 309:2075–2078

    Google Scholar 

  • Lille SH, Pringle JR (1980) Reserve carbohydrate metabolism in Saccharomyces cerevisiae: responses to nutrient limitation. J Bacteriol 143:1384–1394

    Google Scholar 

  • Lin YC, Cornell WC, Jo J, Price-Whelan A, Dietrich LEP (2018) The Pseudomonas aeruginosa complement of lactate dehydrogenases enables use of D- and L-lactate and metabolic cross-feeding. mBio 9:e0096-18

    Google Scholar 

  • Lyssiotis CA, Kimmelman AC (2017) Metabolic interactions in the tumor microenvironment. Trends Cell Biol 27:863–875

    Google Scholar 

  • Mahadevan R, Edwards JS, Doyle FJ 3rd (2002) Dynamic flux balance analysis of diauxic growth in Escherichia coli. Biophys J 83:1331–1340

    Google Scholar 

  • Martin W (2010) Evolutionary origins of metabolic compartmentalization in eukaryotes. Philos Trans R Soc Lond B Biol Sci 365:847–855

    Google Scholar 

  • Miller MB, Bassler BL (2001) Quorum sensing in bacteria. Annu Rev Microbiol 55:165–199

    Google Scholar 

  • Möller P, Liu X, Schuster S, Boley D (2018) Linear programming model can explain respiration of fermentation products. PLoS ONE 13:e0191803

    Google Scholar 

  • Mori M, Schink S, Erickson DW, Gerland U, Hwa T (2017) Quantifying the benefit of a proteome reserve in fluctuating environments. Nat Commun 8:1225

    Google Scholar 

  • Morris BE, Henneberger R, Huber H, Moissl-Eichinger C (2013) Microbial syntrophy: interaction for the common good. FEMS Microbiol Rev 37:384–406

    Google Scholar 

  • Mulcahy LR, Isabella VM, Lewis K (2014) Pseudomonas aeruginosa biofilms in disease. Microb Ecol 68:1–12

    Google Scholar 

  • Nadell CD, Foster KR, Xavier JB (2010) Emergence of spatial structure in cell groups and the evolution of cooperation. PLoS Comput Biol 6:e1000716

    Google Scholar 

  • Nadell CD, Drescher K, Foster KR (2016) Spatial structure, cooperation and competition in biofilms. Nat Rev Microbiol 14:589–600

    Google Scholar 

  • Nakao H, Mikhailov AS (2010) Turing patterns in network-organized activator-inhibitor networks. Nat Phys 6:544–550

    Google Scholar 

  • Nash J (1950) The bargaining problem. Econometrica 18:155–162

    MathSciNet  MATH  Google Scholar 

  • Nelson DL, Cox MM, Michael M (2005) Lehninger principles of biochemistry, 4th edn. W.H, Freeman and Company, New York, p 543

    Google Scholar 

  • Nowak MA, May RM (1992) Evolutionary games and spatial chaos. Nature 359:826–829

    Google Scholar 

  • O’Brien EJ, Utrilla J, Palsson BØ (2016) Quantification and classification of E. coli proteome utilization and unused protein costs across environments. PLoS Comput Biol 12:1004998

    Google Scholar 

  • Oliveira NM, Niehus R, Foster KR (2014) Evolutionary limits to cooperation in microbial communities. Proc Natl Acad Sci USA 111:17941–17946

    Google Scholar 

  • Orth JD, Thiele I, Palsson BØ (2010) What is flux balance analysis? Nat Biotechnol 28:245–248

    Google Scholar 

  • Othmer HG, Scriven LE (1971) Instability and dynamic pattern in cellular networks. J Theor Biol 32:507–537

    Google Scholar 

  • Othmer HG, Scriven LE (1974) Nonlinear aspects of dynamic pattern in cellular networks. J Theor Biol 43:83–112

    Google Scholar 

  • Pacheco AR, Moel M, Segré D (2019) Costless metabolic secretions as drivers of inter-species interactions in microbial ecosystems. Nat Commun 10:103

    Google Scholar 

  • Pattanaik PK (2008) Social welfare function (definition). In: Durlauf SN, Blume LE (eds) The new Palgrave dictionary of economics, 2nd edn. Palgrave Macmillan, Basingstoke

    Google Scholar 

  • Perkins TJ, Swain PS (2009) Strategies for cellular decision-making. Mol Syst Biol 5:326

    Google Scholar 

  • Peterson JR, Cole JA, Luthey-Schulten Z (2017) Parametric studies of metabolic cooperativity in Escherichia coli colonies: strain and geometric confinement effects. PLoS ONE 12:e0182570

    Google Scholar 

  • Pfeiffer T, Schuster S, Bonhoeffer S (2001) Cooperation and competition in the evolution of ATP-producing pathways. Science 292:504–507

    Google Scholar 

  • Popat R, Cornforth DM, McNally L, Brown SP (2015) Collective sensing and collective responses in quorum-sensing bacteria. J R Soc Interface 12:20140882

    Google Scholar 

  • Provost A, Bastin G (2004) Dynamic metabolic modelling under the balanced growth condition. J Process Control 14:717–728

    Google Scholar 

  • Provost A, Bastin G, Agathos SN, Schneider YJ (2006) Metabolic design of macroscopic bioreaction models: application to Chinese hamster ovary cells. Bioprocess Biosyst Eng 29:349–366

    Google Scholar 

  • Rasamiravaka T, Labtani Q, Duez P, El Jaziri M (2015) The formation of biofilms by Pseudomonas aeruginosa: a review of the natural and synthetic compounds interfering with control mechanisms. Biomed Res Int 2015:759348

    Google Scholar 

  • Ridden SJ, Chang HH, Zygalakis KC, MacArthur BD (2015) Entropy, ergodicity, and stem cell multipotency. Phys Rev Lett 115:208103

    Google Scholar 

  • Schink B (1991) Syntrophism among prokaryotes. In: Ballows AT, Dworkin M, Schleifer KH (eds) The prokaryotes. Springer, New York, pp 276–299

    Google Scholar 

  • Schink B (2002) Synergistic interactions in the microbial world. Antonie Van Leeuwenhoek 81:257–261

    Google Scholar 

  • Schuster S, Hilgetag C (1994) On elementary flux modes in biochemical reaction systems at steady state. J Biol Syst 2:165–182

    Google Scholar 

  • Seger J, Brockmann HJ (1987) What is bet-hedging? In: Harvey PH, Partridge L (eds) Oxford surveys in evolutionary biology. Oxford University Press, Oxford, pp 182–211

    Google Scholar 

  • Semenza GL (2008) Tumor metabolism: cancer cells give and take lactate. J Clin Investig 118:3835–3837

    Google Scholar 

  • Shannon CE (1948) A mathematical theory of communication. Bell Syst Tech J 27:379–423

    MathSciNet  MATH  Google Scholar 

  • Shore JE, Johnson RW (1980) Axiomatic derivation of the principle of maximum entropy and the principle of minimum cross-entropy. IEEE Trans Inf Theory 26:26–37

    MathSciNet  MATH  Google Scholar 

  • Smith NW, Shorten PR, Altermann E, Roy NC, McNabb WC (2019) The classification and evolution of bacterial cross-feeding. Front Ecol Evol 7:153

    Google Scholar 

  • Song HS, Ramkrishna D (2010) Prediction of metabolic function from limited data: Lumped Hybrid Cybernetic Modeling (L-HCM). Biotechnol Bioeng 106:271–284

    Google Scholar 

  • Song HS, Ramkrishna D (2011) Cybernetic models based on lumped elementary modes accurately predict strain-specific metabolic function. Biotechnol Bioeng 108:127–140

    Google Scholar 

  • Song H-S, Cannon WR, Beliaev AS, Konopka A (2014) Mathematical modeling of microbial community dynamics: a methodological review. Processes 2:711–752

    Google Scholar 

  • Sonveaux P et al (2008) Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J Clin Investig 118:3930–3942

    Google Scholar 

  • Tourigny DS (2020) Dynamic metabolic resource allocation based on the maximum entropy principle. J Math Biol 80:2395–2430

    MathSciNet  MATH  Google Scholar 

  • Varahan S, Walvekar A, Sinha V, Krishna S, Laxman S (2019) Metabolic constraints drive self-organization of specialized cell groups. Elife 8:e46735

    Google Scholar 

  • Varma A, Palsson BØ (1994) Metabolic flux balancing: basic concepts, scientific and practical use. Nat Biotechnol 12:994–998

    Google Scholar 

  • Wolfsberg E, Long CP, Antoniewicz MR (2018) Metabolism in dense microbial colonies: 13C metabolic flux analysis of E. coli T grown on agar identifies two distinct cell populations with acetate cross-feeding. Metab Eng 49:242–247

    Google Scholar 

  • Woronoff G, Nghe P, Baudry J, Boitard L, Braun E, Griffiths AD, Bibette J (2020) Metabolic cost of rapid adaptation of single yeast cells. Proc Natl Acad Sci USA 117:10660–10666

    Google Scholar 

  • Young JD, Ramkrishna D (2007) On the matching and proportional laws of cybernetic models. Biotechnol Prog 23:83–99

    Google Scholar 

  • Young JD, Henne KL, Morgan JA, Konopka AE, Ramkrishna D (2008) Integrating cybernetic modeling with pathway analysis provides a dynamic, systems-level description of metabolic control. Biotechnol Bioeng 100:542–559

    Google Scholar 

  • Zanghellini J, Gerstl MP, Hanscho M, Nair G, Regensburger G, Müller S, Jungreuthmayer C (2016) Toward Genome-Scale Metabolic Pathway Analysis. In: Wittmann C, Liao JC (eds) Industrial biotechnology

  • Zhang T, Parker A, Carlson RP, Stewart PS (2018) Flux-balance based modeling of biofilm communities. https://doi.org/10.1101/441311

Download references

Acknowledgements

This work benefitted from conversations with HC Causton, L Dietrich, and MR Parsek on microbial biology and quorum sensing. DS Tourigny is a Simons Foundation Fellow of the Life Sciences Research Foundation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David S. Tourigny.

Additional information

Publisher's Note

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

Appendix A: Derivation of the cooperative control law

Appendix A: Derivation of the cooperative control law

Collecting dynamic variables into vectors \({\mathbf {X}}_i = ({\mathbf {m}}_i,x_i)^T\), the system (4) can be represented by the coupled equations \(\dot{{\mathbf {X}}}_i = {\mathbf {F}}_i({\mathbf {X}}, {\mathbf {u}}^i)\) (\(i=1,2,\ldots ,{\mathcal {N}}\)) with \({\mathbf {X}} = ({\mathbf {X}}_1, {\mathbf {X}}_2, \ldots , {\mathbf {X}}_{{\mathcal {N}}})^T\). As in Young and Ramkrishna (2007), Tourigny (2020), the individual controls \({\mathbf {u}}^i\) are determined by using the linearisation of (4) about the state \({\mathbf {X}}(t)\) and reference controls \(\{ {\mathbf {u}}^i_0 \}_{i=1,2,\ldots ,{\mathcal {N}}}\) as a good approximation for the system response at time \(t+\tau \) with \(\tau \in [0,\Delta t]\). The linearised equations are

$$\begin{aligned} \frac{d}{d\tau } \Delta {\mathbf {X}}_i = {\mathbf {A}}_i \Delta {\mathbf {X}} + {\mathbf {B}}_i \Delta {\mathbf {u}}^i + {\mathbf {F}}_i({\mathbf {X}}(t),{\mathbf {u}}^i_0) \end{aligned}$$

where

$$\begin{aligned} {\mathbf {A}}_i = \frac{\partial }{\partial {\mathbf {X}}} {\mathbf {F}}_i ({\mathbf {X}}(t),{\mathbf {u}}^i) \end{aligned}$$

is the ith block row of the full Jacobian matrix \({\mathbf {A}}\) and

$$\begin{aligned} {\mathbf {B}}_i = \frac{\partial }{\partial {\mathbf {u}}^i} {\mathbf {F}}_i ({\mathbf {X}}(t),{\mathbf {u}}^i), \end{aligned}$$

with \(\Delta {\mathbf {X}} = {\mathbf {X}}(t+\tau ) - {\mathbf {X}}(t)\) and \(\Delta {\mathbf {u}}^i = {\mathbf {u}}^i (t+ \tau ) - {\mathbf {u}}^i_0\). When the species at all nodes are identical (i.e., \({\mathbf {S}}_i = {\mathbf {S}}\), \(K_i = K\), \({\mathbf {c}}_i = {\mathbf {c}}\), \(r^i_k = r_k\), and \({\mathbf {Z}}^k_i = {\mathbf {Z}}^k\) for all \(i = 1,2,\ldots , {\mathcal {N}}\)) then each cooperative maximum entropy control \({\mathbf {u}}^i\) is obtained by maximising the objective functional

$$\begin{aligned} {\mathcal {F}}_i({\mathbf {u}}) = \varvec{\lambda }^T_i {\mathbf {B}}_i {\mathbf {u}} + \sigma H({\mathbf {u}}) \end{aligned}$$

with respect to \({\mathbf {u}}\), where

$$\begin{aligned} H({\mathbf {u}} ) = - \sum _{k=1}^K u_k \log (u_k) \end{aligned}$$

is the entropy constraint and \(\varvec{\lambda } = (\varvec{\lambda }_1,\varvec{\lambda }_2,\ldots ,\varvec{\lambda }_{{\mathcal {N}}})^T\) is the \({\mathcal {N}} \times (M+1)\)-dimensional vector of Pontryagin co-state variables obtained by solving the boundary value problem

$$\begin{aligned} - \frac{d}{d \tau } \varvec{\lambda } = {\mathbf {A}}^T \varvec{\lambda }, \quad \varvec{\lambda }(t + \Delta t) = {\mathbf {q}}_i . \end{aligned}$$

The complete solution to the co-state equations is

$$\begin{aligned} \varvec{\lambda } (t + \tau ) = {\mathbf {e}}^{{\mathbf {A}}^T(\Delta t - \tau )} {\mathbf {q}}_i, \quad 0 \le \tau \le \Delta t , \end{aligned}$$

but the effective return-on-investment (6) is obtained by substituting for \( \varvec{\lambda }\) with the heuristic \(\tau =0\) because, ultimately, only the optimal control input at the current time t is of interest (Young and Ramkrishna 2007). Maximisation of \({\mathcal {F}}_i\) proceeds as in Tourigny (2020) to yield the cooperative maximum entropy control (7).

An expression for the greedy effective return-on-investment is obtained by setting \(\Delta t = 0\), in which case \(\varvec{\lambda }_i = \partial \phi _i ({\mathbf {X}})/\partial {\mathbf {X}}_i\). Moreover, \(\partial \phi _i/\partial x_i\) is the only non-zero component of this vector because the cooperative objective \(\phi _i\) does not depend on concentrations of slow metabolites. Differentiation of the generalised mean \(\phi _i = M_p({\mathbf {x}})\) with respect to \(x_i\) yields

$$\begin{aligned} \frac{\partial \phi _i}{\partial x_i} = \frac{x^{p-1}_i}{\sum _{j=1}^{\mathcal {N}} x_j^p} \cdot \left( \frac{1}{{\mathcal {N}}} \sum _{j=1}^{\mathcal {N}} x_j^p \right) ^{1/p} = \frac{x_i^{p-1}}{x_i^p + y^p} \cdot M_p({\mathbf {x}}) \end{aligned}$$

and the kth column of \({\mathbf {B}}_i\) is

$$\begin{aligned} {\mathbf {B}}_i^k = x_i r_k({\mathbf {m}}_i) \begin{pmatrix} {\mathbf {S}} \\ {\mathbf {c}}^T \end{pmatrix} {\mathbf {Z}}^k \end{aligned}$$

so that

$$\begin{aligned} {\mathcal {R}}^{k,i}_0 = \frac{x_i^p}{x_i^p + y^p} \cdot M_p({\mathbf {x}}) \cdot r_k({\mathbf {m}}_i) {\mathbf {c}}^T {\mathbf {Z}}^k . \end{aligned}$$

This provides the greedy effective return-on-investment (8) defined using \(R^k_0({\mathbf {m}}_i)\) in the main text, and the resulting greedy cooperative maximum entropy control is

$$\begin{aligned} u^i_k(t) = \frac{1}{Q_i}\exp \left( {\mathcal {R}}_0^{k,i} /\sigma \right) , \quad k=1,2, \ldots ,K \end{aligned}$$

with \(Q_i = \sum _{k=1}^K \exp ( {\mathcal {R}}_0^{k,i} /\sigma )\). As mentioned in the main text, alternative controls can be obtained in cases where the objectives \(\phi _i\) and/or species are different across nodes in the population network, but are not considered here. To demonstrate the utilitarian objective with \(p=1\) recovers the individualistic greedy maximum entropy control law from Tourigny (2020), note that in this case

$$\begin{aligned} M_1({\mathbf {x}}) = \frac{1}{{\mathcal {N}}} \sum _{j=1}^{\mathcal {N}} x_j , \quad x_i^1 + y^1 = x_i + \sum _{j \ne i }^{\mathcal {N}} x_j = \sum _{j=1}^{\mathcal {N}} x_j \end{aligned}$$

so that substitution for the denominator in \({\mathcal {R}}^{k,i}_0 \) yields

$$\begin{aligned} {\mathcal {R}}^{k,i}_0 = \frac{1}{{\mathcal {N}}} \cdot x_i r_k({\mathbf {m}}_i) {\mathbf {c}}^T {\mathbf {Z}}^k , \end{aligned}$$

which is the individualistic effective return-on-investment (Equation 18 in Tourigny 2020) for the kth EFM at node i, multiplied by a common factor of \(1/{\mathcal {N}}\) that has no effect on the resulting control law.

Although beyond the scope of this paper, it is worth pointing out that use of the maximum entropy principle in Tourigny (2020) and its cooperative extension considered here is of a slightly different nature compared to those discussed previously (De Martino et al. 2017; Fernandez-de-Cossio-Diaz and Mulet 2019) (see also Fleming et al. (2012) for explicit use of entropy in an objective function) because the Boltzmann-like distribution is used to weight EFMs rather than individual reactions. The resulting steady state flux distributions obtained using these approaches are not necessarily the same, since the former maximum entropy principle will emphasise the spread of resource across EFMs while the latter does so on a reaction-by-reaction basis. At least in a simplified case where effective return-on-investment is not considered and only stoichiometric constraints must be satisfied, it could be argued the former is more biologically-relevant because spreading resource equally among metabolic pathways– as fundamental functional units of the cell– would be more meaningful than equal distribution of resource among reactions. The latter, albeit done under stoichiometric constraints, could result in metabolic pathways containing more reactions receiving a greater fraction of resource even though their effective return-on-investments are no larger than others.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tourigny, D.S. Cooperative metabolic resource allocation in spatially-structured systems. J. Math. Biol. 82, 5 (2021). https://doi.org/10.1007/s00285-021-01558-6

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-021-01558-6

Keywords

Mathematics Subject Classification

Navigation