Skip to main content
Log in

Necessary and sufficient conditions for protocell growth

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

Abstract

We consider a generic protocell model consisting of any conservative chemical reaction network embedded within a membrane. The membrane results from the self-assembly of a membrane precursor and is semi-permeable to some nutrients. Nutrients are metabolized into all other species including the membrane precursor, and the membrane grows in area and the protocell in volume. Faithful replication through cell growth and division requires a doubling of both cell volume and surface area every division time (thus leading to a periodic surface area-to-volume ratio) and also requires periodic concentrations of the cell constituents. Building upon these basic considerations, we prove necessary and sufficient conditions pertaining to the chemical reaction network for such a regime to be met. A simple necessary condition is that every moiety must be fed. A stronger necessary condition implies that every siphon must be either fed, or connected to species outside the siphon through a pass reaction capable of transferring net positive mass into the siphon. And in the case of nutrient uptake through passive diffusion and of constant surface area-to-volume ratio, a sufficient condition for the existence of a fixed point is that every siphon be fed. These necessary and sufficient conditions hold for any chemical reaction kinetics, membrane parameters or nutrient flux diffusion constants.

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

Similar content being viewed by others

Notes

  1. For the reference numerical protocell example given in Bigan et al. (2015a), the deficiency is 9 (32) for the threshold (maximum-size) network, respectively (using the reference nutrient and membrane precursor combination). Consistently, numerical simulations on protocells based on random CRNs reveal that the stationary growth states are not complex-balanced.

  2. One intuitive way to grasp the meaning of siphons is the following: some CRNs are so strongly coupled that if one attempts to lower the concentration of one particular species by an appropriate ’sink’ (e.g. incorporation into a structured membrane), then one ’sinks’ all other concentrations as well. This is the case when the only siphon is the full set of species. But for CRNs with a weaker coupling, it is foreseeable to ’sink’ some species while keeping other concentrations positive. A subset of species that can be ’sank’ is a siphon that is shorter than the full set of species.

  3. Condition \(\mathcal {A}\) is stronger (weaker) than condition \(\mathcal {B}\) if \(\mathcal {A}\implies \mathcal {B}\) (\(\mathcal {B}\implies \mathcal {A}\)), respectively. Stronger necessary and weaker sufficient conditions are desirable for a finer delineation of ’working’ protocells.

  4. Moieties are non-negative basis vectors of the left-null space of the stoichiometry matrix. Any mass vector can be decomposed as a positive linear combination of moieties. In contrast, extreme pathways (Schilling et al. 2000) or the closely related elementary flux modes (Schuster and Hilgetag 1994) are non-negative basis vectors of the right-null space of the stoichiometry matrix. Any stationary flux distribution can be decomposed as a positive linear combination of extreme pathways (or elementary flux modes).

  5. It is implicitly assumed that the \((\mathscr {V}, \mathscr {A})\) trajectories are such that at any instant, there is enough membrane surface area \(\mathscr {A}\) to accommodate the volume \(\mathscr {V}\), \(\mathscr {A}\ge \root 3 \of {36\pi \mathscr {V}}\). Else, the protocell may burst. See Mavelli and Ruiz-Mirazo (2013).

  6. It can also be constant, which is a peculiar periodic function.

  7. We were unable to construct a Lyapunov function for this protocell model. As mentioned in Sect. 1.1, the Lyapunov function used in existing deficiency-based CRNT theorems requires complex-balanced equilibria. Consistently, numerical simulations show that protocell stationary growth states are not complex-balanced (see Footnote 1).

  8. The fact that complex bounded trajectories cannot be excluded (in which case the protocell might not be working) is the reason why Definitions 6 and 7 are kept distinct.

  9. This fixed point may depend on the initial conditions. A typical example (considered in Wei 1962) is a closed conservative chemical reaction system: both the trajectory upper bound and the equilibrium point depend on the initial conditions (which determine the total mass in the system). See also Footnote 10.

  10. The original proof in Wei (1962) holds even if the trajectory bound depends on the initial conditions. Whereas the statement in Basener et al. (2006) and Richeson et al. (2002) relies upon the existence of a convex and forward-invariant set, which corresponds to a trajectory bound independent of the initial conditions (after a sufficient time).

  11. An extinction set is defined as follows: if for some particular set of strictly positive initial conditions, the corresponding \(\omega \)-limit set of a bounded concentration vector trajectory has zero concentration for some non-empty subset Z of all species, then Z is an extinction set. Note that this definition as well as the statement of Theorem 4 in Angeli et al. (2011) calls for bounded concentration vector trajectories. However, its proof appears not to rely upon this assumption and only requires continuity (of trajectories and kinetics), which is the reason why we use it here in this extended form.

  12. L or T can only incorporate among already self-assembled L. As in Molenaar et al. (2009), this excludes a growing membrane composed only of T but the corresponding process is modeled mechanistically here, instead of being imposed as en external constraint.

  13. Or equivalently, 1/\(N_{L}\) (1/\(N_{T}\)) is the membrane surface area per self-assembled molecule of L (T), respectively.

  14. As explained in the next discussion section, the transmembrane protein T is distinct from the activator molecule which is a high-energy metabolite, down-converted to a lower-energy molecule in the active transport process.

  15. Even if the proof of the sufficient condition still held, this sufficient condition (which grants the existence of a stationary growth state if every siphon is fed) would obviously not even be met because neither \(Z_{2}\) nor \(Z_{3}\) are fed.

  16. The activator molecule is distinct from the transmembrane protein introduced in the example of Sect. 6.2, which acts as a channel for \(A_{\text {pnu}}\) and which was not included in the generic model of Sect. 3. In the case of the PTS glucose transport system for E. coli, this activator is phosphoenolpyruvate (pep) which is down-converted to pyruvate (pyr) while having at the same time glucose (glu) converted to glucophosphate (g6p) (Jahreis et al. 2008). The corresponding transport reaction is \(glu + pep \rightarrow g6p + pyr\), which only occurs across the membrane so that only glucophosphate (and not glucose) is present inside the cell. This activator molecule should also not be confused with a co-factor, which is typically left unchanged by a biochemical reaction. The activator molecule should rather be seen as an energy currency fueling the active transport process.

  17. Such an assumption does not automatically grant synchronization of the rates of synthesis of membrane and cytoplasmic constituents: a counterexample is given by the simple example of Appendix 1, for which the existence of a fixed point is only granted within some parameter range.

  18. Since the original submission of the present work, we have carried out additional numerical and theoretical work on protocells having a membrane that is so flexible that the osmotic pressure is quasi-instantaneously balanced across the membrane. Filamentation was found to be an emerging growth mode in such a case, but with a filament diameter that depends on the environmental conditions as well as on the membrane and chemical reaction network parameters (Bigan et al. 2015d).

  19. This statement only applies to protocells. Some modern evolved cells have acquired complex regulatory mechanisms to regulate their size or shape, which are presumably fine-tuned. An example is the osmotic stress pathway and osmoadaptation in yeast (Hohmann 2002).

References

  • Angeli D, De Leenheer P, Sontag ED (2007) A Petri net approach to the study of persistence in chemical reaction networks. Math. Biosci. 210(2):598–618

    Article  MathSciNet  MATH  Google Scholar 

  • Angeli D, De Leenheer P, Sontag ED (2011) Persistence results for chemical reaction networks with time-dependent kinetics and no global conservation laws. SIAM J. Appl. Math. 71(1):128–146

    Article  MathSciNet  MATH  Google Scholar 

  • Basener W, Brooks BP, Ross D (2006) The Brouwer fixed point theorem applied to rumour transmission. Appl. Math. Lett. 19(8):841–842

    Article  MathSciNet  MATH  Google Scholar 

  • Bigan E, Steyaert JM, Douady S (2015a) Minimal conditions for protocell stationary growth. Artif. Life 21(2):166–192

    Article  Google Scholar 

  • Bigan E, Steyaert JM, Douady S (2015b) On necessary and sufficient conditions for proto-cell stationary growth. Electron Notes Theor. Comput. Sci. 316:3–15

    Article  Google Scholar 

  • Bigan E, Steyaert JM, Douady S (2015c) Chemical schemes for maintaining different compositions across a semi-permeable membrane with application to proto-cells. Orig. Life Evol. Biosph. 45(4):439–454

    Article  Google Scholar 

  • Bigan E, Steyaert JM, Douady S (2015d) Filamentation as a primitive growth mode? Phys. Biol. 12(6):066024

    Article  Google Scholar 

  • Božič B, Svetina S (2004) A relationship between membrane properties forms the basis of a selectivity mechanism for vesicle self-reproduction. Eur. Biophys. J. 33(7):565–571

    Article  Google Scholar 

  • Busa W, Nuccitelli R (1984) Metabolic regulation via intracellular pH. Am. J. Physiol. Regul. Integr. Comp. Physiol. 246(4):R409–R438

    Google Scholar 

  • Craciun G, Feinberg M (2005) Multiple equilibria in complex chemical reaction networks: I. The injectivity property. SIAM J. Appl. Math. 65(5):1526–1546

    Article  MathSciNet  MATH  Google Scholar 

  • Craciun G, Feinberg M (2006) Multiple equilibria in complex chemical reaction networks: II. The species-reaction graph. SIAM J. Appl. Math. 66(4):1321–1338

    Article  MathSciNet  MATH  Google Scholar 

  • Dickinson J (2008) Filament formation in Saccharomyces cerevisiae—a review. Folia Microbiol. 53(1):3–14

    Article  Google Scholar 

  • Ederer M, Gilles E (2007) Thermodynamically feasible kinetic models of reaction networks. Biophys. J. 92(6):1846–1857

    Article  Google Scholar 

  • Érdi P, Tóth J (1989) Mathematical Models of Chemical Reactions: Theory and Applications of Deterministic and Stochastic Models. Manchester University Press, Manchester, UK

  • Feinberg M (1972) Complex balancing in general kinetic systems. Arch. Ration. Mech. Anal. 49(3):187–194

    Article  MathSciNet  Google Scholar 

  • Feinberg, M.: Lectures on chemical reaction networks. Notes of lectures given at the Mathematics Research Center, University of Wisconsin, Madison, WI. https://crnt.osu.edu/LecturesOnReactionNetworks (1979)

  • Feinberg M (1995) The existence and uniqueness of steady states for a class of chemical reaction networks. Arch. Ration. Mech. Anal. 132(4):311–370

    Article  MathSciNet  MATH  Google Scholar 

  • Gunawardena, J.: Chemical reaction network theory for in-silico biologists. Notes available for download at http://vcp.med.harvard.edu/papers/crnt (2003)

  • Himeoka Y, Kaneko K (2014) Entropy production of a steady-growth cell with catalytic reactions. Phys. Rev. E 90(4):042,714

    Article  Google Scholar 

  • Hohmann S (2002) Osmotic stress signaling and osmoadaptation in yeasts. Microbiol. Mol. Biol. Rev. 66(2):300–372

    Article  Google Scholar 

  • Horn F (1972) Necessary and sufficient conditions for complex balancing in chemical kinetics. Arch. Ration. Mech. Anal. 49(3):172–186

    Article  MathSciNet  Google Scholar 

  • Jahreis K, Pimentel-Schmitt EF, Brückner R, Titgemeyer F (2008) Ins and outs of glucose transport systems in eubacteria. FEMS Microbiol. Rev. 32(6):891–907

    Article  Google Scholar 

  • Jensen RH, Woolfolk CA (1985) Formation of filaments by Pseudomonas putida. Appl. Environ. Microbiol. 50(2):364–372

    Google Scholar 

  • Kondo Y, Kaneko K (2011) Growth states of catalytic reaction networks exhibiting energy metabolism. Phys. Rev. E 84(1):011,927

    Article  Google Scholar 

  • Mavelli F, Ruiz-Mirazo K (2013) Theoretical conditions for the stationary reproduction of model protocells. Integr. Biol. 5(2):324–341

    Article  Google Scholar 

  • Molenaar D, van Berlo R, de Ridder D, Teusink B (2009) Shifts in growth strategies reflect tradeoffs in cellular economics. Mol. Syst. Biol. 5(1):323

    Google Scholar 

  • Morgan JJ, Surovtsev IV, Lindahl PA (2004) A framework for whole-cell mathematical modeling. J. Theor. Biol. 231(4):581–596

    Article  MathSciNet  Google Scholar 

  • Murata T (1989) Petri nets: properties, analysis and applications. Proc. IEEE 77(4):541–580

    Article  Google Scholar 

  • Olson AL, Pessin JE (1996) Structure, function, and regulation of the mammalian facilitative glucose transporter gene family. Annu. Rev. Nutr. 16(1):235–256

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Pawłowski PH, Zielenkiewicz P (2004) Biochemical kinetics in changing volumes. Acta Biochim. Pol. 51:231–243

    Google Scholar 

  • Richeson D, Wiseman J et al (2002) A fixed point theorem for bounded dynamical systems. Ill. J. Math. 46(2):491–495

    MathSciNet  MATH  Google Scholar 

  • Schaechter M, Maaløe O, Kjeldgaard NO (1958) Dependency on medium and temperature of cell size and chemical composition during balanced growth of Salmonella typhimurium. Microbiology 19(3):592–606

    Google Scholar 

  • Schilling CH, Letscher D, Palsson BØ (2000) Theory for the systemic definition of metabolic pathways and their use in interpreting metabolic function from a pathway-oriented perspective. J. Theor. Biol. 203(3):229–248

    Article  Google Scholar 

  • Schlosser PM, Feinberg M (1994) A theory of multiple steady states in isothermal homogeneous CFSTRs with many reactions. Chem. Eng. Sci. 49(11):1749–1767

    Article  Google Scholar 

  • Schuster S, Hilgetag C (1994) On elementary flux modes in biochemical reaction systems at steady state. J. Biol. Syst. 2(02):165–182

    Article  Google Scholar 

  • Schuster S, Höfer T (1991) Determining all extreme semi-positive conservation relations in chemical reaction systems: a test criterion for conservativity. J. Chem. Soc. Faraday Trans. 87(16):2561–2566

    Article  Google Scholar 

  • Scott M, Gunderson CW, Mateescu EM, Zhang Z, Hwa T (2010) Interdependence of cell growth and gene expression: origins and consequences. Science 330(6007):1099–1102

    Article  Google Scholar 

  • Stano P, Luisi PL (2010) Achievements and open questions in the self-reproduction of vesicles and synthetic minimal cells. Chem. Commun. 46(21):3639–3653

    Article  Google Scholar 

  • Surovstev IV, Morgan JJ, Lindahl PA (2007) Whole-cell modeling framework in which biochemical dynamics impact aspects of cellular geometry. J. Theor. Biol. 244(1):154–166

    Article  MathSciNet  Google Scholar 

  • Surovtsev IV, Zhang Z, Lindahl PA, Morgan JJ (2009) Mathematical modeling of a minimal protocell with coordinated growth and division. J. Theor. Biol. 260(3):422–429

    Article  MathSciNet  Google Scholar 

  • Tadmor AD, Tlusty T (2008) A coarse-grained biophysical model of E. coli and its application to perturbation of the rRNA operon copy number. PLoS Comput. Biol. 4(4):e1000,038

    Article  MathSciNet  Google Scholar 

  • Wei J (1962) Axiomatic treatment of chemical reaction systems. J. Chem. Phys. 36(6):1578–1584

    Article  Google Scholar 

  • Weiße AY, Oyarzún DA, Danos V, Swain PS (2015) Mechanistic links between cellular trade-offs, gene expression, and growth. Proc. Natl. Acad. Sci. 112(9):E1038–E1047

    Article  Google Scholar 

  • Wittmann C, Hans M, Van Winden WA, Ras C, Heijnen JJ (2005) Dynamics of intracellular metabolites of glycolysis and TCA cycle during cell-cycle-related oscillation in Saccharomyces cerevisiae. Biotechnol. Bioeng. 89(7):839–847

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to thank Pierre Legrain, Laurent Schwartz and Pierre Plateau for stimulating discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Erwan Bigan.

Appendices

Appendix 1: Analytical treatment of a simple example protocell with nutrient influx by active transport

Consider the same simple CRN example endowed with mass-action kinetics as described in Sect. 6, but where the first reaction is assumed unidirectional and only occurs across the membrane:

$$\begin{aligned}&A+B\rightarrow C\\&\quad 2B\rightleftarrows C \end{aligned}$$

A is the primary nutrient present in the outside growth medium, \(c_{\text {out},A}>0\) and is absent from the protocell, C is the transformed nutrient under activation by B. B and C are the only species present inside. C also plays the role of membrane precursor with incorporation kinetic coefficient \(K_{\text {me}}\) per unit area. We shall assume a constant surface area-to-volume ratio \(\rho _{0}\) so that the surfacic kinetic coefficient of the first surfacic reaction can be transformed in a volumic kinetic coefficient, \(k_{1}^{+}\). As in Sect. 6, the kinetic coefficients for the second reaction are \(k_{2}^{\pm }\).

The protocell ODE system is given by:

$$\begin{aligned} \dot{c_{B}}= & {} -k_{1}^{+}c_{\text {out},A}c_{B}-2k_{2}^{+}{c_{B}}^{2}+2k_{2}^{-}c_{C}-\mu _{\text {inst}}c_{B} \end{aligned}$$
(52)
$$\begin{aligned} \dot{c_{C}}= & {} k_{1}^{+}c_{\text {out},A}c_{B}+k_{2}^{+}{c_{B}}^{2}-k_{2}^{-}c_{C}-\rho _{0}K_{\text {me}}c_{C}-\mu _{\text {inst}}c_{C} \end{aligned}$$
(53)

with the growth rate \(\mu _{\text {inst}}\) given by:

$$\begin{aligned} \mu _{\text {inst}}=\frac{K_{\text {me}}c_{C}}{N_{\text {me}}} \end{aligned}$$
(54)

Looking for a fixed point, we set all time derivatives to zero. Adding Eqs. 52 and 53 with time derivatives set to zero results in the same equation as Eq. 32 which we restate here:

$$\begin{aligned} k_{2}^{+}{c_{B}}^{2}+\mu _{\text {inst}}(c_{B}+c_{C})=(k_{2}^{-}-\rho _{0}K_{\text {me}})c_{C} \end{aligned}$$
(55)

Replacing the growth rate \(\mu _{\text {inst}}\) by its expression given in Eq. 54 gives a quadratic equation that uniquely determines \(c_{B}\) as a function of \(c_{C}\) provided \(k_{2}^{-}>\rho _{0}K_{\text {me}}\) and \(c_{C}<N_{\text {me}}(\frac{k_{2}^{-}}{K_{\text {me}}}-\rho _{0})\):

$$\begin{aligned} c_{B}=\phi _{1}(c_{C})=\frac{-K_{\text {me}}c_{C}+\sqrt{(K_{\text {me}}c_{C})^{2}+4k_{2}^{+}c_{C}N_{\text {me}}((k_{2}^{-}-\rho _{0}K_{\text {me}})N_{\text {me}}-K_{\text {me}}c_{C})}}{2k_{2}^{+}N_{\text {me}}} \end{aligned}$$
(56)

It can be verified that \(\phi _{1}(c_{C})=0\) for \(c_{C}=0\) and for \(c_{C}=N_{\text {me}}(\frac{k_{2}^{-}}{K_{\text {me}}}-\rho _{0})\), that it is strictly positive for \(c_{C}\) in between those two zeros, and that its derivative is infinite at \(c_{C}=0\).

Adding Eqs. 52, 53 multiplied by two (thus obtaining the time derivative of the system density) and setting all time derivatives to zero, we obtain:

$$\begin{aligned} k_{1}^{+}c_{\text {out},A}c_{B}-2\rho _{0}K_{\text {me}}c_{C}-\mu _{\text {inst}}(c_{B}+2c_{C})=0 \end{aligned}$$
(57)

Replacing the growth rate \(\mu _{\text {inst}}\) by its expression given in Eq. 54 and rearranging terms gives the following expression for \(c_{B}\) as a function of \(c_{C}\):

$$\begin{aligned} c_{B}=\phi _{2}(c_{C})=\frac{2K_{\text {me}}c_{C}(\rho _{0}N_{\text {me}}+c_{C})}{k_{1}^{+}c_{\text {out},A}N_{\text {me}}-K_{\text {me}}c_{C}} \end{aligned}$$
(58)

It can be verified that \(\phi _{2}(c_{C})=0\) for \(c_{C}=0\), that \(\phi _{2}(c_{C})\rightarrow +\infty \) with \(c_{C}\rightarrow \frac{k_{1}^{+}c_{\text {out},A}N_{\text {me}}}{K_{\text {me}}}\), that it takes strictly positive values for \(c_{C}\) in between, and that its derivative is finite at \(c_{C}=0\).

Because at \(c_{C}=0\), the derivative of \(\phi _{1}\) is infinite and that of \(\phi _{2}\) is finite, \(\phi _{1}(c_{C})>\phi _{2}(c_{C})\) in the neighbourhood of \(c_{C}=0\). Besides, \(\phi _{2}(c_{C})>\phi _{1}(c_{C})\) in the neighbourhood of some strictly positive \(c_{C}\) value because of the following considerations:

  1. 1.

    If \(k_{1}^{+}c_{\text {out},A}>k_{2}^{-}-\rho _{0}K_{\text {me}}\), then \(\phi _{2}(c_{C})>\phi _{1}(c_{C})=0\) at \(c_{C}=N_{\text {me}}(\frac{k_{2}^{-}}{K_{\text {me}}}-\rho _{0})\).

  2. 2.

    Else if \(k_{1}^{+}c_{\text {out},A}\le k_{2}^{-}-\rho _{0}K_{\text {me}}\), then \(\phi _{2}(c_{C})\rightarrow +\infty \) and \(\phi _{1}(c_{C})\rightarrow \phi _{1}(N_{\text {me}}(\frac{k_{2}^{-}}{K_{\text {me}}}-\rho _{0}))\) (which is finite) when \(c_{C}\rightarrow N_{\text {me}}(\frac{k_{2}^{-}}{K_{\text {me}}}-\rho _{0})\).

This implies that \(c_{B}=\phi _{1}(c_{C})\) and \(c_{B}=\phi _{1}(c_{C})\) intersect for some strictly positive set of concentrations \(({c_{B}}_{0},{c_{C}}_{0})\).

We have thus proved the existence of a fixed point provided \(k_{2}^{-}>\rho _{0}K_{\text {me}}\). This simple example illustrates the fact that in the case of active transport, the protocell may or may not exhibit a fixed point (other than the degenerate zero concentration vector corresponding to an empty protocell) depending on kinetic and membrane parameters.

Appendix 2: Analytical treatment of a mechanistic, coarse-grained whole-cell model inspired by Molenaar et al. (2009)

This appendix proves analytically that the whole-cell model described in Sect. 6.2 exhibits a stationary growth state for any parameter set.

1.1 ODE system

The ODE system governing the cell dynamics is composed of:

  1. 1.

    Three differential equations for metabolites:

    $$\begin{aligned} \dot{c_{S}}= & {} \rho _{0}\alpha \phi _{\text {input},S}-\nu _{P}-\mu c_{S} \end{aligned}$$
    (59)
    $$\begin{aligned} \dot{c_{P}}= & {} \nu _{P}-\nu _{L}-n_{1}\nu _{1}-n_{2}\nu _{2}-n_{R}\nu _{R}-n_{T}\nu _{T}-\mu c_{P} \end{aligned}$$
    (60)
    $$\begin{aligned} \dot{c_{L}}= & {} \nu _{L}-\rho _{0}F_{\text {output},L}-\mu c_{L} \end{aligned}$$
    (61)
  2. 2.

    Four differential equations for proteins:

    $$\begin{aligned} \dot{c_{i}}=\nu _{i}-\mu c_{i} \end{aligned}$$
    (62)

    where \(i=1\), 2 or R, and:

    $$\begin{aligned} \dot{c_{T}}=\nu _{T}-\rho _{0}F_{\text {output},T}-\mu c_{T} \end{aligned}$$
    (63)
  3. 3.

    One differential equation for \(\alpha \) (Eq. 48)

1.2 Expressing all concentrations as functions of \(\alpha \) and of the normalized precursor concentration \(p=c_{P}\)/\(K_{\text {m}_{R}}\)

Looking for a stationary state, we set all time derivatives to zero. Combining Eqs. 41 and 62, both taken for \(i=R\), gives:

$$\begin{aligned} \mu =\frac{k_{{\text {cat}}_{R}}n_{R}p}{1+n_{\text {tot}}p} \end{aligned}$$
(64)

where \(p=c_{P}\)/\(K_{{\text {m}}_{R}}\) is the normalized precursor concentration. Combining Eqs. 41 and 62 for \(j=1\), 2 and R together gives:

$$\begin{aligned} c_{i}=\frac{n_{i}}{n_{R}}c_{R} \end{aligned}$$
(65)

where \(i=1\) or 2.

Combining Eqs. 41, 62 (taken for \(i=R\)) and 63 (taken for \(i=R\) and T) gives:

$$\begin{aligned} c_{R}=\frac{n_{R}}{n_{T}}c_{T}\left( 1+\rho _{0}\frac{K_{T}}{\mu }\right) \end{aligned}$$
(66)

Equation 48 for \(\alpha \) gives (excluding the degenerate case \(\alpha =1\) for which \(\mu =0\)):

$$\begin{aligned} \alpha =\frac{\frac{F_{\text {output},T}}{N_{T}}}{\frac{F_{\text {output},L}}{N_{L}}+\frac{F_{\text {output},T}}{N_{T}}} \end{aligned}$$
(67)

or equivalently:

$$\begin{aligned} 1-\alpha =\frac{\frac{F_{\text {output},L}}{N_{L}}}{\frac{F_{\text {output},L}}{N_{L}}+\frac{F_{\text {output},T}}{N_{T}}} \end{aligned}$$
(68)

We are now in a position to express all concentrations (except \(c_{S}\)) as a function of two unknowns, \(\alpha \) and the normalized precursor concentration p. For simplicity, we shall keep \(\mu \) in these expressions, knowing than it is a simple function of p given by Eq. 64. Combining the above equation with Eq. 46, gives:

$$\begin{aligned} \mu =\frac{F_{\text {output},L}}{N_{L}} \end{aligned}$$
(69)

and:

$$\begin{aligned} \frac{\alpha }{1-\alpha }\mu =\frac{F_{\text {output},T}}{N_{T}} \end{aligned}$$
(70)

From which \(c_{L}\) and \(c_{T}\) are extracted:

$$\begin{aligned} c_{L}= & {} \mu \frac{N_{L}}{K_{L}} \end{aligned}$$
(71)
$$\begin{aligned} c_{T}= & {} \mu \frac{\alpha }{1-\alpha }\frac{N_{T}}{K_{T}} \end{aligned}$$
(72)

Feeding Eq. 72 in Eq. 66, and feeding the result in Eq. 65, gives:

$$\begin{aligned} c_{i}=\frac{\alpha }{1-\alpha }\frac{n_{i}}{n_{T}}N_{T} \left( \rho _{0}+\frac{\mu }{K_{T}}\right) \end{aligned}$$
(73)

where \(i=1\), 2 or R. Finally, we recall that, by definition, \(c_{P}=p\times K_{{\text {m}}_{R}}\).

1.3 Relating \(\alpha \) to p

Setting the time derivative of \(c_{L}\) (given by Eq. 61) to zero, using Eq. 40, and rearranging terms gives:

$$\begin{aligned} \frac{k_{{\text {cat}}_{2}}c_{P}c_{2}}{K_{{\text {m}}_{2}}+c_{P}}=K_{L}\left( \rho _{0}+\frac{\mu }{K_{L}}\right) c_{L} \end{aligned}$$
(74)

Defining \(\theta =K_{{\text {m}}_{R}}\)/\(K_{{\text {m}}_{2}}\), using Eqs. 71 (for \(c_{L}\)), 73 (for \(c_{2}\)) and 64 (for \(\mu \)), and rearranging terms gives:

$$\begin{aligned} \frac{1-\alpha }{\alpha }=\frac{\frac{k_{{\text {cat}}_{2}}\theta }{1+\theta p}}{\frac{k_{{\text {cat}}_{R}}n_{R}}{1+n_{\text {tot}}p}}\frac{n_{2}}{n_{T}}\frac{N_{T}}{N_{L}}\frac{\rho _{0}+\frac{\mu }{K_{T}}}{\rho _{0}+\frac{\mu }{K_{L}}} \end{aligned}$$
(75)

Equation 75 defines a monotonic univocal relation between \(\alpha \) and p, such that \(0<\alpha (p=0)=\alpha _{0}, \,\alpha (p=+\infty )=\alpha _{\infty }<1\). \(\alpha _{0}\) may be lower or greater than \(\alpha _{\infty }\) depending on the specific choice of parameters.

1.4 Determining two independent relations of the kind \(c_{S}=\phi _{1,2}(p)\)

Setting the time derivative of \(c_{P}\) (Eq. 60) to zero, dividing by \(c_{R}\) and rearranging terms, gives:

$$\begin{aligned} \frac{k_{{\text {cat}}_{1}}c_{S}}{K_{{\text {m}}_{1}}+ c_{S}}\frac{n_{1}}{n_{T}}= & {} \frac{k_{{\text {cat}}_{2}}\theta p}{1+\theta p}\frac{n_{2}}{n_{T}}+\mu \frac{({n_{1}}^{2}+{n_{2}}^{2}+{n_{R}}^{2} +{n_{T}}^{2})n_{R}}{n_{T}}\nonumber \\&+ \,\mu \frac{ p K_{{\text {m}}_{R}}}{N_{T} \left( \rho _{0}+\frac{\mu }{K_{T}}\right) }\frac{1-\alpha }{\alpha } \end{aligned}$$
(76)

Combined with Eq. 75, this defines a function \(c_{S}=\phi _{1}(p)\) over the interval \([0;p_{\text {max},1}[\), where \(p_{\text {max},1}\) is the value for which the right-hand side (RHS) of Equation 76 verifies \(RHS(p_{\text {max},1})=k_{{\text {cat}}_{1}}n_{1}\)/\(n_{T}\). This function is such that \(\phi _{1}(p=0)=0\) (and this is the only zero of \(\phi _{1}\)), and \(\phi _{1}(p)\rightarrow +\infty \) if \(p\rightarrow p_{\text {max},1}\).

Setting the sum of the time derivatives of \(c_{S}\), \(c_{P}\) and \(c_{L}\) (sum of Eqs. 59, 60 and 61) to zero and rearranging terms gives:

$$\begin{aligned} \rho _{0}\alpha \phi _{\text {input},S}= & {} \mu c_{S}+\mu \frac{({n_{1}}^{2}+{n_{2}}^{2}+{n_{R}}^{2}+{n_{T}}^{2})}{n_{T}} N_{T}(\rho _{0}+\frac{\mu }{K_{T}})\frac{\alpha }{1-\alpha }\nonumber \\&+\,\mu p K_{{\text {m}}_{R}} + \mu \left( \rho _{0}+\frac{\mu }{K_{L}}\right) N_{L} \end{aligned}$$
(77)

Combined with Eq. 75, this defines a function \(c_{S}=\phi _{2}(p)\). Considering in turn the two different cases for nutrient influx:

  1. 1.

    Facilitated diffusion: \(\phi _{\text {input},S}\) is the function of \(c_{\text {out},S}\) and of \(c_{S}\) given by Eq. 50. If \(p=0\), then the RHS of Eq. 77 is null, and the LHS is also null iff \(c_{S}=c_{\text {out},S}\). We thus have \(\phi _{2}(p=0)=c_{\text {out},S}\). If \(p\rightarrow +\infty \), then \(RHS\rightarrow +\infty \) while the LHS cannot exceed a maximum value \(\rho _{0}\alpha _{\infty }\mathcal {D}c_{\text {out},S}\) which is reached if \(c_{S}=0\). Therefore, there must exist \(p_{\text {max},2}\) such that \(\phi _{2}(p_{\text {max},2})=0\).

  2. 2.

    Active transport, \(\phi _{\text {input},S}\) is only a function of \(c_{\text {out},S}\) and is thus a constant with respect to the internal state variables. It can be verified that \(\phi _{2}(p)\) is a decreasing function of p such that \(\phi _{2}(p=0)=+\infty \) and \(\phi _{2}(p=p_{\text {max},2})=0\) where \(p_{\text {max},2}\) is defined by letting \(c_{S}=0\) in Eq. 77.

In any of the two above cases, we have \(\phi _{2}(p=0)>\phi _{1}(p=0)\) and \(\phi _{2}(p=p_{\text {max}})<\phi _{1}(p=p_{\text {max}})\) where \(p_{\text {max}}=\text {min}\{p_{\text {max},1},p_{\text {max},2}\}\). This implies that \(\phi _{1}(p)\) and \(\phi _{2}(p)\) thus intersect for a finite \(p_{\text {st}}\) which fully defines a stationary growth state.

In the second above case, it is to be expected that if active transport had been more rigorously modeled by taking into account the activator introduced in Sect. 7.2 (equivalent to an energy currency, which would be consumed or down-converted by the active transport process), a stationary growth state might only exist provided the rate of synthesis of this energy currency is sufficient. This would be a situation similar to that described by the simple example of Appendix 1.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bigan, E., Paulevé, L., Steyaert, JM. et al. Necessary and sufficient conditions for protocell growth. J. Math. Biol. 73, 1627–1664 (2016). https://doi.org/10.1007/s00285-016-0998-0

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-016-0998-0

Keywords

Mathematics Subject Classification

Navigation