Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: arydshln
  • failed: chemfig
  • failed: mhchem

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2112.02809v4 [q-bio.PE] 02 Feb 2024
\definearrow

4<U=>\CF_arrowshiftnodes#4[allow upside down](\CF_arrowstartnode)–(\CF_arrowendnode)node[pos=0,yshift=1pt](\CF_arrowstartnodeu0)node[pos=0,yshift=-1pt](\CF_arrowstartnoded0)node[pos=1,yshift=1pt](\CF_arrowstartnodeu1)node[pos=1,yshift=-1pt](\CF_arrowstartnoded1);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnodeu0)–(\CF_arrowstartnodeu1)node[pos=0.4](Uarrowarctangent);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnoded1)–(\CF_arrowstartnoded0);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnodeu0)–(\CF_arrowstartnodeu1)node[pos=0.8,yshift=-9.5pt](UarrowarctangentR);\draw[semithick] (Uarrowarctangent) arc (270:190:.333) node (Uarrowend) ;\node[anchor=south,yshift=2pt] at (Uarrowend.north) #1; \draw[<-,>=stealth,semithick] (UarrowarctangentR) arc (0:90:.333) node (Uarrowend) ;\node[anchor=south,yshift=-22pt] at (Uarrowend.north) #2;

\definearrow

4<V=>\CF_arrowshiftnodes#4[allow upside down](\CF_arrowstartnode)–(\CF_arrowendnode)node[pos=0,yshift=1pt](\CF_arrowstartnodeu0)node[pos=0,yshift=-1pt](\CF_arrowstartnoded0)node[pos=1,yshift=1pt](\CF_arrowstartnodeu1)node[pos=1,yshift=-1pt](\CF_arrowstartnoded1);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnodeu0)–(\CF_arrowstartnodeu1)node[pos=0.4](Uarrowarctangent);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnoded1)–(\CF_arrowstartnoded0);\draw[\CF_arrowcurrentstyle](\CF_arrowstartnodeu0)–(\CF_arrowstartnodeu1)node[pos=0.8,yshift=-9.5pt](UarrowarctangentR);\draw[semithick] (Uarrowarctangent) arc (270:190:.333) node (Uarrowend) ;\node[anchor=south,yshift=2pt] at (Uarrowend.north) #1; \draw[<-,>=stealth,semithick] (UarrowarctangentR) ;

Thermodynamics of Darwinian evolution in molecular replicators

Artemy Kolchinsky artemyk@gmail.com ICREA-Complex Systems Lab, Universitat Pompeu Fabra, 08003 Barcelona, Spain Universal Biology Institute, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
Abstract

We consider the relationship between thermodynamics, fitness, and Darwinian evolution in autocatalytic molecular replicators. We uncover a thermodynamic bound that relates fitness, replication rate, and the Gibbs free energy dissipated per copy. This bound applies to a broad range of systems, including elementary and non-elementary autocatalytic reactions, polymer-based replicators, and certain kinds of autocatalytic sets. In addition, we show that the critical selection coefficient (the minimal fitness difference visible to selection) is bounded by the Gibbs free energy dissipated per replication. Our results imply fundamental thermodynamic bounds on the strength of selection in molecular evolution, complementary to other bounds that arise from finite population sizes and error thresholds. These bounds may be relevant for understanding thermodynamic constraints faced by early replicators at the origin of life. We illustrate our approach on several examples, including a classic model of replicators in a chemostat.

I Introduction

Recent work has uncovered fundamental bounds on the thermodynamic costs of various biomolecular processes, including chemical sensing (mehta2012energetic, ; barato_efficiency_2014, ; govern2014energy, ), copying of polymer-stored information (andrieuxNonequilibriumGenerationInformation2008, ; ouldridge2017fundamental, ; poultonNonequilibriumCorrelationsMinimal2019, ; sartoriKineticEnergeticDiscrimination2013, ), and growth and replication (kondo_growth_2011, ; england2013statistical, ; himeoka_entropy_2014, ; virgo2016complex, ; saakianNonlinearStochasticDynamics2016, ; bishopStochasticBistabilityBifurcation2010, ; pinero_nonequilibrium_2018, ; corominas2019thermodynamics, ). These results are derived from general principles of nonequilibrium thermodynamics — such as flux-force relations and fluctuation theorems (beard2007relationship, ; jarzynski_equalities_2011, ; seifert2012stochastic, ; kondepudiModernThermodynamicsHeat2015, ) — that relate the dynamical and thermodynamic properties of nonequilibrium processes. Due to their generality, these bounds shed light on universal thermodynamic constraints on lifelike systems, including modern and protobiological organisms, synthetic life, and even possible non-terrestrial lifeforms.

One of the most important properties of living systems is that they undergo Darwinian evolution. Generally speaking, Darwinian evolution refers to a process in which organisms with higher fitness are able to outcompete organisms with lower fitness. Darwinian evolution can be exhibited by chemical systems, such as individual replicating molecules or networks of molecules (eigenSelforganizationMatterEvolution1971, ; schusterDynamicsEvolutionaryOptimization1985, ; leeAutocatalyticNetworksTransition1997, ; feistelPhysicsSelfOrganization2011, ; vasasEvolutionGenes2012a, ; takeuchiEvolutionaryDynamicsRNAlike2012, ; vaidyaSpontaneousNetworkFormation2012, ; szilagyi2017ecology, ; peng2020ecological, ; ameta2021self, ; mizuuchiEvolutionaryTransitionSingle2022, ; ametaDarwinianPropertiesTheir2021, ). Synthesizing such systems has been a major focus of research on the origin of life, given that the emergence of Darwinian evolution is considered a crucial point in the transition from nonliving to living matter (bernstein1983darwinian, ; yarusGettingRNAWorld2011, ; vasasEvolutionGenes2012a, ; ameta2021self, ; jeancolasThresholdsOriginLife2020, ).

Here, we consider the relationship between thermodynamics and Darwinian evolution in minimal chemical systems. This relationship may be particularly relevant for understanding thermodynamic constraints on the origin of life (pascal2013towards, ; jeancolasThresholdsOriginLife2020, ; dangerConditionsMimickingNatural2020, ). We consider a nonequilibrium reaction volume which contains autocatalytic replicators. These replicators may copy themselves via elementary reactions, or via more complex multi-step replication mechanisms. We also consider certain types of collectively autocatalytic sets, where replication involves a uniform cycle of cross-catalytic reactions. Our setup includes many types of molecular replicators previously considered in the literature, including self-complementary and complementary templates, polymer-based replicators, and autocatalytic small molecules (patzke2007self, ; bissetteMechanismsAutocatalysis2013a, ; lincolnSelfsustainedReplicationRNA2009, ; leeSelfreplicatingPeptide1996, ). It also encompasses several classic models of molecular replicators, including the chemostat model (schusterDynamicsEvolutionaryOptimization1985, ; feistelPhysicsSelfOrganization2011, ) and Eigen’s quasispecies model (eigenSelforganizationMatterEvolution1971, ).

Each replicator is associated with three quantities. The first is the Gibbs free energy of replication, the free energy dissipated when one copy of the replicator is synthesized, which we write as σ𝜎\sigmaitalic_σ in dimensionless units of kBTsubscript𝑘𝐵𝑇k_{B}Titalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T-per-copy. The second is the replication rate ρ𝜌\rhoitalic_ρ, the number of copies per replicator and unit time under current conditions. Lastly, the fitness f𝑓fitalic_f is the maximum possible replication rate, as achieved in the limit of small concentrations. In Section IV.1, we show that fitness determines a replicator’s ability to invade a given population (metzHowShouldWe1992, ) and to sustain a high dilution rate. The introduction of this operational definition of fitness for molecular replicators is an important contribution of our work.

In Section IV.2, we derive a thermodynamic bound that relates Gibbs free energy of replication, the replication rate, and the fitness as

σln(1ρf).𝜎1𝜌𝑓\sigma\geq-\ln\left(1-\frac{\rho}{f}\right).italic_σ ≥ - roman_ln ( 1 - divide start_ARG italic_ρ end_ARG start_ARG italic_f end_ARG ) . (1)

Thus, the Gibbs free energy of replication increases as the replication rate ρ𝜌\rhoitalic_ρ approaches that f𝑓fitalic_f, the maximum possible. Equivalently said, the Gibbs free energy of replication limits how close the replication rate can approach the maximum.

We also derive a thermodynamic bound on the strength of Darwinian evolution. Observe that a higher fitness replicator is not always able to outcompete a lower fitness one, as this depends on the difference in fitness as well as various environmental and demographic factors (gillespie2004population, ). The strength of Darwinian evolution can be quantified in terms of the smallest fitness difference that can affect evolutionary outcomes in a given population and environment. This is the so-called critical selection coefficient s𝑠sitalic_s, which sets a “resolution limit” for selection beyond which fitness differences are indiscernible. As an example, the strength of Darwinian evolution in finite populations is limited by stochasticity, and a fitter mutant will fixate with significantly higher probability only when s1/Nemuch-greater-than𝑠1subscript𝑁𝑒s\gg 1/N_{e}italic_s ≫ 1 / italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, where Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the effective population size (ewensMathematicalPopulationGenetics2004, ). As another example, Eigen’s work on the “error threshold” shows that the strength of Darwinian evolution is limited by the mutation rate μ𝜇\muitalic_μ, so that a fitter mutant can dominate the population only when s>μ𝑠𝜇s>\muitalic_s > italic_μ (Eq. II-45, eigenSelforganizationMatterEvolution1971, ; smithMajorTransitionsEvolution, ).

We use the inequality (1) to derive a thermodynamic bound on the critical selection coefficient. We suppose that Darwinian evolution is sufficiently strong so that a replicator with fitness f𝑓fitalic_f is present in the steady state of a flow reactor, while another replicator type with fitness f<fsuperscript𝑓𝑓f^{\prime}<fitalic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < italic_f is driven to extinction. We show that s=1f/f𝑠1superscript𝑓𝑓s=1-f^{\prime}/fitalic_s = 1 - italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT / italic_f, the selection coefficient between the two replicators, obeys

seσ*,𝑠superscript𝑒superscript𝜎s\geq e^{-\sigma^{*}},italic_s ≥ italic_e start_POSTSUPERSCRIPT - italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , (2)

where σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is the Gibbs energy dissipated by the fitter replicator in steady state. This bound on the strength of Darwinian evolution applies even in the case of infinite populations and error-free replicators. It implies that selecting for a relative fitness difference of s𝑠sitalic_s requires dissipating more than lns𝑠-\ln s- roman_ln italic_s of free energy, a quantity which diverges for vanishing fitness differences, s0𝑠0s\to 0italic_s → 0. We illustrate this result using a classic model of replicators in a chemostat in Section VI.

In Section VII, we extend our results to autocatalytic sets where replication involves a cycle of cross-catalytic reactions. We first generalize our notion of fitness to autocatalytic sets and then derive generalized versions of the inequalities (1) and (2). In these generalizations, σ𝜎\sigmaitalic_σ refers to the Gibbs free energy dissipated by the average cross-catalytic reaction in the cycle.

Refer to caption
Figure 1: Illustration of our thermodynamic bound (2) for three real-world molecular replicators: a self-replicating prion (baskakovFoldingPrionProtein2001, ), an RNA molecule that copies itself via a single ligation (lincolnSelfsustainedReplicationRNA2009, ), and a peptide that copies itself via “native chemical ligation” (leeSelfreplicatingPeptide1996, ; dawsonSynthesisProteinsNative1994, ).
101101footnotetext: For the prion, we used ΔG=6.073.95=2.12Δsuperscript𝐺6.073.952.12-\Delta G^{\circ}=6.07-3.95=2.12- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT = 6.07 - 3.95 = 2.12 kcal/mol from (Table 1, baskakovFoldingPrionProtein2001, ), which gives ΔG/RT=3.5Δsuperscript𝐺𝑅𝑇3.5-\Delta G^{\circ}/RT=3.5- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = 3.5 at 298superscript298298^{\circ}298 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT K. Note that there is some debate in the literature regarding the precise mechanism of prion replication, such as whether it is first-order, like the replicators considered in this paper, or instead involves higher-order cooperative interactions (eigenPrionicsKineticBasis1996, ; laurentAutocatalyticProcessesCooperative1997a, ; laurentPrionDiseasesProtein1996, ; fematMechanismsPrionDisease2011, ).102102footnotetext: For the self-replicating peptide, we used ΔG=5.41.9=7.3Δsuperscript𝐺5.41.97.3-\Delta G^{\circ}=-5.4-1.9=7.3- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT = - 5.4 - 1.9 = 7.3 kcal/mol from (Fig. 4, wangTheoreticalAnalysisDetailed2011, ), which gives ΔG/RT=12.3Δsuperscript𝐺𝑅𝑇12.3-\Delta G^{\circ}/RT=12.3- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = 12.3 at 298superscript298298^{\circ}298 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT K. This was converted to σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT using Eq. 6.

In principle, our results can also be applied to various real-world molecular replicators, suggesting a route for experimental validation. In Fig. 1, we use published thermodynamic data to illustrate the bound (2) on three real world systems. The first is a self-replicating prion prusiner1991molecular , where σ*3.5superscript𝜎3.5\sigma^{*}\approx 3.5italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≈ 3.5 when native and misfolded concentrations are equal baskakovFoldingPrionProtein2001 ; Note101 . The second is a self-replicating peptide that copies itself using “native chemical ligation”  leeSelfreplicatingPeptide1996 ; wangTheoreticalAnalysisDetailed2011 ; Note102 , where σ*5.9superscript𝜎5.9\sigma^{*}\approx 5.9italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≈ 5.9 at substrates concentration 90μM90𝜇M90\,\mu\mathrm{M}90 italic_μ roman_M and replicator concentration 5μM5𝜇M5\,\mu\mathrm{M}5 italic_μ roman_M (leeSelfreplicatingPeptide1996, ). The third is a self-replicating RNA molecule that copies itself using a single RNA ligation (lincolnSelfsustainedReplicationRNA2009, ), with σ*5superscript𝜎5\sigma^{*}\approx 5italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≈ 5 under in vivo concentrations (julicherMotionRNAPolymerase1998, ; erieSingleNucleotideAdditionCycle1992, ). For example, our result predicts that for the RNA replicator, selection can only discern relative fitness differences of se50.6%𝑠superscript𝑒5percent0.6s\geq e^{-5}\approx 0.6\%italic_s ≥ italic_e start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT ≈ 0.6 %.

II Setup

We consider a reaction volume at constant temperature and pressure. The reaction volume contains an ideal well-mixed solution of replicators X,X,𝑋superscript𝑋X,X^{\prime},\dotsitalic_X , italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , … and other chemical species A1,A2,subscript𝐴1subscript𝐴2A_{1},A_{2},\dotsitalic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … that may serve as substrates and side products. We study this system in terms of deterministic concentrations, assuming that molecular counts are sufficiently large so that stochastic fluctuations can be ignored. We use x𝑥xitalic_x to indicate the concentration of replicator X𝑋Xitalic_X and a=(a1,a2,)𝑎subscript𝑎1subscript𝑎2\vec{a}=(a_{1},a_{2},\dots)over→ start_ARG italic_a end_ARG = ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … ) to indicate the concentrations of substrates/side products (A1,A2,)subscript𝐴1subscript𝐴2(A_{1},A_{2},\dots)( italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … ).

Each replicator X𝑋Xitalic_X undergoes a reversible autocatalytic reaction of the form

X+i=1αiAiX+X+i=1βiAi,𝑋subscript𝑖1subscript𝛼𝑖subscript𝐴𝑖𝑋𝑋subscript𝑖1subscript𝛽𝑖subscript𝐴𝑖X+\sum_{i=1}\alpha_{i}A_{i}\rightleftharpoons X+X+\sum_{i=1}\beta_{i}A_{i},italic_X + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⇌ italic_X + italic_X + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , (3)

where αisubscript𝛼𝑖\alpha_{i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are stoichiometric coefficients of substrate/side products Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. A simple example of Eq. 3 is autocatalysis from a single substrate, X+AX+X𝑋𝐴𝑋𝑋X+A\rightleftharpoons X+Xitalic_X + italic_A ⇌ italic_X + italic_X, but many other schemes are also possible. The autocatalytic reaction may be elementary, or it may proceed via multiple steps. Different types of replicators will generally have different stoichiometric coefficients αi,βisubscript𝛼𝑖subscript𝛽𝑖\alpha_{i},\beta_{i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as well as other thermodynamic and kinetic parameters (discussed below).

We use 𝒥𝒥\mathcal{J}caligraphic_J to indicate the net flux across the autocatalytic reaction in Eq. 3. We define the replication rate ρ𝜌\rhoitalic_ρ of replicator X𝑋Xitalic_X as the net flux per replicator,

ρ:=𝒥x.assign𝜌𝒥𝑥\rho:=\frac{\mathcal{J}}{x}.italic_ρ := divide start_ARG caligraphic_J end_ARG start_ARG italic_x end_ARG . (4)

The replicators may also flow out of the volume with dilution rate ϕ0italic-ϕ0\phi\geq 0italic_ϕ ≥ 0. Accounting for both replication and dilution, the concentration of replicator X𝑋Xitalic_X changes as

x˙=𝒥ϕx=(ρϕ)x.˙𝑥𝒥italic-ϕ𝑥𝜌italic-ϕ𝑥\dot{x}=\mathcal{J}-\phi x=(\rho-\phi)x.over˙ start_ARG italic_x end_ARG = caligraphic_J - italic_ϕ italic_x = ( italic_ρ - italic_ϕ ) italic_x . (5)

We usually leave dependence on time t𝑡titalic_t implicit in our notation. We will use that in steady state, any non-extinct replicator (x>0𝑥0x>0italic_x > 0) must have ρ=ϕ𝜌italic-ϕ\rho=\phiitalic_ρ = italic_ϕ, meaning that dilution and replication balance.

In writing Eq. 5, we assume that the rate of non-autocatalytic formation of the replicator is negligible. We also assume that different replicators do not interact directly by consuming each other as substrates or producing each other as side products, although they may interact indirectly via shared substrates/side products Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. For simplicity, we also ignore the spontaneous degradation of replicators; however, in Appendix A, we show that our results still hold in the presence of degradation reactions.

Some of our results are valid for closed reactors as well as open reactors that exchange matter with their environment (avanzini2022thermodynamics, ). However, our thermodynamic bound on selection (2) applies specifically to a flow reactor in steady state. Different types of flow reactors may be considered. One example is the continuous stirred-tank reactor (CSTR) where the dilution rate and inflow rates are constant, which is often used in chemical (filisettiStochasticModelEmergence2011a, ; semenovAutocatalyticBistableOscillatory2016, ) and biological experiments (dykhuizen1983selection, ; groverResourceCompetition1997, ; hoskisson2005continuous, ), and which can also arise naturally (e.g., in a pond fed by a nutrient-rich stream). To avoid confusion, we note that the CSTR is called a chemostat in the biological literature (harmandChemostat2017, ; smithTheoryChemostatDynamics1995, ), although in nonequilibrium thermodynamics, the term chemostat is usually used instead to refer to an external chemical reservoir (polettini2014irreversible, ). Another possible type of flow reactor is one where the rates of dilution and inflow can vary as a function of the chemical concentrations. An example is provided by Eigen’s quasispecies model, where the dilution rate is adjusted to maintain the total concentration of replicators constant (eigenSelforganizationMatterEvolution1971, ).

The Gibbs free energy of the replication reaction (3) is (kondepudiModernThermodynamicsHeat2015, ; raoNonequilibriumThermodynamicsChemical2016, )

σ=lnx+i=1(αiβi)lnaiΔG/RT,𝜎𝑥subscript𝑖1subscript𝛼𝑖subscript𝛽𝑖subscript𝑎𝑖Δsuperscript𝐺𝑅𝑇\sigma=-\ln x+\sum_{i=1}(\alpha_{i}-\beta_{i})\ln a_{i}-\Delta G^{\circ}/RT,italic_σ = - roman_ln italic_x + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_ln italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T , (6)

where R𝑅Ritalic_R is the gas constant, T𝑇Titalic_T is the temperature, and ΔGΔsuperscript𝐺-\Delta G^{\circ}- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT is the standard Gibbs free energy,

ΔG=RT[lnxeq+i=1(αiβi)lnaieq].Δsuperscript𝐺𝑅𝑇delimited-[]superscript𝑥eqsubscript𝑖1subscript𝛼𝑖subscript𝛽𝑖superscriptsubscript𝑎𝑖eq-\Delta G^{\circ}=RT\Big{[}-\ln x^{\text{eq}}+\sum_{i=1}(\alpha_{i}-\beta_{i})% \ln a_{i}^{\text{eq}}\Big{]}.- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT = italic_R italic_T [ - roman_ln italic_x start_POSTSUPERSCRIPT eq end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_ln italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT eq end_POSTSUPERSCRIPT ] . (7)

Here xeqsuperscript𝑥eqx^{\mathrm{eq}}italic_x start_POSTSUPERSCRIPT roman_eq end_POSTSUPERSCRIPT and aieqsuperscriptsubscript𝑎𝑖eqa_{i}^{\mathrm{eq}}italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_eq end_POSTSUPERSCRIPT are equilibrium concentrations of X𝑋Xitalic_X and Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, as would be reached if the reactor was closed to exchange of matter and allowed to relax completely. We note that we express σ𝜎\sigmaitalic_σ in dimensionless units of entropy per copy. It can be converted to joules per copy as kBTσsubscript𝑘𝐵𝑇𝜎k_{B}T\sigmaitalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_σ, where kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is Boltzmann’s constant, or to joules per mole as RTσ𝑅𝑇𝜎RT\sigmaitalic_R italic_T italic_σ. In the chemistry literature, the Gibbs free energy of reaction is often written as ΔGΔ𝐺-\Delta G- roman_Δ italic_G. The Gibbs free energy of replication can be understood as a fundamental thermodynamic cost that represents lost work potential: a replication reaction that dissipates σ𝜎\sigmaitalic_σ of free energy can be coupled to a thermodynamically disfavored “uphill” reaction and thereby perform up to σ𝜎\sigmaitalic_σ of chemical work.

III Elementary and non-elementary replicators

III.1 Elementary replicators

An elementary replicator refers to the case where Eq. 3 is an elementary reaction. Then, the net flux across the reaction has the mass-action form (kondepudiModernThermodynamicsHeat2015, )

𝒥=rxrx2,𝒥𝑟𝑥superscript𝑟superscript𝑥2\mathcal{J}=rx-r^{-}x^{2},caligraphic_J = italic_r italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (8)

where the forward and backward (pseudo) rate constants are

r=κiaiαi,r=κeΔG/RTiaiβi,formulae-sequence𝑟𝜅subscriptproduct𝑖superscriptsubscript𝑎𝑖subscript𝛼𝑖superscript𝑟𝜅superscript𝑒Δsuperscript𝐺𝑅𝑇subscriptproduct𝑖superscriptsubscript𝑎𝑖subscript𝛽𝑖r=\kappa\prod_{i}a_{i}^{\alpha_{i}},\qquad r^{-}=\kappa e^{\Delta G^{\circ}/RT% }\prod_{i}a_{i}^{\beta_{i}},italic_r = italic_κ ∏ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = italic_κ italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (9)

and κ𝜅\kappaitalic_κ is a baseline rate constant. Note that r𝑟ritalic_r and rsuperscript𝑟r^{-}italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT depend on the concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG, although we always leave this dependence implicit in our notation.

The log-ratio of the forward and backward fluxes in Eq. 8 equals the Gibbs free energy of replication,

σ=lnrxrx2=lnrrx,𝜎𝑟𝑥superscript𝑟superscript𝑥2𝑟superscript𝑟𝑥\sigma=\ln\frac{rx}{r^{-}x^{2}}=\ln\frac{r}{r^{-}x}\,,italic_σ = roman_ln divide start_ARG italic_r italic_x end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = roman_ln divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x end_ARG , (10)

as can be verified by comparing Eqs. 9 and 6. Eq. 10 is an instance of a general result, called the flux-force relation or local detailed balance in the literature (wachtelThermodynamicallyConsistentCoarse2018, ; raoNonequilibriumThermodynamicsChemical2016, ), which relates the Gibbs free energy of an elementary reaction with the forward and reverse fluxes (kondepudiModernThermodynamicsHeat2015, ). The flux-force relation is one of the most important results in nonequilibrium thermodynamics, since it connects the kinetic properties of a chemical reaction with its thermodynamic properties.

Actually, our results will not require that the rate constants r𝑟ritalic_r and rsuperscript𝑟r^{-}italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT have the mass action form of Eq. 9, only that Eq. 8 and Eq. 10 hold. This could be used to study certain non-ideal solutions that exhibit interactions between Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (avanzini2021nonequilibrium, ).

III.2 Non-elementary replicators

Most real-world autocatalytic replicators cannot be treated as elementary reactions. For this reason, we also consider the case where Eq. 3 represents a reaction mechanism that proceeds via a sequence of m𝑚mitalic_m steps,

\setchemfigarrowcoeff=0.6\schemestartX\arrow<U=>[iα1,iAi][iβ1,iAi]Y1\arrow<U=>[iα2,iAi][iβ2,iAi]\arrow<U=>[][]Ym1\arrow<U=>[iαm,iAi][iβm,iAi]X+X\schemestop.\setchemfig{arrowcoeff=0.6}\schemestart X\arrow{<U=>[{\footnotesize\sum_{i}% \alpha_{1,i}A_{i}}][{\footnotesize\sum_{i}\beta_{1,i}A_{i}}]}Y_{1}\arrow{<U=>[% {\footnotesize\sum_{i}\alpha_{2,i}A_{i}}][{\footnotesize\sum_{i}\beta_{2,i}A_{% i}}]}\dots\arrow{<U=>[\cdots][\quad\quad\cdots]}Y_{m-1}\arrow{<U=>[{% \footnotesize\sum_{i}\alpha_{m,i}A_{i}}][{\footnotesize\sum_{i}\beta_{m,i}A_{i% }}]}X+X\schemestop.italic_a italic_r italic_r italic_o italic_w italic_c italic_o italic_e italic_f italic_f = 0.6 italic_X < italic_U = > [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT 1 , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 1 , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] italic_Y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_U = > [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT 2 , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 2 , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] … < italic_U = > [ ⋯ ] [ ⋯ ] italic_Y start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT < italic_U = > [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] italic_X + italic_X . (11)

Each Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is an intermediate chemical species, and each intermediate step is an elementary reversible reaction with mass-action kinetics that may involve substrates/side products Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT with stoichiometric coefficients αk,isubscript𝛼𝑘𝑖\alpha_{k,i}italic_α start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT and βk,isubscript𝛽𝑘𝑖\beta_{k,i}italic_β start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT. The stoichiometry of the overall reaction obeys αi=kαk,isubscript𝛼𝑖subscript𝑘subscript𝛼𝑘𝑖\alpha_{i}=\sum_{k}\alpha_{k,i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT and βi=kβk,isubscript𝛽𝑖subscript𝑘subscript𝛽𝑘𝑖\beta_{i}=\sum_{k}\beta_{k,i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT. We assume that intermediate species are not shared between different types of replicators. For simplicity, we also assumed that degradation reactions are negligible, although our results generalize to the presence of degradation as shown Appendix A.

Refer to caption
Figure 2: Examples of non-elementary autocatalytic replication mechanisms. Left: autocatalysis with binding, conversion, and unbinding steps. Right: templated replication of a self-complementary polymer (shown here using a dimer).

We term this kind of reaction mechanism a non-elementary replicator, although it is also called an “autocatalytic cycle” in the literature (kingAutocatalysis1978, ; virgo2016complex, ; hordijkAutocatalyticConfusionClarified2017, ). A simple example of a non-elementary replicator is a three-step mechanism with binding, conversion, and unbinding steps, see Fig. 2 (left). Another example is the step-by-step replication of a self-complementary dimer, illustrated in Fig. 2 (right), studied in numerous origin-of-life experiments (von1986self, ; patzke2007self, ; bissetteMechanismsAutocatalysis2013a, ). Yet other examples include the formose cycle (bissetteMechanismsAutocatalysis2013a, ) and the autophosphorylation of protein kinase (wangAutophosphorylationKineticsProtein2002, ; bishopStochasticBistabilityBifurcation2010, ).

Since each intermediate reaction k{1,,m}𝑘1𝑚k\in\{1,\dots,m\}italic_k ∈ { 1 , … , italic_m } is elementary with mass-action kinetics, the net flux across it is

Jksubscript𝐽𝑘\displaystyle J_{k}italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT =νkyk1νkyk.absentsubscript𝜈𝑘subscript𝑦𝑘1superscriptsubscript𝜈𝑘subscript𝑦𝑘\displaystyle=\nu_{k}y_{k-1}-\nu_{k}^{-}y_{k}.= italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT - italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (12)

Here yksubscript𝑦𝑘y_{k}italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the concentration of intermediate species Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT for k{1,,m1}𝑘1𝑚1k\in\{1,\dots,m-1\}italic_k ∈ { 1 , … , italic_m - 1 }, and we use the convention y0=x,ym=x2formulae-sequencesubscript𝑦0𝑥subscript𝑦𝑚superscript𝑥2y_{0}=x,y_{m}=x^{2}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_x , italic_y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The terms νksubscript𝜈𝑘\nu_{k}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and νksuperscriptsubscript𝜈𝑘\nu_{k}^{-}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT refer to forward and backward (pseudo) rates constants of intermediate steps, typically defined in a manner analogous to Eq. 9. We note that νksubscript𝜈𝑘\nu_{k}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and νksuperscriptsubscript𝜈𝑘\nu_{k}^{-}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT can depend on the concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG, although we leave this implicit in our notation.

The production of replicator X𝑋Xitalic_X due to the autocatalytic mechanism is

𝒥=2JmJ1,𝒥2subscript𝐽𝑚subscript𝐽1\mathcal{J}=2J_{m}-J_{1},caligraphic_J = 2 italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , (13)

since two copies are produced by the last step and one consumed by the first. The production of intermediate species Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (k{1,,m1}𝑘1𝑚1k\in\{1,\dots,m-1\}italic_k ∈ { 1 , … , italic_m - 1 }) due to the mechanism is JkJk+1subscript𝐽𝑘subscript𝐽𝑘1J_{k}-J_{k+1}italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT, since one Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is produced by intermediate step k𝑘kitalic_k and one is consumed by intermediate step k+1𝑘1k+1italic_k + 1. The rate of change of the concentration of the intermediate species, also accounting for dilution with rate ϕ0italic-ϕ0\phi\geq 0italic_ϕ ≥ 0, is

y˙k=JkJk+1ϕyk.subscript˙𝑦𝑘subscript𝐽𝑘subscript𝐽𝑘1italic-ϕsubscript𝑦𝑘\dot{y}_{k}=J_{k}-J_{k+1}-\phi y_{k}.over˙ start_ARG italic_y end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT - italic_ϕ italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (14)

It will be useful to express the overall flux 𝒥𝒥\mathcal{J}caligraphic_J in a mass-action-like form of Eq. 8. To do so, we introduce the assumption that the relative concentrations of intermediate species Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and replicator X𝑋Xitalic_X is approximately constant,

ddtykx=xy˙kx˙ykx20.𝑑𝑑𝑡subscript𝑦𝑘𝑥𝑥subscript˙𝑦𝑘˙𝑥subscript𝑦𝑘superscript𝑥20\frac{d}{dt}\frac{y_{k}}{x}=\frac{x\dot{y}_{k}-\dot{x}y_{k}}{x^{2}}\approx 0.divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG divide start_ARG italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_x end_ARG = divide start_ARG italic_x over˙ start_ARG italic_y end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - over˙ start_ARG italic_x end_ARG italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ≈ 0 . (15)

This assumption always holds in steady state (y˙k=x˙=0subscript˙𝑦𝑘˙𝑥0\dot{y}_{k}=\dot{x}=0over˙ start_ARG italic_y end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = over˙ start_ARG italic_x end_ARG = 0). It also holds under a separation of timescales, when the relative concentrations yk/xsubscript𝑦𝑘𝑥y_{k}/xitalic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_x quickly reach stationary values — that is, faster than those stationary values are perturbed by changes in other variables like the dilution rate and the absolute concentrations of x𝑥xitalic_x and a𝑎\vec{a}over→ start_ARG italic_a end_ARG. This separation of timescales holds during the exponential growth phase of an initially rare replicator, when the absolute concentrations of x𝑥xitalic_x and yksubscript𝑦𝑘y_{k}italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are very small and have a minimal effect on other variables.

We emphasize that the assumption (15) is different from the quasi-steady-state (QSS) approximation often used in biochemistry (segel1989quasi, ). QSS assumes that the absolute concentrations of intermediate species are approximately stationary, compared to the rate of change of the replicator concentration x𝑥xitalic_x. The QSS assumption can be strongly violated in growing autocatalytic replicators, especially when the intermediate species have large relative concentrations.

Plugging x˙=(ρϕ)x˙𝑥𝜌italic-ϕ𝑥\dot{x}=(\rho-\phi)xover˙ start_ARG italic_x end_ARG = ( italic_ρ - italic_ϕ ) italic_x and y˙k=JkJk+1ϕyksubscript˙𝑦𝑘subscript𝐽𝑘subscript𝐽𝑘1italic-ϕsubscript𝑦𝑘\dot{y}_{k}=J_{k}-J_{k+1}-\phi y_{k}over˙ start_ARG italic_y end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT - italic_ϕ italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT into Eq. 15 and simplifying gives

JkJk+1=ρyk.subscript𝐽𝑘subscript𝐽𝑘1𝜌subscript𝑦𝑘J_{k}-J_{k+1}=\rho y_{k}.italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT = italic_ρ italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (16)

Using Eq. 12, we can express the intermediate concentrations y=(y1,,ym1)𝑦subscript𝑦1subscript𝑦𝑚1\vec{y}=(y_{1},\dots,y_{m-1})over→ start_ARG italic_y end_ARG = ( italic_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_y start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT ) in terms of a linear system

[(M+ρI)y]k=δk,1ν1x+δk,m1νmx2,subscriptdelimited-[]𝑀𝜌𝐼𝑦𝑘subscript𝛿𝑘1subscript𝜈1𝑥subscript𝛿𝑘𝑚1superscriptsubscript𝜈𝑚superscript𝑥2[(M+\rho I)\vec{y}]_{k}=\delta_{k,1}\nu_{1}x+\delta_{k,m-1}\nu_{m}^{-}x^{2},[ ( italic_M + italic_ρ italic_I ) over→ start_ARG italic_y end_ARG ] start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_k , 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x + italic_δ start_POSTSUBSCRIPT italic_k , italic_m - 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (17)

where M(m1)×(m1)𝑀superscript𝑚1𝑚1M\in\mathbb{R}^{(m-1)\times(m-1)}italic_M ∈ blackboard_R start_POSTSUPERSCRIPT ( italic_m - 1 ) × ( italic_m - 1 ) end_POSTSUPERSCRIPT is a matrix of intermediate rate constants,

M=[ν1+ν2ν200ν2ν2+ν3ν300νm100νm1νm1+νm]𝑀delimited-[]superscriptsubscript𝜈1subscript𝜈2superscriptsubscript𝜈200subscript𝜈2superscriptsubscript𝜈2subscript𝜈3superscriptsubscript𝜈300superscriptsubscript𝜈𝑚100subscript𝜈𝑚1superscriptsubscript𝜈𝑚1subscript𝜈𝑚M\!=\!\left[\begin{array}[]{ccccc}\nu_{1}^{-}\!\!+\!\nu_{2}&-\nu_{2}^{-}&0&0&% \dots\\ -\nu_{2}&\nu_{2}^{-}\!\!+\!\nu_{3}&-\nu_{3}^{-}&0&\dots\\ 0&\dots&\dots&\dots&\nu_{m-1}^{-}\\ \dots&0&0&-\nu_{m-1}&\nu_{m-1}^{-}\!\!+\!\nu_{m}\end{array}\right]italic_M = [ start_ARRAY start_ROW start_CELL italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL - italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL … end_CELL end_ROW start_ROW start_CELL - italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + italic_ν start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL start_CELL - italic_ν start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_CELL start_CELL 0 end_CELL start_CELL … end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL … end_CELL start_CELL … end_CELL start_CELL … end_CELL start_CELL italic_ν start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL … end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - italic_ν start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT end_CELL start_CELL italic_ν start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + italic_ν start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_CELL end_ROW end_ARRAY ] (18)

For any ρ0𝜌0\rho\geq 0italic_ρ ≥ 0, M+ρI𝑀𝜌𝐼M+\rho Iitalic_M + italic_ρ italic_I is an “M-matrix” (shivakumar1996two, ), so it is invertible and all of entries of its inverse are nonnegative (shivakumar1996two, ; meyer1978singular, ).

We solve Eq. 17 for y𝑦\vec{y}over→ start_ARG italic_y end_ARG and combine with Eqs. 13 and 12 to write 𝒥𝒥\mathcal{J}caligraphic_J in a mass-action-like form,

𝒥=r(ρ)xr(ρ)x2,𝒥𝑟𝜌𝑥superscript𝑟𝜌superscript𝑥2\mathcal{J}=r(\rho)x-r^{-}(\rho)x^{2},caligraphic_J = italic_r ( italic_ρ ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (19)

with effective rate constants given by

r(ρ):=ν1(ν1G11+2νmGm1,11)r(ρ):=νm(22νmGm1,m1ν1G1,m1),𝑟𝜌assignabsentsubscript𝜈1superscriptsubscript𝜈1subscript𝐺112subscript𝜈𝑚subscript𝐺𝑚111superscript𝑟𝜌assignabsentsuperscriptsubscript𝜈𝑚22subscript𝜈𝑚subscript𝐺𝑚1𝑚1superscriptsubscript𝜈1subscript𝐺1𝑚1\displaystyle\begin{aligned} r(\rho)&:=\nu_{1}(\nu_{1}^{-}G_{11}+2\nu_{m}G_{m-% 1,1}-1)\\ r^{-}(\rho)&:=\nu_{m}^{-}(2-2\nu_{m}G_{m-1,m-1}-\nu_{1}^{-}G_{1,m-1})\,,\end{aligned}start_ROW start_CELL italic_r ( italic_ρ ) end_CELL start_CELL := italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT + 2 italic_ν start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_m - 1 , 1 end_POSTSUBSCRIPT - 1 ) end_CELL end_ROW start_ROW start_CELL italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) end_CELL start_CELL := italic_ν start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 2 - 2 italic_ν start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_m - 1 , italic_m - 1 end_POSTSUBSCRIPT - italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT 1 , italic_m - 1 end_POSTSUBSCRIPT ) , end_CELL end_ROW (20)

where G:=(M+ρI)1assign𝐺superscript𝑀𝜌𝐼1G:=(M+\rho I)^{-1}italic_G := ( italic_M + italic_ρ italic_I ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. These effective rate constants can depend on concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG, since νksubscript𝜈𝑘\nu_{k}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and νksuperscriptsubscript𝜈𝑘\nu_{k}^{-}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT depend on them, although we leave this dependence implicit in our notation.

Below we will use that the first and second derivatives of the effective rate constants obey

ρr(ρ)subscript𝜌𝑟𝜌\displaystyle\partial_{\rho}r(\rho)∂ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT italic_r ( italic_ρ ) 0,ρ2r(ρ)0formulae-sequenceabsent0superscriptsubscript𝜌2𝑟𝜌0\displaystyle\leq 0,\;\partial_{\rho}^{2}r(\rho)\geq 0≤ 0 , ∂ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r ( italic_ρ ) ≥ 0 (21)
ρr(ρ)subscript𝜌superscript𝑟𝜌\displaystyle\partial_{\rho}r^{-}(\rho)∂ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) 0,ρ2r(ρ)0formulae-sequenceabsent0superscriptsubscript𝜌2superscript𝑟𝜌0\displaystyle\geq 0,\;\partial_{\rho}^{2}r^{-}(\rho)\leq 0≥ 0 , ∂ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) ≤ 0

This follows from Eq. 20 by using matrix calculus and the fact that all entries of G=(M+ρI)1𝐺superscript𝑀𝜌𝐼1G=(M+\rho I)^{-1}italic_G = ( italic_M + italic_ρ italic_I ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT are nonnegative.

In addition, for ρ=0𝜌0\rho=0italic_ρ = 0, the effective rate constants can be expressed in closed form,

r(0)𝑟0\displaystyle r(0)italic_r ( 0 ) =[k=1ml=1k1νll=1kνl]1r(0)=r(0)k=1mνkνkformulae-sequenceabsentsuperscriptdelimited-[]superscriptsubscript𝑘1𝑚superscriptsubscriptproduct𝑙1𝑘1superscriptsubscript𝜈𝑙superscriptsubscriptproduct𝑙1𝑘subscript𝜈𝑙1superscript𝑟0𝑟0superscriptsubscriptproduct𝑘1𝑚superscriptsubscript𝜈𝑘subscript𝜈𝑘\displaystyle=\Bigg{[}\sum_{k=1}^{m}\frac{\prod_{l=1}^{k-1}\nu_{l}^{-}}{\prod_% {l=1}^{k}\nu_{l}}\Bigg{]}^{-1}\;\quad r^{-}(0)=r(0)\prod_{k=1}^{m}\frac{\nu_{k% }^{-}}{\nu_{k}}= [ ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT divide start_ARG ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_ARG start_ARG ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) = italic_r ( 0 ) ∏ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT divide start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG (22)

as shown in Appendix B. The ρ=0𝜌0\rho=0italic_ρ = 0 regime corresponds to the limit where the replication rate ρ𝜌\rhoitalic_ρ is much slower than the rate of internal reactions, so that ρ𝜌\rhoitalic_ρ can be neglected when solving the linear system (17).

We finish by discussing the thermodynamics of non-elementary replicators. The Gibbs free energy of intermediate reaction k𝑘kitalic_k in Eq. 11 is

σk=lnyk1yk+i=1(αk,iβk,i)lnaiΔGk/RT,subscript𝜎𝑘subscript𝑦𝑘1subscript𝑦𝑘subscript𝑖1subscript𝛼𝑘𝑖subscript𝛽𝑘𝑖subscript𝑎𝑖Δsuperscriptsubscript𝐺𝑘𝑅𝑇\sigma_{k}=\ln\frac{y_{k-1}}{y_{k}}+\sum_{i=1}(\alpha_{k,i}-\beta_{k,i})\ln a_% {i}-\Delta G_{k}^{\circ}/RT,italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = roman_ln divide start_ARG italic_y start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT ( italic_α start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT ) roman_ln italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T , (23)

where as above we use y0=x,ym=x2formulae-sequencesubscript𝑦0𝑥subscript𝑦𝑚superscript𝑥2y_{0}=x,y_{m}=x^{2}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_x , italic_y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Since intermediate reactions are elementary, the forward and backward fluxes in Eq. 12 obey the flux-force relation,

σk=lnνkyk1νkyk,subscript𝜎𝑘subscript𝜈𝑘subscript𝑦𝑘1superscriptsubscript𝜈𝑘subscript𝑦𝑘\sigma_{k}=\ln\frac{\nu_{k}y_{k-1}}{\nu_{k}^{-}y_{k}},italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = roman_ln divide start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG , (24)

as can be shown explicitly when the rate constants νksubscript𝜈𝑘\nu_{k}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and νksuperscriptsubscript𝜈𝑘\nu_{k}^{-}italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT have mass-action kinetics similar to Eq. 9.

The Gibbs free energy of the overall replication mechanism is the sum of the Gibbs free energy of the individual steps (kondepudiModernThermodynamicsHeat2015, ),

σ=k=1mσk=ln1x+k=1mlnνkνk.𝜎superscriptsubscript𝑘1𝑚subscript𝜎𝑘1𝑥superscriptsubscript𝑘1𝑚subscript𝜈𝑘superscriptsubscript𝜈𝑘\sigma=\sum_{k=1}^{m}\sigma_{k}=\ln\frac{1}{x}+\sum_{k=1}^{m}\ln\frac{\nu_{k}}% {\nu_{k}^{-}}.italic_σ = ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = roman_ln divide start_ARG 1 end_ARG start_ARG italic_x end_ARG + ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT roman_ln divide start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_ARG . (25)

We can express σ𝜎\sigmaitalic_σ in a “flux-force-like” expression by combining Eqs. 25 and 22,

σ=lnr(0)xr(0)=xr(0)x2r(0).𝜎𝑟0𝑥superscript𝑟0𝑥𝑟0superscript𝑥2superscript𝑟0\sigma=\ln\frac{r(0)}{xr^{-}(0)}=\frac{xr(0)}{x^{2}r^{-}(0)}.italic_σ = roman_ln divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_x italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG = divide start_ARG italic_x italic_r ( 0 ) end_ARG start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG . (26)

Comparing to Eq. 19, we see that the numerator and denominator can be interpreted as the forward and backward fluxes in the case of ρ=0𝜌0\rho=0italic_ρ = 0, that is the limit when dilution is much slower than internal reactions. Ref. (wachtelThermodynamicallyConsistentCoarse2018, ) previously derived this flux-force relation for steady state conditions and reaction mechanisms such as Eq. 11, which involve a single reaction cycle.

An elementary replicator is a special case of a non-elementary replicator with a single reaction (m=1𝑚1m=1italic_m = 1) and no intermediate species. In this case, the effective rate constants in Eq. 20 lose dependence on ρ𝜌\rhoitalic_ρ and reduce to Eq. 22 which in turn reduces to Eq. 9. The mass-action form of Eq. 19 reduces to Eq. 8, and the flux-force relation (26) is equivalent to Eq. 10.

IV Fitness, selection, and thermodynamics

IV.1 Fitness and selection coefficient

The concept of fitness can be defined differently in different evolutionary scenarios, and finding an appropriate definition is an important area of research in biology and ecology (metzHowShouldWe1992, ; roffDefiningFitnessEvolutionary2008, ; orrFitnessItsRole2009, ; spaakDifferentMeasuresNiche2023, ). Here, we propose a definition of fitness that is suitable for autocatalytic molecular replicators, both elementary and non-elementary.

Before proceeding, we note that one could simply define fitness as the replication rate ρ𝜌\rhoitalic_ρ at a given point in time. However, this approach runs into problems. For instance, for a reactor in steady state, all present replicators have the same replication rate, i.e., the steady-state dilution rate ϕitalic-ϕ\phiitalic_ϕ, while all non-present replicators have an undefined replication rate. This makes it impossible to ask many natural questions, such as whether higher fitness replicators outcompete lower fitness ones. More generally, replication rate ρ𝜌\rhoitalic_ρ is a statistic of actual performance that does not specify how a replicator would perform in a new environment — which is usually what is desired from a fitness measure (metzHowShouldWe1992, ; roffDefiningFitnessEvolutionary2008, ; orrFitnessItsRole2009, ; spaakDifferentMeasuresNiche2023, ).

Instead, we define the fitness f𝑓fitalic_f of replicator X𝑋Xitalic_X via the equation,

f=r(f)0,𝑓𝑟𝑓0f=r(f)\geq 0,italic_f = italic_r ( italic_f ) ≥ 0 , (27)

where r()𝑟r(\cdot)italic_r ( ⋅ ) is the forward rate rate constant defined in Eq. 20. For an elementary replicator, r(ρ)𝑟𝜌r(\rho)italic_r ( italic_ρ ) does not depend on ρ𝜌\rhoitalic_ρ and f𝑓fitalic_f is simply the rate constant of the elementary reaction in Eq. 9. For a non-elementary replicator, f𝑓fitalic_f is the nonnegative root of the algebraic expression αr(α)𝛼𝑟𝛼\alpha-r(\alpha)italic_α - italic_r ( italic_α ). This expression is strictly increasing in α𝛼\alphaitalic_α, ranging from 0r(0)00𝑟000-r(0)\leq 00 - italic_r ( 0 ) ≤ 0 to r(0)r(r(0))0𝑟0𝑟𝑟00r(0)-r(r(0))\geq 0italic_r ( 0 ) - italic_r ( italic_r ( 0 ) ) ≥ 0, as can be deduced from Eqs. 22 and 21. There is a unique value of 0fr(0)0𝑓𝑟00\leq f\leq r(0)0 ≤ italic_f ≤ italic_r ( 0 ) that satisfies Eq. 27 and that can be found quickly using numerical methods such as bisection.

We note that fitness depends on the concentrations of substrates and side products a𝑎\vec{a}over→ start_ARG italic_a end_ARG, although this is left implicit in our notation. These concentrations specify the “ecological environment” of the reaction volume.

Our definition of fitness is experimentally measurable and operationally meaningful. In particular, f𝑓fitalic_f can be understood as the initial growth rate that is achieved at small concentrations. Imagine a reactor in steady state that does not contain replicator X𝑋Xitalic_X at time t<0𝑡0t<0italic_t < 0, and suppose that X𝑋Xitalic_X is introduced at a small concentration x(0)𝑥0x(0)italic_x ( 0 ) at t=0𝑡0t=0italic_t = 0. Suppose also that X𝑋Xitalic_X is an elementary replicator or a non-elementary replicator that obeys Eq. 15, so that relative concentrations of intermediates are approximately stationary. Assuming for simplicity that there is no dilution or degradation, the replicator’s concentration will initially grow as

x˙=𝒥=ρxr(ρ)x,˙𝑥𝒥𝜌𝑥𝑟𝜌𝑥\dot{x}=\mathcal{J}=\rho x\approx r(\rho)x,over˙ start_ARG italic_x end_ARG = caligraphic_J = italic_ρ italic_x ≈ italic_r ( italic_ρ ) italic_x ,

as follows from Eqs. 4 and 19 and dropping second order terms in x𝑥xitalic_x. This equation ρ=r(ρ)𝜌𝑟𝜌\rho=r(\rho)italic_ρ = italic_r ( italic_ρ ) is uniquely satisfied by ρ=f𝜌𝑓\rho=fitalic_ρ = italic_f, meaning that the concentration will grow as

x(t)etfx(0)𝑥𝑡superscript𝑒𝑡𝑓𝑥0x(t)\approx e^{tf}x(0)italic_x ( italic_t ) ≈ italic_e start_POSTSUPERSCRIPT italic_t italic_f end_POSTSUPERSCRIPT italic_x ( 0 ) (28)

for small t𝑡titalic_t. Thus, f𝑓fitalic_f has a clear operational interpretation: an initially rare mutant with fitness f𝑓fitalic_f will increase in concentration if f>0𝑓0f>0italic_f > 0 and decrease toward extinction if f<0𝑓0f<0italic_f < 0. The same derivation holds for a flow reactor with dilution rate ϕitalic-ϕ\phiitalic_ϕ, in which case it gives x(t)et(fϕ)x(0)𝑥𝑡superscript𝑒𝑡𝑓italic-ϕ𝑥0x(t)\approx e^{t(f-\phi)}x(0)italic_x ( italic_t ) ≈ italic_e start_POSTSUPERSCRIPT italic_t ( italic_f - italic_ϕ ) end_POSTSUPERSCRIPT italic_x ( 0 ). In biology, this type of fitness measure is called invasion fitness (liModelingMicrobialMetabolic2020a, ; arnoldiInvasionsEcologicalCommunities2022a, ; metzHowShouldWe1992, ; spaakDifferentMeasuresNiche2023, ). It has been argued to be a particularly general way to operationalize fitness (metzHowShouldWe1992, ; roffDefiningFitnessEvolutionary2008, ).

The initial growth rate is experimentally accessible as long as the initial concentration x(0)𝑥0x(0)italic_x ( 0 ) is sufficiently small so that one can neglect the reverse flux (second term in Eq. 19) as well as perturbations to steady-state values of a𝑎\vec{a}over→ start_ARG italic_a end_ARG and ϕitalic-ϕ\phiitalic_ϕ over the measurement timescale, but sufficiently large so that stochastic fluctuations can be ignored.

There is also another interpretation of fitness as the critical dilution rate, a quantity that plays an important role in chemostat studies (pirtExtensionTheoryChemostat1970, ; huEngineeringPrinciplesBiotechnology2017, ). Imagine a steady-state flow reactor with substrate concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG and steady-state dilution rate ϕ>0italic-ϕ0\phi>0italic_ϕ > 0. Suppose that the reactor contains replicator X𝑋Xitalic_X at steady-state concentration x*superscript𝑥x^{*}italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT. Using Eq. 19, we can express x*superscript𝑥x^{*}italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT as

x*={r(ϕ)ϕr(ϕ)if r(ϕ)>ϕ0otherwise,superscript𝑥cases𝑟italic-ϕitalic-ϕsuperscript𝑟italic-ϕif 𝑟italic-ϕitalic-ϕ0otherwisex^{*}=\begin{cases}\frac{r(\phi)-\phi}{r^{-}(\phi)}&\text{if\,}r(\phi)>\phi\\ 0&\text{otherwise}\end{cases},italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = { start_ROW start_CELL divide start_ARG italic_r ( italic_ϕ ) - italic_ϕ end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ϕ ) end_ARG end_CELL start_CELL if italic_r ( italic_ϕ ) > italic_ϕ end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL otherwise end_CELL end_ROW , (29)

where we used that ρ=ϕ𝜌italic-ϕ\rho=\phiitalic_ρ = italic_ϕ in steady state. Now imagine slowly increasing the dilution rate ϕitalic-ϕ\phiitalic_ϕ while maintaining the concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG constant. The replicator will be pushed to extinction at the critical dilution rate ϕ^^italic-ϕ\hat{\phi}over^ start_ARG italic_ϕ end_ARG where ϕ^=r(ϕ^)^italic-ϕ𝑟^italic-ϕ\hat{\phi}=r(\hat{\phi})over^ start_ARG italic_ϕ end_ARG = italic_r ( over^ start_ARG italic_ϕ end_ARG ). Given Eq. 27, the critical dilution must equal the fitness f𝑓fitalic_f. The critical dilution rate is experimentally accessible, as long as the dilution rate can be manipulated while maintaining the concentrations a𝑎\vec{a}over→ start_ARG italic_a end_ARG at constant values. Interestingly, the critical dilution rate can be defined for hypercycles and other higher-order replicators for which the replication rate vanishes at small concentrations (schloglChemicalReactionModels1972a, ; szathmarySimpleGrowthLaws1991, ).

Most simply, our definition of fitness can be considered as maximum replication rate that can be achieved by the replicator. That is, the above results can be used to show that ρf𝜌𝑓\rho\leq fitalic_ρ ≤ italic_f always. Thus, f𝑓fitalic_f sets an upper bound on the replication rate which is approached in the limit of low concentrations.

In addition to fitness, we also make use of the notion of the selection coefficient from evolutionary biology. The selection coefficient is a normalized measure of relative fitness difference that ranges from 0 (no difference) to 1 (maximum difference). Given two replicators X𝑋Xitalic_X and Xsuperscript𝑋X^{\prime}italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with fitness values ff𝑓superscript𝑓f\geq f^{\prime}italic_f ≥ italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, we use a common definition of the selection coefficient as (lenski1991long, )

s:=1ff.assign𝑠1superscript𝑓𝑓s:=1-\frac{f^{\prime}}{f}.italic_s := 1 - divide start_ARG italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_f end_ARG . (30)

In real-world experiments, it is often difficult to measure individual concentrations of replicator and intermediate species. Often what is measured is a weighted sum of concentrations,

ω=cXx+i=1m1ciyi,𝜔subscript𝑐𝑋𝑥superscriptsubscript𝑖1𝑚1subscript𝑐𝑖subscript𝑦𝑖\omega=c_{X}x+\sum_{i=1}^{m-1}c_{i}y_{i},italic_ω = italic_c start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_x + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , (31)

given some nonnegative coefficients cXsubscript𝑐𝑋c_{X}italic_c start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT and cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The interpretations of replication rate, fitness, and critical dilution rate generalize to this situation. That is, if the replicator and all intermediate species grow at the same rate ρ𝜌\rhoitalic_ρ, then their weighted sum ω(t)𝜔𝑡\omega(t)italic_ω ( italic_t ) will also grow at the same rate. Similarly, if the replicator and all intermediate species vanish at some critical dilution rate ϕ^^italic-ϕ\hat{\phi}over^ start_ARG italic_ϕ end_ARG, then their weighted sum will also vanish at that dilution rate.

IV.2 Thermodynamic bounds

We now derive our thermodynamic bounds. We consider a reactor containing replicator X𝑋Xitalic_X with a non-negative growth rate ρ0𝜌0\rho\geq 0italic_ρ ≥ 0. Our first bound relates the replicator’s Gibbs free energy of replication σ𝜎\sigmaitalic_σ, fitness f𝑓fitalic_f, and replication rate ρ𝜌\rhoitalic_ρ as

σln(1ρf),𝜎1𝜌𝑓\sigma\geq-\ln\left(1-\frac{\rho}{f}\right),italic_σ ≥ - roman_ln ( 1 - divide start_ARG italic_ρ end_ARG start_ARG italic_f end_ARG ) , (32)

which appeared as inequality (1) in the Introduction. For the special case of a flow reactor in steady state, this bound can be written in terms of the steady-state dilution rate ϕ=ρitalic-ϕ𝜌\phi=\rhoitalic_ϕ = italic_ρ as

σ*ln(1ϕf),superscript𝜎1italic-ϕ𝑓\sigma^{*}\geq-\ln\left(1-\frac{\phi}{f}\right),italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≥ - roman_ln ( 1 - divide start_ARG italic_ϕ end_ARG start_ARG italic_f end_ARG ) , (33)

where σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is Gibbs free energy of replication under steady state concentrations.

To derive the bound above, we combine Eqs. 4 and 19 to express the replicator’s concentration as x=(r(ρ)ρ)/r(ρ)𝑥𝑟𝜌𝜌superscript𝑟𝜌x=(r(\rho)-\rho)/r^{-}(\rho)italic_x = ( italic_r ( italic_ρ ) - italic_ρ ) / italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ). Plugging this into the expression of σ𝜎\sigmaitalic_σ in Eq. 26 gives

σ=ln(r(0)r(0)r(ρ)r(ρ)ρ).𝜎𝑟0superscript𝑟0superscript𝑟𝜌𝑟𝜌𝜌\sigma=\ln\left(\frac{r(0)}{r^{-}(0)}\frac{r^{-}(\rho)}{r(\rho)-\rho}\right).italic_σ = roman_ln ( divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG divide start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) end_ARG start_ARG italic_r ( italic_ρ ) - italic_ρ end_ARG ) . (34)

Inequality (32) is then equivalent to r(0)r(0)r(ρ)r(ρ)ρf/(fρ)𝑟0superscript𝑟0superscript𝑟𝜌𝑟𝜌𝜌𝑓𝑓𝜌\frac{r(0)}{r^{-}(0)}\frac{r^{-}(\rho)}{r(\rho)-\rho}\geq f/(f-\rho)divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG divide start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) end_ARG start_ARG italic_r ( italic_ρ ) - italic_ρ end_ARG ≥ italic_f / ( italic_f - italic_ρ ) for 0ρf0𝜌𝑓0\leq\rho\leq f0 ≤ italic_ρ ≤ italic_f, which in turn is equivalent to the nonnegativity of the function

h(ρ):=r(0)r(0)r(ρ)(fρ)f(r(ρ)ρ).assign𝜌𝑟0superscript𝑟0superscript𝑟𝜌𝑓𝜌𝑓𝑟𝜌𝜌h(\rho):=\frac{r(0)}{r^{-}(0)}r^{-}(\rho)(f-\rho)-f(r(\rho)-\rho)\,.italic_h ( italic_ρ ) := divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) ( italic_f - italic_ρ ) - italic_f ( italic_r ( italic_ρ ) - italic_ρ ) . (35)

Taking second derivatives and using Eq. 21 shows that ρ2h(ρ)0superscriptsubscript𝜌2𝜌0\partial_{\rho}^{2}h(\rho)\leq 0∂ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_h ( italic_ρ ) ≤ 0, so hhitalic_h is concave. Inspection shows that h(ρ)=0𝜌0h(\rho)=0italic_h ( italic_ρ ) = 0 for ρ=0𝜌0\rho=0italic_ρ = 0 and ρ=f𝜌𝑓\rho=fitalic_ρ = italic_f. Therefore, h(ρ)0𝜌0h(\rho)\geq 0italic_h ( italic_ρ ) ≥ 0 over 0ρf0𝜌𝑓0\leq\rho\leq f0 ≤ italic_ρ ≤ italic_f, proving the inequality (32).

Our result is tight for elementary replicators where the rate constants r(ρ)𝑟𝜌r(\rho)italic_r ( italic_ρ ) and r(ρ)superscript𝑟𝜌r^{-}(\rho)italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) do not depend on ρ𝜌\rhoitalic_ρ, so h(ρ)=0𝜌0h(\rho)=0italic_h ( italic_ρ ) = 0 for all ρ𝜌\rhoitalic_ρ. More generally, our bound tends to be tighter for “effectively” elementary replicators, which have fast internal rates so that r(ρ)𝑟𝜌r(\rho)italic_r ( italic_ρ ) and r(ρ)superscript𝑟𝜌r^{-}(\rho)italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) depend weakly on ρ𝜌\rhoitalic_ρ. For general non-elementary replicators, the bound tends to be tighter near ρ=0𝜌0\rho=0italic_ρ = 0 and ρ=f𝜌𝑓\rho=fitalic_ρ = italic_f, corresponding to equilibrium and irreversible regimes, respectively.

We also derive a thermodynamic bound on the selection coefficient, which has implications for the strength of Darwinian evolution. We consider a flow reactor in steady state with dilution rate ϕitalic-ϕ\phiitalic_ϕ. Suppose that the reactor contains some replicator X𝑋Xitalic_X with fitness f>ϕ𝑓italic-ϕf>\phiitalic_f > italic_ϕ, and that there is a second replicator Xsuperscript𝑋X^{\prime}italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with fitness fϕsuperscript𝑓italic-ϕf^{\prime}\leq\phiitalic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≤ italic_ϕ that is pushed to extinction. Plugging into inequality (33) gives

σ*ln(1ff)=lns,superscript𝜎1superscript𝑓𝑓𝑠\sigma^{*}\geq-\ln\left(1-\frac{f^{\prime}}{f}\right)=-\ln s,italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≥ - roman_ln ( 1 - divide start_ARG italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_f end_ARG ) = - roman_ln italic_s , (36)

which appeared as inequality (2) in the Introduction. This result is illustrated on a model of replicators in a chemostat in Section VI.

To build intuition regarding the bound (36), we may consider two extreme situations. The first is equilibrium steady state, where the replication rate ρ𝜌\rhoitalic_ρ and the Gibbs free energy of replication σ𝜎\sigmaitalic_σ vanish for all replicators. Here, all replicators are present in positive equilibrium concentrations which do not depend on fitness, reflecting the fact that Darwinian evolution is impossible in equilibrium (eigenSelforganizationMatterEvolution1971, ; bernstein1983darwinian, ). At the other extreme is the irreversible regime, where each replicator copies itself at its maximum possible rate ρ=f𝜌𝑓\rho=fitalic_ρ = italic_f and σ𝜎\sigmaitalic_σ diverges. Typically, steady states do not exist in this regime even in the presence of dilution, since any replicator with f>ϕ𝑓italic-ϕf>\phiitalic_f > italic_ϕ will grow without bound and any replicator with f<ϕ𝑓italic-ϕf<\phiitalic_f < italic_ϕ will decay to extinction. In the special case that the fittest replicator satisfies f=ϕ𝑓italic-ϕf=\phiitalic_f = italic_ϕ, a non-zero steady state exists which can harbor a single type of replicator (kingAutocatalysis1978, ; lifsonCrucialStagesOrigin1997, ), since a replicator with any other fitness value will go extinct. To summarize, all replicators coexist in equilibrium (σ=0𝜎0\sigma=0italic_σ = 0), while only the fitness replicator can exist in steady state under the irreversible regime (σ𝜎\sigma\to\inftyitalic_σ → ∞). Intermediate values of σ𝜎\sigmaitalic_σ interpolate between these two extremes, permitting steady-state coexistence of some but not all replicators.

The inequalities (32) and (36) are remarkably general, being independent of most details of the chemical system. For instance, they do not depend on the number of coexisting replicators, the substrates/side products involved in replication, whether the replicators copy themselves via elementary or non-elementary reactions, whether the steady state is near or far from equilibrium, etc. They also do not depend on the dynamical mechanism that leads to a particular steady state. For example, they do not depend on whether replicators experience competitive interactions (e.g., different replicators rely on the same substrate) or not (e.g., different replicators do not share substrates but differ in their kinetic parameters).

V Example: Self-complementary dimer

Refer to caption
Figure 3: Fitness and thermodynamic bounds illustrated on Rebek’s self-complementary dimer (tjivikuaSelfreplicatingSystem1990a, ) with 4 reactions, as in Fig. 2 (right). (a) Fitness recovers the initial replication rate given a small starting concentration. We show time-dependent concentrations of replicator x𝑥xitalic_x, bound dimer yXXsubscript𝑦𝑋𝑋y_{XX}italic_y start_POSTSUBSCRIPT italic_X italic_X end_POSTSUBSCRIPT, and weighted sum ω𝜔\omegaitalic_ω from Eq. 37. (b) Fitness recovers the critical dilution rate in a flow reactor. (c) The bound (32) relates replication rate ρ𝜌\rhoitalic_ρ, fitness f𝑓fitalic_f, and Gibbs free energy of replication σ𝜎\sigmaitalic_σ. It is shown on the same system as in (a). (d) The bound (33) relates fitness, steady-state Gibbs free energy σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, and dilution rate ϕitalic-ϕ\phiitalic_ϕ in a steady-state flow reactor. It is shown on the same system as in (b).

We now illustrate our results using a concrete example. We consider a non-elementary replicator that copies itself via the mechanism \setchemfigarrow coeff=0.6 \schemestartX+A𝑋𝐴X+Aitalic_X + italic_A \arrow<=> XA𝑋𝐴XAitalic_X italic_A \arrow<V=>[B𝐵Bitalic_B][] XAB𝑋𝐴𝐵XABitalic_X italic_A italic_B \arrow<=> XX𝑋𝑋XXitalic_X italic_X \arrow<=> X+X𝑋𝑋X+Xitalic_X + italic_X \schemestop

which is also shown in Fig. 2 (right). X𝑋Xitalic_X is self-complementary dimer while A𝐴Aitalic_A and B𝐵Bitalic_B are substrates. The reaction XABXX𝑋𝐴𝐵𝑋𝑋XAB\rightleftharpoons XXitalic_X italic_A italic_B ⇌ italic_X italic_X is a ligation that produces the bound dimer XX𝑋𝑋XXitalic_X italic_X. This type of system was studied in many early experiments on self-replicating chemical systems (von1986self, ; zielinskiAutocatalyticSynthesisTetranucleotide1987a, ; tjivikuaSelfreplicatingSystem1990a, ; rotelloSigmoidalGrowthSelfreplicating1991, ; von_kiedrowski_parabolic_1991, ).

We parameterize the rate constants of the two binding reactions, X+AXA𝑋𝐴𝑋𝐴X+A\rightleftharpoons XAitalic_X + italic_A ⇌ italic_X italic_A and XA+BXAB𝑋𝐴𝐵𝑋𝐴𝐵XA+B\rightleftharpoons XABitalic_X italic_A + italic_B ⇌ italic_X italic_A italic_B, as

ν1=κasubscript𝜈1𝜅𝑎\displaystyle\nu_{1}=\kappa aitalic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_κ italic_a ν1=κeΔG1/RTsuperscriptsubscript𝜈1𝜅superscript𝑒Δsuperscriptsubscript𝐺1𝑅𝑇\displaystyle\qquad\nu_{1}^{-}=\kappa e^{\Delta G_{1}^{\circ}/RT}italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = italic_κ italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T end_POSTSUPERSCRIPT
ν2=κbsubscript𝜈2𝜅𝑏\displaystyle\nu_{2}=\kappa bitalic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_κ italic_b ν2=κeΔG2/RTsuperscriptsubscript𝜈2𝜅superscript𝑒Δsuperscriptsubscript𝐺2𝑅𝑇\displaystyle\qquad\nu_{2}^{-}=\kappa e^{\Delta G_{2}^{\circ}/RT}italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = italic_κ italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T end_POSTSUPERSCRIPT

where a𝑎aitalic_a and b𝑏bitalic_b are concentrations of A𝐴Aitalic_A and B𝐵Bitalic_B. The rate constants of the ligation step are parameterized as

ν3=1subscript𝜈31\displaystyle\nu_{3}=1\qquaditalic_ν start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 1 ν3=eΔG3/RTsuperscriptsubscript𝜈3superscript𝑒Δsuperscriptsubscript𝐺3𝑅𝑇\displaystyle\nu_{3}^{-}=e^{\Delta G_{3}^{\circ}/RT}italic_ν start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T end_POSTSUPERSCRIPT

and of the dimerization step XXX+X𝑋𝑋𝑋𝑋XX\rightleftharpoons X+Xitalic_X italic_X ⇌ italic_X + italic_X as

ν4=κeΔG4/RTν4=κformulae-sequencesubscript𝜈4𝜅superscript𝑒Δsuperscriptsubscript𝐺4𝑅𝑇superscriptsubscript𝜈4𝜅\nu_{4}=\kappa e^{-\Delta G_{4}^{\circ}/RT}\qquad\nu_{4}^{-}=\kappaitalic_ν start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_κ italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = italic_κ

Where possible, we use parameter values from Rebek’s system, one of the first molecules that exhibited self-replication in the lab (tjivikuaSelfreplicatingSystem1990a, ; vonkiedrowskiMinimalReplicatorTheory1993a, ). The binding and unbinding reactions are assumed to be essentially in equilibrium, so we use a fast rate constant κ=109𝜅superscript109\kappa=10^{9}italic_κ = 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT. The standard Gibbs free energies for the binding reactions are ΔG1/RT=ΔG2/RT=ln60Δsuperscriptsubscript𝐺1𝑅𝑇Δsuperscriptsubscript𝐺2𝑅𝑇60-\Delta G_{1}^{\circ}/RT=-\Delta G_{2}^{\circ}/RT=\ln 60- roman_Δ italic_G start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = - roman_Δ italic_G start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = roman_ln 60 (favoring binding), and for dimerization it is ΔG4/RT=ln630Δsuperscriptsubscript𝐺4𝑅𝑇630-\Delta G_{4}^{\circ}/RT=-\ln 630- roman_Δ italic_G start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = - roman_ln 630 (favoring the bound dimer) (tjivikuaSelfreplicatingSystem1990a, ). The ligation step is assumed to be effectively irreversible, so we use a large standard Gibbs free energy of ΔG3/RT=10Δsuperscriptsubscript𝐺3𝑅𝑇10-\Delta G_{3}^{\circ}/RT=10- roman_Δ italic_G start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = 10.

Following experimental setups (vonkiedrowskiMinimalReplicatorTheory1993a, ), we consider the weighted sum of concentrations

ω=x+yXA+yXAB+2yXX,𝜔𝑥subscript𝑦𝑋𝐴subscript𝑦𝑋𝐴𝐵2subscript𝑦𝑋𝑋\omega=x+y_{XA}+y_{XAB}+2y_{XX},italic_ω = italic_x + italic_y start_POSTSUBSCRIPT italic_X italic_A end_POSTSUBSCRIPT + italic_y start_POSTSUBSCRIPT italic_X italic_A italic_B end_POSTSUBSCRIPT + 2 italic_y start_POSTSUBSCRIPT italic_X italic_X end_POSTSUBSCRIPT , (37)

which is the total concentration of replicator and intermediates, with the bound dimer counting as two copies.

We first consider a closed reactor (ϕ=0italic-ϕ0\phi=0italic_ϕ = 0) which starts from nonequilibrium initial concentrations of substrates a(0)=b(0)=8.2mM𝑎0𝑏08.2mMa(0)=b(0)=8.2\,\mathrm{mM}italic_a ( 0 ) = italic_b ( 0 ) = 8.2 roman_mM (tjivikuaSelfreplicatingSystem1990a, ). We choose x(0)=.1μM𝑥0.1𝜇Mx(0)=.1\,\mu\mathrm{M}italic_x ( 0 ) = .1 italic_μ roman_M for the initial replicator concentration. For these parameter values and substrate concentrations, we can use Eq. 20 and Eq. 27 to compute the fitness as

f0.14𝑓0.14f\approx 0.14italic_f ≈ 0.14

Fig. 3 (a) shows the time-dependent concentrations, demonstrating that fitness (dashed line) captures the initial growth rate.

We then consider the same system, but now in a flow reactor in steady state. In Fig. 3 (b), we show steady-state concentrations across different dilution rates, with substrate concentrations maintained at steady-state values a=b=8.2mM𝑎𝑏8.2mMa=b=8.2\,\mathrm{mM}italic_a = italic_b = 8.2 roman_mM. As shown, fitness captures the critical dilution rate at which the replicator goes extinct.

Next, we illustrate our thermodynamic bounds on the same system. Fig. 3 (c) shows the Gibbs free energy of replication versus the bound (32) for the system considered in Fig. 3(a), where the replicator grows from a small initial concentration in a closed reactor. The bound is tightest in the regime of low concentrations, achieving an efficiency of ln(1ρ/f)/σ0.51𝜌𝑓𝜎0.5-\ln(1-\rho/f)/\sigma\approx 0.5- roman_ln ( 1 - italic_ρ / italic_f ) / italic_σ ≈ 0.5 around x.1μM𝑥.1𝜇Mx\approx.1\,\mu\mathrm{M}italic_x ≈ .1 italic_μ roman_M. In Fig. 3 (d), we show the Gibbs free energy of replication versus our bound (33) for the steady-state flow reactor. The bound is tightest near the critical dilution rate, where it achieves an efficiency of ln(1ϕ/f)/σ*0.651italic-ϕ𝑓superscript𝜎0.65-\ln(1-\phi/f)/\sigma^{*}\approx 0.65- roman_ln ( 1 - italic_ϕ / italic_f ) / italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≈ 0.65.

It should be noted that Rebek’s system, like many other self-complementary dimers, is sometimes described as a “parabolic replicator” with growth obeying a square-root law, ω˙ω1/2proportional-to˙𝜔superscript𝜔12\dot{\omega}\propto\omega^{1/2}over˙ start_ARG italic_ω end_ARG ∝ italic_ω start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. Square-root growth results because the autocatalytic mechanism becomes concentrated in the thermodynamically-favored dimerized form XX𝑋𝑋XXitalic_X italic_X. In general, however, the type of growth varies with replicator concentration (vonkiedrowskiMinimalReplicatorTheory1993a, ). At low concentrations, the concentration of the dimerized form is small and growth is first order, as seen in the semi-logarithmic plot Fig. 3 (a). Square-root growth only appears at larger concentrations, once the concentration of the dimerized form (orange line in the figure) is sufficiently large.

VI Example: Darwinian evolution in a chemostat

We now illustrate our thermodynamic bound on selection using a classic model of autocatalytic replicators in a chemostat (continuous stirred-tank flow reactor) (schusterDynamicsEvolutionaryOptimization1985, ).

The reactor may contain up to N𝑁Nitalic_N replicator types, indicated as X1,,XNsubscript𝑋1subscript𝑋𝑁X_{1},\dots,X_{N}italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, where each Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT undergoes an autocatalytic reaction

Xi+AXi+Xisubscript𝑋𝑖𝐴subscript𝑋𝑖subscript𝑋𝑖X_{i}+A\rightleftharpoons X_{i}+X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_A ⇌ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (38)

from a shared substrate A𝐴Aitalic_A. The substrate A𝐴Aitalic_A is supplied at a constant rate ϕitalic-ϕ\phiitalic_ϕ and feed concentration γ𝛾\gammaitalic_γ, and all species flow out with constant dilution rate ϕitalic-ϕ\phiitalic_ϕ.

For simplicity, we suppose that all replicators copy themselves via elementary autocatalytic reactions. The dynamics of concentrations of replicators xisubscript𝑥𝑖x_{i}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and substrate a𝑎aitalic_a are

x˙i=kixi(aeΔGixi)ϕxia˙=ϕ(γa)ikixi(aeΔGixi),subscript˙𝑥𝑖absentsubscript𝑘𝑖subscript𝑥𝑖𝑎superscript𝑒Δsuperscriptsubscript𝐺𝑖subscript𝑥𝑖italic-ϕsubscript𝑥𝑖˙𝑎absentitalic-ϕ𝛾𝑎subscript𝑖subscript𝑘𝑖subscript𝑥𝑖𝑎superscript𝑒Δsuperscriptsubscript𝐺𝑖subscript𝑥𝑖\displaystyle\begin{aligned} \dot{x}_{i}&=k_{i}x_{i}(a-e^{\Delta G_{i}^{\circ}% }x_{i})-\phi x_{i}\\ \dot{a}&=\phi(\gamma-a)-\sum_{i}k_{i}x_{i}(a-e^{\Delta G_{i}^{\circ}}x_{i}),% \end{aligned}start_ROW start_CELL over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_CELL start_CELL = italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_a - italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - italic_ϕ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over˙ start_ARG italic_a end_ARG end_CELL start_CELL = italic_ϕ ( italic_γ - italic_a ) - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_a - italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , end_CELL end_ROW (39)

where kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is a rate constant and ΔGiΔsuperscriptsubscript𝐺𝑖-\Delta G_{i}^{\circ}- roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT is the standard Gibbs free energy of the reaction (38). As usual, we leave dependence on time of x(t)𝑥𝑡x(t)italic_x ( italic_t ) and a(t)𝑎𝑡a(t)italic_a ( italic_t ) implicit.

Although the replicators do not interact directly, they experience an effective interaction due to competition for the shared substrate A𝐴Aitalic_A. This system is related to models of resource competition studied in mathematical ecology and evolutionary biology (dykhuizen1983selection, ; smithTheoryChemostatDynamics1995, ; groverResourceCompetition1997, ; harmandChemostat2017, ). Moreover, this system can be mapped onto a competitive Lotka-Volterra system with an effective interaction (see Appendix C).

This type of dynamical system was considered by Schuster and Sigmund (schusterDynamicsEvolutionaryOptimization1985, ) (see also (feistelPhysicsSelfOrganization2011, )). They showed that for any strictly positive initial conditions, there is a unique steady state which governs the long-term behavior. This steady state is given by a set of coupled equations,

a*=γixi*,xi*=max{0,eΔGi(a*ϕ/ki)}.formulae-sequencesuperscript𝑎𝛾subscript𝑖superscriptsubscript𝑥𝑖superscriptsubscript𝑥𝑖0superscript𝑒Δsuperscriptsubscript𝐺𝑖superscript𝑎italic-ϕsubscript𝑘𝑖a^{*}=\gamma-\sum_{i}x_{i}^{*},\quad x_{i}^{*}=\max\{0,e^{-\Delta G_{i}^{\circ% }}(a^{*}-\phi/k_{i})\}.italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_γ - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_max { 0 , italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) } . (40)

In Appendix C, we show how to solve the coupled equations in Eq. 40 by evaluating at most N𝑁Nitalic_N closed-form expressions.

The strength of selection grows with increasing dilution rate ϕitalic-ϕ\phiitalic_ϕ and/or decreasing substrate feed concentration γ𝛾\gammaitalic_γ, causing the replicators to be driven to extinction one-by-one in order of increasing kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. In Appendix C, we show that replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT becomes extinct once

γϕki1+j:kjkieΔGj(ki1kj1).𝛾italic-ϕsuperscriptsubscript𝑘𝑖1subscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑗1\frac{\gamma}{\phi}\leq k_{i}^{-1}+\sum_{j:k_{j}\geq k_{i}}e^{-\Delta G_{j}^{% \circ}}(k_{i}^{-1}-k_{j}^{-1}).divide start_ARG italic_γ end_ARG start_ARG italic_ϕ end_ARG ≤ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) . (41)
Refer to caption
Figure 4: Steady state behavior of a system of 4 elementary replicators, for varying values of the dilution rate ϕitalic-ϕ\phiitalic_ϕ. Top: steady-state concentrations of the four replicators. As ϕitalic-ϕ\phiitalic_ϕ increases, the replicators are driven to extinction one-by-one (dashed vertical lines). Bottom: As predicted by the bound (36), replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are pushed to extinction once the Gibbs free energy of the fittest replicator (blue line) crosses the selection coefficient lnsi=ln(1fi/f1)subscript𝑠𝑖1subscript𝑓𝑖subscript𝑓1-\ln s_{i}=-\ln(1-f_{i}/f_{1})- roman_ln italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - roman_ln ( 1 - italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ).

We now consider a concrete example of 4 replicators with rate constants (k1,k2,k3,k4)=(4,3,2,1)subscript𝑘1subscript𝑘2subscript𝑘3subscript𝑘44321(k_{1},k_{2},k_{3},k_{4})=(4,3,2,1)( italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) = ( 4 , 3 , 2 , 1 ) and ΔGiΔsuperscriptsubscript𝐺𝑖-\Delta G_{i}^{\circ}- roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT values (1,2,3,2.5)1232.5(1,2,3,2.5)( 1 , 2 , 3 , 2.5 ). Using Eq. 40, we calculate steady-state concentrations xi*superscriptsubscript𝑥𝑖x_{i}^{*}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT of the 4 replicators at different values of the dilution rate ϕitalic-ϕ\phiitalic_ϕ, while substrate feed concentration is set to γ=1𝛾1\gamma=1italic_γ = 1. The steady-state concentrations are shown in Fig. 4 (top). As the dilution rate increases, the replicators go extinct one-by-one in order of increasing kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The critical values of ϕitalic-ϕ\phiitalic_ϕ at which each replicator goes extinct, as specified by Eq. 41, are indicated with dotted vertical lines.

In Fig. 4 (bottom), we show the steady-state Gibbs free energy of replication for each replicator,

σi*=ln(a*/xi*)ΔGi.superscriptsubscript𝜎𝑖superscript𝑎superscriptsubscript𝑥𝑖Δsuperscriptsubscript𝐺𝑖\sigma_{i}^{*}=\ln(a^{*}/x_{i}^{*})-\Delta G_{i}^{\circ}.italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_ln ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT / italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT . (42)

The values of σi*superscriptsubscript𝜎𝑖\sigma_{i}^{*}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT grow with increasing ϕitalic-ϕ\phiitalic_ϕ, diverging to infinity as each replicator approaches extinction. We compare σ1subscript𝜎1\sigma_{1}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, the Gibbs free energy of replication of the fittest replicator X1subscript𝑋1X_{1}italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, to the selection coefficient between X1subscript𝑋1X_{1}italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, si=1fi/f1subscript𝑠𝑖1subscript𝑓𝑖subscript𝑓1s_{i}=1-f_{i}/f_{1}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 - italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Each replicator’s fitness is fi=kia*subscript𝑓𝑖subscript𝑘𝑖superscript𝑎f_{i}=k_{i}a^{*}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, so si=1ki/k1subscript𝑠𝑖1subscript𝑘𝑖subscript𝑘1s_{i}=1-k_{i}/k_{1}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. As predicted by our bound (36), replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT becomes extinct once σ1*superscriptsubscript𝜎1\sigma_{1}^{*}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT crosses lnsisubscript𝑠𝑖-\ln s_{i}- roman_ln italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

We note that fitness values do not determine relative concentrations in steady state. For instance, near equilibrium (small dilution rates), steady-state concentrations are determined by the standard Gibbs free energies ΔGiΔsuperscriptsubscript𝐺𝑖-\Delta G_{i}^{\circ}- roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT rather than fitness values. This can be seen in Fig. 4 (top): replicator X3subscript𝑋3X_{3}italic_X start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT has the largest steady-state concentration at small ϕitalic-ϕ\phiitalic_ϕ values, since it has the largest value of ΔGiΔsuperscriptsubscript𝐺𝑖-\Delta G_{i}^{\circ}- roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT.

We can also consider the overall rate of Gibbs free energy dissipation due to replication. In steady state, it is

Σ˙˙Σ\displaystyle\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG =i𝒥iσi*=ϕixi*[ln(a*/xi*)ΔGi].absentsubscript𝑖subscript𝒥𝑖superscriptsubscript𝜎𝑖italic-ϕsubscript𝑖superscriptsubscript𝑥𝑖delimited-[]superscript𝑎superscriptsubscript𝑥𝑖Δsuperscriptsubscript𝐺𝑖\displaystyle=\sum_{i}\mathcal{J}_{i}\sigma_{i}^{*}=\phi\sum_{i}x_{i}^{*}[\ln(% a^{*}/x_{i}^{*})-\Delta G_{i}^{\circ}].= ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT caligraphic_J start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_ϕ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT [ roman_ln ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT / italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ] . (43)

Here we used Eq. 42 and that the steady-state flux across the autocatalytic reaction of replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is 𝒥i=ϕxi*subscript𝒥𝑖italic-ϕsuperscriptsubscript𝑥𝑖\mathcal{J}_{i}=\phi x_{i}^{*}caligraphic_J start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_ϕ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT.

Refer to caption
Figure 5: Black line shows Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG, the total rate of Gibbs free energy dissipation, Eq. 43, for the 4-replicator model as a function of the dilution rate. Shaded regions indicate contributions from different replicator populations, with colors as in Fig. 4. At the four extinction events (dotted lines), the entropy production rate is continuous but not differentiable, corresponding to second-order nonequilibrium phase transitions.

In Fig. 5, we plot Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG for the 4-replicator system analyzed above. We also plot the contribution from each type of replicator, ϕxi*[ln(a*/xi*)ΔGi]italic-ϕsuperscriptsubscript𝑥𝑖delimited-[]superscript𝑎superscriptsubscript𝑥𝑖Δsuperscriptsubscript𝐺𝑖\phi x_{i}^{*}[\ln(a^{*}/x_{i}^{*})-\Delta G_{i}^{\circ}]italic_ϕ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT [ roman_ln ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT / italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ], as a shaded region. As before, we vary the dilution rate ϕitalic-ϕ\phiitalic_ϕ while holding fixed the substrate feed concentration at γ=1𝛾1\gamma=1italic_γ = 1. Extinction events are marked using vertical dotted lines. Recall from Fig. 4 that the Gibbs free energy of replication σi*subscriptsuperscript𝜎𝑖\sigma^{*}_{i}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT diverges when replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT approaches extinction. However, the concentration xi*superscriptsubscript𝑥𝑖x_{i}^{*}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT vanishes sufficiently fast so that the product xi*σi*0superscriptsubscript𝑥𝑖superscriptsubscript𝜎𝑖0x_{i}^{*}\sigma_{i}^{*}\to 0italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT → 0 at extinction. As we show in Appendix C, Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG is finite and continuous at the extinction events. Thus, under a common classification scheme (schloglChemicalReactionModels1972a, ; mcneilNonequilibriumPhaseTransitions1974, ; zhangCriticalBehaviorEntropy2016, ; tomeEntropyProductionNonequilibrium2012a, ; nguyenExponentialVolumeDependence2020, ), extinction events are second-order nonequilibrium phase transitions.

VII Cross-catalytic cycles

We finish by briefly discussing how our results generalize to certain types of autocatalytic sets (kauffman1986autocatalytic, ). For simplicity, we restrict our attention to autocatalytic sets with a uniform and cyclic organization. A general treatment the thermodynamics of cross-catalytic cycles with arbitrary topologies, kinetics, and thermodynamic parameters is an important direction for future work.

We consider an autocatalytic set that contains n𝑛nitalic_n species (Z1,,Zn)subscript𝑍1subscript𝑍𝑛(Z_{1},\dots,Z_{n})( italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_Z start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) and n𝑛nitalic_n reactions, where each species Zj1subscript𝑍𝑗1Z_{j-1}italic_Z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT catalyzes the formation Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT:

Zj1+iαi(j)AiZj1+Zj+iβi(j)Ai.subscript𝑍𝑗1subscript𝑖superscriptsubscript𝛼𝑖𝑗subscript𝐴𝑖subscript𝑍𝑗1subscript𝑍𝑗subscript𝑖superscriptsubscript𝛽𝑖𝑗subscript𝐴𝑖Z_{j-1}+\sum_{i}\alpha_{i}^{(j)}A_{i}\rightleftharpoons Z_{j-1}+Z_{j}+\sum_{i}% \beta_{i}^{(j)}A_{i}.italic_Z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⇌ italic_Z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT + italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (44)

The indexes are taken as modnmod𝑛\mathrm{mod\,}nroman_mod italic_n, so Z0=Znsubscript𝑍0subscript𝑍𝑛Z_{0}=Z_{n}italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, and αi(j),βi(j)superscriptsubscript𝛼𝑖𝑗superscriptsubscript𝛽𝑖𝑗\alpha_{i}^{(j)},\beta_{i}^{(j)}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT , italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT indicate stoichiometric coefficients of substrates/side products participating in each reaction. Each catalytic reaction in the cycle may be elementary, or it may be a non-elementary mechanism as in Eq. 11 but with initial reactant X𝑋Xitalic_X replaced by Zj1subscript𝑍𝑗1Z_{j-1}italic_Z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT and final products X+X𝑋𝑋X+Xitalic_X + italic_X replaced by Zj1+Zjsubscript𝑍𝑗1subscript𝑍𝑗Z_{j-1}+Z_{j}italic_Z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT + italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

Refer to caption
Figure 6: Examples of cross-catalytic cycles. Left: a 3-element cycle. Right: templated replication of complementary polymers (shown here using dimers).

We term this kind of autocatalytic set a cross-catalytic cycle. A schematic illustration of a 3-member cross-catalytic cycle is shown in Fig. 6 (left). Cross-catalytic cycles have attracted much attention in work on the origin-of-life, both theoretical (eigenSelforganizationMatterEvolution1971, ; bagleyEvolutionMetabolism1992a, ; hordijk2019history, ) and experimental (vaidyaSpontaneousNetworkFormation2012, ). An important example of a two-member cross-catalytic cycle is the templated replication of complementary polymers, illustrated in Fig. 6 (right), which has been investigated in numerous experiments (sievers1994self, ; lincolnSelfsustainedReplicationRNA2009, ; bissetteMechanismsAutocatalysis2013a, ). In biology, a cross-catalytic cycle called the “Hinshelwood cycle” has been proposed as a general model of bacterial growth (hinshelwood136ChemicalKinetics1952, ; iyer2014universality, ).

For simplicity, we assume that the cycle is uniform, in the sense that each cross-catalytic reaction has the same kinetic, stoichiometric, and thermodynamic properties. Although this assumption seems restrictive, it suffices for studying many autocatalytic sets of fundamental interest, such as complementary pairs with two kinetically and thermodynamically similar cross-catalytic reactions, as in Fig. 6 (right).

For an autocatalytic set with a uniform cyclic organization, the cycle members Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT approach equal concentrations zjzksubscript𝑧𝑗subscript𝑧𝑘z_{j}\approx z_{k}italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≈ italic_z start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT after an initial transient. In this regime, we can effectively treat each cycle member Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT as an independent replicator with autocatalytic flux

𝒥j=r(ρ)zjr(ρ)zj2.subscript𝒥𝑗𝑟𝜌subscript𝑧𝑗superscript𝑟𝜌superscriptsubscript𝑧𝑗2\mathcal{J}_{j}=r(\rho)z_{j}-r^{-}(\rho)z_{j}^{2}.caligraphic_J start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_r ( italic_ρ ) italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (45)

Here ρ=𝒥j/zj𝜌subscript𝒥𝑗subscript𝑧𝑗\rho=\mathcal{J}_{j}/z_{j}italic_ρ = caligraphic_J start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the growth rate and r(ρ),r(ρ)𝑟𝜌superscript𝑟𝜌r(\rho),r^{-}(\rho)italic_r ( italic_ρ ) , italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) are effective forward/backward rate constants, neither of which depend on j𝑗jitalic_j due to the assumption of uniformity. When cross-catalytic reactions are elementary, Eq. 45 follows immediately from mass-action kinetics and the assumption zjzj1subscript𝑧𝑗subscript𝑧𝑗1z_{j}\approx z_{j-1}italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≈ italic_z start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT. For non-elementary cross-catalytic reactions with m𝑚mitalic_m elementary steps, we consider the different steps that produce or consume Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT: the m𝑚mitalic_m-th step of cross-catalytic reaction j𝑗jitalic_j produces one Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, the first step of cross-catalytic reaction j+1𝑗1j+1italic_j + 1 consumes one Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, and the m𝑚mitalic_m-th step of cross-catalytic reaction j+1𝑗1j+1italic_j + 1 produces one Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. The overall production of Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT due to the cross-catalytic cycle is

𝒥j=Jm(j)J1(j+1)+Jm(j+1),subscript𝒥𝑗superscriptsubscript𝐽𝑚𝑗superscriptsubscript𝐽1𝑗1superscriptsubscript𝐽𝑚𝑗1\mathcal{J}_{j}=J_{m}^{(j)}-J_{1}^{(j+1)}+J_{m}^{(j+1)},caligraphic_J start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT - italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j + 1 ) end_POSTSUPERSCRIPT + italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j + 1 ) end_POSTSUPERSCRIPT , (46)

where Jm(j)superscriptsubscript𝐽𝑚𝑗J_{m}^{(j)}italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT is the flux across the k𝑘kitalic_k-th intermediate step in cross-catalytic reaction j𝑗jitalic_j, defined similarly to Eq. 12. Assuming uniformity of kinetics and concentrations, the intermediate fluxes are the same for all j𝑗jitalic_j, Ji(j)=Ji(j+1)Jisuperscriptsubscript𝐽𝑖𝑗superscriptsubscript𝐽𝑖𝑗1subscript𝐽𝑖J_{i}^{(j)}=J_{i}^{(j+1)}\equiv J_{i}italic_J start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT = italic_J start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j + 1 ) end_POSTSUPERSCRIPT ≡ italic_J start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, therefore 𝒥j=2JmJ1subscript𝒥𝑗2subscript𝐽𝑚subscript𝐽1\mathcal{J}_{j}=2J_{m}-J_{1}caligraphic_J start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 2 italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT which recovers Eq. 13. We can then derive Eq. 45 using the approach we developed for treating non-elementary replicators in Section III.2. The fitness f𝑓fitalic_f of each cycle member Zjsubscript𝑍𝑗Z_{j}italic_Z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is defined via the relation f=r(f)𝑓𝑟𝑓f=r(f)italic_f = italic_r ( italic_f ), since it can be shown from Eq. 45 that this determines both initial growth rate and critical dilution rate.

The Gibbs free energy of each cross-catalytic reaction j𝑗jitalic_j is σ(j)=ln[r(0)/zjr(0)]subscript𝜎𝑗𝑟0subscript𝑧𝑗superscript𝑟0\sigma_{(j)}=\ln[r(0)/z_{j}r^{-}(0)]italic_σ start_POSTSUBSCRIPT ( italic_j ) end_POSTSUBSCRIPT = roman_ln [ italic_r ( 0 ) / italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ], which can be shown using a similar derivation as Eq. 26. Combining these results and using the derivation from Section IV.2 gives a bound on the Gibbs free energy of each cross-catalytic reaction j𝑗jitalic_j as

σ(j)ln(1ρf),subscript𝜎𝑗1𝜌𝑓\sigma_{(j)}\geq-\ln\left(1-\frac{\rho}{f}\right),italic_σ start_POSTSUBSCRIPT ( italic_j ) end_POSTSUBSCRIPT ≥ - roman_ln ( 1 - divide start_ARG italic_ρ end_ARG start_ARG italic_f end_ARG ) , (47)

thereby recovering the analogue of inequality Eq. 32. In steady state with dilution rate ϕitalic-ϕ\phiitalic_ϕ, this reduces to σ(j)*ln(1ϕ/f)superscriptsubscript𝜎𝑗1italic-ϕ𝑓\sigma_{(j)}^{*}\geq-\ln(1-\phi/f)italic_σ start_POSTSUBSCRIPT ( italic_j ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≥ - roman_ln ( 1 - italic_ϕ / italic_f ), where σ(j)*superscriptsubscript𝜎𝑗\sigma_{(j)}^{*}italic_σ start_POSTSUBSCRIPT ( italic_j ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is the Gibbs free energy of replication under steady-state concentrations.

We can also derive a thermodynamic bound on the strength of selection for cross-catalytic cycles, analogous to Eq. 36. We consider a flow reactor in steady state with dilution rate ϕitalic-ϕ\phiitalic_ϕ that contains a cross-catalytic cycle with fitness f>ϕ𝑓italic-ϕf>\phiitalic_f > italic_ϕ. Suppose there is another replicator Xsuperscript𝑋X^{\prime}italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, which may be either a cross-catalytic cycle or an individual autocatalytic reaction, that has fitness fϕsuperscript𝑓italic-ϕf^{\prime}\leq\phiitalic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≤ italic_ϕ and is therefore extinction in steady state. Plugging fϕsuperscript𝑓italic-ϕf^{\prime}\leq\phiitalic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≤ italic_ϕ into the inequality above gives

σ(j)*ln(1ff)=lns.superscriptsubscript𝜎𝑗1superscript𝑓𝑓𝑠\sigma_{(j)}^{*}\geq-\ln\left(1-\frac{f^{\prime}}{f}\right)=-\ln s.italic_σ start_POSTSUBSCRIPT ( italic_j ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≥ - roman_ln ( 1 - divide start_ARG italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_f end_ARG ) = - roman_ln italic_s . (48)

This result bounds the average Gibbs free energy of a cross-catalytic reaction in terms of the selection coefficient, thereby generalizing the bound (36).

We emphasize that these bounds are all stated in terms of the Gibbs free energy of a single reaction in the cross-catalytic cycle. For example, for a self-replicating complementary polymer as in Fig. 6 (right), these inequalities bound the Gibbs free energy dissipated when making a single complementary copy, that is half of the overall cycle. The Gibbs free energy dissipated in carrying out all n𝑛nitalic_n reactions, as necessary to complete the cycle, is bounded by nln(1ρ/f)𝑛1𝜌𝑓-n\ln(1-\rho/f)- italic_n roman_ln ( 1 - italic_ρ / italic_f ).

VIII Discussion and future work

In this paper, we uncovered a general relationship between dissipated Gibbs free energy, fitness, and selection strength in molecular replicators. This relationship was derived from the principle of local detailed balance, which plays a central role in nonequilibrium thermodynamics. Our results complement other recent work on fundamental stoichiometric and thermodynamic constraints on autocatalytic replication (despons2023structural, ; gagrani2023polyhedral, ; srinivas2023characterizing, ; srinivas2023thermodynamics, ; blokhuisUniversalMotifsDiversity2020, ; pinero_nonequilibrium_2018, ).

Our approach can also be related to competitive exclusion and coexistence theory from microbial ecology (smithTheoryChemostatDynamics1995, ; groverResourceCompetition1997, ; ellnerExpandedModernCoexistence2019, ). Importantly, however, our underlying question differs from the one typically asked in ecology: ecologists usually aim to explain how diversity may be maintained despite competition (ellnerExpandedModernCoexistence2019, ), while we aim to explain how selection may occur (decrease of diversity) despite the drive toward thermodynamic equilibrium. In any case, our work contributes to the study of autocatalytic systems from the perspective of theoretical ecology (leeAutocatalyticNetworksTransition1997, ; vaidyaSpontaneousNetworkFormation2012, ; szilagyi2017ecology, ; peng2020ecological, ) and evolutionary (vasasEvolutionGenes2012a, ; takeuchiEvolutionaryDynamicsRNAlike2012, ; mizuuchiEvolutionaryTransitionSingle2022, ; ametaDarwinianPropertiesTheir2021, ; ameta2021self, ).

Our first inequality (32) may be compared to a bound derived by England (Eq. 10, england2013statistical, ). However, the two bounds apply to different setups, involve different thermodynamic and dynamic quantities, and are not formally equivalent. Our results concern reversible replicators with concentration-dependent fluxes which reach steady state. England’s result concerns irreversible replicators with fixed production and degradation rates which grow exponentially without bound. Our setup may be better suited for studying thermodynamics of Darwinian evolution, since exponentially growing replicators always coexist and do not exhibit selection as traditionally understood (eigenSelforganizationMatterEvolution1971, ; bernstein1983darwinian, ). A deeper comparison of the two approaches remains for future work.

The Gibbs free energy of replication σ𝜎\sigmaitalic_σ plays a central role in our results. We emphasize that this quantity depends both on the intrinsic chemical properties of the replicator and on its concentration x𝑥xitalic_x, which can change over time due to reactions, dilution or inflows, etc. For a replicator in steady state, we may express these two contributions by rewriting Eq. 26 as σ*=ln[r(0)/r(0)]lnx*superscript𝜎𝑟0superscript𝑟0superscript𝑥\sigma^{*}=\ln[r(0)/r^{-}(0)]-\ln x^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_ln [ italic_r ( 0 ) / italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ] - roman_ln italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT. The intrinsic contribution ln[r(0)/r(0)]𝑟0superscript𝑟0\ln[r(0)/r^{-}(0)]roman_ln [ italic_r ( 0 ) / italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ] depends on the replicator’s standard Gibbs free energy ΔGΔsuperscript𝐺-\Delta G^{\circ}- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and stoichiometric coefficients αk,i,βk,isubscript𝛼𝑘𝑖subscript𝛽𝑘𝑖\alpha_{k,i},\beta_{k,i}italic_α start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT, as well as concentrations of substrates/side products a𝑎\vec{a}over→ start_ARG italic_a end_ARG in the reaction volume. This contribution can be increased by using richer substrates (changing stoichiometric coefficients αk,i,βk,isubscript𝛼𝑘𝑖subscript𝛽𝑘𝑖\alpha_{k,i},\beta_{k,i}italic_α start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT), by enriching existing substrates (increasing ΔGΔsuperscript𝐺-\Delta G^{\circ}- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT), as well as in other ways. Such changes to the intrinsic contribution may increase or decrease σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, depending on how they affect the steady-state concentrations of the replicator and substrates/side products. They may also increase, decrease, or leave the fitness f𝑓fitalic_f unchanged, depending on how they affect the effective rate constant of the replication mechanism. In summary, our bounds concern the overall Gibbs free energy of replication, which combines both intrinsic and concentration-dependent contributions. For this reason, they do not directly imply that there is a universal advantage for exploiting richer substrates or otherwise increasing the intrinsic free energy dissipation.

We mention some limitations of our results and directions for future work.

Like many other results derived using local detailed balance, our bound is most meaningful for molecular replicators that are not “too irreversible”, meaning that σ𝜎\sigmaitalic_σ is not too large. For instance, the bound (32) can be rearranged as ρf(1eσ)𝜌𝑓1superscript𝑒𝜎\rho\leq f(1-e^{-\sigma})italic_ρ ≤ italic_f ( 1 - italic_e start_POSTSUPERSCRIPT - italic_σ end_POSTSUPERSCRIPT ), which reduces to the trivial result ρf𝜌𝑓\rho\leq fitalic_ρ ≤ italic_f once σ𝜎\sigmaitalic_σ grows beyond 20 or so (approximately the dissipation produced by ATP hydrolysis). Our bound on the selection coefficient, which appeared as seσ*𝑠superscript𝑒superscript𝜎s\geq e^{-\sigma^{*}}italic_s ≥ italic_e start_POSTSUPERSCRIPT - italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT in (2), also becomes weak in this regime. Interestingly, however, our bound (47) for cross-catalytic cycles refers to the Gibbs free energy of a single reaction in the cycle. This result may be applicable to larger-scale systems that involve large cycles consisting of many reversible cross-catalytic reactions.

We only considered deterministic concentrations, which is justified when molecular counts are large and stochastic fluctuations can be ignored. However, fluctuations cannot be ignored in small systems, nor near extinction events when concentrations are small (mcneilNonequilibriumPhaseTransitions1974, ; nitzanFluctuationsTransitionsChemical1974, ). Future work may extend our analysis to the stochastic regime. It is also interesting to investigate the assumption of ideal solution, which may be violated at sufficiently high concentrations.

We did not consider the effect of mutations. In general, mutations weaken the strength of selection (eigenSelforganizationMatterEvolution1971, ), therefore we expect that mutations can only increase the thermodynamic costs of Darwinian evolution. Future work may verify this prediction and seek stronger bounds on the thermodynamic cost of Darwinian evolution for replicators with mutations. The introduction of mutations leads to other important questions concerning the thermodynamics of evolution, such as the thermodynamic costs of finding new high-fitness replicators rather than merely selecting among existing replicators. This might be considered the thermodynamics of “the arrival of the fittest”, rather than of “the survival of the fittest” (fontanaArrivalFittestTheory1994, ; wagnerArrivalFittestHow2015, ).

Finally, our analysis of autocatalytic sets was restricted to the case where reactions are organized in a single uniform cycle. Future work may consider autocatalytic sets with more general topologies, kinetics, and thermodynamics (jain1998autocatalytic, ; blokhuisUniversalMotifsDiversity2020, ). Similarly, our analysis of multi-step reaction mechanisms was restricted to linear sequences of reactions such as Eq. 11, which may be generalized in future work.

Acknowledgements.
I thank Gülce Kardeş, Nathaniel Virgo, Jenny Poulton, David Saakian, and Jordi Piñero for useful conversations and suggestions. I also thank the Santa Fe Institute for helping to support this work. This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie Grant Agreement No. 101068029.

APPENDIX A Degradation reactions

Here we show that our thermodynamic bounds can be generalized to account for degradation reactions.

Let us consider the case where the replicator X𝑋Xitalic_X undergoes degradation. In the absence of degradation, the replicator’s concentration changes as

x˙=𝒥ϕx=r(ρ)xr(ρ)x2ϕx,˙𝑥𝒥italic-ϕ𝑥𝑟𝜌𝑥superscript𝑟𝜌superscript𝑥2italic-ϕ𝑥\dot{x}=\mathcal{J}-\phi x=r(\rho)x-r^{-}(\rho)x^{2}-\phi x,over˙ start_ARG italic_x end_ARG = caligraphic_J - italic_ϕ italic_x = italic_r ( italic_ρ ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ϕ italic_x , (49)

where 𝒥=r(ρ)xr(ρ)x2𝒥𝑟𝜌𝑥superscript𝑟𝜌superscript𝑥2\mathcal{J}=r(\rho)x-r^{-}(\rho)x^{2}caligraphic_J = italic_r ( italic_ρ ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the net flux across the autocatalytic reaction. Now suppose that the replicator undergoes a degradation reaction,

X+iαiAiiβiAi,𝑋subscript𝑖superscriptsubscript𝛼𝑖subscript𝐴𝑖subscript𝑖superscriptsubscript𝛽𝑖subscript𝐴𝑖X+\sum_{i}\alpha_{i}^{\prime}A_{i}\to\sum_{i}\beta_{i}^{\prime}A_{i}\,,italic_X + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , (50)

where αisuperscriptsubscript𝛼𝑖\alpha_{i}^{\prime}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and βisuperscriptsubscript𝛽𝑖\beta_{i}^{\prime}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT are stoichiometric coefficients. We assume that the degradation is irreversible and that the flux across the degradation reaction is first order with rate constant η0𝜂0\eta\geq 0italic_η ≥ 0, so ηx𝜂𝑥\eta xitalic_η italic_x. As usual, the rate constant η𝜂\etaitalic_η may depend on the concentrations of substrates/side products a𝑎\vec{a}over→ start_ARG italic_a end_ARG, though we leave this implicit.

In the presence of degradation, the replicator’s concentration changes as

x˙=𝒥ϕx=r(ρ)xr(ρ)x2ηxϕx.˙𝑥𝒥italic-ϕ𝑥𝑟𝜌𝑥superscript𝑟𝜌superscript𝑥2𝜂𝑥italic-ϕ𝑥\dot{x}=\mathcal{J}-\phi x=r(\rho)x-r^{-}(\rho)x^{2}-\eta x-\phi x.over˙ start_ARG italic_x end_ARG = caligraphic_J - italic_ϕ italic_x = italic_r ( italic_ρ ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_η italic_x - italic_ϕ italic_x . (51)

We assume that η<r(ρ)𝜂𝑟𝜌\eta<r(\rho)italic_η < italic_r ( italic_ρ ), since otherwise positive growth is impossible. We then define a modified forward rate constant,

r~(ρ):=r(ρ)η0,assign~𝑟𝜌𝑟𝜌𝜂0\tilde{r}(\rho):=r(\rho)-\eta\geq 0,over~ start_ARG italic_r end_ARG ( italic_ρ ) := italic_r ( italic_ρ ) - italic_η ≥ 0 , (52)

so that Eq. 51 becomes

x˙=r~(ρ)xr(ρ)x2ϕx,˙𝑥~𝑟𝜌𝑥superscript𝑟𝜌superscript𝑥2italic-ϕ𝑥\dot{x}=\tilde{r}(\rho)x-r^{-}(\rho)x^{2}-\phi x,over˙ start_ARG italic_x end_ARG = over~ start_ARG italic_r end_ARG ( italic_ρ ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ϕ italic_x , (53)

thereby recovering the form of Eq. 49. The flux-force relation (26) becomes an inequality,

σ=lnr(0)xr(0)lnr~(0)xr(0),𝜎𝑟0𝑥superscript𝑟0~𝑟0𝑥superscript𝑟0\sigma=\ln\frac{r(0)}{xr^{-}(0)}\geq\ln\frac{\tilde{r}(0)}{xr^{-}(0)},italic_σ = roman_ln divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_x italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG ≥ roman_ln divide start_ARG over~ start_ARG italic_r end_ARG ( 0 ) end_ARG start_ARG italic_x italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG , (54)

since r~(0)r(0)~𝑟0𝑟0\tilde{r}(0)\leq r(0)over~ start_ARG italic_r end_ARG ( 0 ) ≤ italic_r ( 0 ). We can then define fitness f~~𝑓\tilde{f}over~ start_ARG italic_f end_ARG in the same way as Eq. 27, except using r~(ρ)~𝑟𝜌\tilde{r}(\rho)over~ start_ARG italic_r end_ARG ( italic_ρ ) rather than r(ρ)𝑟𝜌r(\rho)italic_r ( italic_ρ ). The derivation of our bounds in Section IV.2 proceeds in the same manner as before using these new definitions. Note that the derivative relations in Eq. 21 still apply to r~(ρ)~𝑟𝜌\tilde{r}(\rho)over~ start_ARG italic_r end_ARG ( italic_ρ ), since η𝜂\etaitalic_η does not depend on ρ𝜌\rhoitalic_ρ.

Let us now consider the case of a non-elementary replicator X𝑋Xitalic_X where some intermediate species Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT also undergo irreversible degradation with rate constants πk[0,νk)subscript𝜋𝑘0subscript𝜈𝑘\pi_{k}\in[0,\nu_{k})italic_π start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ [ 0 , italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ). To account for this, the matrix M¯¯𝑀\bar{M}over¯ start_ARG italic_M end_ARG in the linear system (17) turns into

M¯ij:=Mij+δijπi.assignsubscript¯𝑀𝑖𝑗subscript𝑀𝑖𝑗subscript𝛿𝑖𝑗subscript𝜋𝑖\bar{M}_{ij}:=M_{ij}+\delta_{ij}\pi_{i}.over¯ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT := italic_M start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (55)

Matrix calculus shows that

M¯jk1M¯ii=M¯ji1M¯ik1.superscriptsubscript¯𝑀𝑗𝑘1subscript¯𝑀𝑖𝑖superscriptsubscript¯𝑀𝑗𝑖1superscriptsubscript¯𝑀𝑖𝑘1\frac{\partial\bar{M}_{jk}^{-1}}{\partial\bar{M}_{ii}}=-\bar{M}_{ji}^{-1}\bar{% M}_{ik}^{-1}.divide start_ARG ∂ over¯ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ over¯ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT end_ARG = - over¯ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over¯ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_i italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT .

For any set of degradation rates πisubscript𝜋𝑖\pi_{i}italic_π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, M¯¯𝑀\bar{M}over¯ start_ARG italic_M end_ARG is an M-matrix so it is invertible and all of entries of its inverse are nonnegative (shivakumar1996two, ; meyer1978singular, ). Therefore, every entry of M¯1superscript¯𝑀1\bar{M}^{-1}over¯ start_ARG italic_M end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT decreases with increasing πksubscript𝜋𝑘\pi_{k}italic_π start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. This means that the effective rate constants r¯(ρ)¯𝑟𝜌\bar{r}(\rho)over¯ start_ARG italic_r end_ARG ( italic_ρ ) and r¯(ρ)superscript¯𝑟𝜌\bar{r}^{-}(\rho)over¯ start_ARG italic_r end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ), defined as in Eq. 20 using M¯¯𝑀\bar{M}over¯ start_ARG italic_M end_ARG, obey

r¯(ρ)r(ρ)r¯(ρ)r(ρ).formulae-sequence¯𝑟𝜌𝑟𝜌superscript¯𝑟𝜌superscript𝑟𝜌\bar{r}(\rho)\leq r(\rho)\qquad\bar{r}^{-}(\rho)\geq r^{-}(\rho).over¯ start_ARG italic_r end_ARG ( italic_ρ ) ≤ italic_r ( italic_ρ ) over¯ start_ARG italic_r end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) ≥ italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) .

To account for the degradation of the replicator itself with rate η𝜂\etaitalic_η, we can further modify the effective forward rate constant as

r¯~(ρ):=r¯(ρ)η,assign~¯𝑟𝜌¯𝑟𝜌𝜂\tilde{\bar{r}}(\rho):=\bar{r}(\rho)-\eta,over~ start_ARG over¯ start_ARG italic_r end_ARG end_ARG ( italic_ρ ) := over¯ start_ARG italic_r end_ARG ( italic_ρ ) - italic_η ,

similarly to Eq. 52. As above, we have a flux-force inequality,

σ=lnr(0)xr(0)𝜎𝑟0𝑥superscript𝑟0\displaystyle\sigma=\ln\frac{r(0)}{xr^{-}(0)}italic_σ = roman_ln divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_x italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG lnr¯(ρ)xr¯(ρ)lnr¯~(ρ)xr¯(ρ).absent¯𝑟𝜌𝑥superscript¯𝑟𝜌~¯𝑟𝜌𝑥superscript¯𝑟𝜌\displaystyle\geq\ln\frac{\bar{r}(\rho)}{x\bar{r}^{-}(\rho)}\geq\ln\frac{% \tilde{\bar{r}}(\rho)}{x\bar{r}^{-}(\rho)}.≥ roman_ln divide start_ARG over¯ start_ARG italic_r end_ARG ( italic_ρ ) end_ARG start_ARG italic_x over¯ start_ARG italic_r end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) end_ARG ≥ roman_ln divide start_ARG over~ start_ARG over¯ start_ARG italic_r end_ARG end_ARG ( italic_ρ ) end_ARG start_ARG italic_x over¯ start_ARG italic_r end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ρ ) end_ARG .

The derivation of our bounds then proceeds as before, after defining fitness f¯~~¯𝑓\tilde{\bar{f}}over~ start_ARG over¯ start_ARG italic_f end_ARG end_ARG using the forward rate constant r¯~(ρ)~¯𝑟𝜌\tilde{\bar{r}}(\rho)over~ start_ARG over¯ start_ARG italic_r end_ARG end_ARG ( italic_ρ ). Note that the derivative relations in Eq. 21 still apply to r¯~(ρ)~¯𝑟𝜌\tilde{\bar{r}}(\rho)over~ start_ARG over¯ start_ARG italic_r end_ARG end_ARG ( italic_ρ ), since M¯¯𝑀\bar{M}over¯ start_ARG italic_M end_ARG is an M-matrix and η𝜂\etaitalic_η does not depend on ρ𝜌\rhoitalic_ρ.

In general, our thermodynamic bounds will tend to be less tight in the presence of degradation, since the flux-force equality (26) turns into an inequality. At the same time, the Gibbs free energy of replication may increase due to degradation, for instance because the steady-state concentration of replicator decreases due to degradation. Some of our bounds may also be larger, for instance the right side of the steady-state bound (33) will be larger, since fitness will decrease due to degradation.

APPENDIX B Non-elementary replicators

Here we derive closed-form expressions of the effective rate constants in the ϕ=0italic-ϕ0\phi=0italic_ϕ = 0 regime, as appear in Eq. 22. As discussed at the end of Section III.2, this regime applies when the internal reactions are much faster than the dilution rate.

Plugging ϕ=0italic-ϕ0\phi=0italic_ϕ = 0 into Eq. 16 implies that Jk=Jk+1subscript𝐽𝑘subscript𝐽𝑘1J_{k}=J_{k+1}italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT, so all intermediate fluxes are equal:

Jk=Jfor k1..m.J_{k}=J\qquad\text{for }k\in{1..m}.italic_J start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_J for italic_k ∈ 1 . . italic_m . (56)

They also all equal to the rate of production of the replicator, 𝒥=2JmJ1=J𝒥2subscript𝐽𝑚subscript𝐽1𝐽\mathcal{J}=2J_{m}-J_{1}=Jcaligraphic_J = 2 italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_J. We can then rearrange Eq. 12 to write the steady-state concentrations of Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT as

yk=(νk/νk)yk1𝒥/νkk1..m,y_{k}=(\nu_{k}/\nu_{k}^{-})y_{k-1}-\mathcal{J}/\nu_{k}^{-}\qquad k\in{1..m},italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ( italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) italic_y start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT - caligraphic_J / italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_k ∈ 1 . . italic_m ,

where we use the convention y0=xsubscript𝑦0𝑥y_{0}=xitalic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_x, ym=x2subscript𝑦𝑚superscript𝑥2y_{m}=x^{2}italic_y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This is a first-order linear recurrence relation for yksubscript𝑦𝑘y_{k}italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. Using an existing result (Thm. 7.1, charalambidesEnumerativeCombinatorics2002, ), this recurrence can be solved for ym=x2subscript𝑦𝑚superscript𝑥2y_{m}=x^{2}italic_y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT starting from the initial condition y0=xsubscript𝑦0𝑥y_{0}=xitalic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_x to give

x2superscript𝑥2\displaystyle x^{2}italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =(x𝒥k=1ml=1k1νll=1kνl)k=1mνkνkabsent𝑥𝒥superscriptsubscript𝑘1𝑚superscriptsubscriptproduct𝑙1𝑘1superscriptsubscript𝜈𝑙superscriptsubscriptproduct𝑙1𝑘subscript𝜈𝑙superscriptsubscriptproduct𝑘1𝑚subscript𝜈𝑘superscriptsubscript𝜈𝑘\displaystyle=\left(x-\mathcal{J}\sum_{k=1}^{m}\frac{\prod_{l=1}^{k-1}\nu_{l}^% {-}}{\prod_{l=1}^{k}\nu_{l}}\right)\prod_{k=1}^{m}\frac{\nu_{k}}{\nu_{k}^{-}}= ( italic_x - caligraphic_J ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT divide start_ARG ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_ARG start_ARG ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG ) ∏ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT divide start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT end_ARG (57)

where we use the convention l=1k1νl=1superscriptsubscriptproduct𝑙1𝑘1superscriptsubscript𝜈𝑙1\prod_{l=1}^{k-1}\nu_{l}^{-}=1∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = 1 for k=1𝑘1k=1italic_k = 1. Plugging in r(0)𝑟0r(0)italic_r ( 0 ) and r(0)superscript𝑟0r^{-}(0)italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) from Eq. 22 gives

x2=(x𝒥/r(0))r(0)r(0).superscript𝑥2𝑥𝒥𝑟0𝑟0superscript𝑟0x^{2}=\left(x-\mathcal{J}/r(0)\right)\frac{r(0)}{r^{-}(0)}.italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( italic_x - caligraphic_J / italic_r ( 0 ) ) divide start_ARG italic_r ( 0 ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) end_ARG .

This can be rearranged as 𝒥=r(0)xr(0)x2𝒥𝑟0𝑥superscript𝑟0superscript𝑥2\mathcal{J}=r(0)x-r^{-}(0)x^{2}caligraphic_J = italic_r ( 0 ) italic_x - italic_r start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, which recovers the mass-action-like form of Eq. 19.

APPENDIX C Chemostat model

C.1 Steady state

Here we analyze the steady-state behavior of the dynamical system described by Eq. 39. To begin, let ω:=a+ixiassign𝜔𝑎subscript𝑖subscript𝑥𝑖\omega:=a+\sum_{i}x_{i}italic_ω := italic_a + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT indicate the total concentration of substrate and replicators at time t𝑡titalic_t. The first line of Eq. 39 means that kixi[aeΔGixi]=x˙i+ϕxisubscript𝑘𝑖subscript𝑥𝑖delimited-[]𝑎superscript𝑒Δsuperscriptsubscript𝐺𝑖subscript𝑥𝑖subscript˙𝑥𝑖italic-ϕsubscript𝑥𝑖k_{i}x_{i}[a-e^{\Delta G_{i}^{\circ}}x_{i}]=\dot{x}_{i}+\phi x_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [ italic_a - italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] = over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϕ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Plugging this into the second line of Eq. 39 and rearranging gives

ω˙=ϕ(γω).˙𝜔italic-ϕ𝛾𝜔\dot{\omega}=\phi(\gamma-\omega).over˙ start_ARG italic_ω end_ARG = italic_ϕ ( italic_γ - italic_ω ) .

Thus, ω𝜔\omegaitalic_ω converges exponentially fast to the steady-state value ω*=γsuperscript𝜔𝛾\omega^{*}=\gammaitalic_ω start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_γ.

We consider the long-term dynamics of the system restricted to the invariant subspace ω=γ𝜔𝛾\omega=\gammaitalic_ω = italic_γ. Within this subspace, we can rewrite the first line of Eq. 39 as

x˙i=kixi[γjxjeΔGixi]ϕxi.subscript˙𝑥𝑖subscript𝑘𝑖subscript𝑥𝑖delimited-[]𝛾subscript𝑗subscript𝑥𝑗superscript𝑒Δsuperscriptsubscript𝐺𝑖subscript𝑥𝑖italic-ϕsubscript𝑥𝑖\dot{x}_{i}=k_{i}x_{i}[\gamma-\sum_{j}x_{j}-e^{\Delta G_{i}^{\circ}}x_{i}]-% \phi x_{i}.over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [ italic_γ - ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] - italic_ϕ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (58)

Using an appropriate Lyapunov function, Schuster and Sigmund demonstrated that the dynamics in Eq. 58 converge to the steady state in Eq. 40 for any strictly positive initial condition (x1(0),,xN(0))+Nsubscript𝑥10subscript𝑥𝑁0superscriptsubscript𝑁(x_{1}(0),\dots,x_{N}(0))\in\mathbb{R}_{+}^{N}( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( 0 ) ) ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT (schusterDynamicsEvolutionaryOptimization1985, ). A similar global convergence result can also be derived from the theory of Lotka-Volterra dynamics (takeuchi_existence_1980, ). Specifically, Eq. 58 can be written as a competitive Lotka-Volterra system,

x˙=bixi+jRijxixj,˙𝑥subscript𝑏𝑖subscript𝑥𝑖subscript𝑗subscript𝑅𝑖𝑗subscript𝑥𝑖subscript𝑥𝑗\dot{x}=b_{i}x_{i}+\sum_{j}R_{ij}x_{i}x_{j},over˙ start_ARG italic_x end_ARG = italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (59)

where bi:=kiγϕassignsubscript𝑏𝑖subscript𝑘𝑖𝛾italic-ϕb_{i}:=k_{i}\gamma-\phiitalic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT := italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_γ - italic_ϕ and Rij=ki(1+δijeΔGi)subscript𝑅𝑖𝑗subscript𝑘𝑖1subscript𝛿𝑖𝑗superscript𝑒Δsuperscriptsubscript𝐺𝑖R_{ij}=-k_{i}(1+\delta_{ij}e^{\Delta G_{i}^{\circ}})italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 1 + italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ). The matrix R𝑅Ritalic_R can be expressed as R=K(11T+D)𝑅𝐾1superscript1𝑇𝐷R=-K(\vec{1}\vec{1}^{T}+D)italic_R = - italic_K ( over→ start_ARG 1 end_ARG over→ start_ARG 1 end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT + italic_D ), where Kij=δijkisubscript𝐾𝑖𝑗subscript𝛿𝑖𝑗subscript𝑘𝑖K_{ij}=\delta_{ij}k_{i}italic_K start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and D=δijeΔGi𝐷subscript𝛿𝑖𝑗superscript𝑒Δsuperscriptsubscript𝐺𝑖D=\delta_{ij}e^{\Delta G_{i}^{\circ}}italic_D = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT are diagonal matrices and 11\vec{1}over→ start_ARG 1 end_ARG is a vector of all 1s. Note that 11T+D1superscript1𝑇𝐷\vec{1}\vec{1}^{T}+Dover→ start_ARG 1 end_ARG over→ start_ARG 1 end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT + italic_D is positive definite, since 11T1superscript1𝑇\vec{1}\vec{1}^{T}over→ start_ARG 1 end_ARG over→ start_ARG 1 end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is positive semidefinite and D𝐷Ditalic_D is positive definite. It is known that for this type of Lotka-Volterra system, any strictly positive initial condition converges to a unique globally attracting fixed point (takeuchi_existence_1980, ), which is the steady state specified by Eq. 40.

The steady state in Eq. 40 is expressed as a set of coupled equations, which can be solved in the following manner. Assume without loss of generality that the rate constants kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are arranged in decreasing order, k1k2kNsubscript𝑘1subscript𝑘2subscript𝑘𝑁k_{1}\geq k_{2}\geq\dots\geq k_{N}italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≥ ⋯ ≥ italic_k start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. Given Eq. 40, it must then be that xi*=0superscriptsubscript𝑥𝑖0x_{i}^{*}=0italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = 0 implies xj*=0superscriptsubscript𝑥𝑗0x_{j}^{*}=0italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = 0 whenever j>i𝑗𝑖j>iitalic_j > italic_i. Suppose for the moment that the top i{0..N}i\in\{0..N\}italic_i ∈ { 0 . . italic_N } replicators have non-zero steady-state concentrations,

xj*={eΔGj(a*ϕ/kj)ji0j>isubscriptsuperscript𝑥𝑗casessuperscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗𝑗𝑖0𝑗𝑖x^{*}_{j}=\begin{cases}e^{-\Delta G_{j}^{\circ}}(a^{*}-\phi/k_{j})&j\leq i\\ 0&j>i\end{cases}italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = { start_ROW start_CELL italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_CELL start_CELL italic_j ≤ italic_i end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_j > italic_i end_CELL end_ROW (60)

Eq. 40 then gives a*=γj=1ieΔGj(a*ϕ/kj)superscript𝑎𝛾superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗a^{*}=\gamma-\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}(a^{*}-\phi/k_{j})italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_γ - ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ), so

a*superscript𝑎\displaystyle a^{*}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT =γ+ϕj=1ieΔGjkj11+j=1ieΔGj.absent𝛾italic-ϕsuperscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗\displaystyle=\frac{\gamma+\phi\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}k_{j}^{-% 1}}{1+\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}}.= divide start_ARG italic_γ + italic_ϕ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG . (61)

Eqs. 60 and 61 are the solution to Eq. 40 if for all j{0..N}j\in\{0..N\}italic_j ∈ { 0 . . italic_N },

xj*=max{0,eΔGj(a*ϕ/kj)}.superscriptsubscript𝑥𝑗0superscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗x_{j}^{*}=\max\{0,e^{-\Delta G_{j}^{\circ}}(a^{*}-\phi/k_{j})\}.italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_max { 0 , italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) } . (62)

Given Eq. 60, Eq. 62 is satisfied once

a*ϕ/ki0a*ϕ/ki+1,superscript𝑎italic-ϕsubscript𝑘𝑖0superscript𝑎italic-ϕsubscript𝑘𝑖1a^{*}-\phi/k_{i}\geq 0\geq a^{*}-\phi/k_{i+1},italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≥ 0 ≥ italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , (63)

Therefore, to solve Eq. 40, it suffices to evaluate Eqs. 61 and 60 for i=0,1,2,𝑖012i=0,1,2,\dotsitalic_i = 0 , 1 , 2 , …, stopping once the bounds (63) are satisfied.

C.2 Derivation of inequality (41) (condition for extinction)

Given Eq. 40, replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is extinct once

a*ϕ/ki.superscript𝑎italic-ϕsubscript𝑘𝑖a^{*}\leq\phi/k_{i}.italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≤ italic_ϕ / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (64)

If this inequality holds, then any lower fitness replicator Xjsubscript𝑋𝑗X_{j}italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (kjkisubscript𝑘𝑗subscript𝑘𝑖k_{j}\leq k_{i}italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≤ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT) must also be extinct, since then aϕ/kj𝑎italic-ϕsubscript𝑘𝑗a\leq\phi/k_{j}italic_a ≤ italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. Now, combine the equations in Eq. 40 to write

a*superscript𝑎\displaystyle a^{*}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT =γj:xj>0eΔGj(a*ϕ/kj)absent𝛾subscript:𝑗subscript𝑥𝑗0superscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗\displaystyle=\gamma-\sum_{j:x_{j}>0}e^{-\Delta G_{j}^{\circ}}(a^{*}-\phi/k_{j})= italic_γ - ∑ start_POSTSUBSCRIPT italic_j : italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT )
γj:kj>kieΔGj(a*ϕ/kj),absent𝛾subscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗\displaystyle\leq\gamma-\sum_{j:k_{j}>k_{i}}e^{-\Delta G_{j}^{\circ}}(a^{*}-% \phi/k_{j}),≤ italic_γ - ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (65)

where the inequality in the second line reflects that it may be that a*ϕ/kjsuperscript𝑎italic-ϕsubscript𝑘𝑗a^{*}\leq\phi/k_{j}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≤ italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT even for higher fitness replicators (kj>kisubscript𝑘𝑗subscript𝑘𝑖k_{j}>k_{i}italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT). Rearranging inequality (65) gives

aγ+ϕj:kjkieΔGjkj11+j:kjkieΔGj.𝑎𝛾italic-ϕsubscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11subscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗a\leq\frac{\gamma+\phi\sum_{j:k_{j}\geq k_{i}}e^{-\Delta G_{j}^{\circ}}k_{j}^{% -1}}{1+\sum_{j:k_{j}\geq k_{i}}e^{-\Delta G_{j}^{\circ}}}.italic_a ≤ divide start_ARG italic_γ + italic_ϕ ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG . (66)

Given inequality (66), the bound (64) must be satisfied when

γ+ϕj:kjkieΔGjkj11+j:kjkieΔGjϕ/ki.𝛾italic-ϕsubscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11subscript:𝑗subscript𝑘𝑗subscript𝑘𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗italic-ϕsubscript𝑘𝑖\frac{\gamma+\phi\sum_{j:k_{j}\geq k_{i}}e^{-\Delta G_{j}^{\circ}}k_{j}^{-1}}{% 1+\sum_{j:k_{j}\geq k_{i}}e^{-\Delta G_{j}^{\circ}}}\leq\phi/k_{i}.divide start_ARG italic_γ + italic_ϕ ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j : italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG ≤ italic_ϕ / italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

Rearranging gives the inequality (41).

C.3 Nonequilibrium phase transitions at extinctions

Here we show that the overall rate of Gibbs free energy dissipation, Eq. 43, is continuous but not differentiable at extinction events, meaning that extinctions are second-order nonequilibrium phase transitions.

Suppose that the rate constants are strictly ordered as

k1>k2>>kN.subscript𝑘1subscript𝑘2subscript𝑘𝑁k_{1}>k_{2}>\dots>k_{N}.italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > ⋯ > italic_k start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT . (67)

From Eq. 41, the critical dilution rate for replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is

ϕ^(i)subscript^italic-ϕ𝑖\displaystyle\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT =γki1+j=1i1eΔGj(ki1kj1)absent𝛾superscriptsubscript𝑘𝑖1superscriptsubscript𝑗1𝑖1superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑗1\displaystyle=\frac{\gamma}{k_{i}^{-1}+\sum_{j=1}^{i-1}e^{-\Delta G_{j}^{\circ% }}(k_{i}^{-1}-k_{j}^{-1})}= divide start_ARG italic_γ end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) end_ARG (68)
=γki1+j=1ieΔGj(ki1kj1)absent𝛾superscriptsubscript𝑘𝑖1superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑗1\displaystyle=\frac{\gamma}{k_{i}^{-1}+\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}% (k_{i}^{-1}-k_{j}^{-1})}= divide start_ARG italic_γ end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) end_ARG (69)

where in the second line we used ki1kj1=0superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑗10k_{i}^{-1}-k_{j}^{-1}=0italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 0 for j=i𝑗𝑖j=iitalic_j = italic_i. We treat ϕitalic-ϕ\phiitalic_ϕ as the control parameter, while holding γ𝛾\gammaitalic_γ fixed.

We show that the steady-state substrate concentration a*superscript𝑎a^{*}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is continuous as a function of the dilution rate at ϕ^(i)subscript^italic-ϕ𝑖\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT. When ϕ<ϕ^(i)italic-ϕsubscript^italic-ϕ𝑖\phi<\hat{\phi}_{(i)}italic_ϕ < over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT, replicators X1,,Xisubscript𝑋1subscript𝑋𝑖X_{1},\dots,X_{i}italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are not extinct. We consider the limit from below,

limϕϕ^(i)a*subscriptitalic-ϕsubscript^italic-ϕ𝑖superscript𝑎\displaystyle\lim_{\phi\nearrow\hat{\phi}_{(i)}}a^{*}roman_lim start_POSTSUBSCRIPT italic_ϕ ↗ over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT =γ+ϕ^(i)j=1ieΔGjkj11+j=1ieΔGj=ϕ^(i)ki1.absent𝛾subscript^italic-ϕ𝑖superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗subscript^italic-ϕ𝑖superscriptsubscript𝑘𝑖1\displaystyle=\frac{\gamma+\hat{\phi}_{(i)}\sum_{j=1}^{i}e^{-\Delta G_{j}^{% \circ}}k_{j}^{-1}}{1+\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}}=\hat{\phi}_{(i)}% k_{i}^{-1}.= divide start_ARG italic_γ + over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG = over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT .

Here we first used Eq. 61 and then plugged in Eq. 68 and simplified using a bit of tedious algebra. When ϕ>ϕ^(i)italic-ϕsubscript^italic-ϕ𝑖\phi>\hat{\phi}_{(i)}italic_ϕ > over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT, replicators X1,,Xi1subscript𝑋1subscript𝑋𝑖1X_{1},\dots,X_{i-1}italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT are not extinct, but replicator Xisubscript𝑋𝑖X_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is extinct. We consider the limit from above,

limϕϕ^(i)a*subscriptitalic-ϕsubscript^italic-ϕ𝑖superscript𝑎\displaystyle\lim_{\phi\searrow\hat{\phi}_{(i)}}a^{*}roman_lim start_POSTSUBSCRIPT italic_ϕ ↘ over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT =γ+ϕj=1i1eΔGjkj11+j=1i1eΔGj=ϕ^(i)ki1,absent𝛾italic-ϕsuperscriptsubscript𝑗1𝑖1superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11superscriptsubscript𝑗1𝑖1superscript𝑒Δsuperscriptsubscript𝐺𝑗subscript^italic-ϕ𝑖superscriptsubscript𝑘𝑖1\displaystyle=\frac{\gamma+\phi\sum_{j=1}^{i-1}e^{-\Delta G_{j}^{\circ}}k_{j}^% {-1}}{1+\sum_{j=1}^{i-1}e^{-\Delta G_{j}^{\circ}}}=\hat{\phi}_{(i)}k_{i}^{-1},= divide start_ARG italic_γ + italic_ϕ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG = over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ,

where we first used Eq. 61 and then plugged in Eq. 69 and simplified. The two limits match, so a*superscript𝑎a^{*}italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is continuous at ϕ^(i)subscript^italic-ϕ𝑖\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT. This also implies that steady-state replicator concentrations xj*=max{0,eΔGj(a*ϕ/kj)}superscriptsubscript𝑥𝑗0superscript𝑒Δsuperscriptsubscript𝐺𝑗superscript𝑎italic-ϕsubscript𝑘𝑗x_{j}^{*}=\max\{0,e^{-\Delta G_{j}^{\circ}}(a^{*}-\phi/k_{j})\}italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_max { 0 , italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_ϕ / italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) } from Eq. 40 are continuous at ϕ^(i)subscript^italic-ϕ𝑖\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT.

Next, consider the rate of Gibbs free energy dissipation at the critical point,

limϕϕ^(i)Σ˙=limϕϕ^(i)ϕ[j=1i1xj*(lna*xj*ΔGj)+xi*(lna*ΔGi)],subscriptitalic-ϕsubscript^italic-ϕ𝑖˙Σsubscriptitalic-ϕsubscript^italic-ϕ𝑖italic-ϕdelimited-[]superscriptsubscript𝑗1𝑖1superscriptsubscript𝑥𝑗superscript𝑎superscriptsubscript𝑥𝑗Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑥𝑖superscript𝑎Δsuperscriptsubscript𝐺𝑖\lim_{\phi\to\hat{\phi}_{(i)}}\dot{\Sigma}=\\ \lim_{\phi\to\hat{\phi}_{(i)}}\phi\Bigg{[}\sum_{j=1}^{i-1}x_{j}^{*}\Big{(}\ln% \frac{a^{*}}{x_{j}^{*}}-\Delta G_{j}^{\circ}\Big{)}+x_{i}^{*}(\ln a^{*}-\Delta G% _{i}^{\circ})\Bigg{]},start_ROW start_CELL roman_lim start_POSTSUBSCRIPT italic_ϕ → over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT over˙ start_ARG roman_Σ end_ARG = end_CELL end_ROW start_ROW start_CELL roman_lim start_POSTSUBSCRIPT italic_ϕ → over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ϕ [ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( roman_ln divide start_ARG italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ) + italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( roman_ln italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ) ] , end_CELL end_ROW

where we used limϕϕ^(i)xi*lnxi*=limα0αlnα=0.subscriptitalic-ϕsubscript^italic-ϕ𝑖superscriptsubscript𝑥𝑖superscriptsubscript𝑥𝑖subscript𝛼0𝛼𝛼0\lim_{\phi\to\hat{\phi}_{(i)}}x_{i}^{*}\ln x_{i}^{*}=\lim_{\alpha\to 0}\alpha% \ln\alpha=0.roman_lim start_POSTSUBSCRIPT italic_ϕ → over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT roman_ln italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = roman_lim start_POSTSUBSCRIPT italic_α → 0 end_POSTSUBSCRIPT italic_α roman_ln italic_α = 0 . Given Eq. 67, xj*>0superscriptsubscript𝑥𝑗0x_{j}^{*}>0italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT > 0 for j{1..i1}j\in\{1..i-1\}italic_j ∈ { 1 . . italic_i - 1 }, so all the terms on the right side are finite and continuous at ϕ=ϕ^(i)italic-ϕsubscript^italic-ϕ𝑖\phi=\hat{\phi}_{(i)}italic_ϕ = over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT. Thus, Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG is a continuous function of ϕitalic-ϕ\phiitalic_ϕ at ϕ^(i)subscript^italic-ϕ𝑖\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT.

Next, we show that Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG is not differentiable with respect to ϕitalic-ϕ\phiitalic_ϕ at ϕ^(i)subscript^italic-ϕ𝑖\hat{\phi}_{(i)}over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT because it has an infinite left derivative at this point (it is not hard to show that the right derivative is finite). The left derivative is

ϕΣ˙superscriptsubscriptitalic-ϕ˙Σ\displaystyle\partial_{\phi}^{-}\dot{\Sigma}∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT over˙ start_ARG roman_Σ end_ARG =j=1i[ϕxj*(lna*lnxj*1ΔGj)+xj*a*ϕa*].absentsuperscriptsubscript𝑗1𝑖delimited-[]superscriptsubscriptitalic-ϕsuperscriptsubscript𝑥𝑗superscript𝑎superscriptsubscript𝑥𝑗1Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑥𝑗superscript𝑎superscriptsubscriptitalic-ϕsuperscript𝑎\displaystyle=\sum_{j=1}^{i}\left[\partial_{\phi}^{-}x_{j}^{*}(\ln a^{*}-\ln x% _{j}^{*}-1-\Delta G_{j}^{\circ})+\frac{x_{j}^{*}}{a^{*}}\partial_{\phi}^{-}a^{% *}\right].= ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT [ ∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( roman_ln italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - roman_ln italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - 1 - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ) + divide start_ARG italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ] .

All terms in this expression are finite except for (ϕxi*)lnxi*superscriptsubscriptitalic-ϕsuperscriptsubscript𝑥𝑖superscriptsubscript𝑥𝑖-(\partial_{\phi}^{-}x_{i}^{*})\ln x_{i}^{*}- ( ∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) roman_ln italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT. Plugging in Eq. 40 gives

ϕxi*superscriptsubscriptitalic-ϕsuperscriptsubscript𝑥𝑖\displaystyle\partial_{\phi}^{-}x_{i}^{*}∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT =eΔGi(ϕa*ki1)absentsuperscript𝑒Δsuperscriptsubscript𝐺𝑖superscriptsubscriptitalic-ϕsuperscript𝑎superscriptsubscript𝑘𝑖1\displaystyle=e^{-\Delta G_{i}^{\circ}}(\partial_{\phi}^{-}a^{*}-k_{i}^{-1})= italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( ∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT )
=eΔGi[j=1ieΔGjkj11+j=1ieΔGjϕ^(i)ki1ki1]absentsuperscript𝑒Δsuperscriptsubscript𝐺𝑖delimited-[]superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗11superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗subscript^italic-ϕ𝑖superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑖1\displaystyle=e^{-\Delta G_{i}^{\circ}}\left[\frac{\sum_{j=1}^{i}e^{-\Delta G_% {j}^{\circ}}k_{j}^{-1}}{1+\sum_{j=1}^{i}e^{-\Delta G_{j}^{\circ}}}\hat{\phi}_{% (i)}k_{i}^{-1}-k_{i}^{-1}\right]= italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT [ divide start_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ]
=eΔGi[j=1ieΔGj(kj1ki1)ki11+j=1ieΔGj]<0,absentsuperscript𝑒Δsuperscriptsubscript𝐺𝑖delimited-[]superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗superscriptsubscript𝑘𝑗1superscriptsubscript𝑘𝑖1superscriptsubscript𝑘𝑖11superscriptsubscript𝑗1𝑖superscript𝑒Δsuperscriptsubscript𝐺𝑗0\displaystyle=e^{-\Delta G_{i}^{\circ}}\left[\frac{\sum_{j=1}^{i}e^{-\Delta G_% {j}^{\circ}}(k_{j}^{-1}-k_{i}^{-1})-k_{i}^{-1}}{1+\sum_{j=1}^{i}e^{-\Delta G_{% j}^{\circ}}}\right]<0,= italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT [ divide start_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) - italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG ] < 0 ,

where in the second line we used Eq. 61, and in the last line we used that kj>kisubscript𝑘𝑗subscript𝑘𝑖k_{j}>k_{i}italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and ki>0subscript𝑘𝑖0k_{i}>0italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > 0. Hence ϕxi*superscriptsubscriptitalic-ϕsuperscriptsubscript𝑥𝑖\partial_{\phi}^{-}x_{i}^{*}∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is a strictly negative constant, meaning that ϕϕ^(i)italic-ϕsubscript^italic-ϕ𝑖\phi\nearrow\hat{\phi}_{(i)}italic_ϕ ↗ over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT, and xi*0superscriptsubscript𝑥𝑖0x_{i}^{*}\to 0italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT → 0, (ϕxi*)lnxi*superscriptsubscriptitalic-ϕsuperscriptsubscript𝑥𝑖superscriptsubscript𝑥𝑖-(\partial_{\phi}^{-}x_{i}^{*})\ln x_{i}^{*}\to-\infty- ( ∂ start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) roman_ln italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT → - ∞. This shows that the left derivative of Σ˙˙Σ\dot{\Sigma}over˙ start_ARG roman_Σ end_ARG is negative infinite at ϕ=ϕ^(i)italic-ϕsubscript^italic-ϕ𝑖\phi=\hat{\phi}_{(i)}italic_ϕ = over^ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT ( italic_i ) end_POSTSUBSCRIPT.

References

  • (1) P. Mehta and D. J. Schwab, “Energetic costs of cellular computation,” Proceedings of the National Academy of Sciences, vol. 109, no. 44, pp. 17 978–17 982, 2012.
  • (2) A. C. Barato, D. Hartich, and U. Seifert, “Efficiency of cellular information processing,” New Journal of Physics, vol. 16, no. 10, p. 103024, 2014.
  • (3) C. C. Govern and P. R. ten Wolde, “Energy dissipation and noise correlations in biochemical sensing,” Physical Review Letters, vol. 113, no. 25, p. 258102, 2014.
  • (4) D. Andrieux and P. Gaspard, “Nonequilibrium generation of information in copolymerization processes,” Proceedings of the National Academy of Sciences, vol. 105, no. 28, pp. 9516–9521, Jul. 2008.
  • (5) T. E. Ouldridge and P. R. ten Wolde, “Fundamental costs in the production and destruction of persistent polymer copies,” Physical Review Letters, vol. 118, no. 15, p. 158103, 2017.
  • (6) J. M. Poulton, P. R. ten Wolde, and T. E. Ouldridge, “Nonequilibrium correlations in minimal dynamical models of polymer copying,” Proceedings of the National Academy of Sciences, vol. 116, no. 6, pp. 1946–1951, Feb. 2019.
  • (7) P. Sartori and S. Pigolotti, “Kinetic versus Energetic Discrimination in Biological Copying,” Physical Review Letters, vol. 110, no. 18, p. 188101, May 2013.
  • (8) Y. Kondo and K. Kaneko, “Growth states of catalytic reaction networks exhibiting energy metabolism,” Physical Review E, vol. 84, no. 1, p. 011927, Jul. 2011.
  • (9) J. L. England, “Statistical physics of self-replication,” The Journal of chemical physics, vol. 139, no. 12, p. 121923, 2013.
  • (10) Y. Himeoka and K. Kaneko, “Entropy production of a steady-growth cell with catalytic reactions,” Physical Review E, vol. 90, no. 4, p. 042714, Oct. 2014.
  • (11) N. Virgo, T. Ikegami, and S. McGregor, “Complex autocatalysis in simple chemistries,” Artificial life, vol. 22, no. 2, pp. 138–152, 2016.
  • (12) D. B. Saakian and H. Qian, “Nonlinear Stochastic Dynamics of Complex Systems, III: Nonequilibrium Thermodynamics of Self-Replication Kinetics,” IEEE Transactions on Molecular, Biological and Multi-Scale Communications, vol. 2, no. 1, pp. 40–51, 2016.
  • (13) L. M. Bishop and H. Qian, “Stochastic Bistability and Bifurcation in a Mesoscopic Signaling System with Autocatalytic Kinase,” Biophysical Journal, vol. 98, no. 1, pp. 1–11, Jan. 2010.
  • (14) J. Piñero and R. Solé, “Nonequilibrium Entropic Bounds for Darwinian Replicators,” Entropy, vol. 20, no. 2, p. 98, Jan. 2018.
  • (15) B. Corominas-Murtra, “Thermodynamics of duplication thresholds in synthetic protocell systems,” Life, vol. 9, no. 1, p. 9, 2019.
  • (16) D. A. Beard and H. Qian, “Relationship between thermodynamic driving force and one-way fluxes in reversible processes,” PloS one, vol. 2, no. 1, p. e144, 2007.
  • (17) C. Jarzynski, “Equalities and inequalities: irreversibility and the second law of thermodynamics at the nanoscale,” Annu. Rev. Condens. Matter Phys., vol. 2, no. 1, pp. 329–351, 2011.
  • (18) U. Seifert, “Stochastic thermodynamics, fluctuation theorems and molecular machines,” Reports on Progress in Physics, vol. 75, no. 12, p. 126001, 2012.
  • (19) D. Kondepudi and I. Prigogine, Modern Thermodynamics: From Heat Engines to Dissipative Structures, 2nd ed., 2015.
  • (20) M. Eigen, “Selforganization of matter and the evolution of biological macromolecules,” Die Naturwissenschaften, vol. 58, no. 10, pp. 465–523, Oct. 1971.
  • (21) P. Schuster and K. Sigmund, “Dynamics of evolutionary optimization,” Berichte der Bunsengesellschaft für physikalische Chemie, vol. 89, no. 6, pp. 668–682, Jun. 1985.
  • (22) D. H. Lee, K. Severin, and M. R. Ghadiri, “Autocatalytic networks: the transition from molecular self-replication to molecular ecosystems,” Current Opinion in Chemical Biology, vol. 1, no. 4, pp. 491–496, Dec. 1997.
  • (23) R. Feistel and W. Ebeling, Physics of Self-Organization and Evolution.   Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, Sep. 2011.
  • (24) V. Vasas, C. Fernando, M. Santos, S. Kauffman, and E. Szathmáry, “Evolution before genes,” Biology Direct, vol. 7, no. 1, p. 1, Dec. 2012.
  • (25) N. Takeuchi and P. Hogeweg, “Evolutionary dynamics of RNA-like replicator systems: A bioinformatic approach to the origin of life,” Physics of Life Reviews, vol. 9, no. 3, pp. 219–263, Sep. 2012.
  • (26) N. Vaidya, M. L. Manapat, I. A. Chen, R. Xulvi-Brunet, E. J. Hayden, and N. Lehman, “Spontaneous network formation among cooperative RNA replicators,” Nature, vol. 491, no. 7422, pp. 72–77, Nov. 2012.
  • (27) A. Szilágyi, I. Zachar, I. Scheuring, Á. Kun, B. Könnyű, and T. Czárán, “Ecology and evolution in the rna world dynamics and stability of prebiotic replicator systems,” Life, vol. 7, no. 4, p. 48, 2017.
  • (28) Z. Peng, A. M. Plum, P. Gagrani, and D. A. Baum, “An ecological framework for the analysis of prebiotic chemical reaction networks,” Journal of theoretical biology, vol. 507, p. 110451, 2020.
  • (29) S. Ameta, Y. J. Matsubara, N. Chakraborty, S. Krishna, and S. Thutupalli, “Self-reproduction and Darwinian evolution in autocatalytic chemical reaction systems,” Life, vol. 11, no. 4, p. 308, 2021.
  • (30) R. Mizuuchi, T. Furubayashi, and N. Ichihashi, “Evolutionary transition from a single RNA replicator to a multiple replicator network,” Nature Communications, vol. 13, no. 1, p. 1460, Mar. 2022, number: 1 Publisher: Nature Publishing Group.
  • (31) S. Ameta, S. Arsène, S. Foulon, B. Saudemont, B. E. Clifton, A. D. Griffiths, and P. Nghe, “Darwinian properties and their trade-offs in autocatalytic RNA reaction networks,” Nature Communications, vol. 12, no. 1, p. 842, Feb. 2021.
  • (32) H. Bernstein, H. C. Byerly, F. A. Hopf, R. A. Michod, and G. K. Vemulapalli, “The Darwinian dynamic,” The Quarterly Review of Biology, vol. 58, no. 2, pp. 185–207, 1983.
  • (33) M. Yarus, “Getting past the RNA World: the Initial Darwinian Ancestor,” Cold Spring Harbor Perspectives in Biology, vol. 3, no. 4, pp. a003 590–a003 590, Apr. 2011.
  • (34) C. Jeancolas, C. Malaterre, and P. Nghe, “Thresholds in Origin of Life Scenarios,” iScience, vol. 23, no. 11, p. 101756, Nov. 2020.
  • (35) R. Pascal, A. Pross, and J. D. Sutherland, “Towards an evolutionary theory of the origin of life based on kinetics and thermodynamics,” Open biology, vol. 3, no. 11, p. 130156, 2013.
  • (36) G. Danger, L. L. S. d’Hendecourt, and R. Pascal, “On the conditions for mimicking natural selection in chemical systems,” Nature Reviews Chemistry, vol. 4, no. 2, pp. 102–109, Feb. 2020.
  • (37) V. Patzke and G. von Kiedrowski, “Self replicating systems,” Arkivoc, vol. 5, pp. 293–310, 2007.
  • (38) A. J. Bissette and S. P. Fletcher, “Mechanisms of autocatalysis,” Angewandte Chemie International Edition, vol. 52, no. 49, pp. 12 800–12 826, 2013.
  • (39) T. A. Lincoln and G. F. Joyce, “Self-sustained replication of an RNA enzyme,” Science, vol. 323, no. 5918, pp. 1229–1232, 2009.
  • (40) D. H. Lee, J. R. Granja, J. A. Martinez, K. Severin, and M. R. Ghadiri, “A self-replicating peptide,” Nature, vol. 382, no. 6591, pp. 525–528, Aug. 1996.
  • (41) J. A. J. Metz, R. M. Nisbet, and S. A. H. Geritz, “How should we define ‘fitness’ for general ecological scenarios?” Trends in Ecology & Evolution, vol. 7, no. 6, pp. 198–202, Jun. 1992.
  • (42) J. H. Gillespie, Population genetics: a concise guide.   JHU Press, 2004.
  • (43) W. J. Ewens, Mathematical Population Genetics, ser. Interdisciplinary Applied Mathematics.   New York, NY: Springer New York, 2004, vol. 27.
  • (44) J. M. Smith and E. Szathmáry, The Major Transitions in Evolution, 1995.
  • (45) I. V. Baskakov, G. Legname, S. B. Prusiner, and F. E. Cohen, “Folding of Prion Protein to Its Native α𝛼\alphaitalic_α-Helical Conformation Is under Kinetic Control,” Journal of Biological Chemistry, vol. 276, no. 23, pp. 19 687–19 690, Jan. 2001.
  • (46) P. E. Dawson, T. W. Muir, I. Clark-Lewis, and S. B. H. Kent, “Synthesis of Proteins by Native Chemical Ligation,” Science, vol. 266, no. 5186, pp. 776–779, Nov. 1994.
  • (47) S. B. Prusiner, “Molecular biology of prion diseases,” Science, vol. 252, no. 5012, pp. 1515–1522, 1991.
  • (48) For the prion, we used ΔG=6.073.95=2.12Δsuperscript𝐺6.073.952.12-\Delta G^{\circ}=6.07-3.95=2.12- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT = 6.07 - 3.95 = 2.12 kcal/mol from (Table 1, baskakovFoldingPrionProtein2001, ), which gives ΔG/RT=3.5Δsuperscript𝐺𝑅𝑇3.5-\Delta G^{\circ}/RT=3.5- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = 3.5 at 298superscript298298^{\circ}298 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT K. Note that there is some debate in the literature regarding the precise mechanism of prion replication, such as whether it is first-order, like the replicators considered in this paper, or instead involves higher-order cooperative interactions (eigenPrionicsKineticBasis1996, ; laurentAutocatalyticProcessesCooperative1997a, ; laurentPrionDiseasesProtein1996, ; fematMechanismsPrionDisease2011, ).
  • (49) C. Wang, Q.-X. Guo, and Y. Fu, “Theoretical Analysis of the Detailed Mechanism of Native Chemical Ligation Reactions,” Chemistry: An Asian Journal, vol. 6, no. 5, pp. 1241–1251, May 2011.
  • (50) For the self-replicating peptide, we used ΔG=5.41.9=7.3Δsuperscript𝐺5.41.97.3-\Delta G^{\circ}=-5.4-1.9=7.3- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT = - 5.4 - 1.9 = 7.3 kcal/mol from (Fig. 4, wangTheoreticalAnalysisDetailed2011, ), which gives ΔG/RT=12.3Δsuperscript𝐺𝑅𝑇12.3-\Delta G^{\circ}/RT=12.3- roman_Δ italic_G start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT / italic_R italic_T = 12.3 at 298superscript298298^{\circ}298 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT K. This was converted to σ*superscript𝜎\sigma^{*}italic_σ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT using 6.
  • (51) F. Jülicher and R. Bruinsma, “Motion of RNA Polymerase along DNA: A Stochastic Model,” Biophysical Journal, vol. 74, no. 3, pp. 1169–1185, Mar. 1998.
  • (52) D. Erie, T. Yager, and P. von Hippel, “The Single-Nucleotide Addition Cycle in Transcription: A Biophysical and Biochemical Perspective,” Annual review of biophysics and biomolecular structure, vol. 21, pp. 379–415, Feb. 1992.
  • (53) F. Avanzini and M. Esposito, “Thermodynamics of concentration vs flux control in chemical reaction networks,” The Journal of Chemical Physics, vol. 156, no. 1, 2022.
  • (54) A. Filisetti, A. Graudenzi, R. Serra, M. Villani, D. De Lucrezia, R. M. Füchslin, S. A. Kauffman, N. Packard, and I. Poli, “A stochastic model of the emergence of autocatalytic cycles,” Journal of Systems Chemistry, vol. 2, no. 1, pp. 1–10, 2011.
  • (55) S. N. Semenov, L. J. Kraft, A. Ainla, M. Zhao, M. Baghbanzadeh, V. E. Campbell, K. Kang, J. M. Fox, and G. M. Whitesides, “Autocatalytic, bistable, oscillatory networks of biologically relevant organic reactions,” Nature, vol. 537, no. 7622, pp. 656–660, Sep. 2016.
  • (56) D. E. Dykhuizen and D. L. Hartl, “Selection in chemostats,” Microbiological reviews, vol. 47, no. 2, pp. 150–168, 1983.
  • (57) J. P. Grover, Resource Competition.   Boston, MA: Springer US, 1997.
  • (58) P. A. Hoskisson and G. Hobbs, “Continuous culture–making a comeback?” Microbiology, vol. 151, no. 10, pp. 3153–3159, 2005.
  • (59) J. Harmand, The chemostat.   Hoboken, NJ: ISTE Ltd/John Wiley and Sons Inc, 2017.
  • (60) H. L. Smith and P. E. Waltman, The Theory of the Chemostat: Dynamics of Microbial Competition, ser. Cambridge Studies in Mathematical Biology.   Cambridge ; New York, NY: Cambridge University Press, 1995, no. 13.
  • (61) M. Polettini and M. Esposito, “Irreversible thermodynamics of open chemical networks. i. emergent cycles and broken conservation laws,” The Journal of chemical physics, vol. 141, no. 2, p. 07B610_1, 2014.
  • (62) R. Rao and M. Esposito, “Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics,” Physical Review X, vol. 6, no. 4, Dec. 2016.
  • (63) A. Wachtel, R. Rao, and M. Esposito, “Thermodynamically consistent coarse graining of biocatalysts beyond Michaelis–Menten,” New Journal of Physics, vol. 20, no. 4, p. 042002, Apr. 2018.
  • (64) F. Avanzini, E. Penocchio, G. Falasco, and M. Esposito, “Nonequilibrium thermodynamics of non-ideal chemical reaction networks,” The Journal of Chemical Physics, vol. 154, no. 9, p. 094114, 2021.
  • (65) G. a. M. King, “Autocatalysis,” Chemical Society Reviews, vol. 7, no. 2, pp. 297–316, Jan. 1978.
  • (66) W. Hordijk, “Autocatalytic confusion clarified,” Journal of theoretical biology, vol. 435, pp. 22–28, 2017.
  • (67) G. von Kiedrowski, “A self-replicating hexadeoxynucleotide,” Angewandte Chemie International Edition in English, vol. 25, no. 10, pp. 932–935, 1986.
  • (68) Z.-X. Wang and J.-W. Wu, “Autophosphorylation kinetics of protein kinases,” Biochemical Journal, vol. 368, no. 3, pp. 947–952, Dec. 2002.
  • (69) L. A. Segel and M. Slemrod, “The quasi-steady-state assumption: a case study in perturbation,” SIAM review, vol. 31, no. 3, pp. 446–477, 1989.
  • (70) P. Shivakumar, J. J. Williams, Q. Ye, and C. A. Marinov, “On two-sided bounds related to weakly diagonally dominant m-matrices with application to digital circuit dynamics,” SIAM Journal on Matrix Analysis and Applications, vol. 17, no. 2, pp. 298–312, 1996.
  • (71) C. Meyer Jr and M. Stadelmaier, “Singular m-matrices and inverse positivity,” Linear Algebra and its Applications, vol. 22, pp. 139–156, 1978.
  • (72) D. A. Roff, “Defining fitness in evolutionary models,” Journal of Genetics, vol. 87, no. 4, pp. 339–348, Dec. 2008.
  • (73) H. A. Orr, “Fitness and its role in evolutionary genetics,” Nature Reviews Genetics, vol. 10, no. 8, pp. 531–539, Aug. 2009.
  • (74) J. W. Spaak, P.-J. Ke, A. D. Letten, and F. De Laender, “Different measures of niche and fitness differences tell different tales,” Oikos, vol. 2023, no. 4, p. e09573, 2023.
  • (75) Z. Li, B. Liu, S. H.-J. Li, C. G. King, Z. Gitai, and N. S. Wingreen, “Modeling microbial metabolic trade-offs in a chemostat,” PLOS Computational Biology, vol. 16, no. 8, p. e1008156, Aug. 2020, publisher: Public Library of Science.
  • (76) J. Arnoldi, M. Barbier, R. Kelly, G. Barabás, and A. L. Jackson, “Invasions of ecological communities: Hints of impacts in the invader’s growth rate,” Methods in Ecology and Evolution, vol. 13, no. 1, pp. 167–182, Jan. 2022.
  • (77) S. J. Pirt and W. M. Kurowski, “An Extension of the Theory of the Chemostat with Feedback of Organisms. Its Experimental Realization with a Yeast Culture,” Journal of General Microbiology, vol. 63, no. 3, pp. 357–366, Nov. 1970.
  • (78) W.-S. Hu, Engineering Principles in Biotechnology.   John Wiley & Sons, Sep. 2017.
  • (79) F. Schlögl, “Chemical reaction models for non-equilibrium phase transitions,” Zeitschrift für physik, vol. 253, no. 2, pp. 147–161, 1972.
  • (80) E. Szathmáry, “Simple growth laws and selection consequences,” Trends in Ecology & Evolution, vol. 6, no. 11, pp. 366–370, 1991.
  • (81) R. E. Lenski, M. R. Rose, S. C. Simpson, and S. C. Tadler, “Long-term experimental evolution in escherichia coli. i. adaptation and divergence during 2,000 generations,” The American Naturalist, vol. 138, no. 6, pp. 1315–1341, 1991.
  • (82) S. Lifson, “On the Crucial Stages in the Origin of Animate Matter,” Journal of Molecular Evolution, vol. 44, no. 1, pp. 1–8, Jan. 1997.
  • (83) T. Tjivikua, P. Ballester, and J. Rebek, “Self-replicating system,” Journal of the American Chemical Society, vol. 112, no. 3, pp. 1249–1250, Jan. 1990.
  • (84) W. S. Zielinski and L. E. Orgel, “Autocatalytic synthesis of a tetranucleotide analogue,” Nature, vol. 327, no. 6120, pp. 346–347, May 1987.
  • (85) V. Rotello, J. I. Hong, and J. Rebek, “Sigmoidal growth in a self-replicating system,” Journal of the American Chemical Society, vol. 113, no. 24, pp. 9422–9423, Nov. 1991.
  • (86) G. von Kiedrowski, B. Wlotzka, J. Helbing, M. Matzen, and S. Jordan, “Parabolic Growth of a Self-Replicating Hexadeoxynucleotide Bearing a 3’-5’-Phosphoamidate Linkage,” Angewandte Chemie International Edition in English, vol. 30, no. 4, pp. 423–426, Apr. 1991.
  • (87) G. von Kiedrowski, “Minimal Replicator Theory I: Parabolic Versus Exponential Growth,” in Bioorganic Chemistry Frontiers, ser. Bioorganic Chemistry Frontiers, H. Dugas and F. P. Schmidtchen, Eds.   Berlin, Heidelberg: Springer, 1993, pp. 113–146.
  • (88) K. J. McNeil and D. F. Walls, “Nonequilibrium phase transitions in chemical reactions,” Journal of Statistical Physics, vol. 10, no. 6, pp. 439–448, Jun. 1974.
  • (89) Y. Zhang and A. C. Barato, “Critical behavior of entropy production and learning rate: Ising model with an oscillating field,” Journal of Statistical Mechanics: Theory and Experiment, vol. 2016, no. 11, p. 113207, Nov. 2016.
  • (90) T. Tomé and M. J. de Oliveira, “Entropy Production in Nonequilibrium Systems at Stationary States,” Physical Review Letters, vol. 108, no. 2, p. 020601, Jan. 2012.
  • (91) B. Nguyen and U. Seifert, “Exponential volume dependence of entropy-current fluctuations at first-order phase transitions in chemical reaction networks,” Physical Review E, vol. 102, no. 2, p. 022101, Aug. 2020.
  • (92) S. A. Kauffman, “Autocatalytic sets of proteins,” Journal of theoretical biology, vol. 119, no. 1, pp. 1–24, 1986.
  • (93) R. J. Bagley, J. D. Farmer, and W. Fontana, “Evolution of a metabolism,” Artificial life II, vol. 10, pp. 141–158, 1992.
  • (94) W. Hordijk, “A history of autocatalytic sets,” Biological Theory, vol. 14, no. 4, pp. 224–246, 2019.
  • (95) D. Sievers and G. Von Kiedrowski, “Self-replication of complementary nucleotide-based oligomers,” Nature, vol. 369, no. 6477, pp. 221–224, 1994.
  • (96) C. N. Hinshelwood, “136. On the chemical kinetics of autosynthetic systems,” Journal of the Chemical Society (Resumed), pp. 745–755, 1952.
  • (97) S. Iyer-Biswas, G. E. Crooks, N. F. Scherer, and A. R. Dinner, “Universality in stochastic exponential growth,” Physical review letters, vol. 113, no. 2, p. 028101, 2014.
  • (98) A. Despons, Y. De Decker, and D. Lacoste, “Structural constraints limit the regime of optimal flux in autocatalytic reaction networks,” arXiv preprint arXiv:2306.02366, 2023.
  • (99) P. Gagrani, V. Blanco, E. Smith, and D. Baum, “Polyhedral geometry and combinatorics of an autocatalytic ecosystem,” 2023.
  • (100) S. G. M. Srinivas, F. Avanzini, and M. Esposito, “Characterizing the conditions for indefinite growth in open chemical reaction networks,” arXiv preprint arXiv:2310.08345, 2023.
  • (101) ——, “Thermodynamics of growth in open chemical reaction networks,” arXiv preprint arXiv:2310.08336, 2023.
  • (102) A. Blokhuis, D. Lacoste, and P. Nghe, “Universal motifs and the diversity of autocatalytic systems,” Proceedings of the National Academy of Sciences, vol. 117, no. 41, pp. 25 230–25 236, Oct. 2020.
  • (103) S. P. Ellner, R. E. Snyder, P. B. Adler, and G. Hooker, “An expanded modern coexistence theory for empirical applications,” Ecology Letters, vol. 22, no. 1, pp. 3–18, 2019.
  • (104) A. Nitzan, P. Ortoleva, J. Deutch, and J. Ross, “Fluctuations and transitions at chemical instabilities: The analogy to phase transitions,” The Journal of Chemical Physics, vol. 61, no. 3, pp. 1056–1074, Aug. 1974.
  • (105) W. Fontana and L. W. Buss, ““The arrival of the fittest”: Toward a theory of biological organization,” Bulletin of Mathematical Biology, vol. 56, no. 1, pp. 1–64, 1994.
  • (106) A. Wagner, Arrival of the Fittest: How Nature Innovates.   Penguin Group, 2015.
  • (107) S. Jain and S. Krishna, “Autocatalytic sets and the growth of complexity in an evolutionary model,” Physical Review Letters, vol. 81, no. 25, p. 5684, 1998.
  • (108) C. A. Charalambides, Enumerative Combinatorics.   CRC Press, May 2002.
  • (109) Y. Takeuchi and N. Adachi, “The existence of globally stable equilibria of ecosystems of the generalized Volterra type,” Journal of Mathematical Biology, vol. 10, no. 4, pp. 401–415, Dec. 1980.
  • (110) M. Eigen, “Prionics or The kinetic basis of prion diseases,” Biophysical Chemistry, vol. 63, no. 1, pp. A1–A18, Dec. 1996.
  • (111) M. Laurent, “Autocatalytic processes in cooperative mechanisms of prion diseases,” FEBS Letters, vol. 407, no. 1, pp. 1–6, Apr. 1997.
  • (112) ——, “Prion diseases and the ’protein only’ hypothesis: A theoretical dynamic study.” Biochemical Journal, vol. 318, no. Pt 1, pp. 35–39, Aug. 1996.
  • (113) R. Femat and J. Méndez, “Mechanisms of prion disease progression: A chemical reaction network approach,” IET Systems Biology, vol. 5, no. 6, pp. 347–352, Nov. 2011.