This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: These two authors contributed equally.thanks: These two authors contributed equally.

Also at ]Universal Biology Institute, The University of Tokyo, 7-3-1, Hongo, Bunkyo-ku, Tokyo 113-8654, Japan.

Thermodynamic and Stoichiometric Laws Ruling the Fates of Growing Systems

Atsushi Kamimura Institute of Industrial Science, The University of Tokyo, 4-6-1, Komaba, Meguro-ku, Tokyo 153-8505 Japan    Yuki Sughiyama Institute of Industrial Science, The University of Tokyo, 4-6-1, Komaba, Meguro-ku, Tokyo 153-8505 Japan Graduate School of Information Sciences, Tohoku University, Sendai 980-8579, Japan    Tetsuya J. Kobayashi [email protected] [ Institute of Industrial Science, The University of Tokyo, 4-6-1, Komaba, Meguro-ku, Tokyo 153-8505 Japan
Abstract

We delve into growing open chemical reaction systems (CRSs) characterized by autocatalytic reactions within a variable volume, which changes in response to these reactions. Understanding the thermodynamics of such systems is crucial for comprehending biological cells and constructing protocells, as it sheds light on the physical conditions necessary for their self-replication. Building on our recent work, where we developed a thermodynamic theory for growing CRSs featuring basic autocatalytic motifs with regular stoichiometric matrices, we now expand this theory to include scenarios where the stoichiometric matrix has a nontrivial left kernel space. This extension introduces conservation laws, which limit the variations in chemical species due to reactions, thereby confining the system’s possible states to those compatible with its initial conditions. By considering both thermodynamic and stoichiometric constraints, we clarify the environmental and initial conditions that dictate the CRSs’ fate—whether they grow, shrink, or reach equilibrium. We also find that the conserved quantities significantly influence the equilibrium state achieved by a growing CRS. These results are derived independently of specific thermodynamic potentials or reaction kinetics, therefore underscoring the fundamental impact of conservation laws on the growth of the system.

preprint: APS/123-QED

I Introduction

Chemical thermodynamics provides solid physical principles for explaining the energetics and for predicting the fate of chemical reaction processes [1, 2]. Applications of these principles to autocatalytic reactions are essential to elucidate physical constraints on the capability of self-replication [3, 4]. In the self-replication process, both the chemical components and the encapsulating volume of the system have to grow in a coherent manner [5, 6, 7, 8]. This consideration introduces a unique theoretical challenge to establish the thermodynamic consistency between autocatalytic reactions and volume expansion because the conventional chemical thermodynamics is based solely on the density (concentration) of chemicals, presuming a constant volume [9, 10, 11, 12, 13, 14, 15, 16, 17, 18].

We have recently established a thermodynamic theory for growing systems, in which the number of chemicals and the volume are treated based on the rigorous thermodynamic basis [19]. Accordingly, the theory generally formulates physical conditions to realize the growth of the system, identifies several thermodynamic constraints for the possible states of the growing system, and derives the form of entropy production and heat dissipation accompanying growth. However, this theory can only address systems with regular (full-rank) stoichiometric matrices, thus limiting its applicability to a set of minimum autocatalytic motifs [20]. Given that chemical reaction systems (CRSs) are subject to stoichiometric constraints in general, and various biological functions are robustly realized by specific stoichiometric properties [21, 22], it becomes imperative to delve into the influence of stoichiometry to growing systems for comprehensive understanding of the thermodynamics of self-replication.

In this work, we clarify how stoichiometric conservation laws shape the fate of growing systems. The stoichiometry gives rise to linear combinations of the numbers of chemicals being conserved during the dynamics of chemical reactions [9, 23, 24, 25, 12, 26, 27, 13, 14, 16, 15, 28, 29, 30]. These conservation laws stringently restrict possible changes in the number of chemicals within the stoichiometric compatibility class, defined by the stoichiometric matrix’s left kernel and the initial condition. In biological contexts, particularly in metabolic networks, these laws are crucial in linking the dynamic variations of different chemicals (metabolites) and also in providing insights into cells’ production capabilities [31, 32, 33, 34, 35, 36, 37, 38].

This study elucidates the complex interplay between thermodynamics and stoichiometric conservation laws for the growing system and its consequence to the fate of the growing systems. By disentangling the geometric relationship among chemical numbers, densities, and potentials, we establish the conditions for the system to grow, shrink, or equilibrate, while simultaneously satisfying the second law of thermodynamics and the conservation laws. Furthermore, we show that the existence of conservation laws qualitatively alter the fate of the system and the geometric properties of the equilibrium state.

These results are derived solely based on thermodynamic and stoichiometric requirements and thus remain independent of specific thermodynamic potentials or reaction kinetics. This renders our theory universally applicable, enhancing our understanding of the origins of life and the construction of protocells[39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 20, 60, 61, 62], and enabling the search for the universal laws of biological cells [63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76].

This paper is organized as follows. We devote Sec. II to outline our main results accompanying with an illustrative example of a growing CRS. Here, we clarify the environmental and initial conditions that determine the fate of the system: growth, shrinking or equilibration. In Sec. III, we explain the geometric structure of the growing system to characterize the equilibrium state as a maximum of the total entropy function, which the system may achieve. In Sec. IV, the form of the total entropy function is further investigated to determine if the system grows or shrinks by following the second law of thermodynamics. In Sec. V, we provide a mathematical proof of our main results outlined in Sec. II. Finally, we summarize our work with further discussions in Sec. VI. For better readability, we list the symbols and notations in Appendix L.

II Outline of the main results

We outline our main results before presenting their derivations. In Sec. II.1, we give a thermodynamic setup of a growing chemical reaction system (CRS). By introducing the chemical potential space, we analyze candidates of chemical equilibrium states in Sec. II.2. By computing accessible regions of the system, we characterize the equilibrium state as an intersection of the accessible region and the candidates in Sec. II.3. In Sec. II.4, we present our main claims with the above preparation. We comment our claims for a special class of the growing CRS in Sec. II.5. These results are summarized in Table 1. Then, we demonstrate the claims by numerical simulations using a simple example of the growing CRS in Sec. II.6. Finally, in Sec. II.7, we outline the derivations of our results, which will be presented in the subsequent sections.

II.1 Thermodynamic setup

We consider the following thermodynamic setup in the present paper (Fig. 1). A growing open chemical reaction system is surrounded by an environment. We assume that the system is always in a well-mixed state (a local equilibrium state), and therefore, we can completely describe it by extensive variables (E,Ω,N,X)(E,\Omega,N,X). Here, EE and Ω\Omega represent the internal energy and the volume of the system. N={Nm}>0𝒩NN=\{N^{m}\}\in\mathbb{R}_{>0}^{\mathcal{N}_{N}} denotes the number of chemicals that can move across the membrane between the system and the environment, which we call open chemicals hereafter. X={Xi}𝔛=>0𝒩XX=\{X^{i}\}\in\mathfrak{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}} is the number of chemicals confined within the system; the indices mm and ii respectively run from m=1m=1 to 𝒩N\mathcal{N}_{N} and from i=1i=1 to 𝒩X\mathcal{N}_{X}, where 𝒩N\mathcal{N}_{N} and 𝒩X\mathcal{N}_{X} are the number of species of the open and the confined chemicals. The environment is characterized by intensive variables (T~,Π~,μ~)(\tilde{T},\tilde{\Pi},\tilde{\mu}), where T~\tilde{T} and Π~\tilde{\Pi} are the temperature and the pressure; μ~={μ~m}𝒩N\tilde{\mu}=\{\tilde{\mu}_{m}\}\in\mathbb{R}^{\mathcal{N}_{N}} is the chemical potential corresponding to the open chemicals. Also, (E~,Ω~,N~)(\tilde{E},\tilde{\Omega},\tilde{N}) denote the corresponding extensive variables.

In thermodynamics, the entropy function is defined on (E,Ω,N,X)(E,\Omega,N,X) as a concave and smooth function Σ[E,Ω,N,X]\Sigma[E,\Omega,N,X]. We also write the entropy function for the environment as Σ~T~,Π~,μ~[E~,Ω~,N~]\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}[\tilde{E},\tilde{\Omega},\tilde{N}], and the total entropy can be expressed as

Σtot=Σ[E,Ω,N,X]+Σ~T~,Π~,μ~[E~,Ω~,N~],\Sigma^{\mathrm{tot}}=\Sigma[E,\Omega,N,X]+\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}[\tilde{E},\tilde{\Omega},\tilde{N}], (1)

where we use the additivity of the entropy. Further, the entropy function for the system has the homogeneity. Therefore, without loss of generality, we can write it as

Σ[E,Ω,N,X]=Ωσ[ϵ,n,x],\Sigma[E,\Omega,N,X]=\Omega\sigma[\epsilon,n,x], (2)

where σ[ϵ,n,x]\sigma[\epsilon,n,x] is the entropy density and (ϵ,n,x):=(E/Ω,N/Ω,X/Ω)(\epsilon,n,x):=(E/\Omega,N/\Omega,X/\Omega). In this work, we consider only a situation without phase transition, and therefore, we assume that σ[ϵ,n,x]\sigma[\epsilon,n,x] is strictly concave.

Refer to caption
Figure 1: Diagrammatic representation of open CRSs. The chemical reactions occur with the reaction flux J(t)={Jr(t)}J(t)=\{J^{r}(t)\}, the rrth reaction of which is represented as the chemical equation at the bottom. Here, A={Ai}A=\{A_{i}\} is the label of the confined chemicals, and B={Bm}B=\{B_{m}\} is the label of the open chemicals, which can move across the membrane with the diffusion flux JD(t)={JDm(t)}J_{D}(t)=\{J_{D}^{m}(t)\}. In addition, JE(t)J_{E}(t) denotes the energy exchange rate with the environment, and JΩ(t)J_{\Omega}(t) is the volume expansion rate. Let X={Xi}X=\{X^{i}\} and N={Nm}N=\{N^{m}\} denote the numbers of the confined and the open chemicals in the system, respectively. Also, (S+)ri\left(S_{+}\right)^{i}_{r} and (O+)rm\left(O_{+}\right)^{m}_{r} denote stoichiometric coefficients of the reactants in the rrth reaction, whereas (S)ri\left(S_{-}\right)^{i}_{r} and (O)rm\left(O_{-}\right)^{m}_{r} are the ones of the products. The stoichiometric matrices are given as Sri=(S)ri(S+)riS^{i}_{r}=\left(S_{-}\right)^{i}_{r}-\left(S_{+}\right)^{i}_{r} and Orm=(O)rm(O+)rmO^{m}_{r}=\left(O_{-}\right)^{m}_{r}-\left(O_{+}\right)^{m}_{r}. For theoretical simplicity, we ignore the tension of the membrane and assume that the membrane never bursts. The stoichiometric matrix S={Sri}S=\{S^{i}_{r}\} has only a left kernel space. By representing the basis matrix of Ker[ST]\mathrm{Ker}[S^{T}] by UU, L=UXL=UX is conserved under the chemical reaction dynamics.

The dynamics of the number of confined chemicals X(t)X(t) is described by

dXidt=SriJr(t),\displaystyle\frac{dX^{i}}{dt}=S^{i}_{r}J^{r}(t), (3)

where J(t)={Jr(t)}J(t)=\{J^{r}(t)\} represents the chemical reaction flux; S={Sri}S=\{S^{i}_{r}\} denotes the stoichiometric matrix for the confined chemicals (see Fig. 1). The index rr runs from r=1r=1 to 𝒩R\mathcal{N}_{R}, where 𝒩R\mathcal{N}_{R} is the number of reactions. Also, Einstein’s summation convention has been employed in Eq. (3) for notational simplicity. By integrating Eq. (3), we obtain

Xi(t)=X0i+SriΞr(t),\displaystyle X^{i}(t)=X^{i}_{0}+S^{i}_{r}\Xi^{r}(t), (4)

where X0={X0i}X_{0}=\{X^{i}_{0}\} denotes the initial condition and Ξ(t)={Ξr(t)}\Xi(t)=\{\Xi^{r}(t)\} is the integration of J(t)J(t), i.e., the extent of reaction.

In this work, we assume that the stoichiometric matrix SS does not have the right kernel space,

dimKer[S]=0.\displaystyle\dim\mathrm{Ker}[S]=0. (5)

By contrast, the left kernel space exists and generates a set of conserved quantities:

L={Ll=UilX0i},\displaystyle L=\{L^{l}=U^{l}_{i}X^{i}_{0}\}, (6)

i.e., the vector L=dimKer[ST]L\in\mathcal{L}=\mathbb{R}^{\dim\mathrm{Ker}[S^{T}]}. Here, UU is a basis matrix: {Uil}:=(U1,U2,)T\{U_{i}^{l}\}:=(\textbf{U}^{1},\textbf{U}^{2},...)^{T} whose row vectors Ul\textbf{U}^{l} form the basis of Ker[ST]\mathrm{Ker}[S^{T}], i.e., US=0US=0, and ll runs from 11 to dimKer[ST]\dim\mathrm{Ker}[S^{T}]. Since we are mainly interested in the growth of the system, the time evolution of X(t)X(t) should perpetually increase the number of all confined chemicals while satisfying the conservation laws (Eq. (6)). In order to realize this situation, the stoichiometric matrix SS should satisfy

Im[S]>0𝒩X,\mathrm{Im}[S]\cap\mathbb{R}_{>0}^{\mathcal{N}_{X}}\neq\emptyset, (7)

which is known as the condition for the productive SS [20]. The productive SS excludes the mass conservation-type laws, the existence of which trivially precludes the number of confined chemicals to increase continuously.

The number of open chemicals changes through the reactions and the diffusions. By defining the stoichiometric matrix for open chemicals as O={Orm}O=\{O^{m}_{r}\} and the diffusion fluxes as JD(t)={JDm(t)}J_{D}(t)=\{J^{m}_{D}(t)\}, we have

dNmdt=OrmJr(t)+JDm(t).\displaystyle\frac{dN^{m}}{dt}=O^{m}_{r}J^{r}(t)+J_{D}^{m}(t). (8)

Example 1: To give an illustrative example, we consider the following chemical reactions:

R1:\displaystyle R_{1}: A1+A3+B12A2,\displaystyle A_{1}+A_{3}+B_{1}\rightleftarrows 2A_{2},
R2:\displaystyle R_{2}: A2A1+A3+B2.\displaystyle A_{2}\rightleftarrows A_{1}+A_{3}+B_{2}. (9)

Here, three confined chemicals A=(A1,A2,A3)A=(A_{1},A_{2},A_{3}) and two open chemicals B=(B1,B2)B=(B_{1},B_{2}) are involved in two reactions R1R_{1} and R2R_{2} 111One may feel that the order of the reactions in Eq. (9) is high. The reason why we consider this example is just because it is straightforward to visualize our geometric representation in Sec. III. Since 𝒩X=dimKer[ST]+dimIm[S]\mathcal{N}_{X}=\dim\mathrm{Ker}[S^{T}]+\dim\mathrm{Im}[S] and we can visualize up to three dimensional space (𝒩X=3\mathcal{N}_{X}=3), relevant cases are when dimIm[S]\dim\mathrm{Im}[S] is one, two or three. The present example corresponds to the case in which dimIm[S]\dim\mathrm{Im}[S] is two. In Sec. I and II of the Supplementary material, we also consider other examples with which dimIm[S]\dim\mathrm{Im}[S] is one and three, respectively.. The stoichiometric matrices SS and OO for the confined and open chemicals, respectively, are

S=R1R2A1( 11) A221A311, O=R1R2B1( 10) B201.\displaystyle S=\bordermatrix{&R_{1}&R_{2}\cr A_{1}&-1&1\cr A_{2}&2&-1\cr A_{3}&-1&1\cr},\mbox{ }O=\bordermatrix{&R_{1}&R_{2}\cr B_{1}&-1&0\cr B_{2}&0&1\cr}. (10)

Since 𝒩R=2\mathcal{N}_{R}=2 in this case, we can represent the number of confined chemicals by introducing Ξ=(Ξ1,Ξ2)T\Xi=(\Xi^{1},\Xi^{2})^{T} as

(X1X2X3)(X01X02X03)=SΞ=(121)Ξ1+(111)Ξ2.\left(\begin{array}[]{c}X^{1}\\ X^{2}\\ X^{3}\end{array}\right)-\left(\begin{array}[]{c}X_{0}^{1}\\ X_{0}^{2}\\ X_{0}^{3}\end{array}\right)=S\Xi=\left(\begin{array}[]{c}-1\\ 2\\ -1\end{array}\right)\Xi^{1}+\left(\begin{array}[]{c}1\\ -1\\ 1\end{array}\right)\Xi^{2}. (11)

For this example, dimKer[ST]\dim\mathrm{Ker}[S^{T}] is one, and U=(1,0,1)U=(-1,0,1) satisfies US=0US=0. Thus, the reactions have the conserved quantity L=UX=X1+X3L=UX=-X^{1}+X^{3}. We can intuitively see the conserved quantity from the chemical equations in Eq. (9) as follows. The numbers of A1A_{1} and A3A_{3}, i.e., X1X^{1} and X3X^{3}, change in the same manner: each chemical is consumed by one molecule in the forward reaction of R1R_{1} and is produced by one in that of R2R_{2}. Thus, the difference between X1X^{1} and X3X^{3} is conserved when each of the reactions occurs.

\blacksquare

In this paper, we assume that the time scale of the chemical reactions is much slower than that of the others, i.e., JJE,JΩ,JDJ\ll J_{E},J_{\Omega},J_{D}. This means that we consider the isothermal and isobaric process with a balance of chemical potentials for open chemicals. This assumption allows us to separately analyze the slow dynamics of chemical reactions from the fast dynamics of the energy, the volume, and the chemical diffusion. As shown in Appendix A, our dynamics is effectively governed by Eq. (3), and the other variables (E,Ω,N)(E,\Omega,N) are rapidly relaxed to the values at the equilibrium state of the fast dynamics. As a result, (E,Ω,N)=(E(X),Ω(X),N(X))(E,\Omega,N)=(E(X),\Omega(X),N(X)) is obtained as a function of XX.

With the above assumptions, we obtain the following expression for the total entropy (see Eq. (74) in Appendix A),

Σtot(Ξ)=\displaystyle\Sigma^{\mathrm{tot}}(\Xi)= 1T~{Ω(X)φ(XΩ(X))+Ω(X)Π~+μ~mOrmΞr}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\Omega(X)\tilde{\Pi}+\tilde{\mu}_{m}O^{m}_{r}\Xi^{r}\right\}
+const,\displaystyle+\mathrm{const}, (12)

where X=X0+SΞX=X_{0}+S\Xi. Here, φ()\varphi(\cdot) is the partial grand potential density, which is given by a Legendre transformation of σ[ϵ,n,x]\sigma\left[\epsilon,n,x\right]:

φ(x):=φ[T~,μ~;x]=minϵ,n{ϵT~σ[ϵ,n,x]μ~mnm}.\displaystyle\varphi(x):=\varphi\left[\tilde{T},\tilde{\mu};x\right]=\min_{\epsilon,n}\left\{\epsilon-\tilde{T}\sigma\left[\epsilon,n,x\right]-\tilde{\mu}_{m}n^{m}\right\}. (13)

Note that φ(x)\varphi(x) is strictly convex because σ\sigma is strictly concave. The volume Ω(X)\Omega(X) is variationally determined under isobaric conditions as

Ω(X)=argminΩ{Ωφ(XΩ)+ΩΠ~}.\Omega(X)=\arg\min_{\Omega}\left\{\Omega\varphi\left(\frac{X}{\Omega}\right)+\Omega\tilde{\Pi}\right\}. (14)

Example 2: If we assume that the system and the environment are composed of ideal gas or solution, the partial grand potential density φ(x)\varphi(x) is expressed as

φ(x)=xiνio(T~)\displaystyle\varphi(x)=x^{i}\nu_{i}^{o}\left(\tilde{T}\right) +RT~i{xilogxixi}\displaystyle+R\tilde{T}\sum_{i}\left\{x^{i}\log x^{i}-x^{i}\right\}
RT~me{μ~mμmo(T~)}/RT~,\displaystyle-R\tilde{T}\sum_{m}e^{\left\{\tilde{\mu}_{m}-\mu_{m}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}, (15)

where RR is the gas constant, and νo(T~)={νio(T~)}\nu^{o}(\tilde{T})=\{\nu_{i}^{o}(\tilde{T})\} and μo(T~)={μmo(T~)}\mu^{o}(\tilde{T})=\{\mu^{o}_{m}(\tilde{T})\} are the standard chemical potentials of the confined and the open chemicals, respectively[78].

Since we assume that the environment also consists of the ideal gas, the chemical potential μ~\tilde{\mu} in Eq. (15) is represented as

μ~m=μmo(T~)+RT~logn~m,\displaystyle\tilde{\mu}_{m}=\mu_{m}^{o}(\tilde{T})+R\tilde{T}\log\tilde{n}^{m}, (16)

where n~={n~m}\tilde{n}=\{\tilde{n}^{m}\} is the density of the open chemicals in the environment. Furthermore, Eq. (14) gives the following expression for the volume Ω\Omega,

Ω(X)=RT~iXiΠ~RT~mn~m,\displaystyle\Omega(X)=\frac{R\tilde{T}\sum_{i}X^{i}}{\tilde{\Pi}-R\tilde{T}\sum_{m}\tilde{n}^{m}}, (17)

which corresponds to the equation of state.

\blacksquare

From Eq. (14), the density x𝒳=>0𝒩Xx\in\mathcal{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}} is obtained by a nonlinear function ρ𝒳(X)\rho_{\mathcal{X}}(X) of XX as

ρ𝒳(X):=x(X)=X/Ω(X)𝒳.\displaystyle\rho_{\mathcal{X}}(X):=x(X)=X/\Omega(X)\in\mathcal{X}. (18)

Since Ω(X)\Omega(X) is homogeneous, we have ρ𝒳(αX)=ρ𝒳(X)\rho_{\mathcal{X}}(\alpha X)=\rho_{\mathcal{X}}(X) for α>0\alpha>0. This fact implies that ρ𝒳\rho_{\mathcal{X}} induces one-to-one correspondence between rays in the number space 𝔛\mathfrak{X} and points in the density space 𝒳\mathcal{X}. Here, we write the corresponding ray to a density xx, i.e., the fiber of xx for ρ𝒳\rho_{\mathcal{X}}, as

𝔯𝒳(x):\displaystyle\mathfrak{r}^{\mathcal{X}}(x): ={X|ρ𝒳(X)=x,X>0}𝔛,\displaystyle=\Set{X}{\rho_{\mathcal{X}}(X)=x,X>0}\subset\mathfrak{X},
={X|X=αx,α>0}.\displaystyle=\Set{X}{X=\alpha x,\alpha>0}. (19)

II.2 Candidates of chemical equilibrium states

We define the full grand potential density φ(y)=φ[T~,μ~;y]\varphi^{*}(y)=\varphi^{*}[\tilde{T},\tilde{\mu};y] by the Legendre transformation of φ(x)\varphi(x) in Eq. (13):

φ(y):=maxx{yixiφ(x)}.\varphi^{*}(y):=\max_{x}\left\{y_{i}x^{i}-\varphi(x)\right\}. (20)

We mention that φ(y)\varphi^{*}(y) is strictly convex. Because of the one-to-one correspondence of the Legendre transformation by φ(x)\varphi(x) and φ(y)\varphi^{*}(y), a state of the system is equivalently specified either by the density x𝒳=>0𝒩Xx\in\mathcal{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}} of the confined chemicals or by its Legendre dual variable y=φ(x)𝒴=𝒩Xy=\partial\varphi(x)\in\mathcal{Y}=\mathbb{R}^{\mathcal{N}_{X}} 222The inverse transformation is given by x=φ(y)x=\partial\varphi^{*}(y). The existence of one-to-one correspondence between >0𝒩X\mathbb{R}_{>0}^{\mathcal{N}_{X}} and 𝒩X\mathbb{R}^{\mathcal{N}_{X}} is physically reasonable (See Sec. III.1).. The thermodynamic interpretation of yy is the corresponding chemical potential to xx. In addition, φ(y)\varphi^{*}(y) can be interpreted as the pressure of the system at the state yy. Recalling the one-to-one correspondence by ρ𝒳\rho_{\mathcal{X}} between rays in 𝔛\mathfrak{X} and points in 𝒳\mathcal{X}, we notice that there exists the one-to-one correspondence between rays in 𝔛\mathfrak{X} and points in 𝒴\mathcal{Y}. Here, we write the corresponding ray to a chemical potential yy as

𝔯𝒴(y):={X|φρ𝒳(X)=y,X>0}𝔛.\displaystyle\mathfrak{r}^{\mathcal{Y}}(y):=\Set{X}{\partial\varphi\circ\rho_{\mathcal{X}}(X)=y,X>0}\subset\mathfrak{X}. (21)

The chemical equilibrium states are given by the balance of chemical potentials between reactants and products [78]:

yiSri+μ~mOrm=0.y_{i}S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0. (22)

Thus, we can obtain the candidates of chemical equilibrium states by the solutions to Eq. (22):

EQ𝒴(μ~):={y|yiSri+μ~mOrm=0},\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}):=\Set{y}{y_{i}S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0}, (23)

which we term the equilibrium manifold. Here, we note that EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})\neq\emptyset, because we have assumed dimKer[S]=0\dim\mathrm{Ker}[S]=0 in Eq. (5), i.e., dimIm[ST]=𝒩R\dim\mathrm{Im}[S^{T}]=\mathcal{N}_{R}.

II.3 An equilibrium state as the intersection

The reaction dynamics in Eq. (3) with its conserved quantities LL restricts the accessible region in 𝔛\mathfrak{X} as

STO𝔛(L):={X|UilXi=Ll,X>0}.\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L):=\Set{X}{U^{l}_{i}X^{i}=L^{l},X>0}. (24)

This subspace is known as the stoichiometric compatibility class[80]. By using the mappings ρ𝒳\rho_{\mathcal{X}} in Eq. (18) and φ\partial\varphi, the accessible regions in the density space 𝒳\mathcal{X} and the chemical potential space 𝒴\mathcal{Y} are respectively obtained as

STO𝒳(L)\displaystyle\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) =ρ𝒳(STO𝔛(L)),\displaystyle=\rho_{\mathcal{X}}(\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L)), (25)
STO𝒴(L)\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) =φ(STO𝒳(L)).\displaystyle=\partial\varphi(\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L)). (26)

Since the system evolves only on STO𝒴(L)𝒴\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\in\mathcal{Y}, the equilibrium state to which the system converges is represented by the intersection:

STO𝒴(L)EQ𝒴(μ~),\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}), (27)

if it is not empty. If the intersection is empty, the system never converges to the equilibrium state and the growth of the volume may happen.

II.4 Main claims of the present paper

In this subsection, we state our two main claims of this work before presenting their derivations in the following sections.

Our first claim provides the conditions to determine the fate of the system, i.e., growth, shrinking or equilibration. To formulate our claim, we define yminy^{\mathrm{min}} as

ymin:=argminy{φ(y)|yEQ𝒴(μ~)}.\displaystyle y^{\mathrm{min}}:=\arg\min_{y}\Set{\varphi^{*}(y)}{y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})}. (28)

Note that the productivity of SS in Eq. (7) guarantees the existence and the uniqueness of yminy^{\mathrm{min}} (see the latter half of Appendix F). This yminy^{\mathrm{min}} gives the chemical equilibrium state whose pressure φ(ymin)\varphi^{*}(y^{\mathrm{min}}) is the minimum in the candidates, EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}).

Claim 1

The fate of the system is classified by yminy^{\mathrm{min}} and the conserved quantities LL in Eq. (6) as follows (see also Table 1):

  1. 1.

    If φ(ymin)Π~>0\varphi^{*}\left(y^{\mathrm{min}}\right)-\tilde{\Pi}>0, the system grows and finally diverges for any LL.

  2. 2.

    If φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0, the system converges to an equilibrium state when L=0L=0, whereas it grows when L0L\neq 0.

  3. 3.

    If φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, the system shrinks and finally vanishes when L=0L=0, whereas it converges to an equilibrium state when L0L\neq 0.

Here, L=0L=0 represents the case in which Ll=0L^{l}=0 for all the components of L={Ll}L=\{L^{l}\} in Eq. (6), and L0L\neq 0 indicates that at least one of the components is not zero.

We first notice that one of the conditions in the claim is given by the sign of φ(ymin)Π~\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}. Here, φ(ymin)\varphi^{*}\left(y^{\mathrm{min}}\right) represents the minimum pressure of the system within EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (see Eq. (28)), whereas Π~\tilde{\Pi} is the pressure of the environment. The growth of the system occurs independently of the value of LL when φ(ymin)>Π~\varphi^{*}(y^{\mathrm{min}})>\tilde{\Pi}, because the environmental pressure can not prevent the system from growing by counteracting any internal pressures φ(y)\varphi^{*}(y) for yEQ𝒴(μ~)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}).

The equilibration becomes possible only when φ(ymin)Π~0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}\leq 0 is fulfilled. For the singular case of L=0L=0, the equilibration occurs only when the minimum internal pressure perfectly balances with the external one, namely, φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0. When φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, the system shrinks and X(t)X(t) approaches 0 as tt\to\infty because the external pressure dominates the internal one at the equilibrium state. For the generic L0L\neq 0, the fate of the system is qualitatively altered. The system equilibrates even if φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 owing to the non-singular conservation laws, which prevent X(t)X(t) from approaching 0, i.e., shrinking. In fact, the equilibrium state yEQ𝒴(μ~)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) can have a higher pressure in this case than φ(ymin)\varphi^{*}(y^{\mathrm{min}}), and it balances with the external pressure Π~\tilde{\Pi}. In addition, the system grows for φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0.

From claim 1, the system converges to an equilibrium state when (1) φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 and L=0L=0, and (2) φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 and L0L\neq 0. For these cases, our second claim describes how the equilibrium states appear in the number space 𝔛\mathfrak{X}.

Claim 2

In the number space 𝔛\mathfrak{X}, the equilibrium states appear as follows (see also Table 1):

  1. 1.

    When φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 and L=0L=0, the equilibrium state to which the system converges depends on the initial condition X0STO𝔛(L=0)X_{0}\in\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0) and the functional form of the reaction flux J(t)J(t). Such an equilibrium state lies on the ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}) in Eq. (21).

  2. 2.

    When φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 and L0L\neq 0, the equilibrium state is uniquely determined, irrespective of the initial condition X0STO𝔛(L0)X_{0}\in\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L\neq 0) and the form of the reaction flux J(t)J(t).

Table 1: The fate of the system is classified by the sign of φ(ymin)Π~\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi} and the conserved quantities L={Ll}L=\{L^{l}\}. Equilibration is indicated by EQ(D)\mathrm{EQ(D)} and EQ(I)\mathrm{EQ(I)}. For EQ(D)\mathrm{EQ(D)}, the equilibrium state depends on the initial conditions and the functional form of the reaction flux J(t)J(t). For EQ(I)\mathrm{EQ(I)}, it is independent for a given LL. Here, ymin={yimin}y^{\mathrm{min}}=\{y^{\mathrm{min}}_{i}\} is defined in Eq. (28) and it is identical with yiEQ=μ~mOrm(S1)iry^{\mathrm{EQ}}_{i}=-\tilde{\mu}_{m}O^{m}_{r}(S^{-1})^{r}_{i} for regular SS.
L0L\neq 0 L=0L=0 Regular SS
φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0 Growing Growing Growing
φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 Growing EQ (D) EQ (D)
φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 EQ (I) Shrinking Shrinking

II.5 Regular stoichiometric matrix SS

The fates of the system for the singular case of L=0L=0 are basically the same as the case that SS is regular; no conservation laws exist so that dimKer[ST]=0\dim\mathrm{Ker}[S^{T}]=0, and dimKer[S]=0\dim\mathrm{Ker}[S]=0 from Eq. (5) 333Note that regular SS is productive in Eq. (7). (see Sec. V.3 for the reason why regular SS can be interpreted as L=0L=0). Thus, the two claims clarify that the existence of conservation laws is a fundamental determinant of the fate of the growing system. We investigated the case of regular SS in our previous work [19] (see also 444In Sec. II of the Supplementary material, we show an example when SS is regular in which the dimension of Ker[ST]\mathrm{Ker}[S^{T}] is zero and that of Im[S]\mathrm{Im}[S] is three.). For comparison, we summarize the results here (see the cases for L=0L=0 and regular SS in Table 1).

For a regular SS, the candidates of chemical equilibrium states, EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) in Eq. (23), consist of precisely one point yiEQ=μ~mOrm(S1)iry_{i}^{\mathrm{EQ}}=-\tilde{\mu}_{m}O^{m}_{r}(S^{-1})^{r}_{i}. Here, we note that yminy^{\mathrm{min}} is identical with yEQy^{\mathrm{EQ}} by Eq. (28). Then, we obtain

Corollary 1

When the stoichiometric matrix SS is regular, the fate of the system is classified by yiEQ=μ~mOrm(S1)iry_{i}^{\mathrm{EQ}}=-\tilde{\mu}_{m}O^{m}_{r}(S^{-1})^{r}_{i} as follows (see also Table 1):

  1. 1.

    If φ(yEQ)Π~>0\varphi^{*}\left(y^{\mathrm{EQ}}\right)-\tilde{\Pi}>0, the system grows and finally diverges.

  2. 2.

    If φ(yEQ)Π~=0\varphi^{*}(y^{\mathrm{EQ}})-\tilde{\Pi}=0, equilibrium states form the ray 𝔯𝒴(yEQ)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{EQ}}) in Eq. (21). The system converges to one of them, depending on the initial condition X0𝔛X_{0}\in\mathfrak{X} and the functional form of the reaction flux J(t)J(t).

  3. 3.

    If φ(yEQ)Π~<0\varphi^{*}(y^{\mathrm{EQ}})-\tilde{\Pi}<0, the system shrinks and finally vanishes.

This statement corresponds to claim 1 in [19].

II.6 Numerical demonstrations

In this subsection, we demonstrate our claims by using the specific reactions in Example 1 with the ideal gas case (see Example 2).

We first calculate the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) in Eq. (23) to obtain yminy^{\mathrm{min}} in Eq. (28). For the specific reactions in Eq. (9) and the stoichiometric matrices in Eq. (10), the simultaneous equations in Eq. (22) are written as

(y1+2y2y3μ~1y1y2+y3+μ~2)=(00).\displaystyle\begin{pmatrix}-y_{1}+2y_{2}-y_{3}-\tilde{\mu}_{1}\\ y_{1}-y_{2}+y_{3}+\tilde{\mu}_{2}\end{pmatrix}=\begin{pmatrix}0\\ 0\end{pmatrix}. (29)

The solutions y={yi}y=\{y_{i}\} are expressed as

(y1y2y3)=(hμ~1μ~2μ~12μ~2+h),\displaystyle\begin{pmatrix}y_{1}\\ y_{2}\\ y_{3}\end{pmatrix}=\begin{pmatrix}-h\\ \tilde{\mu}_{1}-\tilde{\mu}_{2}\\ \tilde{\mu}_{1}-2\tilde{\mu}_{2}+h\end{pmatrix}, (30)

where hh\in\mathbb{R} is the coordinate of Ker[ST]\mathrm{Ker}[S^{T}]. The set of solutions in Eq. (30) represents the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}).

Second, using Eqs (15) and (20), the full grand potential density φ(y)\varphi^{*}(y) is obtained for the ideal gas or solution as

φ(y)=RT~ie{yiνio(T~)}/RT~+RT~me{μ~mμmo(T~)}/RT~.\displaystyle\varphi^{*}(y)=R\tilde{T}\sum_{i}e^{\left\{y_{i}-\nu_{i}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}+R\tilde{T}\sum_{m}e^{\left\{\tilde{\mu}_{m}-\mu_{m}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}. (31)

Using the solutions in Eq. (30), φ(y)\varphi^{*}(y) is represented on EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) as

φ¯(h)\displaystyle\bar{\varphi}^{*}(h) =RT~[e{hν1o(T~)}/RT~+e{μ~1μ~2ν2o(T~)}/RT~\displaystyle=R\tilde{T}\biggl{[}e^{\left\{-h-\nu_{1}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}+e^{\left\{\tilde{\mu}_{1}-\tilde{\mu}_{2}-\nu_{2}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}
+e{μ~12μ~2+hν3o(T~)}/RT~]\displaystyle+e^{\left\{\tilde{\mu}_{1}-2\tilde{\mu}_{2}+h-\nu_{3}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}\biggr{]}
+RT~[e{μ~1μ1o(T~)}/RT~+e{μ~2μ2o(T~)}/RT~].\displaystyle+R\tilde{T}\left[e^{\left\{\tilde{\mu}_{1}-\mu_{1}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}+e^{\left\{\tilde{\mu}_{2}-\mu_{2}^{o}\left(\tilde{T}\right)\right\}/R\tilde{T}}\right]. (32)

In Fig. 2(a), we numerically plot the pressure φ¯(h)\bar{\varphi}^{*}(h) on EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). Here, we have a unique hmin=(ν1o+ν3oμ~1+2μ~2)/2h^{\mathrm{min}}=(-\nu^{o}_{1}+\nu^{o}_{3}-\tilde{\mu}_{1}+2\tilde{\mu}_{2})/2 that attains the minimum φ¯(h)\bar{\varphi}^{*}(h). Substituting it into Eq. (30), we obtain yminy^{\mathrm{min}} and the minimum pressure φ(ymin)\varphi^{*}(y^{\mathrm{min}}). With the parameters given in the caption, φ(ymin)=14.25\varphi^{*}(y^{\mathrm{min}})=14.25.

To demonstrate that the fate of the system is classified by yminy^{\mathrm{min}} and the conserved quantities LL in claim 1, we show numerical simulations of the dynamics, Eq. (3). From the solution X(t)X(t), the volume Ω(X)\Omega(X) of the system is determined by the equation of state, Eq. (17). To numerically solve Eq. (3) with the stoichiometric matrix SS in Eq. (10), we employ the mass-action kinetics with the local detailed balance condition to specify the flux function J(t)J(t) (see Appendix B).

Claim 1: In Fig. 2(b-d), we show numerical trajectories of the volume Ω(X)\Omega(X) from three initial conditions for different pressures Π~\tilde{\Pi}. The color of curves corresponds to each initial condition. The fate of the system changes with Π~\tilde{\Pi} as follows. When Π~=13<φ(ymin)=14.25\tilde{\Pi}=13<\varphi^{*}(y^{\mathrm{min}})=14.25 (Fig. 2(b)), the volume of the system increases with time from any initial conditions. As Π~\tilde{\Pi} increases to Π~=φ(ymin)=14.25\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}})=14.25 (Fig. 2(c)), the trajectory from an initial condition with L=0L=0 in light blue achieves an equilibrium state, whereas trajectories from the other two with L0L\neq 0 continue growing. As Π~\tilde{\Pi} increases further, Π~=16>φ(ymin)=14.25\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}})=14.25 (Fig. 2(d)), the trajectory with L=0L=0 in light blue decreases and the system shrinks. By contrast, those from the other two with L0L\neq 0 achieve equilibrium states. These numerical results demonstrate the claim 1.

Claim 2: In Fig. 2(e,f), we demonstrate the time evolution of the number XX. When Π~=φ(ymin)=14.25\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}})=14.25 and L=0L=0, the equilibrium state to which the system converges depends on the initial conditions in STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0). Furthermore, it indeed lies on the ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}), which is given by Eq. (21) (see Fig. 2(e)). In our numerical simulation, we compute it as follows. From yminy^{\mathrm{min}}, we obtain the corresponding density xmini=iφ(ymin)=exp{(yiminνio)/RT~}x^{i}_{\mathrm{min}}=\partial^{i}\varphi^{*}(y^{\mathrm{min}})=\exp\{(y^{\mathrm{min}}_{i}-\nu_{i}^{o})/R\tilde{T}\} by using Eq. (31). The ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}) is plotted by the one which includes xminx_{\mathrm{min}}. When Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}) and L=520L=52\neq 0, the system indeed converges to the unique equilibrium state, irrespective of the initial conditions in STO𝔛(L=52)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=52) (see Fig. 2(f)).

Refer to caption
Figure 2: (a) The pressure φ¯(h)\bar{\varphi}^{*}(h) in Eq. (32) is shown by the red curve. The coordinate hh gives each element yEQ𝒴(μ~)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) in Eq. (30). The unique hminh^{\mathrm{min}} for the minimum pressure is obtained as hmin=(ν1o+ν3oμ~1+2μ~2)/2=0.981h^{\mathrm{min}}=(-\nu^{o}_{1}+\nu^{o}_{3}-\tilde{\mu}_{1}+2\tilde{\mu}_{2})/2=0.981. The corresponding yy is ymin=(0.981,0.288,1.099)y^{\mathrm{min}}=(-0.981,-0.288,1.099) and φ¯(hmin)=φ(ymin)=14.25\bar{\varphi}^{*}(h^{\mathrm{min}})=\varphi^{*}(y^{\mathrm{min}})=14.25. The values at φ¯(h)=13,14.25\bar{\varphi}^{*}(h)=13,14.25, and 1616 are shown in dotted, solid, and dashed gray lines, respectively. (b-d) Numerical plot of the time evolution for the reactions in Eq. (9) with the assumptions of ideal gas and mass-action kinetics. The volume Ω(X)\Omega(X) is shown for three different pressures (b) Π~=13\tilde{\Pi}=13, (c) Π~=14.25\tilde{\Pi}=14.25, and (d) Π~=16\tilde{\Pi}=16. In (b-d), different colors indicate different initial conditions: (X01,X02,X03)=(48,1,48)(X_{0}^{1},X_{0}^{2},X_{0}^{3})=(48,1,48) (light blue), (148,25,200)(148,25,200) (orange), (90,5,20)(90,5,20) (green). Here, the initial conditions correspond to the conserved quantities L=0L=0, L=52L=52 and L=70L=-70, respectively, where LL is calculated as L=UX0=X01+X03L=UX_{0}=-X^{1}_{0}+X^{3}_{0}. (e,f) The time evolution of XX is shown in the number space 𝔛\mathfrak{X} for three initial conditions. (e) We set that all the initial conditions have L=0L=0 and are given by (X01,X02,X03)=(48,1,48)(X_{0}^{1},X_{0}^{2},X_{0}^{3})=(48,1,48) (light blue), (25,60,25)(25,60,25) (light green), (10,50,10)(10,50,10) (light purple). Each initial condition is indicated by a circle in the corresponding color. From each initial condition, the curve is constrained in STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0) and finally converges to the square when Π~=14.25\tilde{\Pi}=14.25. All the squares are located on the red ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}). The black and red circles on this ray indicate the origin X=0X=0 and xmini=iφ(ymin)=exp{(yiminνio)/RT~}x^{i}_{\mathrm{min}}=\partial^{i}\varphi^{*}(y^{\mathrm{min}})=\exp\{(y^{\mathrm{min}}_{i}-\nu^{o}_{i})/R\tilde{T}\}, respectively. (f) We set that the three initial conditions have L=52L=52 and are given by (X01,X02,X03)=(148,25,200)(X_{0}^{1},X_{0}^{2},X_{0}^{3})=(148,25,200) (orange), (10,175,62)(10,175,62) (blue), and (1,10,53)(1,10,53) (purple), respectively. Each of them is indicated by a circle in the corresponding color. The curve is constrained in STO𝔛(L=52)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=52) and finally converges to the unique point (square). The parameters are fixed as R=T~=1R=\tilde{T}=1, xo1=8x^{1}_{o}=8, xo2=7x^{2}_{o}=7, xo3=1x^{3}_{o}=1, no1=2n^{1}_{o}=2, n~1=1\tilde{n}^{1}=1, no2=3n^{2}_{o}=3, n~2=2\tilde{n}^{2}=2. Given the parameters, the rate constants w^+r\hat{w}^{r}_{+} and w^r\hat{w}^{r}_{-} are determined by Eq. (80) with w^r=1\hat{w}^{r}_{-}=1 in the Appendix B.

II.7 Outline of the derivations of claims 1 and 2, and Corollary 1

The fates of the system stated in claims 1 and 2, and Corollary 1 are summarized in Table 1. Before closing this section, we outline their derivations, which will be presented in the subsequent sections. They will be derived in two steps. In the first step, we investigate the existence of an equilibrium state, that is the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) in Eq. (27). In Sec. III, we will show that the intersection is not empty in the case 22 for L=0L=0 and in the case 33 for L0L\neq 0 of claim 1, and is empty for the other cases. This will be summarized in Theorem 1. Following Theorem 1, we derive claim 2 from the correspondence in Eq. (21) between the number space 𝔛\mathfrak{X} and the chemical potential space 𝒴\mathcal{Y}. In the second step, when the intersection is empty, we investigate the landscape of the total entropy function Σtot\Sigma^{\mathrm{tot}} to classify whether the system grows or shrinks in Sec. IV. We will show that the system has two possibilities for L=0L=0: it grows when Σtot\Sigma^{\mathrm{tot}} is not bounded above, or it shrinks when the supremum of Σtot\Sigma^{\mathrm{tot}} is at the origin X=0X=0. By contrast, the system always grows for L0L\neq 0 when the intersection is empty because Σtot\Sigma^{\mathrm{tot}} is not bounded above. This will be summarized in Theorem 2. Mathematical proofs of Theorems 1 and 2, and Corollary 1 will be presented in Sec. V.

III Geometric representation of equilibrium states

For the derivation of claims 1 and 2, we analyze the accessible regions in the number space 𝔛\mathfrak{X}, the density space 𝒳\mathcal{X} and the chemical potential space 𝒴\mathcal{Y} by using a geometric technique. In Sec. III.1, we give the assumptions to the full grand potential density φ(y)\varphi^{*}(y) for the following mathematical treatment. In Sec. III.2, we show that the accessible region is restricted by the isobaric condition, which we term the isobaric manifold. This manifold is further partitioned by the conservation laws in Sec. III.3. After defining the equilibrium manifold as the candidates of chemical equilibrium states in Sec. III.4, we characterize the equilibrium state to which the system converges as an intersecting point of the accessible region and the equilibrium manifold in Sec. III.5. Furthermore, by using numerical simulations, we identify the conditions for the existence of the intersecting point, which we state as Theorem 1.

III.1 Assumptions to φ(y)\varphi^{*}(y)

The thermodynamic property of the system is encoded in the full grand potential density φ(y)\varphi^{*}(y) (see Eq. (20)). Here, y𝒴=𝒩Xy\in\mathcal{Y}=\mathbb{R}^{\mathcal{N}_{X}} is the chemical potential of the confined chemicals, and φ(y)\varphi^{*}(y) represents the pressure of the system at the state yy. In this work, we assume that φ(y)\varphi^{*}(y) is a smooth and strictly convex function on 𝒴=𝒩X\mathcal{Y}=\mathbb{R}^{\mathcal{N}_{X}}, and the image of the associated Legendre transformation,

φ(y):y𝒴φ(y)={iφ}={φyi},\displaystyle\partial\varphi^{*}(y):y\in\mathcal{Y}\mapsto\partial\varphi^{*}(y)=\left\{\partial^{i}\varphi^{*}\right\}=\left\{\frac{\partial\varphi^{*}}{\partial y_{i}}\right\}, (33)

is equal to 𝒳=>0𝒩X\mathcal{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}}. In addition, we assume the following properties; (1) φ(y)\varphi^{*}(y) strictly increases with yiy_{i} for an arbitrary fixed {yj}ji\{y_{j}\}_{j\neq i}. (2) iφ(y)0\partial^{i}\varphi^{*}(y)\rightarrow 0 when yiy_{i}\rightarrow-\infty. They are satisfied in most thermodynamic systems such as the ideal gas in Eq. (31).

Furthermore, we assume that the mapping φ:𝒩X>0𝒩X\partial\varphi^{*}:\mathbb{R}^{\mathcal{N}_{X}}\rightarrow\mathbb{R}_{>0}^{\mathcal{N}_{X}} is bijective, and we consider that its inverse map is given by φ:>0𝒩X𝒩X\partial\varphi:\mathbb{R}_{>0}^{\mathcal{N}_{X}}\rightarrow\mathbb{R}^{\mathcal{N}_{X}} (see Eq. (13) for φ(x)\varphi(x)). By using this, the above property (2) is rephrased as yi=iφ(x)y_{i}=\partial_{i}\varphi(x)\rightarrow-\infty when xi0x^{i}\rightarrow 0.

Finally, we assume that the pressure of the environment Π~\tilde{\Pi} is greater than Πmin\Pi_{\mathrm{min}}, to guarantee that the volume Ω(X)\Omega(X) is uniquely determined (see Appendix C). Here, Πmin\Pi_{\mathrm{min}} denotes the minimum pressure that the system can take as

Πmin:=infy𝒴φ(y)=lim{yi}{}φ(y).\displaystyle\Pi_{\mathrm{min}}:=\inf_{y\in\mathcal{Y}}\varphi^{*}(y)=\lim_{\{y_{i}\}\rightarrow\{-\infty\}}\varphi^{*}(y). (34)

The second equality holds from the above property (1).

III.2 The isobaric manifold and the projection ρ𝒳\rho_{\mathcal{X}}

We remind that the density xx is given by the mapping ρ𝒳\rho_{\mathcal{X}} in Eq. (18):

ρ𝒳:=X𝒳ρ𝒳(X)=x(X)={XiΩ(X)}𝒳.\displaystyle\rho_{\mathcal{X}}:=X\in\mathcal{X}\mapsto\rho_{\mathcal{X}}(X)=x(X)=\left\{\frac{X^{i}}{\Omega(X)}\right\}\in\mathcal{X}. (35)

The range of the mapping ρ𝒳\rho_{\mathcal{X}} describes possible states of the density xx under the isobaric condition. To calculate the range, we compute the volume Ω\Omega by solving the minimization in Eq. (14). Its critical equation is given by

φ(XΩ)XiΩiφ(XΩ)+Π~=0.\displaystyle\varphi\left(\frac{X}{\Omega}\right)-\frac{X^{i}}{\Omega}\partial_{i}\varphi\left(\frac{X}{\Omega}\right)+\tilde{\Pi}=0. (36)

Therefore, the possible region in 𝒳\mathcal{X} under the isobaric condition is constrained in

𝒳(Π~,μ~):={x|φ(x)xiiφ(x)+Π~=0,x>0},\displaystyle\mathcal{I}^{\mathcal{X}}\left(\tilde{\Pi},\tilde{\mu}\right):=\Set{x}{\varphi\left(x\right)-x^{i}\partial_{i}\varphi(x)+\tilde{\Pi}=0,x>0}, (37)

which we call the isobaric manifold. Since the range of ρ𝒳\rho_{\mathcal{X}} is represented in Eq. (37), we find that the mapping ρ𝒳(X)\rho_{\mathcal{X}}(X) is a projection of X𝔛X\in\mathfrak{X} into 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) along the ray which includes XX.

Example 3: For the ideal gas case, by substituting φ(x)\varphi(x) in Eq. (15) into Eq. (37), we obtain

𝒳(Π~,μ~)={x|ixi=Π~RT~mn~m,x>0},\displaystyle\mathcal{I}^{\mathcal{X}}\left(\tilde{\Pi},\tilde{\mu}\right)=\Set{x}{\sum_{i}x^{i}=\frac{\tilde{\Pi}}{R\tilde{T}}-\sum_{m}\tilde{n}^{m},x>0}, (38)

where n~m\tilde{n}^{m} is the density of the open chemicals in the environment (see Eq. (16)). Thus, the isobaric manifold in 𝒳\mathcal{X} represents a simplex for the ideal gas case (Fig. 3(a)).

The mapping ρ𝒳\rho_{\mathcal{X}} is the projection of X𝔛=>0𝒩𝒳X\in\mathfrak{X}=\mathbb{R}^{\mathcal{N}_{\mathcal{X}}}_{>0} into this simplex (see Fig. 3(b)).

\blacksquare

Refer to caption
Figure 3: (a) For the ideal gas case, the isobaric manifold 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) is represented by a simplex in gray. (b) The mapping ρ𝒳(X)\rho_{\mathcal{X}}(X) gives the projection of X𝔛X\in\mathfrak{X} into 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) along the ray which includes XX. The thick arrow is pointing the corresponding density x=ρ𝒳(X)x=\rho_{\mathcal{X}}(X). (c) For the reactions in Eq. (9), the stoichiometric manifold STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L) in Eq. (40) represents a two dimensional plane with U=(1,0,1)U=(-1,0,1). By applying the mapping ρ𝒳(X)\rho_{\mathcal{X}}(X) to each point XSTO𝔛(L)X\in\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L), we get the corresponding manifold, STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) (see the upper region of the simplex in (d) colored by orange). (d) In the present case, the isobaric manifold 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) is partitioned by STO𝒳(L>0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L>0), STO𝒳(L<0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L<0), and STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0). Here, STO𝒳(L>0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L>0) and STO𝒳(L<0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L<0) are given by the orange and the green regions, respectively, and their interface is given by STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0) in the light blue line. (e) The stoichiometric manifold STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) is characterized by a ray 𝔯\mathfrak{r} in the space \mathcal{L} of the conserved quantities (see the line below Eq. (6) for \mathcal{L}). Since dimKer[ST]=1\dim\mathrm{Ker}[S^{T}]=1 in the present example, we have three characterizations L=0L=0, L>0L>0 and L<0L<0. In Sec. I of the Supplementary material, we investigate another example with a higher dimension of Ker[ST]\mathrm{Ker}[S^{T}]. (f) Given a ray 𝔯\mathfrak{r}, any plane STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L) for L𝔯L\in\mathfrak{r} is mapped to the same region on the simplex by the mapping ρ𝒳\rho_{\mathcal{X}}. Any plane with L>0L>0 (e.g. in orange) is mapped to the upper region of the simplex (see (d)), whereas, any plane with L<0L<0 (e.g. in green) is mapped to the lower region. For L=0L=0, the plane in light blue is mapped to the interface between the regions for L>0L>0 and L<0L<0 (see the light blue line in (d)).

Finally, by using the mapping φ\partial\varphi, the isobaric manifold in 𝒴\mathcal{Y} is represented as

𝒴(Π~,μ~):=φ(𝒳)={y|φ(y)Π~=0},\displaystyle\mathcal{I}^{\mathcal{Y}}\left(\tilde{\Pi},\tilde{\mu}\right):=\partial\varphi\left(\mathcal{I}^{\mathcal{X}}\right)=\Set{y}{\varphi^{*}\left(y\right)-\tilde{\Pi}=0}, (39)

which is straightforward to interpret that the pressure of the system φ(y)\varphi^{*}(y) at the chemical potential yy is balanced with Π~\tilde{\Pi} in this manifold.

III.3 The stoichiometric manifold as partition of the isobaric manifold

The accessible subspace of the state X(t)X(t) from an initial condition X0X_{0} in Eq. (4) is

STO𝔛(L):={X|UilXi=UilX0i=Ll,X>0},\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L):=\Set{X}{U^{l}_{i}X^{i}=U^{l}_{i}X^{i}_{0}=L^{l},X>0}, (40)

which we call the stoichiometric manifold in 𝔛\mathfrak{X}.

By applying the mapping ρ𝒳\rho_{\mathcal{X}} to STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L), we have the stoichiometric manifold in the density space 𝒳\mathcal{X} as

STO𝒳(L):=\displaystyle\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L):= {x|φ(x)xiiφ(x)+Π~=0,\displaystyle\left.\biggl{\{}x\right|\varphi(x)-x^{i}\partial_{i}\varphi(x)+\tilde{\Pi}=0,
Uilxi=Llα,x>0},\displaystyle U^{l}_{i}x^{i}=\frac{L^{l}}{\alpha},x>0\biggr{\}}, (41)

where α>0\alpha>0 is arbitrary. Using the mapping φ\partial\varphi, the stoichiometric manifold in the chemical potential space 𝒴\mathcal{Y} is given by

STO𝒴(L):={y|φ(y)Π~=0,Uiliφ(y)=Llα}.\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L):=\Set{y}{\varphi^{*}(y)-\tilde{\Pi}=0,U^{l}_{i}\partial^{i}\varphi^{*}(y)=\frac{L^{l}}{\alpha}}. (42)

Example 4: For the specific stoichiometric matrix in Eq. (10), STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) is schematically depicted in Fig. 3(c,d) by applying the mapping ρ𝒳\rho_{\mathcal{X}} to STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L).

\blacksquare

For further investigations, we define the following manifold in 𝒳\mathcal{X} as

^STO𝒳(L^):={x|Uilxi=L^l,x>0},\displaystyle\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}\left(\hat{L}\right):=\Set{x}{U^{l}_{i}x^{i}=\hat{L}^{l},x>0}, (43)

and that in 𝒴\mathcal{Y} as

^STO𝒴(L^)\displaystyle\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}\left(\hat{L}\right) :=φ(^STO𝒳(L^))\displaystyle:=\partial\varphi\left(\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}\left(\hat{L}\right)\right)
={y|Uiliφ(y)=L^l}.\displaystyle=\Set{y}{U^{l}_{i}\partial^{i}\varphi^{*}(y)=\hat{L}^{l}}. (44)

These manifolds correspond to the stoichiometric manifolds with conserved quantities L^\hat{L} under the isochoric condition [78]. Hereafter, we term the isochoric stoichiometric manifolds.

Using these manifolds, the stoichiometric manifold under isobaric condition is represented as follows:

STO𝒳(L)=α>0^STO𝒳(Lα)𝒳(Π~,μ~),\displaystyle\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L)=\bigcup_{\alpha>0}\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}\left(\frac{L}{\alpha}\right)\cap\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}), (45)

and

STO𝒴(L)=α>0^STO𝒴(Lα)𝒴(Π~,μ~).\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)=\bigcup_{\alpha>0}\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}\left(\frac{L}{\alpha}\right)\cap\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}). (46)

For arbitrary c>0c>0, we find STO𝒳(cL)=STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(cL)=\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) and, correspondingly, STO𝒴(cL)=STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(cL)=\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L). Hence, each ray 𝔯\mathfrak{r} in the space \mathcal{L} of conserved quantities characterizes these manifolds (see also Example 5 below and Fig. 3(d-f)). To be more precise, for any L,L𝔯L,L^{\prime}\in\mathfrak{r}, STO𝒳(L)=STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L)=\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L^{\prime}) holds. We denote the stoichiometric manifold corresponding to a ray 𝔯\mathfrak{r} as STO𝒳(L𝔯)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L\in\mathfrak{r}).

When L=0L=0, the stoichiometric manifold in 𝒳\mathcal{X} is written as

STO𝒳(L=0)=^STO𝒳(L^=0)𝒳(Π~,μ~).\displaystyle\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0)=\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}\left(\hat{L}=0\right)\cap\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}). (47)

Since the dimension of α\alpha in Eq. (45) vanishes for STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0), we obtain dimSTO𝒳(L=0)=dimSTO𝒳(L0)1\dim\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0)=\dim\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L\neq 0)-1. In addition, for any given L0L\neq 0, we find that ^STO𝒳(L/α)^STO𝒳(L^=0)\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}\left(L/\alpha\right)\rightarrow\hat{\mathcal{M}}^{\mathcal{X}}_{\mathrm{STO}}(\hat{L}=0) when α\alpha\rightarrow\infty. Thus, the stoichiometric manifold STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0) forms an interface of all the stoichiometric manifolds STO𝒳(L0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L\neq 0). These properties hold similarly in 𝒴\mathcal{Y} by replacing the superscript 𝒳\mathcal{X} with 𝒴\mathcal{Y}.

The above statements are summarized as follows.

Lemma 1

Let 𝔯\mathfrak{r} denote a ray in the space \mathcal{L} of conserved quantities. For every 𝔯\mathfrak{r}, the corresponding stoichiometric manifold STO𝒳(L𝔯)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L\in\mathfrak{r}) exists and the set of the stoichiometric manifolds partitions the isobaric manifold 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}). In addition, their interface is given by the stoichiometric manifold STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0). The isobaric manifold in the chemical potential space 𝒴\mathcal{Y} is similarly partitioned by stoichiometric manifolds (replace the superscript 𝒳\mathcal{X} with 𝒴\mathcal{Y}).

Example 5: In Fig. 3(d-f), we schematically depict the partition of the isobaric manifold 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) by the stoichiometric manifolds STO𝒳(L𝔯)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L\in\mathfrak{r}) for the ideal gas case.

\blacksquare

III.4 The equilibrium manifold

In this subsection, we confirm that the candidates of the equilibrium states are given by Eq. (23).

The equilibrium states of the chemical reaction dynamics are given by the maxima of the total entropy with respect to the extent of reaction Ξ\Xi. From the differentiation of Eq. (12) and using Eq. (14), we have the critical equation as

ΣtotΞr\displaystyle\frac{\partial\Sigma^{\mathrm{tot}}}{\partial\Xi^{r}} =1T~{iφ(XΩ(X))Sri+μ~mOrm}\displaystyle=-\frac{1}{\tilde{T}}\left\{\partial_{i}\varphi\left(\frac{X}{\Omega(X)}\right)S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}\right\}
=0,\displaystyle=0, (48)

where X=X0+SΞX=X_{0}+S\Xi in Eq. (4) (see Appendix D).

Thus, at any equilibrium states, the density xx satisfies

iφ(x)Sri+μ~mOrm=0.\partial_{i}\varphi(x)S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0. (49)

By using the correspondence yi=iφ(x)y_{i}=\partial_{i}\varphi(x), it is represented as

yiSri+μ~mOrm=0,y_{i}S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0, (50)

which confirms the balance of chemical potentials between reactants and products at the chemical equilibrium (see Eq. (22)).

The set of solutions to Eq. (49) represents the candidates of the equilibrium states in the density space 𝒳\mathcal{X}, which we call an equilibrium manifold:

EQ𝒳(T~,μ~):={x|iφ(x)Sri+μ~mOrm=0},\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}):=\Set{x}{\partial_{i}\varphi(x)S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0}, (51)

and in the chemical potential space 𝒴\mathcal{Y} as

EQ𝒴(μ~):={y|yiSri+μ~mOrm=0}.\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}):=\Set{y}{y_{i}S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}=0}. (52)

We note that the equilibrium manifold is an affine subspace in 𝒴\mathcal{Y}, whereas it is generally curved in 𝒳\mathcal{X}. In addition, using the basis matrix UU of Ker[ST]\mathrm{Ker}[S^{T}] and solving Eq. (50), the manifold in Eq. (52) is rewritten as

EQ𝒴(μ~)={y|yi=yiP+hlUil,hdimKer[ST]},\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\Set{y}{y_{i}=y_{i}^{\mathrm{P}}+h_{l}U^{l}_{i},h\in\mathbb{R}^{\mathrm{dim}\mathrm{Ker}[S^{T}]}}, (53)

where yPy^{\mathrm{P}} is a particular solution to Eq. (50) and h={hl}h=\{h_{l}\} represents a coordinate of Ker[ST]\mathrm{Ker}[S^{T}].

III.5 The equilibrium state as an intersection

If the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is not empty, the system converges to the equilibrium state. We can compute the equilibrium state in 𝒴\mathcal{Y} by using Eqs. (42) and (52). Similarly, we obtain the corresponding state in 𝒳\mathcal{X} as STO𝒳(L)EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) by using Eq. (41) and (51). In this subsection, we numerically confirm the convergence in 𝒳\mathcal{X} by the dynamics in Eq. (3) using the specific setup in Sec. II.6.

We first consider the equilibrium manifold. The points in EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) are written in Eq. (30). Here, the second component y2y_{2} is constant p:=μ~1μ~2p:=\tilde{\mu}_{1}-\tilde{\mu}_{2}. Thus, EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is a one-dimensional line located on the plane y2=py_{2}=p in 𝒴\mathcal{Y} (see the left panel of Fig. 4 (a) for the plane y2=py_{2}=p, and the red line in the left panel of Fig. 4 (b) for EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})). By applying the mapping φ\partial\varphi^{*}, we can compute EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}). For the ideal gas case in Eq. (31), the mapping is represented as iφ(y)=exp{(yiνio)/RT~}\partial^{i}\varphi^{*}(y)=\exp\{(y_{i}-\nu_{i}^{o})/R\tilde{T}\}. Thus, the points in EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) are computed as

(x1x2x3)\displaystyle\begin{pmatrix}x^{1}\\ x^{2}\\ x^{3}\end{pmatrix} =(exp{(hν1o)/RT~}exp{(μ~1μ~2ν2o)/RT~}exp{(μ~12μ~2+hν3o)/RT~}),\displaystyle=\begin{pmatrix}\exp\{(-h-\nu_{1}^{o})/R\tilde{T}\}\\ \exp\{(\tilde{\mu}_{1}-\tilde{\mu}_{2}-\nu_{2}^{o})/R\tilde{T}\}\\ \exp\{(\tilde{\mu}_{1}-2\tilde{\mu}_{2}+h-\nu_{3}^{o})/R\tilde{T}\}\end{pmatrix}, (54)

where hh\in\mathbb{R} plays a role of the coordinate of EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}). We note that the second component x2x^{2} is also constant q:=φ(p)=exp{(μ~1μ~2ν2o)/RT~}q:=\partial\varphi^{*}(p)=\exp\{(\tilde{\mu}_{1}-\tilde{\mu}_{2}-\nu_{2}^{o})/R\tilde{T}\}. Thus, EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) is a one-dimensional curve located on the plane x2=qx^{2}=q. (see the red curve in the right panel of Fig. 4(b)).

We next consider whether the intersection 𝒳(Π~,μ~)EQ𝒳(T~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu})\cap\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) is empty or not. As the pressure Π~\tilde{\Pi} increases, the number of intersecting points change as follows. When Π~=13<φ(ymin)=14.25\tilde{\Pi}=13<\varphi^{*}(y^{\mathrm{min}})=14.25, the intersecting point 𝒳(Π~,μ~)EQ𝒳(T~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu})\cap\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) does not exist (see the right panel of Fig. 4(b)). When Π~=φ(ymin)\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}}), the intersection appears as the unique contact point corresponding to yminy^{\mathrm{min}}. When Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}), the intersection consists of two points. The above situation holds similarly in 𝒴\mathcal{Y} (see the left panel of Fig. 4(b)).

We further depict the intersection 𝒳(Π~,μ~)EQ𝒳(T~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu})\cap\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) by the three dimensional plot in Fig. 5. From Fig. 3(d), 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) is partitioned by the stoichiometric manifolds. When Π~=φ(ymin)=14.25\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}})=14.25 (Fig. 5(b)), the intersection lies on STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0). When Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}), the two intersecting points belong to each of STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) for L>0L>0 and L<0L<0, respectively (see Fig. 5(c)). Thus, for a given LL, the intersecting point is uniquely determined and it depends on a ray in the space \mathcal{L} of the conserved quantities (see Fig. 3(e) and Lemma 1).

In Fig. 5(b,c), we also show the time evolution of the system in the numerical simulation. Indeed, the system converges to the intersection STO𝒳(L)EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}).

The above numerical result is an instance of the following theorem.

Theorem 1

The system converges to the intersecting point STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) if it exists. Its existence is classified as follows:

  1. 1.

    φ(ymin)Π~>0\varphi^{*}\left(y^{\mathrm{min}}\right)-\tilde{\Pi}>0 \Rightarrow STO𝒴(L)EQ𝒴(μ~)=\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\emptyset.

  2. 2.

    φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 \Rightarrow

    STO𝒴(L)EQ𝒴(μ~)={{ymin} for L=0, for L0.\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\begin{cases}\{y^{\mathrm{min}}\}&\mbox{ for }L=0,\\ \emptyset&\mbox{ for }L\neq 0.\end{cases}
  3. 3.

    φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 \Rightarrow

    STO𝒴(L)EQ𝒴(μ~)={ for L=0,{yEQ} for L0,\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\begin{cases}\emptyset&\mbox{ for }L=0,\\ \{y^{\mathrm{EQ}}\}&\mbox{ for }L\neq 0,\end{cases}

    where the intersecting point yEQy^{\mathrm{EQ}} is uniquely determined for a given LL, and it depends on a ray in the space \mathcal{L} of conserved quantities.

A mathematical proof of this theorem will be shown in Sec. V.

Refer to caption
Figure 4: (a) For the ideal gas case, the isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) in the chemical potential space 𝒴\mathcal{Y} is a level set for φ(y)\varphi^{*}(y) (left), whereas 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) represents a simplex in the density space 𝒳\mathcal{X} (right). We consider the planes y2=p=μ~1μ~2y_{2}=p=\tilde{\mu}_{1}-\tilde{\mu}_{2} in 𝒴\mathcal{Y} and x2=q=exp{(μ~1μ~2ν2o)/RT~}x^{2}=q=\exp\{(\tilde{\mu}_{1}-\tilde{\mu}_{2}-\nu_{2}^{o})/R\tilde{T}\} in 𝒳\mathcal{X}, respectively. The black curve in 𝒴\mathcal{Y} and the line in 𝒳\mathcal{X} indicate the cross section of the isobaric manifolds by the planes, respectively. (b) In the left panel, we plot EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (the red line) and 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) (gray curves) on the plane y2=py_{2}=p. In the right panel, EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) (the red curve) and 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) (gray lines) are plotted on the plane x2=qx^{2}=q. In 𝒴\mathcal{Y} (left), the three gray curves correspond to the isobaric manifolds 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) with the pressures Π~=13\tilde{\Pi}=13 (dotted), 14.2514.25 (solid) and 1616 (dashed), respectively. When Π~=13\tilde{\Pi}=13, the dotted curve does not have the intersection with EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (red line). When Π~=14.25\tilde{\Pi}=14.25, the solid curve has the unique intersecting point with the red line, denoted by the light blue square. It is located at ymin=(0.981,0.288,1.099)y^{\mathrm{min}}=(-0.981,-0.288,1.099), where yminy^{\mathrm{min}} has been obtained in Fig. 2. When Π~=16\tilde{\Pi}=16, the dashed curve has two intersecting points with the red line. The heatmaps represent the values of φ(y)φ(ymin)\varphi^{*}(y)-\varphi^{*}(y^{\mathrm{min}}). In the right panel, we show the corresponding manifolds in 𝒳\mathcal{X}. Here, the light blue square is located at xmin=(3.00,5.25,3.00)x_{\mathrm{min}}=(3.00,5.25,3.00), which is obtained by xmini=iφ(ymin)=exp{(yiminνio)/RT~}x^{i}_{\mathrm{min}}=\partial^{i}\varphi^{*}(y^{\mathrm{min}})=\exp\{(y^{\mathrm{min}}_{i}-\nu_{i}^{o})/R\tilde{T}\}.

From Theorem 1, we can derive claim 2 as follows. For the case 1 of claim 2 in which φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 and L=0L=0, Theorem 1 states that the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) consists of yminy^{\mathrm{min}}. From Eq. (21), this yminy^{\mathrm{min}} is mapped to a ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}) in the number space 𝔛\mathfrak{X}. In addition, STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0) from Eq. (40) indicates that αXSTO𝔛(L=0)\alpha X\in\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0) if XSTO𝔛(L=0)X\in\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0) for arbitrary α>0\alpha>0. This means that all the points on the ray 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}) is in STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0). Thus, all the points on the ray are equilibrium states and the system converges to one of them. The state is not uniquely determined only by the entropy function and depends on the initial conditions and the functional form of J(t)J(t). For the case 2 of claim 2 in which φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 and L0L\neq 0, Theorem 1 states that the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) exists as yEQy^{\mathrm{EQ}}. This also corresponds to a ray 𝔯𝒴(yEQ)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{EQ}}) in the number space 𝔛\mathfrak{X}. Since STO𝔛(L0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L\neq 0) is an affine subspace which does not include the origin X=0X=0, the intersection between STO𝔛(L0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L\neq 0) and an arbitrary ray is represented by a point. Thus, the equilibrium state is uniquely determined in this case.

Theorem 1 states the cases when the system converges to an equilibrium state in our claim 1. In the next section, we will investigate the case when the system does not converge to an equilibrium state, that is STO𝒴(L)EQ𝒴(μ~)=\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\emptyset.

Refer to caption
Figure 5: (a) When Π~=13<φ(ymin)=14.25\tilde{\Pi}=13<\varphi^{*}(y^{\mathrm{min}})=14.25, EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) (red curve) and 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) (gray simplex) do not have an intersecting point. The black line represents 𝒳(Π~,μ~)\mathcal{I}^{\mathcal{X}}(\tilde{\Pi},\tilde{\mu}) on the plane x2=qx^{2}=q (see also Fig. 4). (b) When Π~=14.25=φ(ymin)\tilde{\Pi}=14.25=\varphi^{*}(y^{\mathrm{min}}), the intersection consists of the unique point xminx_{\mathrm{min}}. It lies on the stoichiometric manifold STO𝒳(L=0)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L=0) (the light blue line on the simplex). In numerical simulations, the time evolution of the system is also shown in a light blue line with arrows. From the initial condition (circle), it converges to the point xminx_{\mathrm{min}} (square). (c) When Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}), the intersection consists of two points (orange and green squares). They are located in the upper and the lower regions of the simplex, which correspond to STO𝒳(L)\mathcal{M}^{\mathcal{X}}_{\mathrm{STO}}(L) for L>0L>0 and L<0L<0, respectively (see Fig. 3(d)). We also show the time evolution of the system from the two initial conditions (circles). They respectively converge to the intersecting points (squares). In the numerical simulation, the initial conditions are given by (X01,X02,X03)=(48,1,48)(X^{1}_{0},X^{2}_{0},X^{3}_{0})=(48,1,48) (light blue circle), (148,25,200)(148,25,200) (orange circle) and (90,5,20)(90,5,20) (green circle).

IV Form of the total entropy function

When the intersection EQ𝒴(μ~)STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) is empty, the system cannot converge to an equilibrium state, and it is expected that the system grows or shrinks. In this section, we observe if the system grows or shrinks by numerical plots of the total entropy function.

If STO𝔛(L)EQ𝔛(T~,μ~)=\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathfrak{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu})=\emptyset, the total entropy function Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) can not have the maximum inside of STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L) (see Sec. III.4). Thus, the functional form of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) has the following two possibilities: (1) Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above. (2) Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is bounded above, but the supremum exists on the boundary of STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L). In the first case, we find |Ξ(t)||\Xi(t)|\rightarrow\infty for tt\rightarrow\infty, because the second law requires that the system must climb up the unbounded landscape of the total entropy function. Since we assume dimKer[S]=0\dim\mathrm{Ker}[S]=0, we get |X||X|\rightarrow\infty for |Ξ||\Xi|\rightarrow\infty. In addition, Ω(X)\Omega(X)\rightarrow\infty for |X||X|\rightarrow\infty (see Appendix E). Thus, if Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above, the volume Ω\Omega grows. In the second case, we can further show that the supremum is possible only at the origin X=0X=0 (see Sec. V.2). Thus, the volume Ω\Omega is shrinking and finally vanishes.

In the following part of this section, we numerically plot the landscape of the total entropy function to determine the fate of the system. In Fig. 6, we plot the total entropy function Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) in Eq. (12) for our example setup (see Example 1 and Example 2). Following Theorem 1, the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is empty for the cases (1) φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0, (2) φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 and L0L\neq 0, and (3) φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 and L=0L=0,

In the case (1) Π~=13<φ(ymin)=14.25\tilde{\Pi}=13<\varphi^{*}(y^{\mathrm{min}})=14.25, the total entropy function Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above for both of L=0L=0 and L0L\neq 0 (see Fig. 6(a)). Indeed, |Ξ(t)||\Xi(t)|\rightarrow\infty in the reaction dynamics of Eq. (3). In the case (2) Π~=φ(ymin)\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}}) and L=520L=52\neq 0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is still not bounded above and |Ξ(t)||\Xi(t)|\rightarrow\infty (see the right panel of Fig. 6(b)). In the case (3) Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}) and L=0L=0, the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is at the origin of the number space 𝔛\mathfrak{X}. The system tends to approach the origin (see the left panel of Fig. 6(c)).

Refer to caption
Figure 6: The entropy landscapes Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) are shown for (a) Π~=13<φ(ymin)=14.25\tilde{\Pi}=13<\varphi^{*}(y^{\mathrm{min}})=14.25, (b) Π~=φ(ymin)\tilde{\Pi}=\varphi^{*}(y^{\mathrm{min}}) and (c) Π~=16>φ(ymin)\tilde{\Pi}=16>\varphi^{*}(y^{\mathrm{min}}). The left and right panels show the cases L=0L=0 and L=520L=52\neq 0, respectively. The heatmaps represent the values of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi). Here, we note that the domain of (Ξ1,Ξ2)(\Xi^{1},\Xi^{2}) is restricted by X>0𝒩XX\in\mathbb{R}^{\mathcal{N}_{X}}_{>0}. The boundaries at which X2=0X^{2}=0 and X1=0X^{1}=0 are represented by dashed and dotted black lines, respectively. The black circles in the left panels denote the origin X=0X=0. The time evolutions of the system are shown by light blue and orange curves for the two initial conditions: (left) X0=(48,1,48)X_{0}=(48,1,48) with L=0L=0 and (right) X0=(148,25,200)X_{0}=(148,25,200) with L=52L=52. (a) The heatmaps in both panels indicate unbounded landscapes and the system goes to infinity. (b) For L=0L=0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) has maxima which form the red ray. The system converges to a point (the light blue square) on the ray. For L0L\neq 0, the heatmap shows unbounded landscape and the system goes to infinity. (c) For L=0L=0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) has the supremum at the origin X=0X=0 and the system is approaching this point. For L0L\neq 0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) has the unique maximum (the orange square) and the system converges to the point. See the caption in Fig. 2 for specific values of the parameters.

The above numerical result demonstrates the following theorem.

Theorem 2

When STO𝒴(L)EQ𝒴(μ~)=\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\emptyset, the landscape of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is given as follows:

  1. 1.

    If φ(ymin)Π~>0\varphi^{*}\left(y^{\mathrm{min}}\right)-\tilde{\Pi}>0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above.

  2. 2.

    If φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0 and L0L\neq 0, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above.

  3. 3.

    If φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0 and L=0L=0, the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is at the origin of the number space 𝔛\mathfrak{X}.

A mathematical proof of this theorem will be shown in Sec. V.

Theorem 1 and 2 lead to the claim 1 and we find the fate of the system under isobaric condition.

Before closing this section, we confirm the claim 2 in the left panel of Fig.6(b) and the right panel of Fig.6(c). In the former panel, the red ray represents the set of the maxima of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi), which is 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}). We can see the system converges to a point on the ray. In the latter case, the maximum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is the unique point and the system converges to it.

V Proof of the Theorems

In this section, we prove Theorems 1 and 2, and Corollary 1.

V.1 Proof of Theorem 1

We first provide a general proof valid for arbitrary dimensions. After concluding the proof, we schematically illustrate it for three dimensional space of 𝒴\mathcal{Y} in Fig. 7.

To investigate the existence of the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}), we introduce Birth’s trajectory.

Let yB(L^)y^{\mathrm{B}}(\hat{L}) represent the unique intersecting point between the isochoric stoichiometric manifold ^STO𝒴(L^)\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(\hat{L}) in Eq. (44) and the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) in Eq. (52):

{yB(L^)}:=^STO𝒴(L^)EQ𝒴(μ~).\displaystyle\left\{y^{\mathrm{B}}\left(\hat{L}\right)\right\}:=\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}\left(\hat{L}\right)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}\left(\tilde{\mu}\right). (55)

The uniqueness of yB(L^)y^{\mathrm{B}}(\hat{L}) is guaranteed by Birch’s theorem and the point is known as Birth’s point [83, 12, 17, 78, 84] (see Appendix F). We define Birch’s trajectory by a collection of Birch’s point along a ray in the space of conserved quantities. For given L0L\neq 0, it is defined as

𝔟(L):={yB(Lα)|α>0}.\displaystyle\mathfrak{b}(L):=\Set{y^{\mathrm{B}}\left(\frac{L}{\alpha}\right)}{\alpha>0}. (56)

We find that yB(L/α)yB(0)y^{\mathrm{B}}(L/\alpha)\rightarrow y^{\mathrm{B}}(0) for α\alpha\rightarrow\infty. Furthermore, from Eq. (94) in Appendix F, we get

{yB(0)}=^STO𝒴(L^=0)EQ𝒴(μ~)={ymin}.\displaystyle\{y^{\mathrm{B}}(0)\}=\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}\left(\hat{L}=0\right)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}\left(\tilde{\mu}\right)=\{y^{\mathrm{min}}\}. (57)

Thus, the point yminy^{\mathrm{min}} is located at one of the end points of Birch’s trajectory, but is not included in the trajectory.

For this Birch’s trajectory, the following lemma is satisfied.

Lemma 2

φ(y)\varphi^{*}(y) is a strictly increasing function along Birch’s trajectory from the starting point yminy^{\mathrm{min}}.

A proof of this lemma is shown in Appendix G.

By employing Birch’s trajectory, we can represent the intersection for L0L\neq 0 as

STO𝒴(L)EQ𝒴(μ~)\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})
=α>0^STO𝒴(Lα)𝒴(Π~,μ~)EQ𝒴(μ~)\displaystyle=\bigcup_{\alpha>0}\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}\left(\frac{L}{\alpha}\right)\cap\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu})\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})
=𝔟(L)𝒴(Π~,μ~),\displaystyle=\mathfrak{b}(L)\cap\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}), (58)

where we use Eqs. (46), (55), and (56). For L=0L=0, the intersection is

STO𝒴(L=0)EQ𝒴(μ~)\displaystyle\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})
=^STO𝒴(L=0)𝒴(Π~,μ~)EQ𝒴(μ~)\displaystyle=\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\cap\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu})\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})
={ymin}𝒴(Π~,μ~),\displaystyle=\{y^{\mathrm{min}}\}\cap\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}), (59)

where we use Eqs. (47) and (57).

From Lemma 2, we can determine the existence of the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) by the position of the starting point yminy^{\mathrm{min}}. When φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0, the starting point yminy^{\mathrm{min}} is located in the superlevel set: {y|φ(y)>Π~}\{y|\varphi^{*}(y)>\tilde{\Pi}\}. Taking into account the fact that 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) is the level set, the intersecting point STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) does not exist for any LL. When φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0, the starting point yminy^{\mathrm{min}} is on 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}). For L=0L=0, the intersecting point uniquely exists as yminy^{\mathrm{min}} because of Eq. (59). For L0L\neq 0, Lemma 2 indicates φ(y)>φ(ymin)=Π~\varphi^{*}(y)>\varphi^{*}(y^{\mathrm{min}})=\tilde{\Pi} at any point yy on Birch’s trajectory 𝔟(L)\mathfrak{b}(L). Thus, from Eq. (58), the intersecting point does not exist. When φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, the starting point yminy^{\mathrm{min}} is located in the sublevel set: {y|φ(y)<Π~}\{y|\varphi^{*}(y)<\tilde{\Pi}\}. For L=0L=0, the intersecting point does not exist. For L0L\neq 0, the intersecting point uniquely exists.

This concludes the proof of Theorem 1.

In Fig. 7, we provide schematic illustrations in three dimensional space of 𝒴\mathcal{Y} for the case φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0. To demonstrate the generality of the proof, we visualize two cases with dimKer[ST]=2\dim\mathrm{Ker}[S^{T}]=2 and dimKer[ST]=1\dim\mathrm{Ker}[S^{T}]=1 in the panels (a) and (b), respectively. From Eq. (6), the case in (a) has two conserved quantities, whereas the case in (b) has one conserved quantity. We also have dimIm[S]=1\dim\mathrm{Im}[S]=1 and dimIm[S]=2\dim\mathrm{Im}[S]=2 for the two cases, respectively, because dim𝒴=dimKer[ST]+dimIm[S]\dim\mathcal{Y}=\dim\mathrm{Ker}[S^{T}]+\dim\mathrm{Im}[S] and dim𝒴=3\dim\mathcal{Y}=3. The dimensions for the example in Eqs (9) and (10) correspond to the case (b), whereas the case (a) corresponds to the one investigated in Sec. I of the Supplementary material.

In the former case (Fig. 7(a)), the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is represented by the two dimensional plane in red, whereas the isochoric stoichiometric manifold ^STO𝒴(L^)\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(\hat{L}) is given by a one-dimensional curve in light blue because dimEQ𝒴(μ~)=dimKer[ST]\dim\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\dim\mathrm{Ker}[S^{T}] and dim^STO𝒴(L/α)=dimIm[S]\dim\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha)=\dim\mathrm{Im}[S]. Accordingly, the black circles represent the intersecting points yB(L^)y^{\mathrm{B}}(\hat{L}) in Eq. (55) for L^=L/α\hat{L}=L/\alpha with varying α\alpha. The green curve indicates Birch’s trajectory 𝔟(L)\mathfrak{b}(L) in Eq. (56), and one of the end points is located at yminy^{\mathrm{min}} in Eq. (57). The isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) is represented by the gray surface, which is a level set for φ(y)\varphi^{*}(y). When φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, yminy^{\mathrm{min}} is in the sublevel set {y|φ(y)<Π~}\{y|\varphi^{*}\left(y\right)<\tilde{\Pi}\}. The stoichiometric manifold STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) is given by the orange curve from Eq. (46). Thus, the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is represented by the red circle, which is also the intersecting point between 𝔟(L)\mathfrak{b}(L) and 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) in Eq. (58).

In the latter case (Fig. 7(b)), EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is represented by the red line, whereas ^STO𝒴(L^)\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(\hat{L}) is given by the two dimensional surface in light blue. The intersecting point yB(L^)y^{\mathrm{B}}(\hat{L}) is indicated by the black circle in Eq. (55) for L^=L/α\hat{L}=L/\alpha with a given α\alpha. In this case, the Birch’s trajectory 𝔟(L)\mathfrak{b}(L) in Eq. (56) overlaps with the red line, and one of the end points is located at yminy^{\mathrm{min}} in Eq. (57). The stoichiometric manifold STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) is given by the orange surface from Eq. (46)555Similar to Fig. 3(d), the isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) is partitioned by L>0L>0 and L<0L<0 in this case.. Thus, the intersection STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is represented by the red circle.

Refer to caption
Figure 7: Schematic illustration of the manifolds in 𝒴\mathcal{Y} for the case φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0. In this illustration, we set dim𝒴=dimKer[ST]+dimIm[S]=3\dim\mathcal{Y}=\dim\mathrm{Ker}[S^{T}]+\dim\mathrm{Im}[S]=3. The conserved quantities LL are generally represented by L={Ll=UilX0i}L=\{L^{l}=U^{l}_{i}X^{i}_{0}\} (l=1,,dimKer[ST]l=1,...,\dim\mathrm{Ker}[S^{T}]). The panels (a) and (b) represent the case for dimKer[ST]=2\dim\mathrm{Ker}[S^{T}]=2 and dimIm[S]=1\dim\mathrm{Im}[S]=1, and the case for dimKer[ST]=1\dim\mathrm{Ker}[S^{T}]=1 and dimIm[S]=2\dim\mathrm{Im}[S]=2, respectively. We note that dimEQ𝒴(μ~)=dimKer[ST]\dim\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\dim\mathrm{Ker}[S^{T}] and dim^STO𝒴(L/α)=dimIm[S]\dim\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha)=\dim\mathrm{Im}[S]. The dimensions in the case (b) correspond to the illustrative example of chemical reactions in Eqs. (9) and (10), whereas those in the case (a) correspond to the one in Sec. I of the Supplementary material. (a) In this case, dimEQ𝒴(μ~)=2\dim\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=2 and dim^STO𝒴(L/α)=1\dim\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha)=1. Thus, EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is denoted by the red plane, and the light blue curves represent ^STO𝒴(L/α)\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha) for α(0,)\alpha\in(0,\infty). The gray surface is 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}). From Eq. (46), STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) is given by the orange curve. Equation (56) enables us to depict Birch’s trajectory 𝔟(L)\mathfrak{b}(L) as the green curve. From Eq. (58), the red circle denotes the intersecting point STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) to which the system converges as the equilibrium state. (b) In this case, we consider dimEQ𝒴(μ~)=1\dim\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=1 and dim^STO𝒴(L/α)=2\dim\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha)=2. Thus, EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is denoted by the red line, and ^STO𝒴(L/α)\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha) represents the two dimensional manifold in light blue. Thus, STO𝒴(L)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L) is given by the orange surface. We note that Birch’s trajectory 𝔟(L)\mathfrak{b}(L) overlaps with the red line of EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). The red circle represents the equilibrium state.

V.2 Proof of Theorem 2

If STO𝔛(L)EQ𝔛(T~,μ~)=\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathfrak{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu})=\emptyset, the system never relaxes to the equilibrium state and have the following two possibilities: (1) Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above, and |Ξ(t)||\Xi(t)|\rightarrow\infty with time. As we noted in Sec. IV, the volume Ω\Omega grows in this setup. (2) Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is bounded above and the supremum exists on the boundary of STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L).

First, we consider the case L0L\neq 0, where STO𝔛(L0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L\neq 0) does not contact with the origin. In this case, Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) is not bounded above. To prove that, we will show that the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) does not exist on the boundary of STO𝔛(L0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L\neq 0) as follows.

The extent of reaction Ξ\Xi provides a coordinate on the stoichiometric manifold STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L). Since 𝔛=>0𝒩X\mathfrak{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}}, the domain of Ξ\Xi has a boundary. We denote the coordinate of a point on a boundary by ΞB\Xi_{B}. Here, at least, one component of the vector XB:=X0+SΞBX_{B}:=X_{0}+S\Xi_{B} is zero. We define the index set by B(XB)={i|XBi=0}\mathfrak{I}_{B}(X_{B})=\{i|X^{i}_{B}=0\}. Since the origin X=0X=0 is excluded, B(XB)\mathfrak{I}_{B}(X_{B}) never has all the indices. Also, the density xB=XB/Ω(XB)x_{B}=X_{B}/\Omega(X_{B}) has zero components as xBi=0x^{i}_{B}=0 for iB(XB)i\in\mathfrak{I}_{B}(X_{B}).

To investigate the gradient of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) on the boundary, we calculate the thermodynamic force as

fr(Ξ):=ΣtotΞr=1T~{iφ(x(X0+SΞ))Sri+μ~mOrm}.\displaystyle f_{r}(\Xi):=\frac{\partial\Sigma^{\mathrm{tot}}}{\partial\Xi^{r}}=-\frac{1}{\tilde{T}}\left\{\partial_{i}\varphi(x(X_{0}+S\Xi))S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}\right\}. (60)

If the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) does not exist on the boundary, we can find a direction from the boundary to the interior of STO𝔛(L)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L) such that Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) increases. To be more precise, by using fr(Ξ)f_{r}(\Xi), this statement is mathematically phrased as follows: If the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) does not exist on the boundary, a vector j={jr}j=\{j^{r}\} exists such that fr(ΞB)jr>0f_{r}(\Xi_{B})j^{r}>0 and Srijr>0S^{i}_{r}j^{r}>0 for all the indices iBi\in\mathfrak{I}_{B}. In the next paragraph, we prove the above statement.

From the assumptions to φ(y)\varphi^{*}(y) in Sec. III.1, for every ii, yi=iφ(x)y_{i}=\partial_{i}\varphi(x)\rightarrow-\infty when xi0x^{i}\rightarrow 0. Thus, iφ(xB)\partial_{i}\varphi(x_{B})\rightarrow-\infty for all iBi\in\mathfrak{I}_{B}. Choose a vector jj satisfying Srijr>0S^{i}_{r}j^{r}>0 for iBi\in\mathfrak{I}_{B}. Then, we can show that, when ΞΞB\Xi\rightarrow\Xi_{B}, fr(Ξ)jrf_{r}(\Xi)j^{r}\rightarrow\infty from Eq. (60). This indicates that the supremum of Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) does not exist on the boundary.

Second, we consider the case L=0L=0. By employing rays on STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0), we investigate whether Σtot(X)\Sigma^{\mathrm{tot}}(X) is bounded above or not.

Since we have assumed dimKer[S]=0\dim\mathrm{Ker}[S]=0, we can rewrite Σtot(Ξ)\Sigma^{\mathrm{tot}}(\Xi) as a function of XX. Using a particular solution yPy^{\mathrm{P}} to Eq. (50), we get

Σtot(X)=\displaystyle\Sigma^{\mathrm{tot}}(X)= 1T~{Ω(X)φ(XΩ(X))+Π~Ω(X)\displaystyle-\frac{1}{\tilde{T}}\biggl{\{}\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\tilde{\Pi}\Omega(X)
yiP(XiX0i)}+const,\displaystyle-y_{i}^{\mathrm{P}}\left(X^{i}-X^{i}_{0}\right)\biggr{\}}+\mathrm{const},
=\displaystyle= Ω(X)T~K𝒴(yP;y(X))yiPX0iT~+const,\displaystyle\frac{\Omega(X)}{\tilde{T}}K^{\mathcal{Y}}(y^{\mathrm{P}};y(X))-\frac{y^{\mathrm{P}}_{i}X^{i}_{0}}{\tilde{T}}+\mathrm{const}, (61)

where

K𝒴(yP;y):=φ(yP)Π~𝒟𝒴[yP||y],K^{\mathcal{Y}}(y^{\mathrm{P}};y):=\varphi^{*}(y^{\mathrm{P}})-\tilde{\Pi}-\mathcal{D}^{\mathcal{Y}}\left[y^{\mathrm{P}}||y\right], (62)

(see Appendix H for the details of the calculation). Here, 𝒟𝒴[yP||y]\mathcal{D}^{\mathcal{Y}}[y^{\mathrm{P}}||y] is the Bregman divergence [86, 87, 88] induced by φ(y)\varphi^{*}(y), which is defined as

𝒟𝒴[y||y]=φ(y)φ(y)iφ(y){yiyi},\mathcal{D}^{\mathcal{Y}}[y||y^{\prime}]=\varphi^{*}(y)-\varphi^{*}(y^{\prime})-\partial^{i}\varphi^{*}(y^{\prime})\left\{y_{i}-y^{\prime}_{i}\right\}, (63)

and y(X)=φρ𝒳(X)y(X)=\partial\varphi\circ\rho_{\mathcal{X}}(X). Although Eq. (61) is apparently a function of the particular solution yPy^{\mathrm{P}}, the value of Σtot\Sigma^{\mathrm{tot}} does not depend on the choice of yPy^{\mathrm{P}} as shown in Appendix I. Since only the first term Ω(X)K𝒴(yP;y(X))/T~\Omega(X)K^{\mathcal{Y}}(y^{\mathrm{P}};y(X))/\tilde{T} depends on XX, we consider its landscape.

The stoichiometric manifold with L=0L=0, STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0), can be described by the collection of rays in 𝔛\mathfrak{X} because it contacts with the origin X=0X=0. Since a ray in the number space 𝔛\mathfrak{X} corresponds to a point in the chemical potential space 𝒴\mathcal{Y} as in Eq. (21), the value of K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is constant on each ray. In addition, Ω(X)\Omega(X) is increasing along each ray from the origin. Therefore, by considering the sign of K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y), we can determine if Σtot(X)\Sigma^{\mathrm{tot}}(X) increases along each ray in STO𝔛(L=0)\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}(L=0). If K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is positive for a ray, Σtot(X)\Sigma^{\mathrm{tot}}(X) increases along the ray. If K𝒴(yP;y)=0K^{\mathcal{Y}}(y^{\mathrm{P}};y)=0, Σtot(X)\Sigma^{\mathrm{tot}}(X) is constant along the ray. If K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is negative for a ray, Σtot(X)\Sigma^{\mathrm{tot}}(X) decreases along the ray.

When φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0, we can find a point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) such that K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is positive (see Appendix J). If such a point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) exists, we can consider a corresponding ray in 𝔛\mathfrak{X} to the point yy. Along the ray, Σtot(X)\Sigma^{\mathrm{tot}}(X) increases and, thus, is not bounded above. When φ(ymin)Π~=0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}=0, K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is negative for any ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) except y=yminy=y^{\mathrm{min}}, and K𝒴(yP;ymin)=0K^{\mathcal{Y}}(y^{\mathrm{P}};y^{\mathrm{min}})=0. Thus, the maxima of Σtot(X)\Sigma^{\mathrm{tot}}(X) forms the ray corresponding to yminy^{\mathrm{min}}, that is 𝔯𝒴(ymin)\mathfrak{r}^{\mathcal{Y}}(y^{\mathrm{min}}) in Eq. (21). When φ(ymin)Π~<0\varphi^{*}({y^{\mathrm{min}}})-\tilde{\Pi}<0, we can show that any ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) gives negative K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) (see Appendix K). Thus, Σtot(X)\Sigma^{\mathrm{tot}}(X) decreases along any ray and the supremum is located at the origin X=0X=0.

Summarizing the cases for L0L\neq 0 and L=0L=0, we obtain Theorem 2.

V.3 Proof of Corollary 1

When SS is regular, the accessible region covers the whole space of 𝔛\mathfrak{X}, that is STO𝔛=𝔛\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}}=\mathfrak{X}. Thus, STO𝔛\mathcal{M}^{\mathfrak{X}}_{\mathrm{STO}} can be described by the collection of rays in 𝔛\mathfrak{X}. This situation coincides with the one for L=0L=0 in the previous subsection. Therefore, by employing a similar proof, we can classify the fate of the system as Corollary 1 (see also [19]).

VI Summary and Discussions

In this work, we have considered chemical thermodynamics of growing CRSs with stoichiometric conservation laws in which the stoichiometric matrix SS has a nontrivial left kernel space. By clarifying the conditions for the fate of the system to grow, shrink, or equilibrate, we show that the existence of conservation laws qualitatively alter the fate of the system. It is emphasized again that our results are derived by general thermodynamic and stoichiometric structures without assuming any specific thermodynamic potentials or reaction kinetics; i.e., they are obtained based solely on the second law of thermodynamics and conservation laws.

Since we are mainly interested in the growth of the system, we have assumed the productivity of the stoichiometric matrix SS: Im[S]>0𝒩X\mathrm{Im}[S]\cap\mathbb{R}_{>0}^{\mathcal{N}_{X}}\neq\emptyset (see Eq. (7)). This condition allows the time evolution of X(t)X(t) so as to perpetually increase the number of all confined chemicals while satisfying the conservation laws. In the following two paragraphs, we comment on the fate of the system with the other classes of the stoichiometric matrix. Here, the complement of the productive SS is defined as Im[S]>0𝒩X=\mathrm{Im}[S]\cap\mathbb{R}_{>0}^{\mathcal{N}_{X}}=\emptyset. We divide it into two cases: (1) Im[S]0𝒩X=\mathrm{Im}[S]\cap\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}=\emptyset, and (2) Im[S]0𝒩X\mathrm{Im}[S]\cap\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}\neq\emptyset and Im[S]>0𝒩X=\mathrm{Im}[S]\cap\mathbb{R}_{>0}^{\mathcal{N}_{X}}=\emptyset 666Note the equality in 0\geq 0 of 0𝒩X\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}..

In the first case, (1) Im[S]0𝒩X=\mathrm{Im}[S]\cap\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}=\emptyset, a vector 𝐯\mathbf{v} exists in Ker[ST]\mathrm{Ker}[S^{T}] such that all the components of 𝐯\mathbf{v} are positive. As a result, the conserved quantity 𝐯iXi=𝐯iX0i\mathbf{v}_{i}X^{i}=\mathbf{v}_{i}X^{i}_{0} implies that the total mass of the confined chemicals are conserved, and the system must converge to the equilibrium state, i.e., the growth of the system is not possible. We refer to this type of SS as non-productive. A numerical example for this case is shown in Sec. III of the Supplementary material.

In the second case, (2) Im[S]0𝒩X\mathrm{Im}[S]\cap\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}\neq\emptyset and Im[S]>0𝒩X=\mathrm{Im}[S]\cap\mathbb{R}_{>0}^{\mathcal{N}_{X}}=\emptyset, a vector 𝐯\mathbf{v} exists in Ker[ST]\mathrm{Ker}[S^{T}] such that all the components of 𝐯\mathbf{v} are non-negative, but Ker[ST]\mathrm{Ker}[S^{T}] does not have a vector 𝐯\mathbf{v}^{\prime} such that all the components of 𝐯\mathbf{v}^{\prime} are positive. This means that the mass conservation law exists for a proper subset of the confined chemicals. Thus, the volume growth of the system can still be allowed with the increase of the confined chemicals that do not participate in the mass conservation law. Although such a situation may not be biologically relevant, we refer to this type of SS as semi-productive. A numerical example for this case is shown in Sec. IV of the Supplementary material. According to the example, it is expected that the claims 1 and 2 still hold for the semi-productive cases.

We can possibly correlate our theoretical findings with empirical observations. Our findings suggest that the fate changes with the initial conditions (conserved quantities LL) even in the same environment. We especially find that, as shown in Table 1, a CRS does not vanish and is more likely to grow when it has non-zero LL than when it has L=0L=0 or a regular SS. If we suppose a population of the CRSs with varying initial conditions, and consider their fates in a slowly changing environment, surviving CRSs can be selective depending on LL, where a CRS with non-zero LL might be advantageous. Identifying a possible mechanism to survive in a varying environment is biologically essential, and our findings may help explain the outcome.

Nevertheless, several extensions of our theory remain to be essential as future work to understand actual experimental situations of growing protocells or biological cells.

The first extension is to consider the case when the system may relax to a state that continuously produces entropy with constant volume, namely, the conventional nonequilibrium steady state (NESS) [16, 14, 15, 10, 17, 13, 12, 9]. It is possible when the matrix SS has a nontrivial right kernel space, i.e., the assumption in Eq. (5) is not satisfied. It is a major challenge to understand how the growth of the system can be characterized and realized in such situations, and compatible with the NESS.

The second one is to consider a different hierarchy of the timescales in which the relaxation of the variables (E,Ω,N)(E,\Omega,N) may not be rapid compared to the timescale of chemical reactions. Although the present work assumes that the chemical reactions are slow (see the assumptions made above Eq. (12)), there may be cases in which the timescale of other processes is slow and/or comparable with that of the chemical reactions. A typical example is when we consider cell transport processes [90]. In this case, the timescale of material exchanges between the system and the environment is relevant, and the processes may also couple with chemical reactions. In addition, the osmotic gradient (pressure difference) can play a driving force to change the volume, whereas isosmotic volume changes are also commonly known and widely investigated as a relevant process by alterations in intracellular solute content [91, 92]. It is of interest to see how the conditions for the growth of the system change with relevant processes of the thermodynamic setup.

The third one is to consider explicitly the effect of the membrane molecules. Although we neglect the tension of the membrane in this paper (see Fig. 1), it could be relevant and changes with time because the membrane molecules themselves are produced and supplied by the CRSs in biological cells. It is important to investigate if and how the effect changes the results by extending our theoretical framework.

Alternatively, it would also be promising to apply our framework to fit into an experimental technique using membrane-free compartments such as droplets based on liquid-liquid phase separation [93]. Since its nature, a rigorous thermodynamic treatment would require an extended framework in which the entropy function is not strictly concave (see the assumption made below Eq. (2)).

In all the above extensions, the present work provides basic results to clarify the role of conservation laws to the growth of protocells and biological cells.

VII Acknowledgments

The authors thank Dimitri Loutchko and Shuhei A. Horiguchi for discussions. This research is supported by JSPS KAKENHI Grant Numbers 19H05799, 21K21308, and 24K00542, and by JST CREST JPMJCR2011 and JPMJCR1927. Y. S. receives financial support from the Public\Private R&D Investment Strategic Expansion PrograM (PRISM) and programs for Bridging the gap between R&D and the IDeal society (society 5.0) and Generating Economic and social value (BRIDGE) from Cabinet Office.

References

  • Callen [1985] H. B. Callen, Thermodynamics and an Introduction to Thermostatistics (John wiley & sons, New York, 1985).
  • Kondepudi and Prigogine [1998] D. Kondepudi and I. Prigogine, Modern thermodynamics: from heat engines to dissipative structures (John Wiley & Sons, New York, 1998).
  • von Neumann and Burks [1966] J. von Neumann and A. W. Burks, Theory of Self-Reproducing Automata (University of Illinois Press, Urbana, IL, 1966).
  • Freitas Jr and Merkle [2004] R. A. Freitas Jr and R. C. Merkle, Kinematic Self-Replication Machines (Landes Bioscience, Georgetown,TX, 2004).
  • De Jong et al. [2017] H. De Jong, S. Casagranda, N. Giordano, E. Cinquemani, D. Ropers, J. Geiselmann, and J.-L. Gouzé, Mathematical modelling of microbes: metabolism, gene expression and growth, Journal of The Royal Society Interface 14, 20170502 (2017).
  • de Groot et al. [2020] D. H. de Groot, J. Hulshof, B. Teusink, F. J. Bruggeman, and R. Planqué, Elementary growth modes provide a molecular description of cellular self-fabrication, PLoS computational biology 16, e1007559 (2020).
  • Müller [2021] S. Müller, Elementary growth modes/vectors and minimal autocatalytic sets for kinetic/constraint-based models of cellular growth, Biorxiv , 2021.02.24.432769 (2021).
  • Müller et al. [2022] S. Müller, D. Széliová, and J. Zanghellini, Elementary vectors and autocatalytic sets for resource allocation in next-generation models of cellular growth, PLOS Computational Biology 18, e1009843 (2022).
  • Horn and Jackson [1972] F. Horn and R. Jackson, General mass action kinetics, Archive for rational mechanics and analysis 47, 81 (1972).
  • Qian and Reluga [2005] H. Qian and T. C. Reluga, Nonequilibrium thermodynamics and nonlinear kinetics in a cellular signaling switch, Physical review letters 94, 028101 (2005).
  • Beard and Qian [2008] D. A. Beard and H. Qian, Chemical biophysics: quantitative analysis of cellular systems, Vol. 126 (Cambridge University Press Cambridge, 2008).
  • Craciun et al. [2009] G. Craciun, A. Dickenstein, A. Shiu, and B. Sturmfels, Toric dynamical systems, Journal of Symbolic Computation 44, 1551 (2009).
  • Pérez Millán et al. [2012] M. Pérez Millán, A. Dickenstein, A. Shiu, and C. Conradi, Chemical reaction systems with toric steady states, Bulletin of mathematical biology 74, 1027 (2012).
  • Polettini and Esposito [2014] M. Polettini and M. Esposito, Irreversible thermodynamics of open chemical networks. i. emergent cycles and broken conservation laws, The Journal of chemical physics 141 (2014).
  • Ge and Qian [2016] H. Ge and H. Qian, Nonequilibrium thermodynamic formalism of nonlinear chemical reaction systems with waage–guldberg’s law of mass action, Chemical Physics 472, 241 (2016).
  • Rao and Esposito [2016] R. Rao and M. Esposito, Nonequilibrium thermodynamics of chemical reaction networks: Wisdom from stochastic thermodynamics, Physical Review X 6, 041064 (2016).
  • Craciun et al. [2019] G. Craciun, S. Muller, C. Pantea, and P. Y. Yu, A generalization of birchs theorem and vertex-balanced steady states for generalized mass-action systems, Mathematical Biosciences and Engineering 16, 8243 (2019).
  • Qian and Ge [2021] H. Qian and H. Ge, Stochastic Chemical Reaction Systems in Biology (Springer, 2021).
  • Sughiyama et al. [2022a] Y. Sughiyama, A. Kamimura, D. Loutchko, and T. J. Kobayashi, Chemical thermodynamics for growing systems, Physical Review Research 4, 033191 (2022a).
  • Blokhuis et al. [2020] A. Blokhuis, D. Lacoste, and P. Nghe, Universal motifs and the diversity of autocatalytic systems, Proceedings of the National Academy of Sciences 117, 25230 (2020).
  • Shinar and Feinberg [2010] G. Shinar and M. Feinberg, Structural sources of robustness in biochemical reaction networks, Science 327, 1389 (2010).
  • Hirono et al. [2023] Y. Hirono, A. Gupta, and M. Khammash, Complete characterization of robust perfect adaptation in biochemical reaction networks, arXiv preprint arXiv:2307.07444  (2023).
  • Schuster and Höfer [1991] S. Schuster and T. Höfer, Determining all extreme semi-positive conservation relations in chemical reaction systems: a test criterion for conservativity, Journal of the Chemical Society, Faraday Transactions 87, 2561 (1991).
  • Schuster and Hilgetag [1995] S. Schuster and C. Hilgetag, What information about the conserved-moiety structure of chemical reaction systems can be derived from their stoichiometry?, The Journal of Physical Chemistry 99, 8017 (1995).
  • De Martino and Marsili [2005] A. De Martino and M. Marsili, Typical properties of optimal growth in the von neumann expanding model for large random economies, Journal of Statistical Mechanics: Theory and Experiment 2005, L09003 (2005).
  • De Martino et al. [2009] A. De Martino, C. Martelli, and F. A. Massucci, On the role of conserved moieties in shaping the robustness and production capabilities of reaction networks, Europhysics Letters 85, 38007 (2009).
  • De Martino et al. [2012] A. De Martino, E. Marinari, and A. Romualdi, Von neumann’s growth model: Statistical mechanics and biological applications, The European Physical Journal Special Topics 212, 45 (2012).
  • Rao and Esposito [2018a] R. Rao and M. Esposito, Conservation laws and work fluctuation relations in chemical reaction networks, The Journal of chemical physics 149 (2018a).
  • Rao and Esposito [2018b] R. Rao and M. Esposito, Conservation laws shape dissipation, New Journal of Physics 20, 023007 (2018b).
  • Pekař [2021] M. Pekař, Non-equilibrium thermodynamics view on kinetics of autocatalytic reactions—two illustrative examples, Molecules 26, 585 (2021).
  • Hofmeyr et al. [1986] J.-H. S. Hofmeyr, H. Kacser, and K. J. van der Merwe, Metabolic control analysis of moiety-conserved cycles, European Journal of Biochemistry 155, 631 (1986).
  • Famili and Palsson [2003] I. Famili and B. O. Palsson, The convex basis of the left null space of the stoichiometric matrix leads to the definition of metabolically meaningful pools, Biophysical journal 85, 16 (2003).
  • Imieliński et al. [2006] M. Imieliński, C. Belta, H. Rubin, and Á. Halász, Systematic analysis of conservation relations in escherichia coli genome-scale metabolic network reveals novel growth media, Biophysical journal 90, 2659 (2006).
  • Shinar et al. [2009] G. Shinar, U. Alon, and M. Feinberg, Sensitivity and robustness in chemical reaction networks, SIAM Journal on Applied Mathematics 69, 977 (2009).
  • Martelli et al. [2009] C. Martelli, A. De Martino, E. Marinari, M. Marsili, and I. Pérez Castillo, Identifying essential genes in escherichia coli from a metabolic optimization principle, Proceedings of the National Academy of Sciences 106, 2607 (2009).
  • De Martino et al. [2014] A. De Martino, D. De Martino, R. Mulet, and A. Pagnani, Identifying all moiety conservation laws in genome-scale metabolic networks, PloS one 9, e100750 (2014).
  • Haraldsdóttir and Fleming [2016] H. S. Haraldsdóttir and R. M. Fleming, Identification of conserved moieties in metabolic networks by graph theoretical analysis of atom transition networks, PLOS Computational Biology 12, e1004999 (2016).
  • Kamei et al. [2023] K.-i. F. Kamei, K. J. Kobayashi-Kirschvink, T. Nozoe, H. Nakaoka, M. Umetani, and Y. Wakamoto, Raman spectra and gene expression correspondences reveal global stoichiometry conservation architecture in cells, bioRxiv , 2023.05.09.539921 (2023).
  • Eigen and Schuster [1979] M. Eigen and P. Schuster, The Hypercycle: A Principle of Natural Self-organization (Springer-Verlag, Berlin, 1979).
  • Kauffman [1986] S. A. Kauffman, Autocatalytic sets of proteins, Journal of theoretical biology 119, 1 (1986).
  • Jain and Krishna [1998] S. Jain and S. Krishna, Autocatalytic sets and the growth of complexity in an evolutionary model, Physical Review Letters 81, 5684 (1998).
  • Segré et al. [2000] D. Segré, D. Ben-Eli, and D. Lancet, Compositional genomes: prebiotic information transfer in mutually catalytic noncovalent assemblies, Proceedings of the National Academy of Sciences 97, 4112 (2000).
  • Andrieux and Gaspard [2008] D. Andrieux and P. Gaspard, Nonequilibrium generation of information in copolymerization processes, Proceedings of the National Academy of Sciences 105, 9516 (2008).
  • Kita et al. [2008] H. Kita, T. Matsuura, T. Sunami, K. Hosoda, N. Ichihashi, K. Tsukada, I. Urabe, and T. Yomo, Replication of genetic information with self-encoded replicase in liposomes, ChemBioChem 9, 2403 (2008).
  • Rasmussen et al. [2008] S. Rasmussen et al.Protocells: Bridging Nonliving and Living Matter (The MIT Press, Cambridge, MA, 2008).
  • Noireaux et al. [2011] V. Noireaux, Y. T. Maeda, and A. Libchaber, Development of an artificial cell, from self-organization to computation and self-reproduction, Proceedings of the National Academy of Sciences 108, 3473 (2011).
  • Kurihara et al. [2011] K. Kurihara, M. Tamura, K.-i. Shohda, T. Toyota, K. Suzuki, and T. Sugawara, Self-reproduction of supramolecular giant vesicles combined with the amplification of encapsulated dna, Nature chemistry 3, 775 (2011).
  • Ichihashi et al. [2013] N. Ichihashi, K. Usui, Y. Kazuta, T. Sunami, T. Matsuura, and T. Yomo, Darwinian evolution in a translation-coupled rna replication system within a cell-like compartment, Nature communications 4, 2494 (2013).
  • Mavelli and Ruiz-Mirazo [2013] F. Mavelli and K. Ruiz-Mirazo, Theoretical conditions for the stationary reproduction of model protocells, Integrative biology 5, 324 (2013).
  • Ruiz-Mirazo et al. [2014] K. Ruiz-Mirazo, C. Briones, and A. de la Escosura, Prebiotic systems chemistry: new perspectives for the origins of life, Chemical reviews 114, 285 (2014).
  • Himeoka and Kaneko [2014] Y. Himeoka and K. Kaneko, Entropy production of a steady-growth cell with catalytic reactions, Physical Review E 90, 042714 (2014).
  • Kurihara et al. [2015] K. Kurihara, Y. Okura, M. Matsuo, T. Toyota, K. Suzuki, and T. Sugawara, A recursive vesicle-based model protocell with a primitive model cell cycle, Nature communications 6, 8352 (2015).
  • Shirt-Ediss et al. [2015] B. Shirt-Ediss, R. V. Solé, and K. Ruiz-Mirazo, Emergent chemical behavior in variable-volume protocells, Life 5, 181 (2015).
  • Pandey and Jain [2016] P. P. Pandey and S. Jain, Analytic derivation of bacterial growth laws from a simple model of intracellular chemical dynamics, Theory in Biosciences 135, 121 (2016).
  • Serra and Villani [2017] R. Serra and M. Villani, Modelling Protocells: The Emergent Synchronization of Reproduction and Molecular Replication (Springer, Berlin, 2017).
  • Joyce and Szostak [2018] G. F. Joyce and J. W. Szostak, Protocells and rna self-replication, Cold Spring Harbor Perspectives in Biology 10, a034801 (2018).
  • Lancet et al. [2018] D. Lancet, R. Zidovetzki, and O. Markovitch, Systems protobiology: origin of life in lipid catalytic networks, Journal of The Royal Society Interface 15, 20180159 (2018).
  • Liu and Sumpter [2018] Y. Liu and D. J. Sumpter, Mathematical modeling reveals spontaneous emergence of self-replication in chemical reaction systems, Journal of Biological Chemistry 293, 18854 (2018).
  • Steel et al. [2019] M. Steel, W. Hordijk, and J. C. Xavier, Autocatalytic networks in biology: structural theory and algorithms, Journal of the Royal Society Interface 16, 20180808 (2019).
  • Pandey et al. [2020] P. P. Pandey, H. Singh, and S. Jain, Exponential trajectories, cell size fluctuations, and the adder property in bacteria follow from simple chemical dynamics and division control, Physical Review E 101, 062406 (2020).
  • Unterberger and Nghe [2022] J. Unterberger and P. Nghe, Stoechiometric and dynamical autocatalysis for diluted chemical reaction networks, Journal of Mathematical Biology 85, 26 (2022).
  • Gagrani et al. [2024] P. Gagrani, V. Blanco, E. Smith, and D. Baum, Polyhedral geometry and combinatorics of an autocatalytic ecosystem, Journal of Mathematical Chemistry , 1 (2024).
  • Scott et al. [2010] M. Scott, C. W. Gunderson, E. M. Mateescu, Z. Zhang, and T. Hwa, Interdependence of cell growth and gene expression: origins and consequences, Science 330, 1099 (2010).
  • Scott and Hwa [2011] M. Scott and T. Hwa, Bacterial growth laws and their applications, Current opinion in biotechnology 22, 559 (2011).
  • Scott et al. [2014] M. Scott, S. Klumpp, E. M. Mateescu, and T. Hwa, Emergence of robust growth laws from optimal regulation of ribosome synthesis, Molecular systems biology 10, 747 (2014).
  • Maitra and Dill [2015] A. Maitra and K. A. Dill, Bacterial growth laws reflect the evolutionary importance of energy efficiency, Proceedings of the National Academy of Sciences 112, 406 (2015).
  • Reuveni et al. [2017] S. Reuveni, M. Ehrenberg, and J. Paulsson, Ribosomes are optimized for autocatalytic production, Nature 547, 293 (2017).
  • Barenholz et al. [2017] U. Barenholz, D. Davidi, E. Reznik, Y. Bar-On, N. Antonovsky, E. Noor, and R. Milo, Design principles of autocatalytic cycles constrain enzyme kinetics and force low substrate saturation at flux branch points, Elife 6, e20667 (2017).
  • Jun et al. [2018] S. Jun, F. Si, R. Pugatch, and M. Scott, Fundamental principles in bacterial physiology—history, recent progress, and the future with focus on cell size control: a review, Reports on Progress in Physics 81, 056601 (2018).
  • Thomas et al. [2018] P. Thomas, G. Terradot, V. Danos, and A. Y. Weiße, Sources, propagation and consequences of stochasticity in cellular growth, Nature communications 9, 4528 (2018).
  • Furusawa and Kaneko [2018] C. Furusawa and K. Kaneko, Formation of dominant mode by evolution in biological systems, Physical Review E 97, 042410 (2018).
  • Kostinski and Reuveni [2020] S. Kostinski and S. Reuveni, Ribosome composition maximizes cellular growth rates in e. coli, Physical review letters 125, 028103 (2020).
  • Dourado and Lercher [2020] H. Dourado and M. J. Lercher, An analytical theory of balanced cellular growth, Nature communications 11, 1226 (2020).
  • Lin et al. [2020] W.-H. Lin, E. Kussell, L.-S. Young, and C. Jacobs-Wagner, Origin of exponential growth in nonlinear reaction networks, Proceedings of the National Academy of Sciences 117, 27795 (2020).
  • Kostinski and Reuveni [2021] S. Kostinski and S. Reuveni, Growth laws and invariants from ribosome biogenesis in lower eukarya, Physical Review Research 3, 013020 (2021).
  • Roy et al. [2021] A. Roy, D. Goberman, and R. Pugatch, A unifying autocatalytic network-based framework for bacterial growth laws, Proceedings of the National Academy of Sciences 118, e2107829118 (2021).
  • Note [1] One may feel that the order of the reactions in Eq. (9) is high. The reason why we consider this example is just because it is straightforward to visualize our geometric representation in Sec. III. Since 𝒩X=dimKer[ST]+dimIm[S]\mathcal{N}_{X}=\dim\mathrm{Ker}[S^{T}]+\dim\mathrm{Im}[S] and we can visualize up to three dimensional space (𝒩X=3\mathcal{N}_{X}=3), relevant cases are when dimIm[S]\dim\mathrm{Im}[S] is one, two or three. The present example corresponds to the case in which dimIm[S]\dim\mathrm{Im}[S] is two. In Sec. I and II of the Supplementary material, we also consider other examples with which dimIm[S]\dim\mathrm{Im}[S] is one and three, respectively.
  • Sughiyama et al. [2022b] Y. Sughiyama, D. Loutchko, A. Kamimura, and T. J. Kobayashi, Hessian geometric structure of chemical thermodynamic systems with stoichiometric constraints, Physical Review Research 4, 033065 (2022b).
  • Note [2] The inverse transformation is given by x=φ(y)x=\partial\varphi^{*}(y). The existence of one-to-one correspondence between >0𝒩X\mathbb{R}_{>0}^{\mathcal{N}_{X}} and 𝒩X\mathbb{R}^{\mathcal{N}_{X}} is physically reasonable (See Sec. III.1).
  • Feinberg [2019] M. Feinberg, Foundations of chemical reaction network theory (Springer, 2019).
  • Note [3] Note that regular SS is productive in Eq. (7).
  • Note [4] In Sec. II of the Supplementary material, we show an example when SS is regular in which the dimension of Ker[ST]\mathrm{Ker}[S^{T}] is zero and that of Im[S]\mathrm{Im}[S] is three.
  • Pachter and Sturmfels [2005] L. Pachter and B. Sturmfels, Algebraic statistics for computational biology, Vol. 13 (Cambridge university press, 2005).
  • Kobayashi et al. [2022] T. J. Kobayashi, D. Loutchko, A. Kamimura, and Y. Sughiyama, Kinetic derivation of the hessian geometric structure in chemical reaction networks, Physical Review Research 4, 033066 (2022).
  • Note [5] Similar to Fig. 3(d), the isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) is partitioned by L>0L>0 and L<0L<0 in this case.
  • Shima [2007] H. Shima, The geometry of Hessian structures (World Scientific, Singapore, 2007).
  • Amari and Nagaoka [2000] S.-i. Amari and H. Nagaoka, Methods of information geometry, Vol. 191 (American Mathematical Soc. Press, Providence, RI, 2000).
  • Bregman [1967] L. M. Bregman, The relaxation method of finding the common point of convex sets and its application to the solution of problems in convex programming, USSR computational mathematics and mathematical physics 7, 200 (1967).
  • Note [6] Note the equality in 0\geq 0 of 0𝒩X\mathbb{R}_{\geq 0}^{\mathcal{N}_{X}}.
  • Koch [1997] A. L. Koch, Microbial physiology and ecology of slow growth, Microbiology and Molecular Biology Reviews 61, 305 (1997).
  • O’Neill [1999] W. C. O’Neill, Physiological significance of volume-regulatory transporters, American Journal of Physiology-Cell Physiology 276, C995 (1999).
  • Strange [2004] K. Strange, Cellular volume homeostasis, Advances in physiology education 28, 155 (2004).
  • Alberti et al. [2019] S. Alberti, A. Gladfelter, and T. Mittag, Considerations and challenges in studying liquid-liquid phase separation and biomolecular condensates, Cell 176, 419 (2019).

Appendix A Calculation of the total entropy function

The complete dynamics for the system is defined as

dEdt=JE(t),dΩdt=JΩ(t),\displaystyle\frac{dE}{dt}=J_{E}(t),\frac{d\Omega}{dt}=J_{\Omega}(t),
dNmdt=OrmJr(t)+JDm(t),dXidt=SriJr(t),\displaystyle\frac{dN^{m}}{dt}=O^{m}_{r}J^{r}(t)+J^{m}_{D}(t),\frac{dX^{i}}{dt}=S^{i}_{r}J^{r}(t), (64)

where JE(t)J_{E}(t), JΩ(t)J_{\Omega}(t), JD(t)={JDm(t)}J_{D}(t)=\{J^{m}_{D}(t)\} and J(t)={Jr(t)}J(t)=\{J^{r}(t)\} represent the energy, the volume, the chemical diffusion, and the chemical reaction fluxes, respectively (see Fig. 1). By contrast, the dynamics for the environment is given as

dE~dt=JE(t),dΩ~dt=JΩ(t),dN~mdt=JDm(t).\frac{d\tilde{E}}{dt}=-J_{E}(t),\frac{d\tilde{\Omega}}{dt}=-J_{\Omega}(t),\frac{d\tilde{N}^{m}}{dt}=-J^{m}_{D}(t). (65)

Since we have assumed that the time scale of the reactions is much slower than that of the others, we can analyze the dynamics, Eqs. (64) and (65), by separating the slow one J(t)J(t) from the fast ones JE(t),JΩ(t),JD(t)J_{E}(t),J_{\Omega}(t),J_{D}(t). Owing to this time-scale separation, the extensive variables (E,Ω,N)(E,\Omega,N) are rapidly relaxed to the values at the equilibrium state of the fast dynamics with a fixed XX. They are computed by the thermodynamic variational problem:

(E(X),Ω(X),N(X))\displaystyle\left(E(X),\Omega(X),N(X)\right)
=argmaxE,Ω,N{Σ[E,Ω,N,X]ET~Π~T~Ω+μ~mT~Nm},\displaystyle=\arg\max_{E,\Omega,N}\left\{\Sigma\left[E,\Omega,N,X\right]-\frac{E}{\tilde{T}}-\frac{\tilde{\Pi}}{\tilde{T}}\Omega+\frac{\tilde{\mu}_{m}}{\tilde{T}}N^{m}\right\}, (66)

(see [78] and [19]). If we use the densities σ\sigma, ϵ\epsilon and nn, this varational problem is equivalently written in

(ϵ(X),Ω(X),n(X))\displaystyle\left(\epsilon(X),\Omega(X),n(X)\right)
=argmaxϵ,Ω,n{Ω(T~σ[ϵ,n,XΩ]ϵΠ~+μ~mnm)}.\displaystyle=\arg\max_{\epsilon,\Omega,n}\left\{\Omega\left(\tilde{T}\sigma\left[\epsilon,n,\frac{X}{\Omega}\right]-\epsilon-\tilde{\Pi}+\tilde{\mu}_{m}n^{m}\right)\right\}. (67)

By defining the partial grand potential density (see Eq. (13)),

φ(x)=φ[T~,μ~;x]:=minϵ,n{ϵT~σ[ϵ,n,x]μ~mnm},\displaystyle\varphi(x)=\varphi\left[\tilde{T},\tilde{\mu};x\right]:=\min_{\epsilon,n}\left\{\epsilon-\tilde{T}\sigma\left[\epsilon,n,x\right]-\tilde{\mu}_{m}n^{m}\right\}, (68)

and taking the maximization with respect to ϵ\epsilon and nn in Eq. (67), we obtain

Ω(X)=argminΩ{Ωφ(XΩ)+ΩΠ~}.\displaystyle\Omega(X)=\arg\min_{\Omega}\left\{\Omega\varphi\left(\frac{X}{\Omega}\right)+\Omega\tilde{\Pi}\right\}. (69)

This determines the volume Ω\Omega at XX (see Eq. (14)).

By substituting (ϵ(X),Ω(X),n(X))(\epsilon(X),\Omega(X),n(X)) in Eq. (67) into Eqs. (64) and (65), we get

dXidt=SriJr(t),dE~dt=dΩ(X)ϵ(X)dt,\displaystyle\frac{dX^{i}}{dt}=S^{i}_{r}J^{r}(t),\frac{d\tilde{E}}{dt}=-\frac{d\Omega(X)\epsilon(X)}{dt},
dΩ~dt=dΩ(X)dt,dN~mdt=OrmJr(t)dΩ(X)nm(X)dt.\displaystyle\frac{d\tilde{\Omega}}{dt}=-\frac{d\Omega(X)}{dt},\frac{d\tilde{N}^{m}}{dt}=O^{m}_{r}J^{r}(t)-\frac{d\Omega(X)n^{m}(X)}{dt}. (70)

Integration of Eq. (70) leads to

Xt\displaystyle X_{t} =X0+SΞt,\displaystyle=X_{0}+S\Xi_{t},
E~(Xt)\displaystyle\tilde{E}(X_{t}) =E~(X0){Ω(Xt)ϵ(Xt)Ω(X0)ϵ(X0)},\displaystyle=\tilde{E}(X_{0})-\left\{\Omega(X_{t})\epsilon(X_{t})-\Omega(X_{0})\epsilon(X_{0})\right\},
Ω~(Xt)\displaystyle\tilde{\Omega}(X_{t}) =Ω~(X0){Ω(Xt)Ω(X0)},\displaystyle=\tilde{\Omega}(X_{0})-\left\{\Omega(X_{t})-\Omega(X_{0})\right\},
N~(Xt)\displaystyle\tilde{N}(X_{t}) =N~(X0)+OΞt{Ω(Xt)n(Xt)Ω(X0)n(X0)},\displaystyle=\tilde{N}(X_{0})+O\Xi_{t}-\left\{\Omega(X_{t})n(X_{t})-\Omega(X_{0})n(X_{0})\right\}, (71)

where Ξt\Xi_{t} is the extent of reaction at time tt.

The total entropy in Eq. (1) is represented as

Σtot=\displaystyle\Sigma^{\mathrm{tot}}= Ω(Xt)σ[ϵ(Xt),n(Xt),XtΩ(Xt)]\displaystyle\Omega(X_{t})\sigma\left[\epsilon(X_{t}),n(X_{t}),\frac{X_{t}}{\Omega(X_{t})}\right]
+Σ~T~,Π~,μ~[E~(Xt),Ω~(Xt),N~(Xt)].\displaystyle+\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}\left[\tilde{E}(X_{t}),\tilde{\Omega}(X_{t}),\tilde{N}(X_{t})\right]. (72)

By substituting Eq. (71) and employing the Taylor expansion for Σ~T~,Π~,μ~\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}, we get

Σtot=Ω(X)[σ[ϵ(X),n(X),XΩ(X)]\displaystyle\Sigma^{\mathrm{tot}}=\Omega(X)\biggl{[}\sigma\left[\epsilon(X),n(X),\frac{X}{\Omega(X)}\right]
ϵ(X)T~Π~T~μ~mT~{OrmΞrΩ(X)nm(X)}]+const.,\displaystyle-\frac{\epsilon(X)}{\tilde{T}}-\frac{\tilde{\Pi}}{\tilde{T}}-\frac{\tilde{\mu}_{m}}{\tilde{T}}\left\{O^{m}_{r}\frac{\Xi^{r}}{\Omega(X)}-n^{m}(X)\right\}\biggr{]}+\mathrm{const.}, (73)

where X=X0+SΞX=X_{0}+S\Xi. Here, we used the thermodynamic relations: Σ~T~,Π~,μ~/E~=1/T~\partial\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}/\partial\tilde{E}=1/\tilde{T}, Σ~T~,Π~,μ~/Ω~=Π~/T~\partial\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}/\partial\tilde{\Omega}=\tilde{\Pi}/\tilde{T} and Σ~T~,Π~,μ~/N~m=μ~m/T~\partial\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}/\partial\tilde{N}^{m}=-\tilde{\mu}_{m}/\tilde{T}. Taking Eq. (68) into account, we obtain

Σtot=\displaystyle\Sigma^{\mathrm{tot}}= 1T~{Ω(X)φ(XΩ(X))+Ω(X)Π~+μ~mOrmΞr}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\Omega(X)\tilde{\Pi}+\tilde{\mu}_{m}O^{m}_{r}\Xi^{r}\right\}
+const.,\displaystyle+\mathrm{const.}, (74)

which is Eq. (12).

Appendix B Mass-action kinetics

To numerically simulate the dynamics, Eq. (3), we assume mass action kinetics for the flux function J(t)={J1(t),J2(t)}J(t)=\{J^{1}(t),J^{2}(t)\}. For Example 1, it is given as

J1(t)\displaystyle J^{1}(t) =w+1Ω(X1Ω)(X3Ω)N1Ωw1Ω(X2Ω)2,\displaystyle=w^{1}_{+}\Omega\left(\frac{X^{1}}{\Omega}\right)\left(\frac{X^{3}}{\Omega}\right)\frac{N^{1}}{\Omega}-w^{1}_{-}\Omega\left(\frac{X^{2}}{\Omega}\right)^{2},
J2(t)\displaystyle J^{2}(t) =w+2Ω(X2Ω)w2Ω(X1Ω)(X3Ω)N2Ω,\displaystyle=w^{2}_{+}\Omega\left(\frac{X^{2}}{\Omega}\right)-w^{2}_{-}\Omega\left(\frac{X^{1}}{\Omega}\right)\left(\frac{X^{3}}{\Omega}\right)\frac{N^{2}}{\Omega}, (75)

where N=(N1,N2)N=(N^{1},N^{2}) denotes the number of B=(B1,B2)B=(B_{1},B_{2}) in the system. The rate constants w+rw^{r}_{+} and wrw^{r}_{-} satisfy the local detailed balance condition [11, 14, 16, 78, 84]:

logw+rwr=1RT~{νio(T~)Sri+μmo(T~)Orm}.\displaystyle\log\frac{w^{r}_{+}}{w^{r}_{-}}=-\frac{1}{R\tilde{T}}\left\{\nu_{i}^{o}(\tilde{T})S_{r}^{i}+\mu_{m}^{o}(\tilde{T})O^{m}_{r}\right\}. (76)

For the ideal gas case, the density N/ΩN/\Omega of the open chemicals in the system is constant and fixed to that of the environment n~\tilde{n}. Also, the volume Ω\Omega is given by the equation of state in Eq. (17). Thus, we can rearrange Eq. (75) as

J1(t)\displaystyle J^{1}(t) =w^+1Ω(X1Ω)(X3Ω)w^1Ω(X2Ω)2,\displaystyle=\hat{w}^{1}_{+}\Omega\left(\frac{X^{1}}{\Omega}\right)\left(\frac{X^{3}}{\Omega}\right)-\hat{w}^{1}_{-}\Omega\left(\frac{X^{2}}{\Omega}\right)^{2},
J2(t)\displaystyle J^{2}(t) =w^+2Ω(X2Ω)w^2Ω(X1Ω)(X3Ω),\displaystyle=\hat{w}^{2}_{+}\Omega\left(\frac{X^{2}}{\Omega}\right)-\hat{w}^{2}_{-}\Omega\left(\frac{X^{1}}{\Omega}\right)\left(\frac{X^{3}}{\Omega}\right), (77)

where we absorb the constant densities of the open chemicals into the rate constants as w^+r\hat{w}^{r}_{+} and w^r\hat{w}^{r}_{-}. For these effective rate constants, the local detailed balance condition in Eq. (76) is written as

logw^+rw^r=1RT~{νio(T~)Sri+μ~mOrm},\displaystyle\log\frac{\hat{w}^{r}_{+}}{\hat{w}^{r}_{-}}=-\frac{1}{R\tilde{T}}\left\{\nu_{i}^{o}(\tilde{T})S_{r}^{i}+\tilde{\mu}_{m}O^{m}_{r}\right\}, (78)

where

μ~m=μmo(T~)+RT~logn~m,\displaystyle\tilde{\mu}_{m}=\mu^{o}_{m}(\tilde{T})+R\tilde{T}\log\tilde{n}^{m}, (79)

in Eq. (16). For our specific case, it is given as

w^+1w^1=xo2xo2n~1xo1xo3no1,w^+2w^2=xo1xo3no2xo2n~2,\displaystyle\frac{\hat{w}^{1}_{+}}{\hat{w}^{1}_{-}}=\frac{x^{2}_{o}x^{2}_{o}\tilde{n}^{1}}{x^{1}_{o}x^{3}_{o}n^{1}_{o}},\hskip 14.22636pt\frac{\hat{w}^{2}_{+}}{\hat{w}^{2}_{-}}=\frac{x^{1}_{o}x^{3}_{o}n^{2}_{o}}{x^{2}_{o}\tilde{n}^{2}}, (80)

where xoi:=eνio(T~)/RT~x^{i}_{o}:=e^{-\nu^{o}_{i}(\tilde{T})/R\tilde{T}} and nom:=eμmo(T~)/RT~n^{m}_{o}:=e^{-\mu^{o}_{m}(\tilde{T})/R\tilde{T}}.

Appendix C The assumption for the pressure Π~\tilde{\Pi}

The assumption Π~>Πmin\tilde{\Pi}>\Pi_{\mathrm{min}} is made to guarantee that, for a given XX, the volume Ω(X)\Omega(X) is uniquely determined and the system can relax to the equilibrium state of the fast dynamics.

The volume Ω(X)\Omega(X) is variationally determined in Eq. (14) as

Ω(X)=argminΩ{Ωφ(XΩ)+Π~Ω}.\displaystyle\Omega(X)=\arg\min_{\Omega}\left\{\Omega\varphi\left(\frac{X}{\Omega}\right)+\tilde{\Pi}\Omega\right\}. (81)

For a given XX, the critical equation is computed as

h(Ω)\displaystyle h(\Omega) :=φ(XΩ)XiΩiφ(XΩ)+Π~=0.\displaystyle:=\varphi\left(\frac{X}{\Omega}\right)-\frac{X^{i}}{\Omega}\partial_{i}\varphi\left(\frac{X}{\Omega}\right)+\tilde{\Pi}=0. (82)

The differentiation of h(Ω)h(\Omega) is given as

dhdΩ=Ω3Xi[ijφ(XΩ)]Xj.\displaystyle\frac{dh}{d\Omega}=\Omega^{-3}X^{i}\left[\partial_{i}\partial_{j}\varphi\left(\frac{X}{\Omega}\right)\right]X^{j}. (83)

Since φ\varphi is strictly convex, its Hessian ijφ\partial_{i}\partial_{j}\varphi is positive definite. Thus, the function h(Ω)h(\Omega) is a strictly increasing function for Ω>0\Omega>0. In addition, h(Ω)h(\Omega) is further calculated as

h(Ω)=φ(φ(XΩ))+Π~.\displaystyle h(\Omega)=-\varphi^{*}\left(\partial\varphi\left(\frac{X}{\Omega}\right)\right)+\tilde{\Pi}. (84)

If Ω0\Omega\rightarrow 0, φ(X/Ω)\partial\varphi(X/\Omega)\rightarrow\infty and h(Ω)h(\Omega)\rightarrow-\infty. By contrast, if Ω\Omega\rightarrow\infty, φ(X/Ω)\partial\varphi(X/\Omega)\rightarrow-\infty and h(Ω)Π~Πminh(\Omega)\rightarrow\tilde{\Pi}-\Pi_{\mathrm{min}}, because φ(y)\varphi^{*}(y) is strictly increasing and its infimum is given by Πmin\Pi_{\mathrm{min}}. Thus, if Π~>Πmin\tilde{\Pi}>\Pi_{\mathrm{min}}, the critical equation has a unique solution with respect to Ω\Omega, and, therefore, the system can relax to the equilibrium state of the fast dynamics.

If Π~Πmin\tilde{\Pi}\leq\Pi_{\mathrm{min}}, the critical equation h(Ω)=0h(\Omega)=0 in Eq. (82) does not have a solution, and the volume Ω\Omega diverges from the variational form in Eq. (81). This is consistent with our intuition; if the pressure Π~\tilde{\Pi} of the environment is smaller than the minimum pressure of the system, the pressures cannot be balanced and the volume of the system would diverge.

Appendix D Derivation of Eq. (48)

The total entropy function is given in Eq. (12) as

Σtot(Ξ)=\displaystyle\Sigma^{\mathrm{tot}}(\Xi)= 1T~{Ω(X)φ(XΩ(X))+Ω(X)Π~+μ~mOrmΞr}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\Omega(X)\tilde{\Pi}+\tilde{\mu}_{m}O^{m}_{r}\Xi^{r}\right\}
+const,\displaystyle+\mathrm{const}, (85)

where X=X0+SΞX=X_{0}+S\Xi. It is rewritten as a function of Ω\Omega, XX and Ξ\Xi:

Σ¯tot(Ω,X,Ξ)=\displaystyle\bar{\Sigma}^{\mathrm{tot}}(\Omega,X,\Xi)= 1T~{Ωφ(XΩ)+ΩΠ~+μ~mOrmΞr}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega\varphi\left(\frac{X}{\Omega}\right)+\Omega\tilde{\Pi}+\tilde{\mu}_{m}O^{m}_{r}\Xi^{r}\right\}
+const.\displaystyle+\mathrm{const}. (86)

Its total derivative is calculated as

dΣ¯tot\displaystyle d\bar{\Sigma}^{\mathrm{tot}} =Σ¯totΩdΩ+Σ¯totXidXi+Σ¯totΞrdΞr.\displaystyle=\frac{\partial\bar{\Sigma}^{\mathrm{tot}}}{\partial\Omega}d\Omega+\frac{\partial\bar{\Sigma}^{\mathrm{tot}}}{\partial X^{i}}dX^{i}+\frac{\partial\bar{\Sigma}^{\mathrm{tot}}}{\partial\Xi^{r}}d\Xi^{r}. (87)

The first coefficient Σ¯tot/Ω\partial\bar{\Sigma}^{\mathrm{tot}}/\partial\Omega vanishes at Ω=Ω(X)\Omega=\Omega(X) because of Eq. (14). The second and third coefficients are computed as Σ¯tot/Xi=iφ(X/Ω)/T~\partial\bar{\Sigma}^{\mathrm{tot}}/\partial X^{i}=-\partial_{i}\varphi(X/\Omega)/\tilde{T} and Σ¯tot/Ξr=μ~mOrm/T~\partial\bar{\Sigma}^{\mathrm{tot}}/\partial\Xi^{r}=-\tilde{\mu}_{m}O^{m}_{r}/\tilde{T}. Since dXi=SridΞrdX^{i}=S^{i}_{r}d\Xi^{r}, we obtain

dΣ¯tot\displaystyle d\bar{\Sigma}^{\mathrm{tot}} =1T~{iφ(XΩ(X))Sri+μ~mOrm}dΞr.\displaystyle=-\frac{1}{\tilde{T}}\left\{\partial_{i}\varphi\left(\frac{X}{\Omega(X)}\right)S^{i}_{r}+\tilde{\mu}_{m}O^{m}_{r}\right\}d\Xi^{r}. (88)

This corresponds to Eq. (48).

Appendix E Proof of Ω(X)\Omega(X)\rightarrow\infty when |X||X|\rightarrow\infty

In this Appendix, we show that the volume Ω(X)\Omega(X)\rightarrow\infty when |X||X|\rightarrow\infty. For a given XX, the volume Ω(X)\Omega(X) is variationally determined by Eq. (14).

In the number space 𝔛\mathfrak{X}, we can show that the volume function Ω(X)\Omega(X) satisfies the homogeneity: Ω(αX)=αΩ(X)\Omega(\alpha X)=\alpha\Omega(X) for α>0\alpha>0. Thus, for a given X𝔯X_{\mathfrak{r}}, we have

Ω(αX𝔯)α=α{αΩ(X𝔯)}=Ω(X𝔯).\displaystyle\frac{\partial\Omega(\alpha X_{\mathfrak{r}})}{\partial\alpha}=\frac{\partial}{\partial\alpha}\left\{\alpha\Omega(X_{\mathfrak{r}})\right\}=\Omega(X_{\mathfrak{r}}). (89)

This implies that the volume function Ω(X)\Omega(X) increases with the rate Ω(X𝔯)\Omega(X_{\mathfrak{r}}) along the ray including X𝔯X_{\mathfrak{r}}. Therefore, the volume Ω(X)\Omega(X) goes to infinity for |X||X|\rightarrow\infty.

Appendix F Birch’s theorem

In this Appendix, we introduce Birch’s theorem and its extension.

Consider the following variational problem:

argminy{φ(y)yix0i|yEQ𝒴(μ~)},\displaystyle\arg\min_{y}\Set{\varphi^{*}(y)-y_{i}x^{i}_{0}}{y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})}, (90)

where x0𝒳=>0𝒩Xx_{0}\in\mathcal{X}=\mathbb{R}_{>0}^{\mathcal{N}_{X}} is an arbitrary constant such that Ux0=L^Ux_{0}=\hat{L} for a given L^\hat{L}; it corresponds to the initial condition in the main text. Also, φ(y)\varphi^{*}(y) is a strictly increasing function as introduced in Sec. III.1. From Eq. (53), the variational problem is rewritten as

argminy{φ(y)yix0i|yyP+Im[UT]}={yB(L^)}.\displaystyle\arg\min_{y}\Set{\varphi^{*}(y)-y_{i}x^{i}_{0}}{y\in y^{\mathrm{P}}+\mathrm{Im}[U^{T}]}=\{y^{\mathrm{B}}(\hat{L})\}. (91)

The point yB(L^)y^{\mathrm{B}}(\hat{L}) to attain the minimum uniquely exists, because φ(y)yix0i\varphi^{*}(y)-y_{i}x^{i}_{0} is strictly convex and x0>0x_{0}>0. To be more precise, we have φ(y)yix0i\varphi^{*}(y)-y_{i}x^{i}_{0}\rightarrow\infty as |y||y|\rightarrow\infty. By the directional derivative, the left hand side of Eq. (91) can be represented as

argminy{φ(y)yix0i|yyP+Im[UT]}\displaystyle\arg\min_{y}\Set{\varphi^{*}(y)-y_{i}x^{i}_{0}}{y\in y^{\mathrm{P}}+\mathrm{Im}[U^{T}]}
={y|φ(y)x0+Ker[U],yyP+Im[UT]}\displaystyle=\Set{y}{\partial\varphi^{*}(y)\in x_{0}+\mathrm{Ker}[U],y\in y^{\mathrm{P}}+\mathrm{Im}[U^{T}]}
={y|Uiliφ(y)=L^l,yyP+Im[UT]}.\displaystyle=\Set{y}{U^{l}_{i}\partial^{i}\varphi^{*}(y)=\hat{L}^{l},y\in y^{\mathrm{P}}+\mathrm{Im}[U^{T}]}. (92)

Taking Eq. (44) into account, we obtain

^STO𝒴(L^)EQ𝒴(μ~)={yB(L^)},\displaystyle\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(\hat{L})\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\{y^{\mathrm{B}}(\hat{L})\}, (93)

which is called Birch’s theorem.

Next, we extend Birch’s theorem to the case x00x_{0}\rightarrow 0. In this case, Eq. (90) is formally rewritten as

{ymin}\displaystyle\{y^{\mathrm{min}}\} =argminy{φ(y)|yEQ𝒴(μ~)}\displaystyle=\arg\min_{y}\Set{\varphi^{*}(y)}{y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})}
=^STO𝒴(L^=0)EQ𝒴(μ~)={yB(0)}.\displaystyle=\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(\hat{L}=0)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu})=\{y^{\mathrm{B}}(0)\}. (94)

However, the existence of the point yminy^{\mathrm{min}} is not trivial because φ(y)\varphi^{*}(y) is strictly increasing and yminy^{\mathrm{min}} may goes to infinity. In the following, we show that the productivity of SS guarantees its existence.

When we write {Uil}=(U1,U2,)T\{U_{i}^{l}\}=(\textbf{U}^{1},\textbf{U}^{2},...)^{T}, the productivity of SS guarantees that any vector Ul\textbf{U}^{l} has both positive and negative components. The directional derivative on EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) leads to

φηl=Uiliφ(yP+ηU),\displaystyle\frac{\partial\varphi^{*}}{\partial\eta_{l}}=U^{l}_{i}\partial^{i}\varphi^{*}(y^{\mathrm{P}}+\eta U), (95)

where η\eta is a coordinate of Im[UT]\mathrm{Im}[U^{T}]. For a given ll, as ηl\eta_{l}\rightarrow\infty, the following holds:

iφ(yP+ηU){ for i s.t. Uil>0,0 for i s.t. Uil<0,iφ(yP) for i s.t. Uil=0.\displaystyle\partial^{i}\varphi^{*}(y^{\mathrm{P}}+\eta U)\rightarrow\begin{cases}\infty&\mbox{ for }i\mbox{ }s.t.\mbox{ }U^{l}_{i}>0,\\ 0&\mbox{ for }i\mbox{ }s.t.\mbox{ }U^{l}_{i}<0,\\ \partial^{i}\varphi^{*}(y^{\mathrm{P}})&\mbox{ for }i\mbox{ }s.t.\mbox{ }U^{l}_{i}=0.\end{cases}

Therefore, by taking Eq. (95) and the productivity of SS into account, we conclude that, for any ll, φ/ηl\partial\varphi^{*}/\partial\eta_{l}\rightarrow\infty as ηl\eta_{l}\rightarrow\infty. In a similar manner, we can show that, for any ll, φ/ηl\partial\varphi^{*}/\partial\eta_{l}\rightarrow-\infty as ηl\eta_{l}\rightarrow-\infty. As a result, φ(y)\varphi^{*}(y) restricted on EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) has a unique minimum. Note that, if SS is not productive, the above statement is not satisfied.

Appendix G Proof of Lemma 2

The starting point yminy^{\mathrm{min}} of Birch’s trajectory gives the infimum of φ(y)\varphi^{*}(y) in the trajectory because yminy^{\mathrm{min}} gives the minimum φ(y)\varphi^{*}(y) in the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (see Eq. (28)) and any points in the trajectory is in the manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). Thus, the statement of the Lemma 2 can be proved if we show that φ(y)\varphi^{*}(y) is a strictly monotonic function along Birch’s trajectory. We prove that by contradiction.

For a given L0L\neq 0, we pick arbitrary two different Birch’s points on Birch’s trajectory: yB1,yB2𝔟(L)y^{\mathrm{B1}},y^{\mathrm{B2}}\in\mathfrak{b}(L). We can find α1\alpha_{1} and α2\alpha_{2} such that yB1^STO𝒴(L/α1)y^{\mathrm{B1}}\in\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha_{1}) and yB2^STO𝒴(L/α2)y^{\mathrm{B2}}\in\hat{\mathcal{M}}^{\mathcal{Y}}_{\mathrm{STO}}(L/\alpha_{2}) where α1α2\alpha_{1}\neq\alpha_{2} and α1,α2>0\alpha_{1},\alpha_{2}>0. From Eq. (44), we have

Uiliφ(yB1)=Llα1, Uiliφ(yB2)=Llα2.\displaystyle U^{l}_{i}\partial^{i}\varphi^{*}(y^{\mathrm{B1}})=\frac{L^{l}}{\alpha_{1}},\mbox{ }U^{l}_{i}\partial^{i}\varphi^{*}(y^{\mathrm{B2}})=\frac{L^{l}}{\alpha_{2}}. (96)

Since UU is a basis matrix of Ker[ST]\mathrm{Ker}[S^{T}]: {Uil}:=(U1,U2,)T\{U_{i}^{l}\}:=(\textbf{U}^{1},\textbf{U}^{2},...)^{T}, and yB1,yB2EQ𝒴(μ~)y^{\mathrm{B1}},y^{\mathrm{B2}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}), we can choose U1\textbf{U}^{1} as yB2yB1y^{\mathrm{B2}}-y^{\mathrm{B1}}. Then, we obtain

Ui1iφ(yB1)=α1α2Ui1iφ(yB2),\displaystyle U^{1}_{i}\partial^{i}\varphi^{*}(y^{\mathrm{B1}})=\frac{\alpha_{1}}{\alpha_{2}}U^{1}_{i}\partial^{i}\varphi^{*}(y^{\mathrm{B2}}), (97)

which implies that the signs of Ui1iφ(y)U^{1}_{i}\partial^{i}\varphi^{*}(y) are the same at yB1y^{\mathrm{B1}} and yB2y^{\mathrm{B2}}.

If φ(y)\varphi^{*}(y) is not strictly monotonic along Birch’s trajectory, we can find yB1yB2y^{\mathrm{B1}}\neq y^{\mathrm{B2}} such that φ(yB1)=φ(yB2)\varphi^{*}(y^{\mathrm{B1}})=\varphi^{*}(y^{\mathrm{B2}}). Since the vector U1\textbf{U}^{1} points in the direction from yB1y^{\mathrm{B1}} to yB2y^{\mathrm{B2}}, and φ(y)\varphi^{*}(y) is strictly convex, the signs of Ui1iφ(y)U^{1}_{i}\partial^{i}\varphi^{*}(y) are different at yB1y^{\mathrm{B1}} and yB2y^{\mathrm{B2}}. It contradicts with Eq. (97), and therefore, φ(y)\varphi^{*}(y) should be strictly monotonic along Birch’s trajectory.

Appendix H The form of the total entropy function using a particular solution yPy^{\mathrm{P}}

Substituting a particular solution yPy^{\mathrm{P}} to Eq. (50) into Eq. (12) (also Eq. (74)), we get

Σtot=\displaystyle\Sigma^{\mathrm{tot}}= 1T~{Ω(X)φ(XΩ(X))+Ω(X)Π~yiPSriΞr}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\Omega(X)\tilde{\Pi}-y_{i}^{\mathrm{P}}S^{i}_{r}\Xi^{r}\right\}
+const.\displaystyle+\mathrm{const}. (98)

Since SriΞr=XiX0iS^{i}_{r}\Xi^{r}=X^{i}-X^{i}_{0} from Eq. (4), it can be represented as a function of the number of the confined chemicals XX:

Σtot=\displaystyle\Sigma^{\mathrm{tot}}= 1T~{Ω(X)φ(XΩ(X))+Ω(X)Π~yiP(XiX0i)}\displaystyle-\frac{1}{\tilde{T}}\left\{\Omega(X)\varphi\left(\frac{X}{\Omega(X)}\right)+\Omega(X)\tilde{\Pi}-y_{i}^{\mathrm{P}}\left(X^{i}-X^{i}_{0}\right)\right\}
+const.\displaystyle+\mathrm{const}. (99)

This is further calculated as

Σtot=\displaystyle\Sigma^{\mathrm{tot}}= Ω(X)T~{φ(XΩ(X))yiPXiΩ(X)+Π~}yiPX0iT~\displaystyle-\frac{\Omega(X)}{\tilde{T}}\left\{\varphi\left(\frac{X}{\Omega(X)}\right)-y^{\mathrm{P}}_{i}\frac{X^{i}}{\Omega(X)}+\tilde{\Pi}\right\}-\frac{y^{\mathrm{P}}_{i}X^{i}_{0}}{\tilde{T}}
=\displaystyle= Ω(X)T~[φ(φ(XΩ(X)))\displaystyle-\frac{\Omega(X)}{\tilde{T}}\biggl{[}-\varphi^{*}\left(\partial\varphi\left(\frac{X}{\Omega(X)}\right)\right)
{yiPiφ(XΩ(X))}XiΩ(X)+Π~]yiPX0iT~,\displaystyle-\left\{y^{\mathrm{P}}_{i}-\partial_{i}\varphi\left(\frac{X}{\Omega(X)}\right)\right\}\frac{X^{i}}{\Omega(X)}+\tilde{\Pi}\biggr{]}-\frac{y^{\mathrm{P}}_{i}X^{i}_{0}}{\tilde{T}}, (100)

where we omit the constant term in Eq. (99) for simplicity. In addition, by defining y(X):=φ(X/Ω(X))=φρ𝒳(X)y(X):=\partial\varphi(X/\Omega(X))=\partial\varphi\circ\rho_{\mathcal{X}}(X), we get

Σtot=Ω(X)T~[φ(yP)φ(y(X))\displaystyle\Sigma^{\mathrm{tot}}=-\frac{\Omega(X)}{\tilde{T}}\biggl{[}\varphi^{*}(y^{\mathrm{P}})-\varphi^{*}(y(X))
iφ(y(X)){yiPyi(X)}+Π~φ(yP)]yiPX0iT~\displaystyle-\partial^{i}\varphi^{*}(y(X))\left\{y_{i}^{\mathrm{P}}-y_{i}(X)\right\}+\tilde{\Pi}-\varphi^{*}(y^{\mathrm{P}})\biggr{]}-\frac{y^{\mathrm{P}}_{i}X^{i}_{0}}{\tilde{T}}
=Ω(X)T~[φ(yP)Π~𝒟𝒴[yP||y]]yiPX0iT~.\displaystyle=\frac{\Omega(X)}{\tilde{T}}\biggl{[}\varphi^{*}(y^{\mathrm{P}})-\tilde{\Pi}-\mathcal{D}^{\mathcal{Y}}\left[y^{\mathrm{P}}||y\right]\biggr{]}-\frac{y^{\mathrm{P}}_{i}X^{i}_{0}}{\tilde{T}}. (101)

This corresponds to Eq. (61).

Appendix I The entropy function does not depend on the choice of a particular solution yPy^{\mathrm{P}}

We show that the value of the total entropy function in Eq. (61) (i.e., Eq. (99)) does not depend on the particular choice of yiPy_{i}^{\mathrm{P}}. The general solution to Eq. (50) is represented by using the particular solution yPy^{\mathrm{P}} as

yi=yiP+hlUil.y_{i}=y^{\mathrm{P}}_{i}+h_{l}U^{l}_{i}. (102)

Then, the term including yiPy^{\mathrm{P}}_{i} in Eq. (99) is written as

yiP(XiX0i)\displaystyle y_{i}^{\mathrm{P}}\left(X^{i}-X_{0}^{i}\right) =yi(XiX0i)hl(UilXiUilX0i)\displaystyle=y_{i}\left(X^{i}-X_{0}^{i}\right)-h_{l}\left(U^{l}_{i}X^{i}-U^{l}_{i}X^{i}_{0}\right)
=yi(XiX0i).\displaystyle=y_{i}\left(X^{i}-X_{0}^{i}\right). (103)

Here, the second equality holds because the quantities UilXiU^{l}_{i}X^{i} is conserved in the time evolution, Eq. (3). Thus, the value in Eq. (61) does not depend on the choice of yiPy^{\mathrm{P}}_{i}.

Appendix J Proof of the existence of a point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) with positive K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) when φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0

We show that, when φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0, we can find ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) such that K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is positive.

When the stoichiometric matrix SS is productive (see Eq. (7)), we first show that STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\neq\emptyset, i.e., there exists y𝒴(Π~,μ~)y\in\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) such that Uiliφ(y)=0U^{l}_{i}\partial^{i}\varphi^{*}(y)=0. From the productivity of SS, we can construct a vector 𝒗={vi}\bm{v}=\{v^{i}\} such that its components are positive and satisfies Uilvi=0U^{l}_{i}v^{i}=0. In addition, we can find y𝒴(Π~,μ~)y\in\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) such that the vector φ(y)\partial\varphi^{*}(y) has the same direction with 𝒗\bm{v} because the range of iφ(y)\partial^{i}\varphi^{*}(y) covers the positive orthant. This point yy is on the stoichiometric manifold STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0). Thus, we get STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\neq\emptyset.

For ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0), we have φ(y)=Π~\varphi^{*}(y)=\tilde{\Pi} because STO𝒴(L=0)𝒴(Π~,μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\subset\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}). Thus, the expression in Eq. (62) is rearranged for ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) as

K𝒴(yP;y)\displaystyle K^{\mathcal{Y}}(y^{\mathrm{P}};y) =φ(yP)Π~𝒟𝒴[yP||y]\displaystyle=\varphi^{*}(y^{\mathrm{P}})-\tilde{\Pi}-\mathcal{D}^{\mathcal{Y}}\left[y^{\mathrm{P}}||y\right]
=iφ(y)(yiPyi).\displaystyle=\partial^{i}\varphi^{*}(y)\left(y^{\mathrm{P}}_{i}-y_{i}\right). (104)

The vector φ(y)={iφ(y)}\partial\varphi^{*}(y)=\{\partial^{i}\varphi^{*}(y)\} represents a gradient of the convex function φ(y)\varphi^{*}(y), which is a normal vector at a point yy of the level set {y|φ(y)=Π~}\{y|\varphi^{*}(y)=\tilde{\Pi}\}. Note that the orientation of the normal vector points to the superlevel set (see Fig. 8). Also, (yPy)(y^{\mathrm{P}}-y) is a vector from the point yy on the level set to the point yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) represents the inner product of the two vectors.

To investigate the sign of the inner product, we consider the tangent plane of the level set at the point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) (see the dashed line in Fig. 8). This plane divides the space of 𝒴\mathcal{Y} into two regions. If the orientations of the two vectors, φ(y)\partial\varphi^{*}(y) and (yPy)(y^{\mathrm{P}}-y), point to the same side of the tangent plane, their inner product is positive. If they point to the opposite sides, the inner product is negative.

When φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0, the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is located in the superlevel set. In addition, the tangent plane at the point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) is parallel to EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). Then, we can find ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) such that EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is located on the same side of the tangent plane with the orientation of the normal vector φ(y)\partial\varphi^{*}(y) at the point yy (see Fig. 8). In such a case, the inner product between φ(y)\partial\varphi^{*}(y) and (yiPyi)(y^{\mathrm{P}}_{i}-y_{i}) is positive for yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}). Thus, ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) exists such that K𝒴(yP;y)>0K^{\mathcal{Y}}(y^{\mathrm{P}};y)>0.

Refer to caption
Figure 8: Illustration of the proof for the existence of ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) such that K𝒴(yP;y)>0K^{\mathcal{Y}}(y^{\mathrm{P}};y)>0 when φ(ymin)Π~>0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}>0. The isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) (solid black curve) represents the level set {y|φ(y)=Π~}\{y|\varphi^{*}(y)=\tilde{\Pi}\}, which divides the space 𝒴\mathcal{Y} into the sublevel set (lower left) and the superlevel set (upper right). The equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (red line) is located in the superlevel set. The black vector is the normal vector φ(y)\partial\varphi^{*}(y) of the level set. Since the range of φ(y)=x\partial\varphi^{*}(y)=x covers the positive orthant, y𝒴(Π~,μ~)y\in\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) exists such that it satisfies Uφ(y)=0U\partial\varphi^{*}(y)=0, shown by the black circle. This point is on the stoichiometric manifold STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0). The orange vector represents (yPy)(y^{\mathrm{P}}-y), and the dashed line expresses the tangent plane of the level set at the point yy. One can find ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) such that any yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is located on the same side of the tangent plane with the black vector φ(y)\partial\varphi^{*}(y). For such ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0), the inner product φ(y)(yPy)\partial\varphi^{*}(y)(y^{\mathrm{P}}-y) is positive.

Appendix K Proof of negative K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) for any ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) when φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0

We show that, when φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is negative for ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0). Here, we assume that the stoichiometric matrix SS is productive, therefore, STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0)\neq\emptyset (see Appendix J).

Consider any point ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0). At the point yy, we can consider the tangent plane of the level set {y|φ(y)=Π~}\{y|\varphi^{*}(y)=\tilde{\Pi}\}, which divides the space of 𝒴\mathcal{Y} into two regions. Since K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) represents the inner product of the two vectors φ(y)\partial\varphi^{*}(y) and (yPy)(y^{\mathrm{P}}-y) in Eq. (104), the sign of K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is determined from the sides of the tangent plane to which the two vectors point. We also note that any point yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is on the one side of the tangent plane because the plane is parallel to the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (see Fig. 9).

When φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0, the equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (the red line in Fig. 9) is located both in the sublevel and the superlevel sets. From the convexity of the function φ(y)\varphi^{*}(y), any point in the sublevel set is located on the opposite side of the tangent plane to the orientation of the normal vector φ(y)\partial\varphi^{*}(y) at the point yy (see Fig. 9). Since any point yEQ𝒴(μ~)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is on the one side of the tangent plane, we can show that the particular solution yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is located on the opposite side of the tangent plane to the orientation of φ(y)\partial\varphi^{*}(y). Thus, for yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}), the inner product between φ(y)\partial\varphi^{*}(y) and (yiPyi)(y^{\mathrm{P}}_{i}-y_{i}) is negative. Therefore, for ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0), K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) is always negative.

Refer to caption
Figure 9: Illustration of the proof that K𝒴(yP;y)=φ(y)(yPy)K^{\mathcal{Y}}(y^{\mathrm{P}};y)=\partial\varphi^{*}(y)(y^{\mathrm{P}}-y) is negative for ySTO𝒴(L=0)y\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0) when φ(ymin)Π~<0\varphi^{*}(y^{\mathrm{min}})-\tilde{\Pi}<0. The isobaric manifold 𝒴(Π~,μ~)\mathcal{I}^{\mathcal{Y}}(\tilde{\Pi},\tilde{\mu}) (solid black curve) represents the level set {y|φ(y)=Π~}\{y|\varphi^{*}(y)=\tilde{\Pi}\}, which divides the space 𝒴\mathcal{Y} into the sublevel set (lower left) and the superlevel set (upper right). The equilibrium manifold EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) (red line) is located both in the sublevel and the superlevel sets. The black vector is the normal vector φ(y)\partial\varphi^{*}(y) of the level set, which satisfies Uφ(y)=0U\partial\varphi^{*}(y)=0. Thus, the point yy shown by the black circle is on the stoichiometric manifold STO𝒴(L=0)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L=0). The orange vector represents (yPy)(y^{\mathrm{P}}-y) and the dashed line expresses the tangent plane of the level set at the point yy. Any yPEQ𝒴(μ~)y^{\mathrm{P}}\in\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) is located on the opposite side of the tangent plane to the orientation of the black vector. Thus, the inner product φ(y)(yPy)\partial\varphi^{*}(y)(y^{\mathrm{P}}-y) is negative.

Appendix L Symbols and Notations

Table 2: List of symbols and notations for open chemical reaction systems (CRSs)
Symbol Description/definition First appearance
X={Xi}X=\{X^{i}\} The number of confined chemicals (ii = 1,,𝒩X1,\ldots,\mathcal{N}_{X}). Sec. II.1
X0={X0i}X_{0}=\{X^{i}_{0}\} The initial condition of XX Sec. II.1/Eq. (4)
N={Nm}N=\{N^{m}\} The number of open chemicals (mm = 1,,𝒩N1,\ldots,\mathcal{N}_{N}) Sec. II.1
𝒩X\mathcal{N}_{X} The number of species of confined chemicals Sec. II.1
𝒩N\mathcal{N}_{N} The number of species of open chemicals Sec. II.1
EE The energy of the CRS Sec. II.1
Ω\Omega The volume of the CRS Sec. II.1
μ~={μ~m}\tilde{\mu}=\{\tilde{\mu}_{m}\} The chemical potential of open chemicals in the environment Sec. II.1
T~\tilde{T} The temperature of the environment Sec. II.1
Π~\tilde{\Pi} The pressure of the environment Sec. II.1
N~={N~m}\tilde{N}=\{\tilde{N}^{m}\} The number of open chemicals in the environment Sec. II.1
E~\tilde{E} The energy of the environment Sec. II.1
Ω~\tilde{\Omega} The volume of the environment Sec. II.1
Σ[E,Ω,N,X]\Sigma[E,\Omega,N,X] The entropy function for the CRS Sec. II.1
Σ~T~,Π~,μ~[E~,Ω~,N~]\tilde{\Sigma}_{\tilde{T},\tilde{\Pi},\tilde{\mu}}[\tilde{E},\tilde{\Omega},\tilde{N}] The entropy function for the environment Sec. II.1
Σtot\Sigma^{\mathrm{tot}} The total entropy function Sec. II.1/Eq. (1)
σ[ϵ,n,x]\sigma[\epsilon,n,x] The entropy density Sec. II.1/Eq. (2)
(ϵ,n,x)(\epsilon,n,x) (E/Ω,N/Ω,X/Ω)(E/\Omega,N/\Omega,X/\Omega) Sec. II.1
S={Sri}S=\{S^{i}_{r}\} The stoichiometric matrix for confined chemicals Sec. II.1/Fig. 1
O={Ori}O=\{O^{i}_{r}\} The stoichiometric matrix for open chemicals Sec. II.1/Fig. 1
J(t)={Jr(t)}J(t)=\{J^{r}(t)\} The chemical reaction flux (rr = 1,,𝒩R1,\ldots,\mathcal{N}_{R}). Sec. II.1/Eq. (3)
𝒩R\mathcal{N}_{R} The number of reactions Sec. II.1
JE(t)J_{E}(t) The energy exchange rate with the environment Sec. II.1/Fig. 1
JΩ(t)J_{\Omega}(t) The volume expansion rate Sec. II.1/Fig. 1
JD(t)={JDm(t)}J_{D}(t)=\{J^{m}_{D}(t)\} The diffusion flux for open chemicals Sec. II.1/Fig. 1
Ξ(t)={Ξr(t)}\Xi(t)=\{\Xi^{r}(t)\} The extent of reaction Sec. II.1/Eq. (4)
L={Ll=UilX0i}L=\{L^{l}=U^{l}_{i}X^{i}_{0}\} The conserved quantities (ll = 1,,dimKer[ST]1,\ldots,\dim\mathrm{Ker}[S^{T}]). Sec. II.1/Eq. (6)
U={Uil}U=\{U^{l}_{i}\} A basis matrix whose row vectors form the basis of Ker[ST]\mathrm{Ker}[S^{T}] Sec. II.1/Eq. (6)
L^\hat{L} The conserved quantities under the isochoricisochoric condition Sec. III.3
φ(x)\varphi(x) The partial grand potential density Sec. II.1/Eq. (13)
yy The Legendre dual variable of xx Sec. II.2
φ(y)\varphi^{*}(y) The full grand potential density Sec. II.2/Eq. (20)
Πmin\Pi_{\mathrm{min}} The minimum pressure that the CRS can take Sec. III.1/Eq. (34)
fr(Ξ)f_{r}(\Xi) The thermodynamic force Sec. V.2/Eq. (60)
Table 3: List of symbols and notations for geometric descriptions and manifolds
Symbol Description/definition First appearance
𝔛\mathfrak{X} The number space >0𝒩X\mathbb{R}_{>0}^{\mathcal{N}_{X}} Sec. II.1
𝒳\mathcal{X} The density space >0𝒩X\mathbb{R}_{>0}^{\mathcal{N}_{X}} Sec. II.1
𝒴\mathcal{Y} The chemical potential space 𝒩X\mathbb{R}^{\mathcal{N}_{X}} Sec. II.2
\mathcal{L} The space of conserved quantities dimKer[ST]\mathbb{R}^{\dim\mathrm{Ker}[S^{T}]} Sec. II.1
EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) The equilibrium manifold in 𝒴\mathcal{Y} Sec. II.2/Eq. (23)
EQ𝒳(T~,μ~)\mathcal{M}^{\mathcal{X}}_{\mathrm{EQ}}(\tilde{T},\tilde{\mu}) The equilibrium manifold in 𝒳\mathcal{X} Sec. III.4/Eq.(51)
#(Π~,μ~)\mathcal{I}^{\#}(\tilde{\Pi},\tilde{\mu}) The isobaric manifold in #=𝒳\#=\mathcal{X} or 𝒴\mathcal{Y} Sec. III.2/Eq. (37), Eq. (39)
STO#(L)\mathcal{M}^{\#}_{\mathrm{STO}}(L) The stoichiometric manifold/compatibility class in #=𝔛\#=\mathfrak{X}, 𝒳\mathcal{X}, or 𝒴\mathcal{Y} Sec. II.3/Eq. (24)-(26)
^STO#(L^)\hat{\mathcal{M}}^{\#}_{\mathrm{STO}}(\hat{L}) The stoichiometric manifold in #=𝒳\#=\mathcal{X} or 𝒴\mathcal{Y} under the isochoricisochoric condition Sec. III.3/Eq. (43), Eq. (44)
ρ𝒳(X)\rho_{\mathcal{X}}(X) A nonlinear function of XX defined as ρ𝒳(X)=X/Ω(X)\rho_{\mathcal{X}}(X)=X/\Omega(X) Sec. II.1/Eq. (18)
𝔯𝒳(x)\mathfrak{r}^{\mathcal{X}}(x) The corresponding ray to xx in the space 𝔛\mathfrak{X} Sec. II.1/Eq. (19)
𝔯𝒴(y)\mathfrak{r}^{\mathcal{Y}}(y) The corresponding ray to yy in the space 𝔛\mathfrak{X} Sec. II.2/Eq. (21)
yminy^{\mathrm{min}} Eq. (28) Sec. II.4
xminx_{\mathrm{min}} φ(ymin)\partial\varphi^{*}(y^{\mathrm{min}}) Sec. II.6/Fig. 2
yPy^{\mathrm{P}} A particular solution to Eq. (50) Sec. III.4/Eq. (53)
h={hl}h=\{h_{l}\} A coordinate of Ker[ST]\mathrm{Ker}[S^{T}] Sec. III.4/Eq. (53)
yEQy^{\mathrm{EQ}} The intersecting point STO𝒴(L)EQ𝒴(μ~)\mathcal{M}^{\mathcal{Y}}_{\mathrm{STO}}(L)\cap\mathcal{M}^{\mathcal{Y}}_{\mathrm{EQ}}(\tilde{\mu}) Sec. III.5/Theorem 1
For regular SS, yiEQ=μ~mOrm(S1)iry^{\mathrm{EQ}}_{i}=-\tilde{\mu}_{m}O^{m}_{r}(S^{-1})^{r}_{i}. Sec. II.5
yB(L^)y^{\mathrm{B}}(\hat{L}) Birch’s point Sec. V.1/Eq. (55)
𝔟(L)\mathfrak{b}(L) Birch’s trajectory Sec. V.1/Eq. (56)
K𝒴(yP;y)K^{\mathcal{Y}}(y^{\mathrm{P}};y) Eq. (62) Sec. V.2
𝒟𝒴[y||y]\mathcal{D}^{\mathcal{Y}}[y^{\prime}||y] The Bregman divergence Sec. V.2/Eq. (63)
ΞB\Xi_{B} The extent of reaction Ξ\Xi at the boundary of the domain Sec. V.2
XBX_{B} X0+SΞBX_{0}+S\Xi_{B} Sec. V.2
B(XB)\mathfrak{I}_{B}(X_{B}) The index set such that XBi=0X^{i}_{B}=0 Sec. V.2
Table 4: List of symbols and notations used for describing the ideal gas and mass action kinetics
Symbol Description/definition First appearance
RR The gas constant Sec.II.1
νo(T~)={νio(T~)}\nu^{o}(\tilde{T})=\{\nu^{o}_{i}(\tilde{T})\} The standard chemical potentials of confined chemicals Sec.II.1
μo(T~)={μmo(T~)}\mu^{o}(\tilde{T})=\{\mu^{o}_{m}(\tilde{T})\} The standard chemical potentials of open chemicals Sec.II.1
n~={n~m}\tilde{n}=\{\tilde{n}^{m}\} The density of open chemicals in the environment Sec.II.1/Eq. (16)
xoix^{i}_{o} exp[νio(T~)/RT~]\exp[-\nu^{o}_{i}(\tilde{T})/R\tilde{T}] Fig. 2/Appendix B
nomn^{m}_{o} exp[μmo(T~)/RT~]\exp[-\mu^{o}_{m}(\tilde{T})/R\tilde{T}] Fig. 2/Appendix B
w+rw^{r}_{+}, wrw^{r}_{-} The rate constants of rr-th forward and backward reactions Appendix B
w^+r\hat{w}^{r}_{+}, w^r\hat{w}^{r}_{-} The rate constants where we absorb the constant densities of open chemicals Appendix B