22
Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics Riccardo Rao and Massimiliano Esposito Complex Systems and Statistical Mechanics, Physics and Materials Science Research Unit, University of Luxembourg, L-1511 Luxembourg, Luxembourg (Dated: January 13, 2017. Published in Phys. Rev. X, DOI: 10.1103/PhysRevX.6.041064) We build a rigorous nonequilibrium thermodynamic description for open chemical reaction networks of elementary reactions (CRNs). Their dynamics is described by deterministic rate equations with mass action kinetics. Our most general framework considers open networks driven by time-dependent chemostats. The energy and entropy balances are established and a nonequilibrium Gibbs free energy is introduced. The difference between this latter and its equilibrium form represents the minimal work done by the chemostats to bring the network to its nonequilibrium state. It is minimized in nondriven detailed-balanced networks (i.e. networks which relax to equilibrium states) and has an interesting information-theoretic interpretation. We further show that the entropy production of complex balanced networks (i.e. networks which relax to special kinds of nonequilibrium steady states) splits into two non-negative contributions: one characterizing the dissipation of the nonequilibrium steady state and the other the transients due to relaxation and driving. Our theory lays the path to study time-dependent energy and information transduction in biochemical networks. PACS numbers: 05.70.Ln, 87.16.Yc I. INTRODUCTION Thermodynamics of chemical reactions has a long his- tory. The second half of the 19th century witnessed the dawn of the modern studies on thermodynamics of chem- ical mixtures. It is indeed at that time that J. W. Gibbs introduced the concept of chemical potential and used it to define the thermodynamic potentials of non-interacting mixtures [1]. Several decades later, this enabled T. de Don- der to approach the study of chemical reacting mixtures from a thermodynamic standpoint. He proposed the concept of affinity to characterize the chemical force ir- reversibly driving chemical reactions and related it to the thermodynamic properties of mixtures established by Gibbs [2]. I. Prigogine, who perpetuated the Brussels School founded by de Donder, introduced the assumption of local equilibrium to describe irreversible processes in terms of equilibrium quantities [3, 4]. In doing so, he pioneered the connections between thermodynamics and kinetics of chemical reacting mixtures [5]. During the second half of the 20th century, part of the attention moved to systems with small particle numbers which are ill-described by “deterministic” rate equations. The Brussels School, as well as other groups, produced various studies on the nonequilibrium thermodynamics of chemical systems [611] using a stochastic description based on the (Chemical) Master Equation [12, 13]. These studies played an important role during the first decade of the 21st century for the development of Stochastic Thermodynamics, a theory that systematically establishes a nonequilibrium thermodynamic description for systems obeying stochastic dynamics [1417], including chemical reaction networks (CRNs) [1822]. Another significant part of the attention moved to the thermodynamic description of biochemical reactions in terms of deterministic rate equations [23, 24]. This is not so surprising since living systems are the paramount ex- ample of nonequilibrium processes and they are powered by chemical reactions. The fact that metabolic processes involve thousands of coupled reactions also emphasized the importance of a network description [2527]. While complex dynamical behaviors such as oscillations were analyzed in small CRNs [28, 29], most studies on large bio- chemical networks focused on the steady-state dynamics. Very few studies considered the thermodynamic properties of CRNs [3033]. One of the first nonequilibrium thermo- dynamic description of open biochemical networks was proposed in Ref. [34]. However, it did not take advantage of Chemical Reaction Network Theory which connects the network topology to its dynamical behavior and which was extensively studied by mathematicians during the seventies [3537] (this theory was also later extended to stochastic dynamics [3841]). As far as we know, the first and single study that related the nonequilibrium thermodynamics of CRNs to their topology is Ref. [22], still restricting itself to steady states. In this paper, we consider the most general setting for the study of CRNs, namely open networks driven by chemostatted concentrations which may change over time. To the best of our knowledge, this was never con- sidered before. In this way, steady-state properties as well as transient ones are captured. Hence, in the same way that Stochastic Thermodynamics is built on top of stochastic dynamics, we systematically build a nonequi- librium thermodynamic description of CRNs on top of deterministic chemical rate equations. In doing so, we establish the energy and entropy balance and introduce the nonequilibrium entropy of the CRN as well as its nonequilibrium Gibbs free energy. We show this latter to bear an information-theoretical interpretation similar to that of Stochastic Thermodynamics [4245] and to be related to the dynamical potentials derived by mathemati- arXiv:1602.07257v3 [cond-mat.stat-mech] 11 Jan 2017

Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

  • Upload
    others

  • View
    10

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

Nonequilibrium Thermodynamics of Chemical Reaction Networks:Wisdom from Stochastic Thermodynamics

Riccardo Rao and Massimiliano EspositoComplex Systems and Statistical Mechanics, Physics and Materials ScienceResearch Unit, University of Luxembourg, L-1511 Luxembourg, Luxembourg

(Dated: January 13, 2017. Published in Phys. Rev. X, DOI: 10.1103/PhysRevX.6.041064)

We build a rigorous nonequilibrium thermodynamic description for open chemical reaction networksof elementary reactions (CRNs). Their dynamics is described by deterministic rate equations withmass action kinetics. Our most general framework considers open networks driven by time-dependentchemostats. The energy and entropy balances are established and a nonequilibrium Gibbs free energyis introduced. The difference between this latter and its equilibrium form represents the minimal workdone by the chemostats to bring the network to its nonequilibrium state. It is minimized in nondrivendetailed-balanced networks (i.e. networks which relax to equilibrium states) and has an interestinginformation-theoretic interpretation. We further show that the entropy production of complexbalanced networks (i.e. networks which relax to special kinds of nonequilibrium steady states) splitsinto two non-negative contributions: one characterizing the dissipation of the nonequilibrium steadystate and the other the transients due to relaxation and driving. Our theory lays the path to studytime-dependent energy and information transduction in biochemical networks.

PACS numbers: 05.70.Ln, 87.16.Yc

I. INTRODUCTION

Thermodynamics of chemical reactions has a long his-tory. The second half of the 19th century witnessed thedawn of the modern studies on thermodynamics of chem-ical mixtures. It is indeed at that time that J. W. Gibbsintroduced the concept of chemical potential and used itto define the thermodynamic potentials of non-interactingmixtures [1]. Several decades later, this enabled T. de Don-der to approach the study of chemical reacting mixturesfrom a thermodynamic standpoint. He proposed theconcept of affinity to characterize the chemical force ir-reversibly driving chemical reactions and related it tothe thermodynamic properties of mixtures established byGibbs [2]. I. Prigogine, who perpetuated the BrusselsSchool founded by de Donder, introduced the assumptionof local equilibrium to describe irreversible processes interms of equilibrium quantities [3, 4]. In doing so, hepioneered the connections between thermodynamics andkinetics of chemical reacting mixtures [5].

During the second half of the 20th century, part of theattention moved to systems with small particle numberswhich are ill-described by “deterministic” rate equations.The Brussels School, as well as other groups, producedvarious studies on the nonequilibrium thermodynamicsof chemical systems [6–11] using a stochastic descriptionbased on the (Chemical) Master Equation [12, 13]. Thesestudies played an important role during the first decadeof the 21st century for the development of StochasticThermodynamics, a theory that systematically establishesa nonequilibrium thermodynamic description for systemsobeying stochastic dynamics [14–17], including chemicalreaction networks (CRNs) [18–22].

Another significant part of the attention moved to thethermodynamic description of biochemical reactions interms of deterministic rate equations [23, 24]. This is not

so surprising since living systems are the paramount ex-ample of nonequilibrium processes and they are poweredby chemical reactions. The fact that metabolic processesinvolve thousands of coupled reactions also emphasizedthe importance of a network description [25–27]. Whilecomplex dynamical behaviors such as oscillations wereanalyzed in small CRNs [28, 29], most studies on large bio-chemical networks focused on the steady-state dynamics.Very few studies considered the thermodynamic propertiesof CRNs [30–33]. One of the first nonequilibrium thermo-dynamic description of open biochemical networks wasproposed in Ref. [34]. However, it did not take advantageof Chemical Reaction Network Theory which connects thenetwork topology to its dynamical behavior and whichwas extensively studied by mathematicians during theseventies [35–37] (this theory was also later extended tostochastic dynamics [38–41]). As far as we know, thefirst and single study that related the nonequilibriumthermodynamics of CRNs to their topology is Ref. [22],still restricting itself to steady states.

In this paper, we consider the most general settingfor the study of CRNs, namely open networks drivenby chemostatted concentrations which may change overtime. To the best of our knowledge, this was never con-sidered before. In this way, steady-state properties aswell as transient ones are captured. Hence, in the sameway that Stochastic Thermodynamics is built on top ofstochastic dynamics, we systematically build a nonequi-librium thermodynamic description of CRNs on top ofdeterministic chemical rate equations. In doing so, weestablish the energy and entropy balance and introducethe nonequilibrium entropy of the CRN as well as itsnonequilibrium Gibbs free energy. We show this latterto bear an information-theoretical interpretation similarto that of Stochastic Thermodynamics [42–45] and to berelated to the dynamical potentials derived by mathemati-

arX

iv:1

602.

0725

7v3

[co

nd-m

at.s

tat-

mec

h] 1

1 Ja

n 20

17

Page 2: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

2

cians. We also show the relation between the minimalchemical work necessary to manipulate the CRNs farfrom equilibrium and the nonequilibrium Gibbs free en-ergy. Our theory embeds both the Prigoginian approachto thermodynamics of irreversible processes [5] and thethermodynamics of biochemical reactions [23]. Makingfull use of the mathematical Chemical Reaction NetworkTheory, we further analyze the thermodynamic behav-ior of two important classes of CRNs: detailed-balancednetworks and complex-balanced networks. In absence oftime-dependent driving, the former converges to thermo-dynamic equilibrium by minimizing their nonequilibriumGibbs free energy. In contrast, the latter converges to aspecific class of nonequilibrium steady states and alwaysallow for an adiabatic–nonadiabatic separation of theirentropy production, which is analogous to that found inStochastic Thermodynamics [46–50]. We note that whilefinalizing this paper, a result similar to the latter wasindependently found in Ref. [51].

Outline and Notation

The paper is organized as follows. After introducingthe necessary concepts in chemical kinetics and ChemicalReaction Network Theory, sec. II, the nonequilibriumthermodynamic description is established, sec. III. As inStochastic Thermodynamics, we build it on top of thedynamics and formulate the entropy and energy balance,§ IIID and § III E. Chemical work and nonequilibriumGibbs free energy are also defined and the information-theoretic content of the latter is discussed. The specialproperties of detailed-balance and of complex-balancednetworks are considered in sec. V and IV, respectively.Conclusions and perspectives are drawn in sec. VI, whilesome technical derivations are detailed in the appendices.

We now proceed by fixing the notation. We consider asystem composed of reacting chemical species Xσ, eachof which is identified by an index σ ∈ S, where S is theset of all indices/species. The species populations changedue to elementary reactions, i.e. all reacting species andreactions must be resolved (none can be hidden), and allreactions must be reversible, i.e. each forward reaction +ρhas a corresponding backward reaction −ρ. Each pair offorward–backward reactions is a reaction pathway denotedby ρ ∈ R. The orientation of the set of reaction pathwaysR is arbitrary. Hence, a generic CRN is represented as∑

σ

∇σ+ρ Xσk+ρ

k−ρ

∑σ

∇σ−ρ Xσ . (1)

The constants k+ρ (k−ρ) are the rate constants of theforward (backward) reactions. The stoichiometric coeffi-cients −∇σ+ρ and ∇σ−ρ identify the number of moleculesof Xσ involved in each forward reaction +ρ (the stoi-chiometric coefficients of the backward reactions haveopposite signs). Once stacked into two non-negative ma-trices ∇+ = {∇σ+ρ} and ∇− = {∇σ−ρ}, they define the

system

environment

k+1

k-1Xa 2Xb + Xc

k+2

k-2Xc + Xd Xe

FIG. 1: Representation of a closed CRN. The chemicalspecies are {Xa, · · · ,Xe}. The two reaction pathways arelabeled by 1 and 2. The nonzero stoichiometric coefficientsare −∇a

+1 = −1, ∇b−1 = 2 and ∇c

−1 = 1 for the firstforward reaction and −∇c

+2 = −1, −∇d+2 = −1 and

∇e−2 = 1 for the second one. Since the network is closed,

no chemical species is exchanged with the environment.

integer-valued stoichiometric matrix

∇ ≡ ∇− −∇+ . (2)

The reason for the choice of the symbol “∇” will becomeclear later.

Example 1. The stoichiometric matrix of the CRN de-picted in fig. 1 is

∇ =

−1 02 01 −10 −10 1

. (3)

Physical quantities associated to species and reactionsare represented in upper–lower indices vectorial nota-tion. Upper and lower indexed quantities have the samephysical values, e.g. Zi = Zi,∀ i. We use the Einsteinsummation notation: repeated upper–lower indices im-plies the summation over all the allowed values for thoseindices—e.g. σ ∈ S for species and ρ ∈ R for reactions.Given two arbitrary vectorial quantities a = {ai} andb = {bi}, the following notation will be used

aibi ≡

∏i

aibi.

Finally, given the matrix C, whose elements are {Cij}, theelements of the transposed matrix CT are {Cji }.

The time derivative of a physical quantity A is denotedby dtA, its steady state value by an overbar, A, and itsequilibrium value by Aeq or Aeq. We reserve the overdot,A, to denote the rate of change of quantities which arenot exact time derivatives.

Page 3: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

3

II. DYNAMICS OF CRNS

In this section, we formulate the mathematical descrip-tion of CRNs [52, 53] in a suitable way for a thermo-dynamic analysis. We introduce closed and open CRNsand show how to drive these latter in a time-dependentway. We then define conservation laws and cycles and re-view the dynamical properties of two important classes ofCRNs: detailed-balanced networks and complex-balancednetworks.We consider a chemical system in which the reacting

species {Xσ} are part of a homogeneous and ideal di-lute solution: the reactions proceed slowly compared todiffusion and the solvent is much more abundant thanthe reacting species. Temperature T and pressure p arekept constant. Since the volume of the solution V isoverwhelmingly dominated by the solvent, it is assumedconstant. The species abundances are large enough sothat the molecules discreteness can be neglected. Thus,at any time t, the system state is well-described by themolar concentration distribution {Zσ ≡ Nσ/V }, whereNσ is the molarity of the species Xσ.

The reaction kinetics is controlled by the reaction ratefunctions J±ρ

({Zσ}

), which measure the rate of occur-

rence of reactions and satisfy the mass action kinetics[52, 54, 55]

J±ρ ≡ J±ρ({Zσ}

)= k±ρZσ∇

±ρσ . (4)

The net concentration current along a reaction pathwayρ is thus given by

Jρ ≡ J+ρ − J−ρ = k+ρZσ∇+ρσ − k−ρZσ∇

−ρσ . (5)

Example 2. For the CRN in fig. 1 the currents are

J1 = k+1Za − k−1(Zb)2Zc

J2 = k+2ZcZd − k−2Ze .(6)

A. Closed CRNs

A closed CRN does not exchange any chemical specieswith the environment. Hence, the species concentrationsvary solely due to chemical reactions and satisfy the rateequations

dtZσ = ∇σρ Jρ, ∀σ ∈ S . (7)

Since rate equations are nonlinear, complex dynamicalbehaviors may emerge [29]. The fact that the rate equa-tions (7) can be thought of as a continuity equation forthe concentration, where the stoichiometric matrix ∇ (2)acts as a discrete differential operator, explains the choiceof the symbol “∇” for the stoichiometric matrix [56].

system

environment

k+1

k-1Ya 2Xb + Xc

k+2

k-2Xc + Xd Ye

FIG. 2: Representation of an open CRN. With respect tothe CRN in fig. 1, the species Xa and Xe are chemostatted,hence represented as Ya and Ye. The green boxes asiderepresent the reservoirs of chemostatted species.

B. Driven CRNs

In open CRNs, matter is exchanged with the environ-ment via reservoirs which control the concentrations ofsome specific species, fig. 2. These externally controlledspecies are said to be chemostatted, while the reservoirscontrolling them are called chemostats. The chemostat-ting procedure may mimic various types of controls bythe environment. For instance, a direct control could beimplemented via external reactions (not belonging to theCRN) or via abundant species whose concentrations arenegligibly affected by the CRN reactions within relevanttime scales. An indirect control may be achieved viasemipermeable membranes or by controlled injection ofchemicals in continuous stirred-tank reactors.

Among the chemical species, the chemostatted ones aredenoted by the indices σy ∈ Sy, and the internal ones byσx ∈ Sx (S ≡ Sx∪Sy). Also, the part of the stoichiometricmatrix related to the internal (resp. chemostatted) speciesis denoted by ∇X =

{∇σxρ

}(resp. ∇Y =

{∇σyρ

}).

Example 3. When chemostatting the CRN in fig. 1 asin fig. 2 the stoichiometric matrix (3) splits into

∇X =

2 01 −10 −1

, ∇Y =(−1 00 1

). (8)

In nondriven open CRNs the chemostatted species haveconstant concentrations, i.e. {dtZσy = 0}. In drivenopen CRNs the chemostatted concentrations change overtime according to some time-dependent protocol π(t):{Zσy ≡ Zσy(π(t))}. The changes of the internal speciesare solely due to reactions and satisfy the rate equations

dtZσx = ∇σxρ J

ρ , ∀σx ∈ Sx . (9)

Instead, the changes of chemostatted species {dtZσy} arenot only given by the species formation rates {∇σy

ρ Jρ}

Page 4: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

4

but must in addition contain the external currents {Iσy},which quantify the rate at which chemostatted speciesenter into the CRN (negative if chemostatted species leavethe CRN),

dtZσy = ∇σyρ J

ρ + Iσy , ∀σy ∈ Sy . (10)

This latter equation is not a differential equation sincethe chemostatted concentrations {Zσy} are not dynam-ical variables. It shows that the external control of thechemostatted concentration is not necessarily direct, viathe chemostatted concentrations, but can also be indi-rectly controlled via the external currents. We note that(10) is the dynamical expression of the decomposition ofchanges of species populations in internal–external intro-duced by de Donder (see [57, §§ 4.1 and 15.2]).A steady-state distribution {Zσx}, if it exists, must

satisfy

∇σxρ Jρ = 0 , ∀σx ∈ Sx , (11a)

∇σyρ Jρ + Iσy = 0 , ∀σy ∈ Sy , (11b)

for given chemostatted concentrations {Zσy}.

C. Conservation Laws

In a closed CRN, a conservation law ` = {`σ} is a leftnull eigenvector of the stoichiometric matrix ∇ [23, 25]

`σ∇σρ = 0 , ∀ ρ ∈ R . (12)

Conservation laws identify conserved quantities L ≡`σ Z

σ, called components [23, 25], which satisfy

dtL = `σ dtZσ = 0 . (13)

We denote a set of independent conservation laws ofthe closed network by

{`λ}and the corresponding com-

ponents by{Lλ ≡ `λσ Zσ

}. The choice of this set is not

unique, and different choices have different physical mean-ings. This set is never empty since the total mass isalways conserved. Physically, conservation laws are oftenrelated to parts of molecules, called moieties [58], whichare exchanged between different species and/or subjectto isomerization (see example 4).

In an open CRN, since only {Zσx} are dynamical vari-ables, the conservation laws become the left null eigenvec-tors of the stoichiometric matrix of the internal species∇X. Stated differently, when starting from the closedCRN, the chemostatting procedure may break a subsetof the conservation laws of the closed network {`λ} [56].E.g. when the first chemostat is introduced the total massconservation law is always broken. Within the set

{`λ},

we label the broken ones by λb and the unbroken ones byλu. The broken conservation laws are characterized by

`λbσx∇σxρ︸ ︷︷ ︸

6= 0

+`λbσy∇σyρ = 0 , ∀ ρ ∈ R , (14)

system

environment

k+1

k-1H2O 2H + O

k+2

k-2O + C CO

FIG. 3: Specific implementation of the CRN in fig. 2.

where the first term is nonvanishing for at least one ρ ∈ R.The broken components {Lλb ≡ `λb

σ Zσ} are no longerconstant over time. On the other hand, the unbrokenconservation laws are characterized by

`λuσx∇σxρ︸ ︷︷ ︸

= 0

+`λuσy∇σyρ = 0 , ∀ ρ ∈ R , (15)

where the first term vanishes for all ρ ∈ R. Therefore, theunbroken components {Lλu ≡ `λu

σ Zσ} remain constantover time. Without loss of generality, we choose the set{`λ} such that the entries related to the chemostattedspecies vanish, `λu

σy= 0, ∀λu, σy.

Example 4. For the CRN in fig. 1, an independent setof conservation laws is

`1 =(2 1 0 0 0

),

`2 =(0 0 0 1 1

)and

`3 =(0 1

2 −1 1 0).

(16)

When chemostatting as in fig. 2, the first two conservationlaws break while the last one remains unbroken. Wealso note that this set is chosen so that the unbrokenconservation law satisfies `3a = `3e = 0. When consideringthe specific implementation in fig. 3 of the CRN in fig. 2,we see that the first two conservation laws in (16) representthe conservation of the concentrations of the moiety Hand C, respectively. Instead, the third conservation lawin (16) does not have a straightforward interpretation. Itis related to the fact that when the species H or C areproduced also O must be produced and vice versa.

D. Detailed-Balanced Networks

A steady state (11) is said to be an equilibrium state,{Zσeq}, if it satisfies the detailed balance property [57, § 9.4],i.e. all concentration currents (5) vanish,

Jρeq ≡ Jρ({Zσeq}

)= 0, ∀ ρ ∈ R . (17)

Page 5: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

5

For open networks, this means that the external currents,eq. (11b), must also vanish {Iσy

eq = 0}. By virtue of massaction kinetics, eq. (4), the detailed balance property (17)can be rewritten as

k+ρ

k−ρ= Zσeq

∇ρσ , ∀ ρ ∈ R . (18)

A CRN is said to be detailed-balanced if, for givenkinetics {k±ρ} and chemostatting {Zσy}, its dynamicsexhibits an equilibrium steady state (17). For each setof unbroken components {Lλu}—which are given by theinitial condition and constrain the space where the dynam-ics dwells—the equilibrium distribution is globally stable[59]. Equivalently, detailed-balanced networks always re-lax to an equilibrium state, which for a given kinetics andchemostatting is unique and depends on the unbrokencomponents only, see also sec. V.Closed CRNs must be detailed-balanced. This state-

ment can be seen as the zeroth law for CRNs. Conse-quently, rather than considering eq. (18) as a propertyof the equilibrium distribution, we impose it as a prop-erty that the rate constants must satisfy and call it localdetailed-balance property. It is a universal property ofelementary reactions which holds regardless of the net-work state. Indeed, while the equilibrium distributiondepends on the components, the r.h.s. of eq. (18) doesnot. This point will become explicit after introducing thethermodynamic structure, eq. (88) in sec.V. The localdetailed-balance property will be rewritten in a thermo-dynamic form in § III B, eq. (50).

In open nondriven CRNs, the chemostatting proceduremay prevent the system from reaching an equilibriumstate. To express this scenario algebraically we now in-troduce the concepts of emergent cycle and cycle affinity.A cycle c = {cρ} is a right null eigenvector of the

stoichiometric matrix [56], namely

∇σρ cρ = 0 , ∀σ ∈ S . (19)

Since ∇ is integer-valued, c can always be rescaled toonly contain integer coefficients. In this representation,its entries denote the number of times each reaction occurs(negative signs identify reactions occurring in backwarddirection) along a transformation which overall leaves theconcentration distributions {Zσ} unchanged, see example5. We denote by {cα} a set of linearly independent cycles.An emergent cycle c = {cρ} is defined algebraically as[56] {

∇σxρ c

ρ = 0 , ∀σx ∈ Sx

∇σyρ c

ρ 6= 0 , for at least one σy ∈ Sy .(20)

In its integer-valued representation, the entries of c de-note the number of times each reaction occurs along atransformation which overall leaves the concentrations ofthe internal species {Zσx} unchanged while changing theconcentrations of the chemostatted species by an amount∇σyρ cρ. These latter are however immediately restored to

their prior values due to the injection of −∇σyρ cρ molecules

of Xσyperformed by the chemostats. Emergent cycles are

thus pathways transferring chemicals across chemostatswhile leaving the internal state of the CRN unchanged.We denote by {cε} a set of linearly independent emergentcycles.

When chemostatting an initially closed CRN, for eachspecies which is chemostatted, either a conservation lawbreaks—as mentioned in § II C—or an independent emer-gent cycle arises [56]. This follows from the rank nul-lity theorem for the stoichiometric matrices ∇ and ∇X,which ensures that the number of chemostatted species|Sy| equals the number of broken conservation laws |λb|plus the number of independent emergent cycles |ε|:|Sy| = |λb|+ |ε|. Importantly, the rise of emergent cyclesis a topological feature: it depends on the species whichare chemostatted, but not on the chemostatted concentra-tions. We also note that emergent cycles are modeled as“flux modes” in the context of metabolic networks [60–62].

Example 5. To illustrate the concepts of cycles andemergent cycles, we use the following CRN [56]

Y1 +Xa Xb

Xc Y2 +Xa

k+1

k−4 k+2

k−1

k+4

k−3

k+3

k−2

(21)

whose Y1 and Y2 species are chemostatted. The stoichio-metric matrix decomposes as

∇X =

−1 1 −1 11 −1 0 00 0 1 −1

∇Y =(−1 0 0 10 1 −1 0

).

(22)

The set of linearly independent cycles (19) consists of onlyone cycle, which can be written as

c =(1 1 1 1

)T. (23)

As the CRN is chemostatted one linearly independentemergent cycle (20) arises

c =(1 1 −1 −1

)T. (24)

We now see that if each reaction occurs a number of timesgiven by the entry of the cycle (23), the CRN goes back tothe initial state, no matter which one it is. On the otherhand, when the emergent cycle (24) is performed, thestate of the internal species does not change, while twomolecules of Y1 are annihilated and two of Y2 are created.However, since the chemostats restore their initial values,the overall result of c is to transfer two Y1, transformedin Y2, from the first to the second chemostat.

Page 6: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

6

The closed version of this CRN has two independentconservation laws,

`1 =(0 1 1 1 1

)`2 =

(1 1 1 0 0

) (25)

the first of which, `1, is broken following the chemostattingof any of the two species Y1 or Y2. The other chemostat-ted species, instead, gives rise to the emergent cycle (24),so that the relationship |Sy| = |λb|+ |ε| is satisfied.

Any cycle cα and emergent cycle cε bears a cycle affinity[56]

Aα = cραRT ln J+ρ

J−ρ, (26)

Aε = cρε RT ln J+ρ

J−ρ. (27)

From the definition of cycle (19) and current (5), andthe local detailed balance (18), it follows that the cycleaffinities along the cycles (19) vanish, {Aα = 0}, and thatthe cycle affinities along the emergent cycles only dependon the chemostatted concentrations

Aε = cρε RT ln k+ρ

k−ρZ−∇σy

ρσy . (28)

Since emergent cycles are pathways connecting differentchemostats, the emergent affinities quantify the chemicalforces acting along the cycles. This point will becomeclearer later, when the thermodynamic expressions of theemergent cycle affinities {Aε} will be given, eq. (49).

A CRN is detailed-balanced if and only if all the emer-gent cycle affinities {Aε} vanish. This condition is equiv-alent to the Wegscheider’s condition [59]. This happenswhen the chemostatted concentrations fit an equilibriumdistribution. As a special case, unconditionally detailed-balanced networks are open CRNs with no emergent cycle.Therefore, they are detailed-balanced for any choice of thechemostatted concentrations. Consequently, even when atime-dependent driving acts on such a CRN and preventsit from reaching an equilibrium state, a well-defined equi-librium state exists at any time: the equilibrium state towhich the CRN would relax if the time-dependent drivingwere stopped.

Example 6. Any CRN with one chemostatted speciesonly (|Sy| = 1) is unconditionally detailed-balanced. In-deed, as mentioned in § II C, the first chemostatted speciesalways breaks the mass conservation law, |λb| = 1 andthus no emergent cycle arises, |ε| = |Sy| − |λb| = 0.

The open CRN in fig. 2 is an example of unconditionallydetailed-balanced network with two chemostatted species,since the chemostatting breaks two conservation laws,see example 4. Indeed, a nonequilibrium steady statewould require a continuous injection of Ya and ejection ofYe (or vice versa). But this would necessary result in acontinuous production of Xb and consumption of Xd whichis in contradiction with the steady state assumption.

Finally, a tacit assumption in the above discussion isthat the network involves a finite number of species andreactions, i.e. the CRN is finite-dimensional. Infinite-dimensional CRNs can exhibits long-time behaviors differ-ent from equilibrium even in absence of emergent cycles[63].

E. Complex-Balanced Networks

To discuss complex-balanced networks and complex-balanced distributions, we first introduce the notion ofcomplex in open CRNs.A complex is a group of species which combines in a

reaction as products or as reactants. Each side of eq. (1)defines a complex but different reactions might involvethe same complex. We label complexes by γ ∈ C, whereC is the set of complexes.

Example 7. Let us consider the following CRN [64]

Xak+1

−−−⇀↽−−−k−1

Xb

Xa + Xbk+2

−−−⇀↽−−−k−2

2 Xbk+3

−−−⇀↽−−−k−3

Xc

. (29)

The set of complexes is C = {Xa,Xb,Xa + Xb, 2Xb,Xc},and the complex 2Xb is involved in both the second andthird reaction.

The notion of complex allows us to decompose thestoichiometric matrix ∇ as

∇σρ = Γσγ ∂γρ . (30)

We call Γ = {Γσγ} the composition matrix [35, 37]. Itsentries Γσγ are the stoichiometric number of species Xσ

in the complex γ. The composition matrix encodes thestructure of each complex in terms of species, see example8. The matrix ∂ = {∂γρ } denotes the incidence matrix ofthe CRN, whose entries are given by

∂γρ =

1 if γ is the product complex of +ρ ,−1 if γ is the reactant complex of +ρ ,0 otherwise.

(31)

The incidence matrix encodes the structure of the networkat the level of complexes, i.e. how complexes are con-nected by reactions. If we think of complexes as networknodes, the incidence matrix associates an edge to eachreaction pathway and the resulting topological structureis a reaction graph, e.g. fig. 1 and eqs. (21) and (29). Thestoichiometric matrix instead encodes the structure of thenetwork at the level of species. If we think of species asthe network nodes, the stoichiometric matrix does notdefine a graph since reaction connects more than a pair ofspecies, in general. The structure originating is rather ahyper-graph [56, 65], or equivalently a Petri net [66, 67].

Page 7: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

7

Example 8. The composition matrix and the incidencematrix of the CRN in (29) are

Γ =

1 0 1 0 00 1 1 2 00 0 0 0 1

, ∂ =

−1 0 01 0 00 −1 00 1 −10 0 1

,

(32)where the complexes are ordered as in example 7. Thecorresponding reaction hyper-graph is

Xa Xb Xc

k+1

k+2

k+3

(33)where only the forward reactions are depicted.

In an open CRN, we regroup all complexes γ ∈ C ofthe closed CRN which have the same stoichiometry forthe internal species (i.e. all complexes with the sameinternal part of the composition matrix ΓX

γ regardlessof the chemostatted part ΓY

γ ) in sets denoted by Cj , forj = 1, 2, . . . . Complexes of the closed network made solelyof chemostatted species in the open CRN are all regroupedin the same complex C0. This allows to decompose theinternal species stoichiometric matrix as

∇σxρ = Γσx

j ∂jρ . (34)

where {Γσxj ≡ Γσx

γ , for γ ∈ Cj} are the entries of thecomposition matrix corresponding to the internal species,and {∂jρ ≡

∑γ∈Cj ∂

γρ } are the entries of the incidence

matrix describing the network of regrouped complexes.This regrouping corresponds to the—equivalent—CRNonly made of internal species with effective rate constant{k±ρZσy∇

±ρσy } ruling each reaction.

Example 9. Let us consider the CRN (29) where thespecies Xa and Xc are chemostatted. The five complexesof the closed network, see example 7, are regrouped asC0 = {Xa,Xc}, C1 = {Xb,Xb + Xa} and C2 = {2 Xb}.In terms of these groups of complexes, the compositionmatrix and incidence matrix are

ΓX =(0 1 2

), ∂C =

−1 0 11 −1 00 1 −1

, (35)

which corresponds to the effective representation

C0

2Xb Xb

k−3 Z

c

k +1Za

k+3

k−2

k −1

k+2Za (36)

A steady-state distribution {Zσx} (11) is said to becomplex-balanced if the net current flowing in each groupof complexes Cj vanishes, i.e. if the currents {Jρ} satisfy

∂jρJρ ≡

∑γ∈Cj

∂γρ Jρ = 0 , ∀j . (37)

Complex-balance steady states are therefore a subclass ofsteady states (11a) which include equilibrium ones (17)as a special case

Γσxj ∂jρ Jρ︸︷︷︸

= 0 iff DB︸ ︷︷ ︸= 0 iff CB︸ ︷︷ ︸

= 0 for generic SS

. (38)

While for generic steady states only the internal speciesformation rates vanish, for complex-balanced ones thecomplex formation rates also vanish.

For a fixed kinetics ({k±ρ}) and chemostatting (Sy and{Zσy}), a CRN is complex-balanced if its dynamics ex-hibits a complex-balanced steady state (37) [35, 36]. Thecomplex-balanced distribution (37) depends on the unbro-ken components {Lλu}, which can be inferred from theinitial conditions, and is always globally stable [68]. Hence,complex-balanced networks always relax to a—complex-balanced—steady state. Detailed-balanced networks area subclass of complex-balanced networks.

Whether or not a CRN is complex-balanced depends onthe network topology (∇), the kinetics ({k±ρ}) and thechemostatting (Sy and {Zσy}). For any given networktopology and set of chemostatted species Sy, one canalways find a set of effective rate constants {k±ρZσy∇

±ρσy }

which makes that CRN complex-balanced [37]. However,for some CRNs, this set coincides with the one whichmakes the CRN detailed-balanced [69]. A characterizationof the set of effective rate constants which make a CRNcomplex balanced is reported in Refs. [37, 69].Deficiency-zero CRNs are a class of CRN which is

complex-balanced irrespective of the effective kinetics{k±ρZσy∇

±ρσy } [35–37]. The network deficiency is a topo-

logical property of the CRN which we briefly discuss inapp. D, see Refs. [22, 52, 53] for more details. Conse-quently, regardless of the way in which a deficiency-zeroCRN is driven in time, it will always remain complex-balanced. Throughout the paper, we will refer to theseCRNs as unconditionally complex-balanced, as in the sem-inal work [35].

Example 10. The open CRN (36) has a single steadystate Zb for any given set of rate constants and chemostat-ted concentrations Za and Zc [64], defined by (11a)

dtZb = J1 + J2 − 2J3

= k+1Za − k−1Zb + k+2ZaZb − k−2(Zb)2++ 2k−3Zc − 2k+3(Zb)2 = 0 .

Page 8: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

8

(39)

If the stronger condition (37) holds

J3 − J1 = 0 , (group C0) ,J1 − J2 = 0 , (group C1) ,J2 − J3 = 0 , (group C2) ,

(40)

which is equivalent to

k+1Za − k−1Zb = k+2ZaZb − k−2(Zb)2

= k+3(Zb)2 − k−3Zc , (41)

the steady state is complex-balanced. Yet, if the steady-state currents are all independently vanishing,

J1 = J2 = J3 = 0 , (42)

i.e., eq. (41) is equal to zero, then the steady state isdetailed-balanced.When, for simplicity, all rate constants are taken as

1, the complex-balanced set of quadratic equations (41)admits a positive solution Zb only if Za = 2− Zc (0 <Zc < 2) or Za =

√Zc. The former case corresponds to

a genuine complex-balanced state, Zb = 1 with currentsJ1 = J2 = J3 = 1− Zc, while the second to a detailed-balance state Zb =

√Zc with vanishing currents. When,

for example, Za = 1 and Zc = 4, neither of the twoprevious conditions holds: the nonequilibrium steady stateis Zb =

√3 with currents J1 = 1 −

√3, J2 = −3 +

√3,

and J3 = −1.

Example 11. Let us now consider the following openCRN [22]

Yak+1

−−−⇀↽−−−k−1

Xbk+2

−−−⇀↽−−−k−2

Xc + Xdk+3

−−−⇀↽−−−k−3

Ye (43)

where the species Ya and Ye are chemostatted. Out ofthe four complexes of the closed network, {Ya,Xb,Xc +Xd,Ye}, two are grouped into C0 = {Ya,Ye} and theother two remain C1 = {Xb} and C2 = {Xc + Xd}. Theeffective representation of this open CRN is

C0

Xc +Xd Xb

k−3 Z

e

k +1Z a

k+3

k−2

k −1

k+2 (44)

This network is deficiency-zero and hence unconditionallycomplex-balanced [22]. Therefore, given any set of rateconstants k±1, k±2, and k±3, and the chemostatted con-centrations Za and Ze, the steady state of this CRN iscomplex-balanced, i.e. the steady state always satisfies aset of condition like those in eq. (40). Indeed, contrary toexample 10, steady state currents {J1, J2, J3} differentfrom each other cannot exist since they would induce agrowth or decrease of some concentrations.

III. THERMODYNAMICS OFCHEMICAL NETWORKS

Using local equilibrium, we here build the connectionbetween the dynamics and the nonequilibrium thermo-dynamics for arbitrary CRNs. In the spirit of StochasticThermodynamics, we derive an energy and entropy bal-ance, and express the dissipation of the CRN as thedifference between the chemical work done by the reser-voirs on the CRN and its change in nonequilibrium freeenergy. We finally discuss the information-theoreticalcontent of the nonequilibrium free energy and its relationto the dynamical potentials used in Chemical ReactionNetwork Theory.

A. Local Equilibrium

Since we consider homogeneous reaction mixtures inideal dilute solutions, the assumption of local equilibrium[57, § 15.1] [70] means that the equilibration followingany reaction event is much faster than any reaction timescale. Thus, what is assumed is that the nonequilibriumnature of the thermodynamic description is solely dueto the reaction mechanisms. If all reactions could beinstantaneously shut down, the state of the whole CRNwould immediately become an equilibrated ideal mixtureof species. As a result, all the intensive thermodynamicvariables are well-defined and equal everywhere in thesystem. The temperature T is set by the solvent, whichacts as a thermal bath, while the pressure p is set by theenvironment the solution is exposed to. As a result, eachchemical species is characterized by a chemical potential[23, § 3.1]

µσ = µ◦σ +RT ln ZσZtot

, ∀σ ∈ S , (45)

where R denotes the gas constant and {µ◦σ ≡ µ◦σ(T )}are the standard-state chemical potentials, which dependon the temperature and on the nature of the solvent.The total concentration of the solution is denoted byZtot =

∑σ Z

σ +Z0, where Z0 is the concentration of thesolvent. We assume for simplicity that the solvent doesnot react with the solutes. In case it does, our resultsstill hold provided one treats the solvent as a nondrivenchemostatted species, as discussed in app. A. Since thesolvent is much more abundant than the solutes, the totalconcentration is almost equal to that of the solvent whichis a constant, Ztot ' Z0. Without loss of generality, theconstant term −RT lnZtot ' −RT lnZ0 in eq. (45) isabsorbed in the standard-state chemical potentials. Con-sequently, many equations appear with non-matchingdimensions. We also emphasize that standard state quan-tities, denoted with “◦”, are defined as those measuredin ideal conditions, at standard pressure (p◦ = 100 kPa)and molar concentration (Z◦σ = 1 mol/dm3), but not at astandard temperature [71, p. 61].

Page 9: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

9

Due to the assumption of local equilibrium and homo-geneous reaction mixture, the densities of all extensivethermodynamic quantities are well-defined and equal ev-erywhere in the system. With a slight abuse of notation,we use the same symbol and name for densities as for theircorresponding extensive quantity. E.g. S is the molarentropy divided by the volume of the solution, but wedenote it as entropy. We apply the same logic to rates ofchange. E.g. we call entropy production rate the molarentropy production density rate.

B. Affinities, Emergent Affinities andLocal Detailed Balance

The thermodynamic forces driving reactions are givenby differences of chemical potential (45)

∆rGρ ≡ ∇σρ µσ , (46)

also called Gibbs free energies of reaction [57, § 9.3][23, § 3.2]. Since these must all vanish at equilibrium,∇σρ µeq

σ = 0, ∀ρ, we have

∆rGρ = −RT ∇σρ ln ZσZeqσ. (47)

The local detailed-balance (18) allows us to express thesethermodynamic forces in terms of reaction affinities,

Aρ ≡ RT ln J+ρ

J−ρ= −∆rGρ (48)

which quantify the kinetic force acting along each reactionpathway [57, § 4.1.3].

The change of Gibbs free energy along emergent cycles,

Aε = −cρε ∆rGρ = −cρε ∇σyρ µσy , (49)

gives the external thermodynamic forces the network iscoupled to, as we shall see in eq. (61), and thus providea thermodynamic meaning to the cycle affinities (28).

Combining the detailed-balance property (18) and theequilibrium condition on the affinities Aeq

ρ = 0 (46), wecan relate the Gibbs free energies of reaction to the rateconstants

k+ρ

k−ρ= exp

{−

∆rG◦ρ

RT

}, (50)

where ∆rG◦ρ ≡ ∇σρ µ◦σ. This relation is the thermody-

namic counterpart of the local detailed balance (18). Itplays the same role as in Stochastic Thermodynamics,namely connecting the thermodynamic description to thestochastic dynamics. We emphasize that the local detailedbalance property as well as the local equilibrium assump-tion by no mean imply that the CRN operates close toequilibrium. Their importance is to assign well definedequilibrium potentials to the states of the CRN, whichare then connected by the nonequilibrium mechanisms,i.e. reactions.

C. Enthalpies and Entropies of Reaction

To identify the heat produced by the CRN, we needto distinguish the enthalpic change produced by eachreaction from the entropic one. We consider the decompo-sition of the standard state chemical potentials [23, § 3.2]:

µ◦σ = h◦σ − Ts◦σ . (51)

The standard enthalpies of formation {h◦σ} take into ac-count the enthalpic contributions carried by each species[23, § 3.2] [72, § 10.4.2]. Enthalpy changes caused byreactions give the enthalpies of reaction [23, § 3.2] [57,§ 2.4]

∆rHρ = ∇σρ h◦σ , (52)

which at constant pressure measure the heat of reaction.This is the content of the Hess’ Law (see e.g. [72, § 10.4.1]).The standard entropies of formation {s◦σ} take into ac-count the internal entropic contribution carried by eachspecies under standard-state conditions [23, § 3.2]. Using(51), the chemical potentials (45) can be rewritten as

µσ = h◦σ − T (s◦σ −R lnZσ)︸ ︷︷ ︸≡ sσ

. (53)

The entropies of formation {sσ ≡ s◦σ −R lnZσ} accountfor the entropic contribution of each species in the CRN[23, § 3.2]. Entropy changes along reactions are given by

∆rSρ = ∇σρ sσ , (54)

called entropies of reaction [23, § 3.2].

D. Entropy Balance

1. Entropy Production Rate

The entropy production rate is a non-negative measureof the break of detailed balance in each chemical reaction.Its typical form is given by [57, § 9.5] [8]

T Si ≡ RT (J+ρ − J−ρ) ln J+ρ

J−ρ≥ 0 , (55)

because

1. It is non-negative and vanishes only at equilibrium,i.e. when the detailed balance property (17) issatisfied;

2. It vanishes to first order around equilibrium, thusallowing for quasi-static reversible transformations.Indeed, defining

Zσ − Zσeq

Zσeq= εσ , |εσ| � 1 , ∀σ ∈ S , (56)

Page 10: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

10

we find that

Si = Eσσ′ εσ′ εσ + O

(ε3), (57)

where E ≡ {Eσσ′} is a positive semidefinite symmet-ric matrix.

Furthermore it can be rewritten in a thermodynamicallyappealing way using (48):

T Si = −Jρ ∆rGρ . (58)

It can be further expressed as the sum of two distinctcontributions [56]:

T Si = −µσxdtZσx︸ ︷︷ ︸≡ T Sx

−µσy (dtZσy − Iσy)︸ ︷︷ ︸≡ T Sy

. (59)

The first term is due to changes in the internal speciesand thus vanishes at steady state. The second term isdue to the chemostats. It takes into account both theexchange of chemostatted species and the time-dependentdriving of their concentration. If the system reaches anonequilibrium steady state, the external currents {Iσy}do not vanish and the entropy production reads

T ¯Si = Iσy µσy . (60)

This expression can be rewritten as a bilinear form ofemergent cycle affinities {Aε} (49) and currents alongemergent cycle {J ε ≡ cερJρ} [56]

T ¯Si = J εAε , (61)

which clearly emphasizes the crucial role of emergentcycles in steady state dissipation.

2. Entropy Flow Rate

The entropy flow rate measures the reversible entropychanges in the environment due to exchange processeswith the system [57]. Using the expressions for the en-thalpy of reaction (52) and entropy of formation (53), weexpress the entropy flow rate as

T Se ≡ Jρ ∆rHρ︸ ︷︷ ︸≡ Q

+Iσy Tsσy . (62)

The first contribution is the heat flow rate (positive if heatis absorbed by the system). When divided by temperature,it measures minus the entropy changes in the thermalbath. The second contribution accounts for minus theentropy change in the chemostats.

3. System Entropy

The entropy of the ideal dilute solution constitutingthe CRN is given by (see app. A)

S = Zσ sσ +RZS + S0 . (63)

The total concentration term

ZS ≡∑σ∈S

Zσ (64)

and the constant S0 together represent the entropic con-tribution of the solvent. S0 may also account for theentropy of chemical species not involved in the reactions.We also prove in app. B that the entropy (63) can beobtained as a large particle limit of the stochastic entropyof CRNs.S would be an equilibrium entropy if the reactions

could be all shut down. But in presence of reactions, itbecomes the nonequilibrium entropy of the CRN. Indeed,using eqs. (53), (58) and (62), we find that its changecan be expressed as

dtS = sσ dtZσ + Zσ dtsσ +R dtZS

= sσ dtZσ

= Jρ ∆rSρ + Iσy sσy

= Si + Se .

(65)

This relation is the nonequilibrium formulation of theSecond Law of Thermodynamics for CRNs. It demon-strates that the non-negative entropy production (55)measures the entropy changes in the system plus those inthe reservoirs (thermal and chemostats) [57].

E. Energy Balance

1. First Law of Thermodynamics

Since the CRN is kept at constant pressure p, its en-thalpy

H = Zσ h◦σ +H0 (66)

is equal to the CRN internal energy, up to a constant.Indeed, the enthalpyH is a density which, when written interms of the internal energy (density) U , reads H = U+p.

Using the rate equations (9) and (10), the enthalpy rateof change can be expressed as the sum of the heat flowrate, defined in eq. (62), and the enthalpy of formationexchange rate

dtH = h◦σ dtZσ = Q+ Iσy h◦σy. (67)

Equivalently, it can be rewritten in terms of the entropyflow rate (62) as [57, § 4.1.2]

dtH = T Se + Iσy µσy . (68)

Page 11: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

11

The last term on the r.h.s. of eq. (68) is the free energyexchanged with the chemostats. It represents the chemicalwork rate performed by the chemostats on the CRN [21,23]

Wc ≡ Iσy µσy . (69)

Either eq. (67) or (68) may be considered as the nonequi-librium formulation of the First Law of Thermodynamicsfor CRNs. The former has the advantage to solely focuson energy exchanges. The latter contains entropic contri-butions but is appealing because it involves the chemicalwork (69).

2. Nonequilibrium Gibbs Free Energy

We are now in the position to introduce the thermody-namic potential regulating CRNs. The Gibbs free energyof ideal dilute solutions reads

G ≡ H − TS = Zσ µσ −RT ZS +G0 . (70)

As for entropy, the total concentration term −RT ZS andthe constant G0 represent the contribution of the solvent(see app. A). Furthermore, in presence of reactions, Gbecomes the nonequilibrium Gibbs free energy of CRNs.We will now show that the nonequilibrium Gibbs free

energy of a closed CRN is always greater or equal thanits corresponding equilibrium form. A generic nonequi-librium concentration distribution {Zσ} is characterizedby the set of components {Lλ = `λσ Z

σ}. Let {Zσeq} bethe corresponding equilibrium distribution defined by thedetailed-balance property (18) and characterized by thesame set of components {Lλ} (a formal expression forthe equilibrium distribution will be given in eq. (88)). Atequilibrium, the Gibbs free energy (70) reads

Geq = Zσeq µeqσ −RT ZSeq +G0 . (71)

As discussed in § III B, the equilibrium chemical potentialsmust satisfy ∇σρ µeq

σ = 0. We deduce that µeqσ must be a

linear combination of the closed system conservation laws(12)

µeqσ = fλ `

λσ , (72)

where {fλ} are real coefficients. Thus, we can write theequilibrium Gibbs free energy as

Geq = fλ Lλ −RT ZSeq +G0 . (73)

In this form, the first term of the Gibbs free energyappears as a bilinear form of components {Lλ} and con-jugated generalized forces {fλ} [23, § 3.3], which canbe thought of as chemical potentials of the components.From eq. (72) and the properties of components (13), theequality Zσeq µ

eqσ = Zσ µeq

σ follows. Hence, using the defini-tion of chemical potential (45), the nonequilibrium Gibbs

free energy G of the generic distribution {Zσ} definedabove is related to Geq (73) by

G = Geq +RTL({Zσ}|{Zσeq}

), (74)

where we introduced the relative entropy for non-normalized concentration distributions, also called ShearLyapunov Function, or pseudo-Helmholtz function [35, 73,74]

L ({Zσ}|{Z ′σ}) ≡ Zσ ln ZσZ ′σ−(ZS − Z ′S

)≥ 0 . (75)

This quantity is a natural generalization of the relativeentropy, or Kullback–Leibler divergence, used to com-pare two normalized probability distributions [75]. Forsimplicity, we still refer to it as relative entropy. It quan-tifies the distance between two distributions: it is alwayspositive and vanishes only if the two distributions areidentical, {Zσ} = {Z ′σ}. Hence, eq. (74) proves that thenonequilibrium Gibbs free energy of a closed CRN is al-ways greater or equal than its corresponding equilibriumform, G ≥ Geq.

We now proceed to show that the nonequilibrium Gibbsfree energy is minimized by the dynamics in closed CRNs,viz. G—or equivalently L

({Zσ}|{Zσeq}

)[59, 76]—acts as

a Lyapunov function in closed CRNs. Indeed, the timederivative of G (70) always reads

dtG = µσ dtZσ + Zσ dtµσ +R dtZS

= µσ dtZσ .(76)

When using the rate equation for closed CRNs (7) wefind that dtG = −Jρ∇σρ µσ. Using eq. (74) together witheqs. (46) and (58), we get

dtG = RT dtL({Zσ}|{Zσeq}

)= −T Si ≤ 0 , (77)

which proves the aforementioned result.

3. Chemical Work

In arbitrary CRNs, the rate of change of nonequilibriumGibbs free energy (76) can be related to the entropyproduction rate (59) using the rate equations of openCRN (9) and (10) and the chemical work rate (69),

T Si = Wc − dtG ≥ 0 . (78)

This important results shows that the positivity of theentropy production sets an intrinsic limit on the chemicalwork that the chemostats must perform on the CRN tochange its concentration distribution. The equality signis achieved for quasi-static transformations (Si ' 0).If we now integrate eq. (78) along a transformation

generated by an arbitrary time dependent protocol π(t),which drives the CRN from an initial concentration dis-tribution {Zσi } to a final one {Zσf }, we find

T∆iS = Wc −∆G ≥ 0 , (79)

Page 12: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

12

where ∆G = Gf −Gi is the difference of nonequilibriumGibbs free energies between the final and the initial state.Let us also consider the equilibrium state {Zσeqi} (resp.{Zσeqf}) obtained from {Zσi } (resp. {Zσf }) if one closesthe network (i.e. interrupt the chemostatting procedure)and let it relax to equilibrium, as illustrated in fig. 4. TheGibbs free energy difference between these two equilibriumdistributions, ∆Geq = Geqf −Geqi, is related to ∆G viathe difference of relative entropies, eq. (74),

∆G = ∆Geq +RT ∆L , (80)

where

∆L ≡ L({Zσf }|{Zσeqf}

)− L

({Zσi }|{Zσeqi}

). (81)

Thus, the chemical work (79) can be rewritten as

Wc −∆Geq = RT∆L+ T∆iS , (82)

which is a key result of our paper. ∆Geq representsthe reversible work needed to reversibly transform theCRN from {Zσeqi} to {Zσeqf}. Implementing such a re-versible transformation may be difficult to achieve inpractice. However, it allows us to interpret the differenceW irr

c ≡Wc −∆Geq in eq. (82) as the chemical work dissi-pated during the nonequilibrium transformation, i.e. theirreversible chemical work. The positivity of the entropyproduction implies that

W irrc ≥ RT ∆L . (83)

This relation sets limits on the irreversible chemical workinvolved in arbitrary far-from-equilibrium transformations.For transformations connecting two equilibrium distribu-tions, we get the expected inequality W irr

c ≥ 0. Moreinterestingly, eq. (83) tells us how much chemical workthe chemostat need to provide to create a nonequilibriumdistribution from an equilibrium one. It also tells ushow much chemical work can be extracted from a CRNrelaxing to equilibrium.

The conceptual analogue of (82) in Stochastic Thermo-dynamics (where probability distributions replace non-normalized concentration distributions) is called thenonequilibrium Landauer’s principle [42, 43] (see also[77–79]). It has been shown to play a crucial role to ana-lyze the thermodynamic cost of information processing(e.g. for Maxwell demons, feedback control or proofread-ing). The inequality (83) is therefore a nonequilibriumLandauer’s principle for CRN.

IV. THERMODYNAMICS OFCOMPLEX-BALANCED NETWORKS

In this section, we focus on unconditionally complex-balanced networks. We shall see that the thermodynam-ics of these networks bears remarkable similarities withStochastic Thermodynamics.

relaxation*

relaxation*

equilibrium driving

non-e

quilib

riumdri

ving{Zσ

i }

{Zσf }

{Zσeqi

} {Zσeqf

}

FIG. 4: Pictorial representation of the transformation be-tween two nonequilibrium concentration distributions. Thenonequilibrium transformation (blue line) is comparedwith the equilibrium one (green line). The equilibriumtransformation depends on the equilibrium states corre-sponding to the initial and final concentration distribu-tions. In § III E 3, for an arbitrary CRN, these equilibriumstates are obtained by first closing the network and thenletting it relax to equilibrium. Instead, in sec. V, for a de-tailed balance CRN, the equilibrium states are obtained bysimply stopping the time-dependent driving and letting thesystem spontaneously relax to equilibrium.

Let us first observe that whenever a CRN displays awell-defined steady-state distribution {Zσx}, the entropyproduction rate (55) can be formally decomposed as thesum of an adiabatic and a nonadiabatic contribution

T Si = JρRT ln J+ρ

J−ρ︸ ︷︷ ︸≡ T Sa

−dtZσx RT ln Zσx

Zσx︸ ︷︷ ︸≡ T Sna

, (84)

in analogy to what was done in Stochastic Thermody-namics [46–50]. As discussed in § II E, unconditionallycomplex-balanced networks have a unique steady-statedistribution

{Zσx ≡ Zσx(π(t))

}, eq. (37), for any value of

the chemostatted concentrations {Zσy ≡ Zσy(π(t))} andof the fixed unbroken components {Lλu}. The decomposi-tion (84) is thus well-defined at any time, for any protocolπ(t). As a central result, we prove in app. C that theadiabatic and nonadiabatic contribution are non-negativefor unconditionally complex-balanced networks as well asfor complex-balanced networks without time-dependentdriving.The adiabatic entropy production rate encodes the

dissipation of the steady state{Zσx

}. It can be rewritten

in terms of the steady state Gibbs free energy of reaction{∆rGρ} (48) as

T Sa = −Jρ ∆rGρ ≥ 0 . (85)

This inequality highlights the fact that the transientdynamics—generating the currents {Jρ}—is constrained

Page 13: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

13

by the thermodynamics of the complex balanced steadystate, i.e. by {∆rGρ}.

The nonadiabatic entropy production rate characterizesthe dissipation of the transient dynamics. It can bedecomposed as

T Sna = −RT dtL({Zσx}|{Zσx}

)+

+ RT dtZSx − Zσxdtµσx︸ ︷︷ ︸≡ T Sd

≥ 0 , (86)

where ZSx =∑σx∈Sx

Zσx (see Refs. [46, 48] for the anal-ogous decomposition in the stochastic context). The firstterm is proportional to the time derivative of the relativeentropy (75) between the nonequilibrium concentrationdistribution at time t and the corresponding complex-balanced steady-state distribution. Hence, it describesthe dissipation of the relaxation towards the steady state.The second term, T Sd, is related to the time-dependentdriving performed via the chemostatted species and thusdenoted driving entropy production rate [46]. It vanishesin nondriven networks where we obtain

Sna = −R dtL({Zσx}|{Zσx}

)≥ 0 . (87)

This result shows the role of the relative entropyL({Zσx}|{Zσx}

)as a Lyapunov function in nondriven

complex-balanced networks with mass action kinetics. Itwas known in the mathematical literature [35, 80], butwe provide a clear thermodynamic interpretation to thisresult by demonstrating that it derives from the nonadia-batic entropy production rate.We mention that an alternative derivation of the

adiabatic–nonadiabatic decomposition for nondrivencomplex-balanced networks with mass action kinetics wasfound in Ref. [51], while we were finalizing our paper.

V. THERMODYNAMICS OF OPENDETAILED-BALANCED NETWORKS

We finish our study by considering detailed-balancednetworks. We discuss the equilibrium distribution, intro-duce a new class of nonequilibrium potentials and derivea new work inequality.Let us also emphasize that open detailed-balanced

CRNs are a special class of open complex-balanced CRNsfor which the adiabatic entropy production rate vanishes(since the steady state is detailed balanced) and thus thenonadiabatic entropy production characterizes the entiredissipation.

A. Equilibrium Distribution

As discussed in § IID, for given kinetics {k±ρ},chemostatting {Zσy} and unbroken components {Lλu},

detailed-balanced networks always relax to a unique equi-librium distribution. Since the equilibrium chemical po-tentials can be expressed as a linear combination of con-servation laws, eq. (72), we can express the equilibriumdistribution as

Zeqσ = exp

{−µ◦σ − fλ `λσRT

}, (88)

inverting the expression for the chemical potentials (45).Since the independent set of unbroken conservation laws,{`λu}, are such that `λuσy

= 0, ∀λu, σy, see § II C, we havethat

µeqσy

= fλb `λbσy, ∀σy ∈ Sy . (89)

We thus conclude that the |λb| broken generalized forces{fλb} only depend on the chemostatted concentrations{Zσy}. Instead, the remaining |λu| unbroken generalizedforces fλu can be determined by inverting the nonlinearset of equations Lλu = `λu

σxZσx

eq . They therefore dependon both {Zσy} and {Lλu}.

One can easily recover the local detailed-balanced prop-erty (50) and (18) using eq. (88).

B. Open nondriven networks

As a consequence of the break of conservation laws,the nonequilibrium Gibbs free energy G (70) is no longerminimized at equilibrium in open detailed-balanced net-works. In analogy to equilibrium thermodynamics [23],the proper thermodynamic potential is obtained from Gby subtracting the energetic contribution of the brokenconservation laws. This transformed nonequilibrium Gibbsfree energy reads

G ≡ G− fλb Lλb

= Zσ(µσ − fλb `

λbσ

)−RT ZS +G0 .

(90)

We proceed to show that G is minimized by the dy-namics in nondriven open detailed-balanced networks.Let {Zσx} be a generic concentration distribution in adetailed-balanced network characterized by {Lλu} and{Zσy}, and let {Zσx

eq } be its corresponding equilibrium.Using the relation between equilibrium chemical poten-tials and conservation laws (72), the transformed Gibbsfree energy (90) at equilibrium reads

Geq = fλu Lλu −RT ZSeq +G0 . (91)

Yet, combinig eq. (72) and the properties of unbroken com-ponents, one can readily show that Zσeq

(µeqσ − fλb `

λbσ

)=

Zσ(µeqσ − fλb `

λbσ

). The relation between the nonequi-

librium G and the corresponding equilibrium value thusfollows

G = Geq +RTL({Zσx}|{Zσx

eq })

(92)

Page 14: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

14

(we show in app. A the derivation of the latter in presenceof reacting solvent). The non-negativity of the relativeentropy for concentration distributions L

({Zσx}|{Zσx

eq })

ensures that the nonequilibrium transformed Gibbs freeenergy is always greater or equal to its equilibrium value,G ≥ Geq. Since entropy production and nonadiabaticentropy production coincide, using eqs. (87) and (92), weobtain

dtG = RT dtL({Zσx}|{Zσx

eq })

= −T Si ≤ 0 , (93)

which demonstrates the role of G as a Lyapunov func-tion. The relative entropy L ({Zσx}|{Z ′σx}) was knownto be a Lyapunov function for detailed-balanced networks[76, 81], but we provided its clear connection to the trans-formed nonequilibrium Gibbs free energy. To summarize,instead of minimizing the nonequilibrium Gibbs free en-ergy G (70) as in closed CRNs, the dynamics minimizesthe transformed nonequilibrium Gibbs free energy G inopen nondriven detailed-balanced CRNs.

C. Open driven networks

We now consider unconditionally detailed-balancedCRNs. As discussed in § IID, they are characterizedby a unique equilibrium distribution

{Zσx

eq ≡ Zσxeq (π(t))

},

defined by eq. (18), for any value of the chemostattedconcentrations {Zσy = Zσy(π(t))}.We start by showing that the external fluxes {Iσy}

can be expressed as influx rate of moieties. Since theCRN is open and unconditionally detailed balanced, eachchemostatted species broke a conservation law (no emer-gent cycle is created, § IID). Therefore, the matrix whoseentries are {`λb

σy} in eq. (89) is square and also nonsingular

[82]. We can thus invert eq. (89) to get

fλb = µeqσy

ˆσyλb, (94)

where {ˆσyλb} denote the entries of the inverse matrix of

that with entries {`λbσy}. Hence, using the definition of

broken component, {Lλb ≡ `λbσ Zσ}, we obtain that

fλb Lλb = µeq

σyˆσyλb`λbσ Zσ︸ ︷︷ ︸

≡Mσy

. (95)

From the rate equations for the chemostatted concentra-tions (10), we find that

dtMσy = Iσy , ∀σy ∈ Sy . (96)

We can thus interpret Mσy as the concentration of amoiety which is exchanged with the environment onlythrough the chemostatted species Xσy . Eq. (95) showsthat the energetic contribution of the broken componentscan be expressed as the Gibbs free energy carried by thesespecific moieties.

Example 12. A simple implementation of this scenariois the thermodynamic description of CRNs at constantpH [23, Ch. 4] where the chemostatted species becomesthe ion H+ and MH+ is the total amount of H+ ionsin the system. The transformed Gibbs potential thusbecome G′ = G−µH+M

H+ and the transformed chemicalpotentials can be written in our formalism as µ′σx

=µσx−µH+

ˆH+b `bσ, where `bσ is the conservation law broken

by chemostatting H+.

Example 13. For the CRN in fig. 2, whose conserva-tion laws given in example 4, the concentrations of theexchanged moieties are

M1 = Za + 12Z

b

M2 = Zd + Ze .(97)

For the specific implementation of that CRN, fig. 3, thefirst term (resp. second term) is the total number of moi-ety 2H (resp. C) in the system, which can be exchangedwith the environment only via the chemostatted speciesH2O (resp. CO).

We now turn to the new work relation. From thegeneral work relation (78), using (90) and (95), we find

T Si = Wd − dtG ≥ 0 , (98)

where the driving work due to the time-dependent drivingof the chemostatted species is obtained using the chemicalwork rate (69) together with eqs. (95) and (96)

Wd ≡ Wc − dt(fλb L

λb)

= µeqσy

dtMσy − dt(µeqσyMσy

)= −dtµeq

σyMσy .

(99)

Equivalently, the driving work rate (99) can be defined asthe rate of change of the transformed Gibbs free energy(90) due to the time dependent driving only, i.e.

Wd ≡∂G∂t≡ dtµeq

σy

∂G∂µeq

σy

. (100)

To relate this alternative definition to eq. (99), all {Zσy}must be expressed in terms of {µeq

σy} using the definition

of chemical potential (45).The driving work rate Wd vanishes in nondriven CRNs,

where (98) reduces to (93). After demonstrating thatthe entropy production rate is always proportional thedifference between the chemical work rate and the changeof nonequilibrium Gibbs free energy in eq. (79), we showedthat, for unconditionally detailed-balanced CRNs, it isalso proportional the difference between the driving workrate and the change in transformed nonequilibrium Gibbsfree energy, eq. (98).We end by formulating a nonequilibrium Landauer’s

principle for the driving work instead the chemical workdone in § III E 3. We consider a time-dependent trans-formation driving the unconditionally detailed-balanced

Page 15: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

15

CRN from {Zσi } to {Zσf }. The distribution {Zσeqi} (resp.{Zσeqf}) denotes the equilibrium distribution obtainedfrom {Zσi } (resp. {Zσf }) by stopping the time-dependentdriving and letting the system relax towards the equilib-rium, fig. 4. Note that this reference equilibrium state isdifferent from the one obtained by closing the network in§ III E 3. Integrating (98) over time and using (92), weget

Wd −∆Geq = RT∆L+ T∆iS , (101)

where

∆L ≡ L({Zσx

f }|{Zσxeq f}

)−L({Zσx

i }|{Zσxeq i}

). (102)

∆Geq represents the reversible driving work, and the irre-versible driving work satisfies the inequality

W irrd ≡Wd −∆Geq ≥ RT∆L . (103)

This central relation sets limits on the irreversible workspent to manipulate nonequilibrium distributions. It is anonequilibrium Landauer’s principle for the driving workby the same reasons why inequality (83) is a nonequilib-rium Landauer’s principle for the chemical work. Thekey difference is that the choice of the reference equi-librium state is different in the two cases. The abovediscussed inequality (103) only holds for unconditionallydetailed-balanced CRNs while eq. (83) is valid for anyCRNs.

VI. CONCLUSIONS AND PERSPECTIVES

Following a strategy reminiscent of Stochastic Ther-modynamics, we systematically build a nonequilibriumthermodynamic description for open driven CRNs madeof elementary reactions in homogeneous ideal dilute so-lutions. The dynamics is described by deterministic rateequations whose kinetics satisfies mass action law. Ourframework in not restricted to steady states and allowsto treat transients as well as time-dependent drivingsperformed by externally controlled chemostatted concen-trations. Our theory embeds the nonequilibrium thermo-dynamic framework of irreversible processes establishedby the Brussels School of Thermodynamics.We now summarize our results. Starting from the ex-

pression for the entropy production rate, we establisheda nonequilibrium formulation of the first and second lawof thermodynamics for CRNs. The resulting expressionfor the system entropy is that of an ideal dilute solution.The clear separation between chemostatted and internalspecies allowed us to identify the chemical work doneby the chemostats on the CRN and to relate it to thenonequilibrium Gibbs potential. We were also able toexpress the minimal chemical work necessary to changethe nonequilibrium distribution of species in the CRNas a difference of relative entropies for non-normalizeddistributions. These latter measure the distance of the

initial and final concentration distributions from theircorresponding equilibrium ones, obtained by closing thenetwork. This result is reminiscent of the nonequilibriumLandauer’s principle derived in Stochastic Thermodynam-ics [43] and which proved very useful to study the energeticcost of information processing [45]. We also highlightedthe deep relationship between the topology of CRNs, theirdynamics and their thermodynamics. Closed CRNs (resp.nondriven open detailed-balanced networks) always relaxto a unique equilibrium by minimizing their nonequilib-rium Gibbs free energy (resp. transformed nonequilibriumGibbs free energy). This latter is given, up to a con-stant, by the relative entropy between the nonequilibriumand equilibrium concentration distribution. Non-drivencomplex-balanced networks relax to complex-balancednonequilibrium steady states by minimizing the relativeentropy between the nonequilibrium and steady stateconcentration distribution. In all these cases, even inpresence of driving, we showed how the rate of changeof the relative entropy relates to the CRN dissipation.For complex-balanced networks, we also demonstratedthat the entropy production rate can be decomposed,as in Stochastic Thermodynamics, in its adiabatic andnonadiabatic contributions quantifying respectively thedissipation of the steady state and of the transient dy-namics.Our framework could be used to shed new light on a

broad range of problems. We mention only a few.Stochastic thermodynamics has been successfully used

to study the thermodynamics cost of information process-ing in various synthetic and biological systems [44, 83–87].However, most of these are modeled as few state sys-tems or linear networks [8, 9]—e.g. quantum dots [88],molecular motors [89, 90] and single enzyme mechanisms[91, 92]—while biochemical networks involve more com-plex descriptions. The present work overcomes this limi-tation. It could be used to study biological information-handling processes such as kinetic proofreading [93–99]or enzyme-assisted copolymerization [92, 100–105] whichhave currently only been studied as single enzyme mecha-nisms.

Our theory could also be used to study metabolic net-works. However, these require some care, since complexenzymatic reaction mechanisms are involved [106]. Nev-ertheless our framework provides a basis to build effec-tive coarse-graining procedures for enzymatic reactions[107]. For instance, proofreading mechanisms operatingin metabolic processes could be considered [108]. Weforesee an increasing use of thermodynamics to improvethe modeling of metabolic networks, as recently shown inRefs. [30, 32, 33].

Since our framework accounts for time-dependent driv-ings and transient dynamics, it could be used to repre-sent the transmission of signals through CRNs or theirresponse to external modulations in the environment.These features become crucial when considering prob-lems such as signal transduction and biochemical switches[24, 109, 110], biochemical oscillations [28, 111], growth

Page 16: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

16

and self-organization in evolving bio-systems [112, 113],or sensory mechanisms [85, 87, 114–117]. Also, since tran-sient signals in CRNs can be used for computation [118]and have been shown to display Turing-universality [119–122], one should be able to study the thermodynamic costof chemical computing [123].Finally, one could use our framework to study any

process that can be described as nucleation or reversiblepolymerization [124–129] (see also Ref. [130, ch. 5–6])since these processes can be described as CRNs [63].As closing words, we believe that our results consti-

tute an important contribution to the theoretical studyof CRNs. It does for nonlinear chemical kinetics whatStochastic Thermodynamics has done for stochastic dy-namics, namely build a systematic nonequilibrium ther-modynamics on top of the dynamics. It also opens manynew perspectives and builds bridges between approachesfrom different communities studying CRNs: mathemati-cians who study CRNs as dynamical systems, physicistswho study them as nonequilibrium complex systems, andbiochemists as well as bioengineers who aim for accuratemodels of metabolic networks.

ACKNOWLEDGMENTS

The present project was supported by the NationalResearch Fund, Luxembourg (project FNR/A11/02 andAFR PhD Grant 2014-2, No. 9114110) as well as by theEuropean Research Council (project 681456).

Appendix A: Thermodynamics of Ideal DiluteSolutions

We show that the nonequilibrium Gibbs free energy(70) is the Gibbs free energy of an ideal dilute solution[131, Ch. 7] (see also [51]). We also show that in opendetailed-balanced networks in which the solvent reactswith the solutes, the expression of the transformed Gibbsfree energy (92) is recovered by treating the solvent as aspecial chemostatted species.

The Gibbs free energy (density) of an ideal dilute mix-ture of chemical compounds kept at constant temperatureand pressure reads

G = Zσµσ + Z0µ0 , (A1)

where the labels σ ∈ S refer to the solutes and 0 to thesolvent. The chemical potentials of each species (45) read

µσ = µ◦σ +RT ln ZσZtot

, ∀σ ∈ S

µ0 = µ◦0 +RT ln Z0

Ztot.

(A2)

Since the solution is dilute, Ztot =∑σ∈S Z

σ + Z0 ' Z0and the standard state chemical potentials {µ◦σ} depend

on the nature of the solvent. Hence, the chemical poten-tials of the solutes read

µσ ' µ◦σ +RT ln ZσZ0

, ∀σ ∈ S , (A3)

while that of the solvent

µ0 ' µ◦0 −RTZS

Z0, (A4)

where ZS ≡∑σ∈S Z

σ. Therefore, the Gibbs free energy(A1) reads

G ' Zσ µσ + Z0 µ◦0 −RT ZS , (A5)

which is eq. (70) in the main text, where G0 is equal toZ0 µ◦0 plus possibly the Gibbs free energy of solutes whichdo not react.

We now consider the case where the solvent reacts withthe solutes. We assume that both the solutes and thesolvent react according to the stoichiometric matrix

∇ =

∇0

∇X

∇Y

, (A6)

where the first row refers to the solvent, the secondblock of rows to the internal species and the last oneto the chemostatted species. The solvent is treated as achemostatted species such that dtZ0 = 0.In order to recover the expression for the transformed

Gibbs free energy (92) in unconditionally detailed bal-anced networks, we observe that, at equilibrium

∇σρ µeqσ +∇0

ρ µeq0 = 0 . (A7)

Therefore, the equilibrium chemical potentials are a linearcombination of the conservation laws of ∇ (A6)

µeqσ = fλ `

λσ

µeq0 = fλ `

λ0 .

(A8)

As mentioned in the main text, § IIC, the chemostat-ting procedure breaks some conservation laws, which arelabeled by λb. The unbroken ones are labeled by λu.The transformed Gibbs free energy is defined as in

eq. (90), reported here for convenience

G ≡ G− fλb Lλb , (A9)

where G reads as in eq. (A1), {Lλb} are the broken com-ponents and {fλb} are here interpreted as the conjugatedgeneralized forces. Adding and subtracting the termZσµeq

σ + Z0µeq0 from the last equation and using eq. (A8)

we obtain

G = Geq + Zσ (µσ − µeqσ ) + Z0 (µ0 − µeq

0 ) , (A10)

where

Geq = fλu Lλu . (A11)

Page 17: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

17

From eqs. (A3) and (A4) and the fact that Zσy = Zσyeq

and Z0 = Zeq0 we obtain

G ' Geq + Zσx(µσx − µeq

σx

)−RT

(ZSx − ZSx

eq)

= Geq + Zσx RT ln Zσx

Zσxeq−RT

(ZSx − ZSx

eq)

≡ Geq +RT L({Zσx}|{Zσxeq })

(A12)

in agreement with the expression derived in the main text,eq. (92).

Appendix B: Entropy of CRNs

We show how the nonequilibrium entropy (63) canbe obtained as a large particle limit of the stochasticentropy. We point out that while finalizing the papersimilar derivations for other thermodynamic quantitieshave obtained in Refs. [51, 132].In the stochastic description of CRNs, the state is

characterized by the population vector n = {nσ}. Theprobability to find the network is in state n at time t isdenoted pt(n). The stochastic entropy of that state reads[21, 107], up to constants,

S(n) = −kB ln pt(n) + s(n) . (B1)

The first term is a Shannon-like contribution while thesecond term is the configurational entropy

s(n) ≡ nσ s◦σ − kB∑σ

ln nσ!nn

σ

0. (B2)

s◦σ is the standard entropy of one single Xσ molecule, andn0 is the very large number of solvent molecules.We now assume that the probability becomes very

narrow in the large particle limit nσ � 1 and behaves asa discrete delta function pt(n) ' δ(n− n(t)). The vectorn(t) ≡ {nσ} denotes the most probable and macroscopicamount of chemical species, such that Zσ = nσ/(V NA).Hence, the average entropy becomes

〈S〉 =∑

npt(n)S(n) ' s(n) . (B3)

When using the Stirling approximation (lnm! ' m lnm−m for m� 1), we obtain

s(n) ' nσ s◦σ − nσ kB ln nσ

n0+ kB

∑σ

= nσ(s◦σ + kB ln n0

V NA

)+

− nσ kB ln nσ

V NA+ kB

∑σ

≡ nσ (s◦σ + kB lnZ0) +

− nσ kB lnZσ + kB∑σ

nσ .

(B4)

Dividing by V and using the relation R = NAkB we finallyget the macroscopic entropy density (63)

〈S〉/V ' Zσ s◦σ − Zσ R lnZσ +RZS , (B5)

where the (molar) standard entropies of formation s◦σreads

s◦σ = NA (s◦σ + kB lnZ0) . (B6)

Mindful of the information-theoretical interpretationof the entropy [133], we note that the uncertainty due tothe stochasticity of the state disappears (the first termon the r.h.s. of eq. (B1)). However, the uncertaintydue to the indistinguishability of the molecules of thesame species—quantified by the configurational entropy(B2)—remains and contributes to the whole deterministicentropy function (63).

Appendix C: Adiabatic–NonadiabaticDecomposition

We prove the positivity of the adiabatic and nonadia-batic entropy production rates (84) using the theory ofcomplex-balanced networks, see § II E.We first rewrite the mass action kinetics currents (5)

as [53, 81]: Jρ = Kργ′ ψ

γ′ , where ψγ ≡ ZσΓγ′σ and K =

{Kργ ≡ K+ρ

γ −K−ργ } is the rate constants matrix whoseentries are defined by

Kργ =

k+ρ if γ is the reactant complex of +ρ ,−k−ρ if γ is the product complex of +ρ ,0 otherwise .

(C1)

Hence, the definition of complex-balanced network (37)reads ∑

γ∈Cj

Wγγ′ ψ

γ′ = 0 , ∀j . (C2)

where W ≡ ∂ K = {∂γρ Kργ′} ≡ {W

γγ′} is the so-called

kinetic matrix [35], and ψγ ≡ ZσΓγσ .

The kinetic matrix W is a Laplacian matrix [76, 81]:any off-diagonal term is equal to the rate constant of thereaction having γ′ as a reactant and γ as product if thereaction exists, and it is zero otherwise. Also, it satisfies∑

γ∈CWγγ′ = 0 , (C3)

which is a consequence of the fact that the diagonal termsare equal to minus the sum of the off-diagonal terms alongthe columns. The detailed balanced property (18) impliesthat

Wγγ′ ψ

eqγ′ =Wγ′

γ ψeqγ , ∀ γ, γ′ , (C4)

Page 18: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

18

where ψeqγ ≡ Zeq

σΓσγ .

In order to prove the non-negativity of the adiabaticterm (84), we rewrite it as

Sa ≡ Jρ ln J+ρ

J−ρ= Kρ

γ′ ψγ′ ln

(ZσZeqσ

)−∇σρ= −Wγ

γ′ ψγ′ ln ψγ

ψeqγ.

(C5)

The detailed balance property is used in the first equality,and the decomposition of the stoichiometric matrix (30)in the second one. Also, the constant RT is taken equalto one. Using (C3), eq. (C5) can be rewritten as

Sa = −Wγγ′ ψ

γ′ lnψγψ

γ′

eq

ψeqγ ψγ

′ . (C6)

From the log inequality − ln x ≥ 1− x and the detailedbalance property (C4), we obtain

Sa ≥ Wγγ′ ψ

γ′

(1−

ψγψγ′

eq

ψeqγ ψγ

)

= −Wγγ′ ψ

γ′

eqψγψ

γ′

ψeqγ ψγ

′ = −Wγ′

γ ψγψγ′

ψγ′= 0 .

(C7)

The last equality follows from the assumption of complex-balanced steady state (C2), the properties of the groupsof complexes {Cj} (§ II E), and the fact that {Zσy = Zσy}.Indeed

Wγ′

γ ψγψγ′

ψγ′=∑j

∑γ′∈Cj

Wγ′

γ ψγ

(Zσx

Zσx

)Γσxγ′

=∑j

(Zσx

Zσx

)Γσxj ∑

γ′∈Cj

Wγ′

γ ψγ = 0 .

(C8)

Concerning the nonadiabatic term (84), using the rateequations (9) and the fact that {Zσy = Zσy}, we canrewrite it as

Sna ≡ −dtZσ ln ZσZσ

= −Wγγ′ ψ

γ′ ln ψγψγ

. (C9)

Because of (C3), we further get that

Sna = −Wγγ′ ψ

γ′ ln ψγψγ′

ψγψγ′ . (C10)

From the log inequality − ln x ≥ 1− x and from (C4)

Sna ≥ Wγγ′ ψ

γ′

(1− ψγψ

γ′

ψγψγ′

)= −Wγ

γ′ ψγ′ ψγ

ψγ= 0 .

(C11)

The last equality again follows from the assumption ofcomplex-balance steady state (C2) as in eq. (C8).

Appendix D: Deficiency of CRNs

The deficiency of an open CRN is defined as [22]

δ = dim ker∇X − dim ker ∂C ≥ 0 , (D1)

where ∂C = {∂jρ ≡∑γ∈Cj ∂

γρ }. Other equivalent defini-

tions can be found in [52] and [53]. The kernel of ∇X

identifies the set of cycles, eqs. (19) and (20), while thekernel of the incidence matrix ∂ identifies the set of cyclesof the reaction graph. Hence, the deficiency measures thedifference between the number of cyclic transformationson chemical species and how many of them can be repre-sented as cycles on the reaction graph. Deficiency-zeronetworks are defined by δ = 0, i.e. they exhibit a one toone correspondence between the two. This topologicalproperty has many dynamical consequences, the mostimportant of which is that deficiency-zero networks areunconditionally complex-balanced [36, 37]. As shown inRef. [22], deficiency has also implications on the stochasticthermodynamic description of networks: the stochasticentropy production of a deficiency-zero network convergesto the deterministic entropy production in the long-timelimit. Linear networks are the simplest class of deficiency-zero networks. Since only one internal species appears ineach complex with a stoichiometric coefficient equal toone, ∇X ≡ ∂C and thus δ = 0.

[1] J.W. Gibbs, The Scientific Papers of J. Willard Gibbs,Vol.1: Thermodynamics (Dover, 1961).

[2] T. de Donder, L’affinité, Mémoires de la Classe dessciences No. vol. 1 (Gauthier-Villars, 1927).

[3] I. Prigogine, Etude Thermodynamique des PhénomènesIrréversibles (Desoer, Liège, 1947).

[4] I. Prigogine, Introduction to thermodynamics of irre-versible processes (John Wiley & Sons, 1967).

[5] I. Prigogine and R. Defay, Chemical Thermodynamics(Longmans, Green & Co., 1954).

[6] George F. Oster, Alan S. Perelson, and Aharon Katchal-sky, “Network thermodynamics: dynamic modelling ofbiophysical systems,” Q. Rev. Biophys. 6, 1–134 (1973).

[7] M. Malek-Mansour and G. Nicolis, “A master equationdescription of local fluctuations,” J. Stat. Phys. 13, 197–217 (1975).

[8] Jürgen Schnakenberg, “Network theory of microscopicand macroscopic behavior of master equation systems,”Rev. Mod. Phys. 48, 571–585 (1976).

Page 19: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

19

[9] T.L. Hill, Free energy transduction in biology (AcademicPress, New York, 1977).

[10] C. Y. Mou, J. Luo, and G. Nicolis, “Stochastic thermo-dynamics of nonequilibrium steady states in chemicalreaction systems,” J. Chem. Phys. 84, 7011–7017 (1986).

[11] John Ross, Thermodynamics and Fluctuations far fromEquilibrium (Springer, 2008).

[12] Donald A. McQuarrie, “Stochastic approach to chemicalkinetics,” J. Appl. Probab. 4, 413 (1967).

[13] Daniel T. Gillespie, “A rigorous derivation of the chemi-cal master equation,” Physica A 188, 404–425 (1992).

[14] Ken Sekimoto, Stochastic Energetics, 1st ed., LectureNotes in Physics, Vol. 799 (Springer-Verlag Berlin Hei-delberg, 2010).

[15] Christopher Jarzynski, “Equalities and inequalities: Ir-reversibility and the second law of thermodynamics atthe nanoscale,” Annu. Rev. Condens. Matter Phys. 2,329–351 (2011).

[16] Udo Seifert, “Stochastic thermodynamics, fluctuationtheorems and molecular machines,” Rep. Prog. Phys. 75,126001 (2012).

[17] C. Van den Broeck and M. Esposito, “Ensemble and tra-jectory thermodynamics: A brief introduction,” PhysicaA 418, 6–16 (2015).

[18] Pierre Gaspard, “Fluctuation theorem for nonequilib-rium reactions,” J. Chem. Phys. 120, 8898–8905 (2004).

[19] David Andrieux and Pierre Gaspard, “Fluctuation theo-rem and onsager reciprocity relations,” J. Chem. Phys.121, 6167 (2004).

[20] Tim Schmiedl, Thomas Speck, and Udo Seifert, “En-tropy production for mechanically or chemically drivenbiomolecules,” J. Stat. Phys. 128, 77–93 (2007).

[21] Tim Schmiedl and Udo Seifert, “Stochastic thermody-namics of chemical reaction networks,” J. Chem. Phys.126, 044101 (2007).

[22] M. Polettini, A. Wachtel, and M. Esposito, “Dissipationin noisy chemical networks: The role of deficiency,” J.Chem. Phys. 143, 184103 (2015).

[23] Robert A Alberty, Thermodynamics of biochemical reac-tions (Wiley-Interscience, 2003).

[24] Daniel A. Beard and Hong Qian, Chemical Biophysics.Quantitative Analysis of Cellular Systems (CambridgeUniversity Press, 2008).

[25] Bernhard Ø. Palsson, Systems Biology: Properties ofReconstructed Networks (Cambridge University Press,2006).

[26] Bernhard Ø. Palsson, Systems Biology: Simulation ofDynamic Network States (Cambridge University Press,2011).

[27] Bernhard Ø. Palsson, Systems Biology: Constraint-basedNetwork Reconstruction and Analysis (Cambridge Uni-versity Press, 2015).

[28] Albert Goldbeter, Biochemical Oscillations And CellularRhythms (Cambridge University Press, 1996).

[29] I.R. Epstein and J.A. Pojman, An Introduction to Non-linear Chemical Dynamics: Oscillations, Waves, Pat-terns, and Chaos, Topics in Physical Chemistry (OxfordUniversity Press, 1998).

[30] Daniel A. Beard, Shou-dan Liang, and Hong Qian, “En-ergy balance for analysis of complex metabolic networks,”Biophys. J. 83, 79–86 (2002).

[31] Hong Qian, Daniel A. Beard, and Shou-dan Liang, “Sto-ichiometric network theory for nonequilibrium biochemi-cal systems,” Eur. J. Biochem. 270, 415–421 (2003).

[32] Daniel A. Beard, Eric Babson, Edward Curtis, andHong Qian, “Thermodynamic constraints for biochemicalnetworks,” J. Theor. Biol. 228, 327–333 (2004).

[33] Anirikh Chakrabarti, Ljubisa Miskovic, Keng Cher Soh,and Vassily Hatzimanikatis, “Towards kinetic modelingof genome-scale metabolic networks without sacrificingstoichiometric, thermodynamic and physiological con-straints,” Biotechnol. J. 8, 1043–1057 (2013).

[34] Hong Qian and Daniel A. Beard, “Thermodynamics ofstoichiometric biochemical networks in living systems farfrom equilibrium,” Biophys. Chem. 114, 213–220 (2005).

[35] F. Horn and R. Jackson, “General mass action kinetics,”Arch. Ration. Mech. An. 47, 81–116 (1972).

[36] Martin Feinberg, “Complex balancing in general kineticsystems,” Arch. Ration. Mech. An. 49, 187–194 (1972).

[37] F. Horn, “Necessary and sufficient conditions for complexbalancing in chemical kinetics,” Arch. Ration. Mech. An.49, 172–186 (1972).

[38] Thomas G. Kurtz, “The relationship between stochas-tic and deterministic models for chemical reactions,” J.Chem. Phys. 57, 2976 (1972).

[39] David F. Anderson, Gheorghe Craciun, and Thomas G.Kurtz, “Product-form stationary distributions for de-ficiency zero chemical reaction networks,” Bull. Math.Biol. 72, 1947–1970 (2010).

[40] David F. Anderson, Gheorghe Craciun, Manoj Gopalkr-ishnan, and Carsten Wiuf, “Lyapunov functions, sta-tionary distributions, and non-equilibrium potential forreaction networks,” Bull. Math. Biol. 77, 1744–1767(2015).

[41] Daniele Cappelletti and Carsten Wiuf, “Product-formpoisson-like distributions and complex balanced reactionsystems,” SIAM J. Appl. Math. 76, 411–432 (2016).

[42] K. Takara, H.-H. Hasegawa, and D.J. Driebe, “Gen-eralization of the second law for a transition betweennonequilibrium states,” Phys. Lett. A 375, 88–92 (2010).

[43] M. Esposito and C. Van den Broeck, “Second law andlandauer principle far from equilibrium,” Europhys. Lett.95, 40004 (2011).

[44] Takahiro Sagawa, Thermodynamics of Information Pro-cessing in Small Systems (Springer Japan, 2013) doctoralThesis.

[45] Juan M. R. Parrondo, Jordan M. Horowitz, andTakahiro Sagawa, “Thermodynamics of information,”Nature Phys. 11, 131–139 (2015).

[46] M. Esposito, U. Harbola, and S. Mukamel, “Entropyfluctuation theorems in driven open systems: Applicationto electron counting statistics,” Phys. Rev. E 76, 031132(2007).

[47] R. J. Harris and G. M. Schütz, “Fluctuation theoremsfor stochastic dynamics,” J. Stat. Mech. Theor. Exp. ,P07020 (2007).

[48] M. Esposito and C. Van den Broeck, “Three faces of thesecond law. i. master equation formulation,” Phys. Rev.E 82, 011143 (2010).

[49] C. Van den Broeck and M. Esposito, “Three faces of thesecond law. ii. fokker-planck formulation,” Phys. Rev. E82, 011144 (2010).

[50] Hao Ge and Hong Qian, “Physical origins of entropy pro-duction, free energy dissipation, and their mathematicalrepresentations,” Phys. Rev. E 81, 051133 (2010).

[51] Hao Ge and Hong Qian, “Nonequilibrium thermody-namic formalism of nonlinear chemical reaction systemswith Waage–Guldberg’s law of mass action,” Chem. Phys.

Page 20: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

20

472, 241–248 (2016).[52] Martin Feinberg, “Lectures on chemical reaction net-

works,” (1979), lecture Notes.[53] Jeremy Gunawardena, “Chemical reaction network the-

ory for in-silico biologists contents,” (2003), unpub-lished.

[54] S. R. de Groot and P. Mazur, Non-equilibrium Thermo-dynamics (Dover, 1984).

[55] Miloslav Pekař, “Thermodynamics and foundations ofmass-action kinetics,” Prog. React. Kinet. Mech. 30,3–113 (2005).

[56] Matteo Polettini and Massimiliano Esposito, “Irre-versible thermodynamics of open chemical networks. i.emergent cycles and broken conservation laws,” J. Chem.Phys. 141, 024117 (2014).

[57] D. Kondepudi and I. Prigogine, Modern Thermodynam-ics: From Heat Engines to Dissipative Structures, 2nded. (Wiley, 2014).

[58] Moiety, in Compendium of Chemical Terminology (IU-PAC Gold Book) (2014).

[59] Stefan Schuster and Ronny Schuster, “A generalizationof wegscheider’s condition. implications for properties ofsteady states and for quasi-steady-state approximation,”J. Math. Chem. 3, 25–42 (1989).

[60] J. Leiser and J. J. Blum, “On the analysis of substratecycles in large metabolic systems,” Cell Biophys. 11,123–138 (1987).

[61] S. Schuster and C. Hilgetag, “On elementary flux modesin biochemical reaction systems at steady state,” J. Biol.Syst. 02, 165–182 (1994).

[62] Steffen Klamt and Jörg Stelling, “Two approaches formetabolic pathway analysis?” Trends Biotechnol. 21,64–69 (2003).

[63] Riccardo Rao, David Lacoste, and Massimiliano Espos-ito, “Glucans monomer-exchange dynamics as an openchemical network,” J. Chem. Phys. 143, 244903 (2015).

[64] F. Horn, “Stability and complex balancing in mass-actionsystems with three short complexes,” Proc. R. Soc. A334, 331–342 (1973).

[65] S. Klamt, U.-U. Haus, and F. Theis, “Hypergraphsand cellular networks,” PLoS Comput. Biol. 5, e1000385(2009).

[66] C. Adam Petri and W. Reisig, “Petri net,” Scholarpedia3, 6477 (2008), revision 91646.

[67] J. C. Baez and B. Fong, “Quantum techniques for study-ing equilibrium in reaction networks,” J. Complex Net-works 3, 22–34 (2014).

[68] Gheorghe Craciun, “Toric differential inclusions and aproof of the global attractor conjecture,” ArXiv e-prints(2015).

[69] Alicia Dickenstein and Mercedes Pérez Millán, “Howfar is complex balancing from detailed balancing?” Bull.Math. Biol. 73, 811–828 (2011).

[70] I. Prigogine, “Le domaine de validité de la thermody-namique des phénomènes irréversibles,” Physica 15, 272–284 (1949).

[71] E.R. Cohen, T. Cvitas, J.G. Frey, B. Holmström, K. Ku-chitsu, R. Marquardt, I. Mills, F. Pavese, M. Quack,J. Stohner, H.L. Strauss, M. Takami, and A.J. Thor,Quantities, Units and Symbols in Physical Chemistry(IUPAC Green Book), 3rd ed. (IUPAC & RSC Publish-ing, 2008) 2nd Printing.

[72] R. Holyst and A. Poniewierski, Thermodynamics forChemists, Physicists and Engineers (Springer, 2012).

[73] David Shear, “An analog of the boltzmann h-theorem(a liapunov function) for systems of coupled chemicalreactions,” J. Theor. Biol. 16, 212–228 (1967).

[74] J. Higgins, “Some remarks on shear’s liapunov functionfor systems of chemical reactions,” J. Theor. Biol. 21,293–304 (1968).

[75] Thomas M. Cover and Joy A. Thomas, Elements ofInformation Theory, 2nd ed. (Wiley-Interscience, 2006).

[76] Arjan van der Schaft, Shodhan Rao, and Bayu Jayaward-hana, “On the mathematical structure of balanced chem-ical reaction networks governed by mass action kinetics,”SIAM J. Appl. Math. 73, 953–973 (2013).

[77] I. Procaccia and R. D. Levine, “Potential work: Astatistical-mechanical approach for systems in disequi-librium,” J. Chem. Phys. 65, 3357 (1976).

[78] S. Vaikuntanathan and C. Jarzynski, “Dissipation andlag in irreversible processes,” Europhys. Lett. 87, 60005(2009).

[79] H.-H. Hasegawa, J. Ishikawa, K. Takara, and D.J.Driebe, “Generalization of the second law for a nonequi-librium initial state,” Phys. Lett. A 374, 1001–1004(2010).

[80] Manoj Gopalkrishnan, “On the Lyapunov function forcomplex-balanced mass-action systems,” ArXiv e-prints(2013).

[81] Arjan van der Schaft, Shodhan Rao, and Bayu Jayaward-hana, “Complex and detailed balancing of chemical reac-tion networks revisited,” J. Math. Chem. 53, 1445–1458(2015).

[82] The fact that the matrix whose entries are {`λbσy} is

nonsingular follows from `λu ∈ ker∇ , ∀λu, and fromthe linear independence of {`λ}.

[83] Massimiliano Esposito and Gernot Schaller, “Stochasticthermodynamics for “Maxwell demon” feedbacks,” E.P.L.99, 30003 (2012).

[84] G. Diana, G.B. Bagci, and M. Esposito, “Finite-timeerasing of information stored in fermionic bits,” Phys.Rev. E 87, 012111 (2013).

[85] Jordan M. Horowitz and Massimiliano Esposito, “Ther-modynamics with continuous information flow,” Phys.Rev. X 4, 031015 (2014).

[86] A. C. Barato and U. Seifert, “Unifying three perspectiveson information processing in stochastic thermodynamics,”Phys. Rev. Lett. 112, 090601 (2014).

[87] S. Bo, M. Del Giudice, and A. Celani, “Thermodynamiclimits to information harvesting by sensory systems,” J.Stat. Mech. Theor. Exp. , P01014 (2015).

[88] P. Strasberg, G. Schaller, T. Brandes, and M. Esposito,“Thermodynamics of a physical model implementing aMaxwell demon,” Phys. Rev. Lett. 110, 040601 (2013).

[89] N. Golubeva and A. Imparato, “Efficiency at maximumpower of interacting molecular machines,” Phys. Rev.Lett. 109, 190602 (2012).

[90] Bernhard Altaner, Artur Wachtel, and Jürgen Vollmer,“Fluctuating currents in stochastic thermodynamics. ii.energy conversion and nonequilibrium response in kinesinmodels,” Phys. Rev. E 92, 042133 (2015).

[91] U. Seifert, “Stochastic thermodynamics of single enzymesand molecular motors,” Eur. Phys. J. E 34, 26 (2011).

[92] Riccardo Rao and Luca Peliti, “Thermodynamics of ac-curacy in kinetic proofreading: dissipation and efficiencytrade-offs,” J. Stat. Mech. Theor. Exp. , P06001 (2015).

Page 21: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

21

[93] John J. Hopfield, “Kinetic proofreading: a new mecha-nism for reducing errors in biosynthetic processes requir-ing high specificity,” Proc. Natl. Acad. Sci. U.S.A. 71,4135–4139 (1974).

[94] Jacques Ninio, “Kinetic amplification of enzyme discrim-ination,” Biochimie 57, 587–595 (1975).

[95] J. J. Hopfield, “The energy relay: a proofreading schemebased on dynamic cooperativity and lacking all character-istic symptoms of kinetic proofreading in dna replicationand protein synthesis,” Proc. Natl. Acad. Sci. U.S.A. 77,5248–5252 (1980).

[96] E.D. Sontag, “Structure and stability of certain chemicalnetworks and applications to the kinetic proofreadingmodel of t-cell receptor signal transduction,” IEEE Trans-actions on Automatic Control 46, 1028–1047 (2001).

[97] Hao Ge, Min Qian, and Hong Qian, “Stochastic theoryof nonequilibrium steady states. part II: Applications inchemical biophysics,” Phys. Rep. 510, 87–118 (2012).

[98] A. Murugan, D. H. Huse, and S. Leibler, “Speed, dis-sipation, and error in kinetic proofreading,” Proc. Natl.Acad. Sci. U.S.A. 109, 12034 (2012).

[99] A. Murugan, D. H. Huse, and S. Leibler, “Discriminatoryregimes in non-equilibrium systems,” Phys. Rev. X 4,021016 (2014).

[100] Charles H. Bennett, “Dissipation–error tradeoff in proof-reading,” Biosystems 11, 85–91 (1979).

[101] David Andrieux and Pierre Gaspard, “Nonequilibriumgeneration of information in copolymerization processes,”Proc. Natl. Acad. Sci. U.S.A. 105, 9516–9521 (2008).

[102] David Andrieux and Pierre Gaspard, “Molecular infor-mation processing in nonequilibrium copolymerizations,”J. Chem. Phys. 130, 014901 (2009).

[103] J. Ricardo Arias-Gonzalez, “Entropy involved in fidelityof DNA replication,” PLoS ONE 7, e42272 (2012).

[104] P. Sartori and S. Pigolotti, “Kinetic versus energeticdiscrimination in biological copying,” Phys. Rev. Lett.110, 188101 (2013).

[105] P. Sartori and S. Pigolotti, “Thermodynamics of errorcorrection,” Phys. Rev. X 5, 041039 (2015).

[106] Athel Cornish-Bowden, Fundamentals of Enzyme Kinet-ics, 4th ed. (Wiley-Blackwell, 2012).

[107] Massimiliano Esposito, “Stochastic thermodynamics un-der coarse graining,” Phys. Rev. E 85, 041125 (2012).

[108] C. L. Linster, E. van Schaftingen, and A. D. Hanson,“Metabolite damage and its repair or pre-emption,” Nat.Chem. Biol. 9, 72–80 (2013).

[109] Hong Qian and Timothy C. Reluga, “Nonequilibriumthermodynamics and nonlinear kinetics in a cellularsignaling switch,” Phys. Rev. Lett. 94, 028101 (2005).

[110] H. Qian and S. Roy, “An information theoretical analysisof kinase activated phosphorylation dephosphorylationcycle,” IEEE Trans. Nanobiosci. 11, 289–295 (2012).

[111] José M. G. Vilar, Hao Yuan Kueh, Naama Barkai, andStanislas Leibler, “Mechanisms of noise-resistance ingenetic oscillators,” Proc. Natl. Acad. Sci. U.S.A. 99,5988–5992 (2002).

[112] G. Nicolis and I. Prigogine, Self-organization in Nonequi-librium Systems: From Dissipative Structures to OrderThrough Fluctuations (Wiley-Blackwell, 1977).

[113] R. Feistel and W. Ebeling, Physics of Self-Organizationand Evolution (Wiley VCH, 2011).

[114] Pankaj Mehta and David J. Schwab, “Energetic costsof cellular computation,” Proc. Natl. Acad. Sci. U.S.A.109, 17978–17982 (2012).

[115] Pablo Sartori, Léo Granger, Chiu Fan Lee, and Jor-dan M. Horowitz, “Thermodynamic costs of informationprocessing in sensory adaptation,” PLoS Comput Biol10, e1003974 (2014).

[116] David Hartich, Andre C. Barato, and Udo Seifert,“Nonequilibrium sensing and its analogy to kinetic proof-reading,” New J. Phys. 17, 055026 (2015).

[117] Shou-Wen Wang, Yueheng Lan, and Lei-Han Tang,“Energy dissipation in an adaptive molecular circuit,” J.Stat. Mech. Theor. Exp. , P07025 (2015).

[118] Yasubumi Sakakibara and Yongli Yongli Mi, eds., DNAComputing and Molecular Programming (Springer BerlinHeidelberg, 2011); Luca Cardelli and William Shih, eds.,DNA Computing and Molecular Programming (SpringerBerlin Heidelberg, 2011); Darko Stefanovic and An-drew Turberfield, eds., DNA Computing and Molecu-lar Programming (Springer Berlin Heidelberg, 2012);David Soloveichik and Bernard Yurke, eds., DNA Com-puting and Molecular Programming (Springer Interna-tional Publishing, 2013); Satoshi Murata and SatoshiKobayashi, eds., DNA Computing and Molecular Pro-gramming (Springer International Publishing, 2014); An-drew Phillips and Peng Yin, eds., DNA Computing andMolecular Programming (Springer International Publish-ing, 2015).

[119] A. Hjelmfelt, E.D. Weinberger, and J. Ross, “Chemicalimplementation of neural networks and turing machines,”Proc. Natl. Acad. Sci. U.S.A. 88, 10983–10987 (1991).

[120] A. Hjelmfelt, E.D. Weinberger, and J. Ross, “Chemicalimplementation of finite-state machines,” Proc. Natl.Acad. Sci. U.S.A. 89, 383–387 (1992).

[121] Marcelo O. Magnasco, “Chemical kinetics is turing uni-versal,” Phys. Rev. Lett. 78, 1190–1193 (1997).

[122] David Soloveichik, Matthew Cook, Erik Winfree, andJehoshua Bruck, “Computation with finite stochasticchemical reaction networks,” Nat. Comp. 7, 615–633(2008).

[123] Charles H. Bennett, “The thermodynamics ofcomputation—a review,” Int. J. Theor. Phys. 21,905–940 (1982).

[124] Tuomas P. J. Knowles, Wenmiao Shu, Glyn L. Devlin,Sarah Meehan, Stefan Auer, Christopher M. Dobson,and Mark E. Welland, “Kinetics and thermodynamics ofamyloid formation from direct measurements of fluctua-tions in fibril mass,” Proc. Natl. Acad. Sci. U.S.A. 104,10016–10021 (2007).

[125] Tuomas P. J. Knowles, Christopher A. Waudby, Glyn L.Devlin, Samuel I. A. Cohen, Adriano Aguzzi, MicheleVendruscolo, Eugene M. Terentjev, Mark E. Welland,and Christopher M. Dobson, “An analytical solutionto the kinetics of breakable filament assembly,” Science326, 1533–1537 (2009).

[126] Samuel I. A. Cohen, Sara Linse, Leila M. Luheshi, ErikHellstrand, Duncan A. White, Luke Rajah, Daniel E.Otzen, Michele Vendruscolo, Christopher M. Dobson,and Tuomas P. J. Knowles, “Proliferation of amyloid-β42 aggregates occurs through a secondary nucleationmechanism,” Proc. Natl. Acad. Sci. U.S.A. 110, 9758–9763 (2013).

[127] Zoe Budrikis, Giulio Costantini, Caterina A. M. La Porta,and Stefano Zapperi, “Protein accumulation in the endo-plasmic reticulum as a non-equilibrium phase transition,”Nat. Commun. 5, 3620 (2014).

Page 22: Complex Systems and Statistical Mechanics, Physics and ... › pdf › 1602.07257.pdf · Nonequilibrium Thermodynamics of Chemical Reaction Networks: Wisdom from Stochastic Thermodynamics

22

[128] Önder Kartal, Sebastian Mahlow, Alexander Skupin,and Oliver Ebenhöh, “Carbohydrate-active enzymes ex-emplify entropic principles in metabolism,” Mol. Syst.Biol. 7, 542–542 (2011).

[129] Sourabh Lahiri, Yang Wang, Massimiliano Esposito, andDavid Lacoste, “Kinetics and thermodynamics of re-versible polymerization in closed systems,” New J. Phys.17, 085008 (2015).

[130] Paul Krapivsky, Sidney Redner, and Eli Ben-Naim, AKinetic View of Statistical Physics (Cambridge Univer-

sity Press, 2010).[131] Enrico Fermi, Thermodynamics (Dover, 1956).[132] Hao Ge and Hong Qian, “Mesoscopic kinetic basis of

macroscopic chemical thermodynamics: A mathematicaltheory,” ArXiv e-prints (2016).

[133] E. T. Jaynes, “Information theory and statistical me-chanics,” Phys. Rev. 106, 620–630 (1957).