+ All Categories
Home > Documents > with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response...

with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response...

Date post: 01-Jun-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
21
Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M. Horowitz 3, 4, 5 1 Physics of Living Systems Group, Department of Physics, Massachusetts Institute of Technology, 400 Technology Square, Cambridge, MA 02139 2 Department of Chemistry, Northwestern University, Evanston, IL 60208 3 Department of Biophysics, University of Michigan, Ann Arbor, Michigan, 48109, USA 4 Center for the Study of Complex Systems, University of Michigan, Ann Arbor, Michigan 48104, USA 5 Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA (Dated: February 11, 2020) Diverse physical systems are characterized by their response to small perturbations. Near ther- modynamic equilibrium, the fluctuation-dissipation theorem provides a powerful theoretical and experimental tool to determine the nature of response by observing spontaneous equilibrium fluc- tuations. In this spirit, we derive here a collection of equalities and inequalities valid arbitrarily far from equilibrium that constrain the response of nonequilibrium steady states in terms of the strength of nonequilibrium driving. Our work opens new avenues for characterizing nonequilibrium response. As illustrations, we show how our results rationalize the energetic requirements of two common biochemical motifs. I. INTRODUCTION One of the basic characteristics of any physical sys- tem is its response to small perturbations [1]. For in- stance, response is used to quantify everything from ma- terial properties—such as conductivity [2] and viscoelas- ticity [3]—to the sensing capability of cells [4, 5] and the accuracy of biomolecular processes [6–8]. Near thermo- dynamic equilibrium, response is completely determined by the nature of spontaneous fluctuations, according to the fluctuation-dissipation theorem (FDT) [2]. This deep connection between response and fluctuations is not only of theoretical interest, but also finds practical applica- tion. The FDT forms the basis of powerful experimen- tal techniques, such as microrheology, spectroscopy, and dynamic light scattering [1]. It additionally has impli- cations for the design of mesoscopic devices: highly- responsive equilibrium devices are always plagued by noise. To combine low fluctuations with high sensitiv- ity, a device must be driven away from equilibrium. The great utility of the FDT near equilibrium [1] has led to significant interest in expanding its validity and developing generalizations for nonequilibrium situa- tions. Generically, response can be related to some for- mal nonequilibrium correlation functions [9–13]. While these predictions offer fundamental theoretical insight, the necessary correlations are often prohibitively difficult to measure except in simple single-particle systems [14– 16]. In certain special cases, such as under stalling condi- tions, the FDT holds unmodified [17]. More commonly, however, the study of nonequilibrium response has fo- cused on how the FDT is violated. For example, the violation of the velocity FDT for Brownian particles can be related to the steady-state heat dissipation through the Harada-Sasa equality [18, 19]; a useful prediction that has been utilized to measure dissipation and ef- ficiency in molecular motors [20]. Alternatively, viola- tions of the FDT can be used to fit model parameters, as has been suggested for models of biomolecular pro- cesses [21, 22]. More often FDT violations are framed in terms of system-specific “effective temperatures” [23– 25], whose time-dependence under some circumstances can reveal information about effective equilibrium de- scriptions. Inspired by the recent demonstration of thermody- namic bounds on far-from-equilibrium dynamical fluc- tuations [26, 27], we show here that generic nonequilib- rium steady-state response can be constrained in terms of experimentally-accessible thermodynamic quantities. In particular, we present equalities and inequalities—akin to the FDT but valid arbitrarily far from equilibrium—that link static response to the strength of nonequilibrium driving. Our results open new possibilities to experimen- tally characterize away-from-equilibrium response and suggest design principles for high-sensitivity, low-noise devices. As illustrations, we show how our results ratio- nalize the energetic requirements of biochemical switches, biochemical sensing, and kinetic proofreading. II. MODELING NONEQUILIBRIUM STEADY STATES Nonequilibrium steady states are characterized by the constant and irreversible exchange of energy and mat- ter between a system and its environment. These flows are driven by thermodynamic affinities—quantities like temperature gradients, chemical potential differences and nonconservative mechanical forces. The underlying dy- namics leading to the establishment of such steady states are often well-modeled as a continuous-time Markov jump process on a finite set of states i =1,...,N , which represent (coarse-grained) physical configurations. The probability p i (t) to find the system in state i at time t arXiv:1905.07449v2 [cond-mat.stat-mech] 10 Feb 2020
Transcript
Page 1: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

Universal thermodynamic bounds on nonequilibrium responsewith biochemical applications

Jeremy A. Owen,1 Todd R. Gingrich,2 and Jordan M. Horowitz3, 4, 5

1Physics of Living Systems Group, Department of Physics,Massachusetts Institute of Technology, 400 Technology Square, Cambridge, MA 02139

2Department of Chemistry, Northwestern University, Evanston, IL 602083Department of Biophysics, University of Michigan, Ann Arbor, Michigan, 48109, USA

4Center for the Study of Complex Systems, University of Michigan, Ann Arbor, Michigan 48104, USA5Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA

(Dated: February 11, 2020)

Diverse physical systems are characterized by their response to small perturbations. Near ther-modynamic equilibrium, the fluctuation-dissipation theorem provides a powerful theoretical andexperimental tool to determine the nature of response by observing spontaneous equilibrium fluc-tuations. In this spirit, we derive here a collection of equalities and inequalities valid arbitrarilyfar from equilibrium that constrain the response of nonequilibrium steady states in terms of thestrength of nonequilibrium driving. Our work opens new avenues for characterizing nonequilibriumresponse. As illustrations, we show how our results rationalize the energetic requirements of twocommon biochemical motifs.

I. INTRODUCTION

One of the basic characteristics of any physical sys-tem is its response to small perturbations [1]. For in-stance, response is used to quantify everything from ma-terial properties—such as conductivity [2] and viscoelas-ticity [3]—to the sensing capability of cells [4, 5] and theaccuracy of biomolecular processes [6–8]. Near thermo-dynamic equilibrium, response is completely determinedby the nature of spontaneous fluctuations, according tothe fluctuation-dissipation theorem (FDT) [2]. This deepconnection between response and fluctuations is not onlyof theoretical interest, but also finds practical applica-tion. The FDT forms the basis of powerful experimen-tal techniques, such as microrheology, spectroscopy, anddynamic light scattering [1]. It additionally has impli-cations for the design of mesoscopic devices: highly-responsive equilibrium devices are always plagued bynoise. To combine low fluctuations with high sensitiv-ity, a device must be driven away from equilibrium.

The great utility of the FDT near equilibrium [1]has led to significant interest in expanding its validityand developing generalizations for nonequilibrium situa-tions. Generically, response can be related to some for-mal nonequilibrium correlation functions [9–13]. Whilethese predictions offer fundamental theoretical insight,the necessary correlations are often prohibitively difficultto measure except in simple single-particle systems [14–16]. In certain special cases, such as under stalling condi-tions, the FDT holds unmodified [17]. More commonly,however, the study of nonequilibrium response has fo-cused on how the FDT is violated. For example, theviolation of the velocity FDT for Brownian particles canbe related to the steady-state heat dissipation throughthe Harada-Sasa equality [18, 19]; a useful predictionthat has been utilized to measure dissipation and ef-ficiency in molecular motors [20]. Alternatively, viola-tions of the FDT can be used to fit model parameters,

as has been suggested for models of biomolecular pro-cesses [21, 22]. More often FDT violations are framedin terms of system-specific “effective temperatures” [23–25], whose time-dependence under some circumstancescan reveal information about effective equilibrium de-scriptions.

Inspired by the recent demonstration of thermody-namic bounds on far-from-equilibrium dynamical fluc-tuations [26, 27], we show here that generic nonequilib-rium steady-state response can be constrained in terms ofexperimentally-accessible thermodynamic quantities. Inparticular, we present equalities and inequalities—akin tothe FDT but valid arbitrarily far from equilibrium—thatlink static response to the strength of nonequilibriumdriving. Our results open new possibilities to experimen-tally characterize away-from-equilibrium response andsuggest design principles for high-sensitivity, low-noisedevices. As illustrations, we show how our results ratio-nalize the energetic requirements of biochemical switches,biochemical sensing, and kinetic proofreading.

II. MODELING NONEQUILIBRIUM STEADYSTATES

Nonequilibrium steady states are characterized by theconstant and irreversible exchange of energy and mat-ter between a system and its environment. These flowsare driven by thermodynamic affinities—quantities liketemperature gradients, chemical potential differences andnonconservative mechanical forces. The underlying dy-namics leading to the establishment of such steady statesare often well-modeled as a continuous-time Markovjump process on a finite set of states i = 1, . . . , N , whichrepresent (coarse-grained) physical configurations. Theprobability pi(t) to find the system in state i at time t

arX

iv:1

905.

0744

9v2

[co

nd-m

at.s

tat-

mec

h] 1

0 Fe

b 20

20

Page 2: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

2

FIG. 1. Transition graphs and cycles. (a) Representa-tive transition graph for a 4-state system, which will act asa recurring illustrative example. (b) Cycles around which thecycle forces FC drive the system out of equilibrium.

then evolves according to the master equation

pi(t) =

N∑j=1

Wijpj(t), (1)

where the off-diagonal entries of the transition rate ma-trix Wij specify the probability per unit time to jumpfrom j to i, and diagonal entries Wii = −

∑j 6=iWji

are fixed by the conservation of probability. Time-reversibility of the underlying microscopic dynamics im-plies that Wij 6= 0 only if Wji 6= 0 [28]. We will addi-tionally suppose that for any two states, there is somesequence of allowed transitions (Wij 6= 0) connectingthem, a property known as irreducibility. Under thisassumption, no matter the initial condition, the solu-tion of the master equation (1) converges at long timesto the unique steady state distribution π that satisfies∑Nj=1Wijπj = 0. This distribution π, and in particular

its dependence on physical quantities through the tran-sition rates Wij , serves as a general model of a nonequi-librium steady state.

Many key properties of this nonequilibrium steadystate, including its thermodynamics, come to light bypicturing the stochastic dynamics described by (1) play-ing out on a transition graph—a weighted directed graphG, as in Fig. 1(a), where the vertices {i} represent thestates and directed edges {emn} represent possible transi-tions, weighted by the rates Wmn. Note that, by assump-tion, every edge in G has a reverse, so we will representand discuss the transition graph as if it were an undi-rected graph, with the understanding that every undi-rected edge represents two opposing directed edges.

The cycles in the graph, like those in Fig. 1(b), play acentral role in the thermodynamics of the steady state. Acycle is a sequence of directed edges and vertices connect-ing the initial vertex to itself without self-intersecting,

C = {i0ei1i0−−−→ i1 → · · · → im

ei0im−−−→ i0}. The asymmetryof the rates around these cycles then encodes the thermo-dynamic affinities driving the system out of equilibriumthrough the cycle forces—the log of the product of ratesaround the cycle divided by the product of rates in the

reverse orientation [29, 30]:

FC = ln

(Wi0im · · ·Wi2i1Wi1i0

Wimi0 · · ·Wi1i2Wi0i1

). (2)

These cycle forces are linear combinations of thermody-namic affinities multiplied by their conjugate distances—for example a chemical potential gradient times a changein particle number. As such, the cycle forces equal thedissipation (entropy production) in the environment ac-crued every time the system flows around the cycle C.This means that the cycle forces depend on macroscop-ically tunable parameters—such as environmental tem-perature or chemical potential—that characterize howstrongly the system is driven away from equilibrium. Ifall the cycle forces vanish, the system satisfies detailedbalance, a statistical time-reversal symmetry [31] charac-teristic of thermodynamic equilibrium.

III. STATIC RESPONSE TO PERTURBATIONS

Now, suppose the transition rates Wij(λ) depend ona control parameter λ, which could represent, say, thestrength of an applied electric field, a temperature, oreven a microscopic kinetic parameter such as a reactionbarrier. In this work, we focus on the response to staticperturbations, that is how steady-state averages 〈Q〉π =∑j Qjπj respond to small changes in λ:

∂λ〈Q〉π =∑i

Qi∂λπi =∑i

Qi∑kl

∂Wkl

∂λ

∂πi∂Wkl

. (3)

At thermal equilibrium, the steady state πeqi ∝ e−βεi(λ)

depends only on the underlying (free) energy landscapeεi(λ), irrespective of the precise form of the transitionrates, where β = 1/kBT with kB Boltzmann’s constantand T temperature. This simplifying fact immediatelyimplies the static FDT, which equates the static responseto an equilibrium correlation function,

∂λ〈Q〉eq = βCeq(Q,V ), (4)

where Ceq(Q,V ) = 〈QV 〉eq − 〈Q〉eq〈V 〉eq and the sub-script “eq” emphasizes that averages are taken with re-spect to the equilibrium distribution [2]. Here, V = −∂λεis known as the coordinate conjugate to λ and representsthe displacement induced by λ—for example, volume isconjugate to pressure and particle number is conjugateto chemical potential. The FDT’s utility in part stemsfrom the fact that we often know the conjugate coordi-nate from basic physical reasoning and it is easily mea-sured.

Away from equilibrium, the steady-state distribu-tion generally has a complicated dependence on therates. Nevertheless, response can always be related toa nonequilibrium correlation function [9, 10],

∂λ〈Q〉π = Cπ(Q, ∂λ lnπ), (5)

Page 3: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

3

but the relevant “conjugate coordinate” ∂λ lnπ requiresknowledge of the parameter-dependence of the fullnonequilibrium steady-state distribution, which can bechallenging to calculate or measure. Still, this responseformula has been given thermodynamic meaning by re-lating the nonequilibrium conjugate coordinate to thestochastic entropy production rate [13] as well as to thetime-reversal symmetry properties of the path action [11].

IV. PARAMETERIZING PERTURBATIONS

Here, we turn our attention to the variations ofthe steady-state distribution with the transition rates,∂πi/∂Wkl, constraining them in terms of π and the cy-cle forces FC . We accomplish this goal by breaking anygeneral perturbation into a linear combination of threespecial types of perturbations that change rates in a co-ordinated way. By focusing on these subclasses of pertur-bations, we will be able to provide clear and measurablethermodynamic constraints on static response.

To classify perturbations, it will prove fruitful to pa-rameterize the rate matrix as

Wij = exp [−(Bij − Ej − Fij/2)] , (6)

introducing the vertex parameters Ej , (symmetric) edgeparameters Bij = Bji, and asymmetric edge parametersFij = −Fji, which can all be varied independently. Anyrate matrix can be cast in this form, albeit non-uniquely.To see this, consider the following program for identifyingsuch a parameterization: choose the vertex parameters{E1, . . . , EN} arbitrarily, then set

Bij =1

2[(Ej − lnWij) + (Ei − lnWji)] (7)

Fij = − [(Ej − lnWij)− (Ei − lnWji)] . (8)

The non-uniqueness of the parameterization is manifestin this construction because of the freedom to choosethe Ei. For example, we could choose Ei = 0 for all i.We emphasize, however, that no matter the choice, thederivatives with respect to Ej , Bij or Fij are independentof the values of the parameterization. Variations of avertex parameter, say Ej , is equivalent to scaling all ofthe rates out of state j

∂πk∂Ej

=∑i

Wij∂πk∂Wij

. (9)

Similarly, derivatives with respect to Bij and Fij multi-plicatively scale transition rates associated with a singleedge

∂πk∂Bij

= −Wij∂πk∂Wij

−Wji∂πk∂Wji

(10)

∂πk∂Fij

=1

2

(Wij

∂πk∂Wij

−Wji∂πk∂Wji

). (11)

FIG. 2. Parameterizing perturbations. Any perturbationof rates can be decomposed into the variation of some combi-nation of (a) vertex parameters, (b) edge parameters, and (c)asymmetric edge parameters. Affected rates are highlighted.

These facts are illustrated in Fig. 2. We note that theright hand sides of equations (9)–(11) do not depend onthe choice of parameterization (6), even though the pa-rameterization is not unique.

Our parameterization is reminiscent of the Arrheniusexpression for transition rates for a system evolving in anenergy landscape with wells of depth Ei and barriers ofheight Bij driven by forces Fij . While we stress that (6)will not in general support such an interpretation, theanalogy is suggestive in several ways. For example, theasymmetric edge parameters Fij are the sole contributorsto the cycle forces (affinities) FC =

∑eij∈C Fij . Further-

more, if all the Fij = 0, the steady-state distribution hasthe Boltzmann form πi ∝ exp(−Ei), with the Ei actingas a dimensionless energy.

Our main results are a series of simple thermodynamicequalities and inequalities for how the steady state re-sponds to perturbations of the Ej , Bij and Fij . By com-bining these results, we can constrain the response to anyarbitrary perturbation of the rates through a decompo-sition of the form

∂πk∂λ

=∑j

∂Ej∂λ

∂πk∂Ej

+∑i>j

∂Bij∂λ

∂πk∂Bij

+∑i>j

∂Fij∂λ

∂πk∂Fij

(12)The freedom in the rate parameterization makes this de-composition non-unique, and the tightness of our inequal-ities will depend on the specific decomposition. For ex-ample, one choice could force ∂λEi = 0 for all i. Weare not presently aware of a good strategy to identify thedecomposition which yields optimally tight inequalities.In this work, we show that simple decompositions cannevertheless yield interesting bounds.

In deriving our response results, the basic mathemat-ical tool we rely on is the matrix-tree theorem (MTT),presented in Appendix A, which gives an exact algebraicexpression for the steady-state probabilities πi in termsof the rates Wij [29, 32]. All our results—presented inthe following sections—are obtained by reasoning aboutthe result of differentiating the expression given by theMTT. Proofs are given in the appendices.

Page 4: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

4

V. VERTEX PERTURBATIONS

Our first main result is the exact expression for theresponse to a vertex perturbation (Appendix B)

∂πi∂Ek

=

{−πi(1− πi) if i = k

πiπk if i 6= k. (13)

We stress that the Bij and Fij are unrestricted, so thisequality holds even for nonequilibrium steady states. Asan immediate consequence, we also find that for i 6= j,

∂ ln (πi/πj)

∂Ek=

−1 if i = k

1 if j = k

0 otherwise

, (14)

which implies that the relative probability between twostates is insensitive to the vertex parameters elsewherein the graph.

Remarkably, these equalities are exactly equivalentto the response of a Boltzmann distribution to energyperturbations, which leads to the surprising conclusionthat far-from-equilibrium response has an equilibrium-like structure if the perturbation leaves the Bij and Fijfixed. To leverage this observation, let us assume thatwe vary the rates only through the system’s energy func-tion εi(λ) and that the rates depend on the energy asWij = ωije

βεj , with arbitrary energy-independent ωij .Comparison with (6), shows that variations in the energyεi in this case can be parameterized as vertex parametersEi. Then (13) implies that arbitrarily far from equilib-rium the response maintains the equilibrium-like form ofthe FDT,

∂λ〈Q〉π = βCπ(Q,V ), (15)

with the response proportional to the nonequilibriumsteady-state correlation with the coordinate conjugate tothe energy V = −∂λε [cf. (4) and (5)]. This predictionimplies that experimental verification of the static FDTis not sufficient to conclude that a system is in equilib-rium.

VI. SYMMETRIC EDGE PERTURBATIONS

More generally, a perturbation will modify not onlythe vertex parameters Ei, but also the edge parametersBij . While at equilibrium the steady state is independentof edge parameters Bij , this is generically not the caseout of equilibrium. In this section, we demonstrate thatin fact response to edge perturbations is constrained bythermodynamic affinities through the cycle forces. Forproofs of the results in this section, see Appendix C.

A. Single edge

Our second main result is that the response to theperturbation of a symmetric edge parameter Bmn, asso-

ciated to a single edge emn, is constrained by the cycleforces: ∣∣∣∣ ∂πi∂Bmn

∣∣∣∣ ≤ πi(1− πi) tanh (Fmax/4) (16)∣∣∣∣∂(πi/πj)

∂Bmn

∣∣∣∣ ≤ (πiπj)

tanh (Fmax/4) , (17)

where

Fmax = maxC3emn

|FC | (18)

is the maximum cycle force over all cycles that con-tain the (undirected) perturbed edge emn (illustrated inFig. 3). If the cycle forces all equal zero—as they must atequilibrium—then Fmax = 0, and the response is zero, asexpected. In addition, only perturbations of an edge con-tained in a cycle can induce a response: perturbations ofedges whose removal would disconnect G therefore can-not alter the steady state. Equation (16), furthermore,has the character of the FDT, once we recognize πi(1−πi)as the variance of the occupation fluctuations of statei; thus, we see a manifestation of how thermodynamicsshapes the interplay between response and fluctuations.These inequalities, applying to all discrete stochastic dy-namics, significantly generalize a bound for two-state sys-tems derived by Hartich et al. in a model of nonequilib-rium sensing [33].

The conditions for equality in (16) and (17) would sug-gest methods for designing optimized or highly respon-sive devices. As detailed in Appendix E, we can exhibitat least one scenario that does saturate (17): a single cy-cle with strong time-scale separation so that the systemeffectively only has two states. This limiting scenariosuggests that small single cycle systems are ideal for op-timizing response. Systematically deducing the systemparameters that saturate our inequalities in general re-mains for future work.

B. Multiple edges

The response to a perturbation of multiple edge param-eters can be bounded by combining (17) with the triangleinequality. For example, for any set S of |S| edges,∣∣∣∣∣ ∑

emn∈S

∂ ln(πi/πj)

∂Bmn

∣∣∣∣∣ ≤ ∑emn∈S

∣∣∣∣∂ ln(πi/πj)

∂Bmn

∣∣∣∣≤ |S| tanh(Fmax/4).

(19)

It is clear, however, that this inequality is not always thebest we can do. Consider for example the case where Sconsists of every edge in G. In this case, increasing allthe edge parameters by the same amount (which is whatthe sum above amounts to) is like rescaling time, whichcannot affect π and therefore has zero response.

In this section, we provide a different bound on theresponse to a perturbation of multiple edge parameters

Page 5: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

5

FIG. 3. Thermodynamics and topology bound re-sponse. (a) Transition graph for our representative 4-statesystem with a single perturbed edge connecting states 1 and2, e12, highlighted in blue. (b) Only cycle forces FC aroundcycles that contain the perturbed edge, e12, highlighted inblue, constrain the static response. (c) The maximum re-sponse maxj |∂πj/∂B12| to the perturbation of the edge pa-rameter B12 as a function of maximum cycle force aroundcycles containing e12 for 15000 randomly sampled rate matri-ces (grey dots). All samples fall below the predicted bound(1/4) tanh(Fmax/4) (red line).

that in many cases improves on (19). Suppose we varythe edge parameters associated to the edges S of a sub-graph H 6= G. Let W be the set of vertices of H thatconnect it to the rest of the graph (i.e. the set of verticesof H incident to an edge not in H). Then,

∣∣∣∣∣ ∑emn∈S

∂Bmnln

(πiπj

)∣∣∣∣∣ ≤ (|W | − 1) tanh

(Fi↔j

4

).

(20)where Fi↔j , defined precisely in Appendix C, can bephysically identified as the largest possible entropy pro-duced in the environment when the system goes from i toj and back again (along paths without self-intersection).Whenever there is only one path through state space be-tween i and j, and in all cases at thermodynamic equi-librium, Fi↔j = 0.

We finally note that our results (16), (17), and (20)admit generalization to the response of a ratio of positiveobservables 〈A〉/〈B〉. In this general case, the boundsremain unchanged, except that Fi↔j is replaced by itsmaximum over all pairs of vertices i, j.

VII. ASYMMETRIC EDGE PERTURBATIONS

Lastly, the MTT allows us to bound the response toasymmetric edge perturbations as (Appendix F)∣∣∣∣ ∂πi∂Fmn

∣∣∣∣ ≤ πi(1− πi) ≤ 1

4, (21)

which is related to, but distinct from, inequalities estab-lished in [34]. This result is a consequence of an iden-tical inequality that holds for general rate perturbations∂πi/∂ lnWjk.

Any perturbation of rates can be decomposed into alinear combination of perturbations of the vertex andedge parameters Ei, Bij , and Fij we have introduced,and so the response can be bounded using our inequal-ities (via the triangle inequality). What our results inthis section show is that even for a general perturbations,there is a universal bound—the response to the variationof a single rate is bounded by a constant independent ofthe structure of G or rates of other transitions, mean-ing that high sensitivity always requires many differenttransitions to be perturbed, their cumulative effect gen-erating a response that can greatly exceed 1/4.

VIII. BIOCHEMICAL APPLICATIONS

In this section, we illustrate the use of our main resultsby detailing applications to well-studied motifs appearingin biochemical networks.

A. Covalent modification cycle

First, we consider a well-studied model [35–38] of abiological switch—the modification/demodification cycledepicted in Fig. 4, also known as the Goldbeter–Koshlandloop [35, 37], or “push–pull” network [39].

The network consists of a substrate with two forms, an“unmodified” S and “modified” S∗, along with enzymes,E1 and E2, that actively catalyze its modification anddemodification, respectively. For example, if E1 is a ki-nase, E2 a phosphatase, and S∗ a singly-phosphorylatedform of S, then the system is driven by the chemicalpotential gradient ∆µ = µATP − µADP − µPi for ATPhydrolysis. In the limit in which the substrate is veryabundant compared to its modifying enzymes, it is wellknown that such a system can exhibit unlimited sensi-tivity to changes in the ratio of the concentrations of themodifying and demodifying enzymes [35].

In the other limit—that of a single substratemolecule—our results (17) limit the sensitivity of the ra-tio πS∗/πS for a particular substrate molecule to changesin the enzyme concentration (Appendix G):

s =

∣∣∣∣∂ ln (πS∗/πS)

∂ ln[E1]

∣∣∣∣ ≤ tanh(∆µ/4kBT ), (22)

Page 6: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

6

0.00.10.20.30.40.50.6

FIG. 4. Sensitivity of a biochemical switch. Modi-fication/demodification cycle with (a) a single intermediateor (b) general enzymology. Perturbations of [E1]-dependentrates (red arrows) can be parameterized by blue vertex andedge perturbations. (c) Single-cycle switch-like behavior ofln(πS∗/πS) as a function of ln([E1]), with redder curves rep-resenting larger chemical driving ∆µ. (d) Sensitivity s =|∂ ln(πS∗/πS)/∂ ln([E1])| as a function of ln([E1]) for samevalues of ∆µ, compared to predicted bound (22) (dashedlines).

where Fmax = ∆µ/kBT is the single chemical drivingforce. For the simple cycle in Fig. 4(a) where each enzymehas a single intermediate and we assume mass-action ki-netics, this result arises from unraveling a change in [E1]as a change in the vertex parameter associated to E1S,together with changes in the parameters of the edges con-necting E1S to S and E1S to S∗.

In fact, inequality (22) turns out to hold under as-sumptions much more general than those of Fig. 4(a).It remains true even if catalysis by E1 and E2 proceedsvia any number of intermediate complexes with arbitraryrates as in Fig. 4(b), as long as there is no irreversibleformation of a dead-end complex and the chemical driv-ing is the same around every cycle in which E1 makesthe modification of S and E2 removes it [37, 38]. In thisgeneral case, the many perturbed vertex and edge param-eters (Fig. 4(b)) form a subgraph that acts effectively asa single edge perturbation. Our multi-edge bound (20)then applies with FS↔S∗ = ∆µ/kBT being the maxi-mum entropy produced to go from the unmodified S tothe modified S∗ form and back again.

In the absence of nonequilibrium drive (∆µ = 0), itis clear this switch cannot work, because it operates byvarying the kinetics via an enzyme concentration, and atequilibrium the steady state is independent of kinetics.It has long been known that switches require energy [36,40, 41]. Our results provide a general quantification ofthis requirement.

B. Biochemical sensing

The covalent modification cycle, discussed in the pre-vious section, is an integral component of numerous bio-chemical models for cellular sensing [5, 39, 42–44]. So far,we have described a single substrate molecule stochasti-cally switching between its two forms due to the actionof abundant enzymes E1 and E2. Here, we imagine thereare N substrate molecules, which act as N independentcopies of the system studied above, as long as the num-bers of both enzymes greatly exceeds N . Then the num-ber s of modified substrate molecules S∗ can be inter-preted as a noisy readout of the enzyme concentration[E1]. The random variable s is binomially distributed,with mean µs = NπS∗ and variance σ2

s = NπS∗(1−πS∗),which implicitly depend on [E1]. Thus, this scenario pro-vides a mechanism to measure a chemical concentration,by exploiting the relation between s and [E1].

Now suppose one makes the observation at some timethat there are s molecules of S∗. Supposing [E2] and allrate constants assume known, fixed values, one can pro-

duce an estimate ˆ[E1] of [E1] by choosing ˆ[E1] to be the

value of [E1] that gives µs( ˆ[E1]) = s. The variance of the

estimate ˆ[E1] so constructed is often well-approximated,when the noise is small, by [5, 39]

σ2e ≈

σ2s

(∂µs/∂[E1])2, (23)

where the quantities on the right hand side should beevaluated at the true concentration [E1]. Our result (16)combined with the probabilistic inequality πS∗(1−πS∗) ≤1/4 then leads to the following bound on the relative error(

σe[E1]

)2

&4

N tanh(∆µ/4kBT )2. (24)

This result interpolates between bounds on error estab-lished by Govern and ten Wolde [5, 39] in two limits of re-source limitation in cellular sensing systems. That workstudies a model of sensing in which cell surface recep-tors bind to an extracellular ligand whose concentrationthe cell needs to determine. The ligand-bound receptorsthen participate in a modification/demodification cyclelike the one we study here, playing the role of E1. SeeAppendix G 2 for details.

C. Kinetic proofreading

As a third application, we turn to the effectiveness ofkinetic proofreading [45, 46]. A common challenge facedby biomolecular processes is that of discriminating be-tween two very similar chemical species. At equilibrium,the probability of an enzyme E being bound to a sub-strate S, divided by the probability of that enzyme be-ing free is exp(−∆), where kBT∆ is the binding (free)

Page 7: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

7

FIG. 5. Bounding the discriminatory index. (a) Graphof a single-cycle kinetic proofreading network. Perturbationsin the (dimensionless) binding energy ∆ can be unraveledas blue vertex and edge perturbations. (b) Discriminatoryindex ν plotted against the thermodynamic affinity |∆µ|/kBTfor the single-cycle network generated from 30000 randomlysampled transition rates. All samples fall within the predictedbound (26) (red line).

energy of the enzyme-substrate complex ES. Two sub-strates with very similar binding energies are constrainedto be bound by the enzyme a similar fraction of time.

Kinetic proofreading is a scheme to use nonequilibriumdriving to improve discrimination based on binding en-ergy. One way to quantify the discriminatory ability of akinetic network is using the discriminatory index intro-duced by Murugan et al. [47],

ν = −∂ ln(πE/πES)

∂∆. (25)

At equilibrium, ν = 1. The simplest nonequilibriumscheme to improve on this is the single-cycle network il-lustrated in Fig. 5(a). Note that we have supposed thebinding energy ∆ appears exclusively in the unbindingrates. Hopfield observed that in a certain nonequilib-rium limit of the rates, ν → 2 [45]. Our results lead to aconstraint on ν that interpolates between the equilibriumcase and this limit.

In the single-cycle network, the variation of the bindingenergy ∆ is equivalent to the variation of two vertex pa-rameters (that of ES and ES∗) and a single edge param-eter (ES ↔ ES∗), leading to the inequality (AppendixG),

|ν − 1| ≤ tanh(∆µ/4kBT ), (26)

where Fmax = ∆µ/kBT is the chemical driving aroundthe cycle. This bound, which can be saturated, reducescorrectly to ν = 1 at equilibrium and is consistent withν → 2 in the limit of strong driving ∆µ→∞.

We can also bound ν in the case of a more general ki-netic proofreading scheme [47, 48] in which there are mcomplexes that can dissociate. Each of the dissociationtransitions can be thought of as crossing a “discrimina-tory fence” [47], its rate depending on the binding energy∆, as in Fig. 6. We suppose that the dissociation transi-tions are the only ones that depend on ∆. We make noassumptions about the structure of the transition graphon either side of the fence. In such a network, perturbing

ESkmeΔ

k1eΔE

*

**

FIG. 6. Multi-step kinetic proofreading scheme. Thecollection of edges with rates that depend on the binding en-ergy ∆ specify a “discriminatory fence”. Perturbing ∆ isequivalent to perturbing vertex and edge parameters of thesubgraph labeled in blue.

∆ is equivalent to perturbing the edge and vertex pa-rameters on one side of the “fence” forming the subgraphhighlighted in blue in Fig. 6. We then have (AppendixG)

|ν − 1| ≤ (m− 1) tanh(FE↔ES/4), (27)

where FE↔ES is the maximum entropy produced to gofrom E to ES and back again. Notably, we recover theresult |ν − 1| ≤ m− 1 of [47].

IX. CONCLUSION

In this work, we have developed a series of univer-sal bounds on nonequilibrium response in terms of thestrength of the nonequilibrium driving. We show thatfor a large class of static perturbations, a result equiv-alent to the FDT continues to hold out of equilibrium.For many other perturbations, we bound the response interms of the dimensionless thermodynamic forces, whichquantify departure from equilibrium.

The illustrations detailed in the previous sectiondemonstrate the potential of our results to unify long-standing observations about the importance of energy“expenditure” in many different models. The tasks ofmaking a sharp molecular switch, a good sensor, or dis-criminating between two similar ligands, all have in com-mon the need for a large response to a small perturbation.We find new bounds interpolating between known limitsin these systems, and show how they all descend fromour results on vertex and edge perturbations.

A more detailed analysis of the conditions under whichour bounds are saturated would lead to design principlesfor optimal response. Our preliminary investigation iden-tified single cycles as ideal when a single edge parameteris varied. We expect that for more complex perturba-tions, the most highly responsive systems may have amore complicated structure.

An important theme highlighted by our work is thatsensitivity is limited not only by nonequilibrium driving,but also, very strictly, by network size and structure. Thetotal number of transitions in a biochemical network lim-its response, because the response to the scaling of any

Page 8: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

8

one rate is bounded by 1/4. At the same time, our multi-edge results show how many enlargements or complica-tions of networks (e.g. departure from Michaelis-Mentenassumptions in the covalent modification cycle), do notconfer any advantage. In this sense, our results build onthe work of others who studied similar questions in thecontext of kinetic proofreading [47] and biochemical copyprocesses [49].

Our results point to numerous other extensions, includ-ing bounds on the response of currents with implicationsfor the Green-Kubo and Einstein relations [50–52]. Wehave also focused on results that hold in general, nottaking into account possible characteristic structures inthe graph of states and transitions, which are presentfor example in many natural examples, such as chemi-cal reaction networks. The study of such extensions andspecial cases strike us as promising directions for futurework.

ACKNOWLEDGMENTS

We would like to thank Alexandre Solon and MatteoPolettini for very useful discussions. JAO would like tothank Jeremy England for advice and support.

Appendix A: Matrix-tree theorem

The key tool that we apply in our analysis of nonequi-librium response is the matrix-tree theorem (MTT). Tostate the theorem, we must introduce some additionalnotation and concepts.

For any set of directed edges S = {i → j, k → l, . . . },we define the weight w(S) to be the product of the weightsof the edges,

w(S) = WjiWlk · · · . (A1)

The weight w(H) of any subgraph H we define to be theweight of its edge set.

We also need to introduce spanning trees, which areconnected subgraphs of a graph G that contain everyvertex, but have no cycles, see Fig. 7. Every graph that

FIG. 7. Spanning trees. All spanning trees for our 4-stateillustrative graph.

is connected (as is, by assumption, the transition graph

of our system) has at least one spanning tree. For anyspanning tree T and vertex r of G, there is a unique wayto direct the edges of T so that they all “point towards”r, which we then call the “root”. The resulting directedgraph, which we write Tr, is a rooted spanning tree ofG. The steady-state distribution π is given explicitly bythe matrix-tree theorem (MTT) [29, 53–57] in terms ofweights of rooted spanning trees of G.

Theorem (matrix-tree theorem). Let W be the tran-sition rate matrix of an irreducible continuous-timeMarkov chain with N states. Then the unique steady-state distribution is

πk =1

N∑

spanning treesT of G

w(Tk), (A2)

where N =∑Nk=1

∑T w(Tk) is the normalization con-

stant.

This theorem, also known as the Markov chain tree the-orem, is a consequence of a result of Tutte [53], and hasbeen rediscovered repeatedly in different literatures, seee.g. [29, 54–56] and [57] for further discussion.

The MTT offers a graphical representation of thesteady-state distribution that provides a convenientmethod for organizing the structure of the solution. Weillustrate this result in Fig. 8.

FIG. 8. Matrix-tree theorem. Graphical representationof the steady-state probability π1 as the sum of all spanningtrees rooted at 1 (blue vertex).

Appendix B: Vertex perturbations

Theorem 1.

∂πi∂Ek

=

{−πk(1− πk) if i = k

πkπi if i 6= k. (B1)

Proof. The matrix-tree theorem implies that πi can beexpressed as the ratio of sums of weights of rooted span-ning trees. So to evaluate ∂πi/∂Ek, we need to under-stand in which spanning trees, and in what form, Ekappears. The only rates that depend on Ek are rates oftransitions out of k, W∗k = exp(Ek − B∗k + F∗k/2), seeFig. 2. Furthermore, any rooted spanning tree has ex-actly one edge directed out of k, unless the tree is rootedat k, in which case it has none. These observations allowus to group spanning trees in the MTT expression forthe steady-state distribution in a convenient manner asillustrated in Fig 9.

Page 9: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

9

Thus, for i = k, the matrix-tree theorem implies that

πk =a

a+ beEk, (B2)

where

a =∑T

w(Tk), beEk =∑j 6=k

∑T

w(Tj). (B3)

Here, a is the sum of weights of all spanning trees rootedat k—these do not depend on Ek since they have no edgedirected out of k—and beEk is the sum of weights of allspanning trees not rooted at k—each of these has exactlyone factor of Ek, making b independent of Ek.

If i 6= k, the MTT yields by a similar argument

πi =ceEk

a+ beEk, (B4)

with

ceEk =∑T

w(Ti). (B5)

The theorem now follows by differentiating these ex-pressions. When i 6= k,

∂πi∂Ek

=ceEk

a+ beEk− ceEkbeEk

(a+ beEk)2

=ceEk

a+ beEk

(a

a+ beEk

)= πiπk,

(B6)

and similarly for i = k.

Corollary 1. If i 6= j,

∂ ln (πi/πj)

∂Ek=

−1 if i = k

1 if j = k, and

0 otherwise.

(B7)

Proof. First, note that

∂ ln (πi/πj)

∂Ek=

1

πi

∂πi∂Ek

− 1

πj

∂πj∂Ek

. (B8)

Now we apply Theorem 1. If i = k, then j 6= k, and∂ ln(πi/πj)

∂Ek= −(1 − πk) − πk = −1. If j = k, then i 6= k,

and∂ ln(πi/πj)

∂Ek= πk + (1− πk) = 1. And if neither i nor

j equal k, then∂ ln(πi/πj)

∂Ek= πk − πk = 0.

Appendix C: Symmetric edge perturbations

1. Single edge

In this appendix, we bound the response to the per-turbation of a single symmetric edge parameter in termsof the cycle forces driving the system out of equilibrium.

First, we prove a general bound on the response of aratio of observables. Equations (16) and (17) will thenfollow as corollaries by choosing suitable observables.

Theorem 2. Consider any two observables A,B ∈ RN≥0with at least one positive entry. Then,∣∣∣∣ ∂

∂Bmnln〈A〉〈B〉

∣∣∣∣ ≤ tanh

(Fmax

4

)(C1)

where Fmax is the magnitude of the cycle force that islargest in magnitude, among all those associated to cy-cles containing the distinguished edge m ↔ n (in eitherdirection).

Our proof relies on the following technical lemma,which we prove in Appendix D.

Lemma 1 (“Tree surgery”). Let Emn be the set of span-ning trees of G containing the distinguished (undirected)edge m ↔ n. Then for any two distinct vertices i, j ofG, ∣∣∣∣∣

∑T∈Emn

∑S/∈Emn

w(Ti)w(Sj)∑T∈Emn

∑S/∈Emn

w(Tj)w(Si)

∣∣∣∣∣ ≤ exp(Fmax). (C2)

Proof of Theorem 2. The matrix-tree theorem offers agraphical representation of the steady-state distributionin terms of rooted spanning trees. This observation sug-gests that we can segregate those contributions to steady-state averages that contain Bmn by selecting those (undi-rected) spanning trees in G that contain the edge emn.Let us call this set Emn.

Then by the matrix-tree theorem, we can write

〈A〉〈B〉

=

∑iAiπi∑j Bjπj

=a1 + a0b1 + b0

(C3)

where

a1 =∑i

∑T∈Emn

Aiw(Ti) a0 =∑i

∑S/∈Emn

Aiw(Si)

(C4)

b1 =∑i

∑T∈Emn

Biw(Ti) b0 =∑i

∑S/∈Emn

Biw(Si),

(C5)

where a1 and b1 are linear in exp(−Bmn), since they con-tain edge emn, whereas a0 and b0 are independent ofBmn.An illustrative example is presented in Fig. 10.

This implies

∂Bmnln〈A〉〈B〉

=b0a1 − a0b1

(b0 + b1)(a0 + a1). (C6)

Now note that by the AM-GM inequality the denomina-tor is bounded as

(b0 + b1)(a0 + a1) = b0a0 + b1a0 + b1a1 + b0a1

≥ b0a1 + b1b0 + 2√a0b0a1b1 =

(√b0a1 +

√a0b1

)2.

(C7)

Page 10: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

10

FIG. 9. Vertex perturbation. Groupings of spanning trees with roots labeled in blue for π1 (i = 1) utilized in the vertexperturbation derivation when the perturbed vertex is k = 4, labeled in red. Affected rates labeled in red.

FIG. 10. Symmetric edge perturbation. Groupings ofspanning trees used in the derivation of the symmetric edgeperturbation bound for the ratio of observables π1/π4 withperturbed edge parameter B12. Roots labeled in blue andaffected rates highlighted in red.

Since the numerator b0a1 − a0b1 =(√b0a1 −

√a0b1

) (√b0a1 +

√a0b1

), the bound (C7)

implies,∣∣∣∣ ∂

∂Bmnln〈A〉〈B〉

∣∣∣∣ ≤ ∣∣∣∣√b0a1 −√a0b1√b0a1 +

√a0b1

∣∣∣∣ = tanh

(1

4

∣∣∣∣ln b0a1a0b1

∣∣∣∣) .(C8)

To complete the proof, we need to bound the ratiob0a1/a0b1 by exp (Fmax). To do this, we match up termsabove and below, writing the fraction as

b0a1a0b1

=

∑i

∑j

(AiBj

∑T∈Emn

∑S/∈Emn

w(Ti)w(Sj))∑

i

∑j

(AiBj

∑T∈Emn

∑S/∈Emn

w(Si)w(Tj)) .

(C9)The desired result is now a consequence of the inequality∑n

i=1 xi∑ni=1 yi

=

∑ni=1 (xi/yi) yi∑n

i=1 yi≤ max

i

(xiyi

), (C10)

to give∣∣∣∣b0a1a0b1

∣∣∣∣ ≤ maxi,j

∣∣∣∣∣∑T∈Emn

∑S/∈Emn

w(Ti)w(Sj)∑T∈Emn

∑S/∈Emn

w(Si)w(Tj)

∣∣∣∣∣ , (C11)

followed by Lemma 1.

From Theorem 2 we readily obtain our bounds onsteady-state response. For (17):

Corollary 2.∣∣∣∣∂ ln (πi/πj)

∂Bmn

∣∣∣∣ ≤ tanh

(Fmax

4

)(C12)

Proof. Choose the observables in Theorem 2 to be Al =δil and Bl = δkl, where δij is the Kronecker delta.

We also have:

Corollary 3. Let be πX =∑k∈X πk be the total proba-

bility of a set of states X. Then,∣∣∣∣ ∂πX∂Bmn

∣∣∣∣ ≤ πX(1− πX) tanh

(Fmax

4

). (C13)

Proof. Choose the observables in Theorem 2 to be Ai =δi(X) and Bi = 1−δi(X), where the indicator δi(X) = 1if i ∈ X and δi(X) = 0 otherwise. Note that we thenhave 〈A〉 = πX and 〈B〉 = 1− πX .

If X = {i} consists of only a single state we recover thebound (16).

2. Multiple edges

In this section, we derive our inequality for the re-sponse to perturbations by multiple edge parameters(20). The proof proceeds in two steps. We first provea bound on an arbitrary set of edges S from which (20)and other results are ready corollaries.

Here, the magnitude of response is bounded by a dif-ferent function Fi↔j of cycle forces. The quantity Fi↔jis defined for any graph G and vertices i and j to be

Fi↔j = maxPi→j ,Pj→i

∣∣∣∣∣ln w(Pi→j ∪ Pj→i)w(P ∗i→j ∪ P ∗j→i)

∣∣∣∣∣ . (C14)

where Pi→j is a (non-self-intersecting) path from i to j,Pj→i is a (non-self-intersecting) path from j to i, and thesuperscript ‘∗’ denotes the reverse path.

Page 11: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

11

i

j

i

j00

FIG. 11. Illustration of the identity (C18).

Theorem 3. Let S be a set of edges, and define cmax

to be the size of the largest intersection S has with anyspanning tree of G. Similarly, define cmin to be the sizeof the smallest such intersection. Then,∣∣∣∣∣ ∑emn∈S

∂Bmnln

(πiπj

)∣∣∣∣∣ ≤ (cmax − cmin) tanh

(Fi↔j

4

).

(C15)

The appearance of Fi↔j in this result stems from thislemma, that we rely on here and prove in Appendix D.

Lemma 2 (“Cycle flip only”). For any spanning treesT, S and vertices i, j of G,

w(Ti)w(Sj)

w(Tj)w(Si)≤ exp(Fi↔j). (C16)

We will also rely on the following lemma, which gen-eralizes the first part of the proof of Theorem 2.

Lemma 3. For any symbols {an}, {bn},∣∣∣∣∣∣(∑j

n=i nan

)∑jn=i bn −

(∑jn=i nbn

)∑jn=i an∑j

n=i an∑jn=i bn

∣∣∣∣∣∣≤

j∑m=i+1

tanh

(1

4ln

∣∣∣∣∣∑jn=m an

∑m−1n=i bn∑j

n=m bn∑m−1n=i an

∣∣∣∣∣)

(C17)

Proof. First note we can rearrange the sum as

j∑n=i

nan =

j∑n=i

n∑m=1

an = i

j∑n=i

an +

j∑m=i+1

j∑n=m

an, (C18)

which is illustrated in Fig. 11. As a result, we have

(∑jn=i nan

)∑jn=i bn −

(∑jn=i nbn

)∑jn=i an∑j

n=i an∑jn=i bn

=

j∑m=i+1

(∑jn=m an

)∑jn=i bn −

(∑jn=m bn

)∑jn=i an∑j

n=i an∑jn=i bn

=

j∑m=i+1

(∑jn=m an

)∑m−1n=i bn −

(∑jn=m bn

)∑m−1n=i an

(∑m−1n=i an +

∑jn=m an)(

∑m−1n=i bn +

∑jn=m bn)

.

(C19)

Comparison with equations (C6) through (C8) in theproof of Theorem 2, together with the triangle inequality,establishes the desired result.

Now we are ready to proceed with the proof of thetheorem.

Proof of Theorem 3. Define

ac =∑

T :|T∩S|=c

w(Ti), bc =∑

T :|T∩S|=c

w(Tj). (C20)

so that we have, for all c∑emn∈S

∂ac∂Bmn

= cac ,∑

emn∈S

∂bc∂Bmn

= cbc. (C21)

By the matrix-tree theorem, the derivative we wish tobound can be written in terms of these quantities as∑emn∈S

∂Bmnln

(πiπj

)=

∑emn∈S

∂Bmnln

∑T w(Ti)∑T w(Tj)

=∑

emn∈S

∂Bmnln

∑cmax

c=cminac∑cmax

c=cminbc.

(C22)

Expanding the derivative and applying Lemma 3 yields∣∣∣∣∣ ∑emn∈S

∂Bmnln

(πiπj

)∣∣∣∣∣≤

cmax∑m=cmin+1

tanh

(1

4ln

∣∣∣∣∣∑cmax

n=m an∑m−1n=cmin

bn∑cmax

n=m bn∑m−1n=cmin

an

∣∣∣∣∣).

(C23)

To prove the theorem, all that remains is to demon-strate that∑cmax

n=m an∑m−1n=cmin

bn∑cmax

n=m bn∑m−1n=cmin

an≤ exp(Fi↔j) (C24)

holds for all m. This follows by an application of Lemma2. So we have∣∣∣∣∣ ∑

emn∈S

∂Bmnlnπiπj

∣∣∣∣∣ ≤cmax∑

m=cmin+1

tanh

(Fi↔j

4

)= (cmax − cmin) tanh

(Fi↔j

4

)(C25)

Page 12: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

12

as desired.

Theorem 3 has a number of simple corollaries.

Corollary 4. If G has r independent cycles, then forany set S of edges,∣∣∣∣∣ ∑

emn∈S

∂Bmnln

(πiπj

)∣∣∣∣∣ ≤ r tanh

(Fi↔j

4

). (C26)

Proof. Let m be the number of edges in G. The largestpossible intersection of a spanning tree and S cannot ex-ceed |S| in size, so we have cmax ≤ |S|. Furthermore,each spanning tree of G has exactly m − r edges. Sothe smallest possible intersection is realized if all r edgesa spanning tree excludes are edges in the set S, whichmeans cmin ≥ |S| − r. Therefore, cmax − cmin ≤ r, andthe corollary follows from Theorem 3.

Corollary 5. Let H be a subgraph of G, and write Wfor the set of vertices of H incident to an edge not in H.Let S be the edge set of H. Then,∣∣∣∣∣ ∑

emn∈S

∂Bmnln

(πiπj

)∣∣∣∣∣ ≤ (|W | − 1) tanh

(Fi↔j

4

).

(C27)

Proof. Consider a spanning tree T of G. Viewed as asubgraph of H, T is still at least a spanning forest (i.e. itmay no longer be connected, but still has no cycles andincludes every vertex of H), with no more than |W | com-ponent trees. To see this, suppose it had |W | + 1 com-ponent trees. In this case, one component would have tobe disconnected from all the vertices in W (if every com-ponent is connected to a vertex in |W |, there can be atmost |W |, as components cannot be connected to eachother). But in that case, T (as a subgraph of G) wasdisconnected—it was never a spanning tree at all.

Let n be the number of vertices in H. The numberof edges in a spanning forest is always the number ofvertices in the forest minus the number of components(trees in the forest). This means that for our graph G,the size of the intersection of S and the edge set of T isrestricted to lie between n− 1 = cmax or n− |W | = cmin.By Theorem 3, this implies the result.

Appendix D: Proofs of the root-swapping lemmas

In the course of proving our results above we cameacross ratios of products of spanning tree weights, suchas ∑

T∈Emn

∑S/∈Emn

w(Ti)w(Sj)∑T∈Emn

∑S/∈Emn

w(Tj)w(Si), (D1)

which we bounded using Lemmas 1 and 2, yielding ourtheorems. Here, we present proofs of these key lemmas.The arguments will depend on the existence of invertible

mappings between the pairs of spanning trees in the nu-merator to pairs of spanning trees in the denominator,which have their roots “swapped”: (Ti, Sj) → (T ′j , S

′i).

We will construct these mappings explicitly, but first weset out some relevant notation and definitions.

First, we will find it helpful in this section to use thestandard notation s(e) (the source) for the vertex at thetail of a directed edge e and t(e) (the target) for thevertex at the head of e, where the arrow points. In ad-dition, the graph formed by the removal of the edge hfrom a graph H, i.e. by the deletion of h, will be denotedH \ h, and the graph formed by adding an edge h to Hwill be denoted H ∪ h.

Second, we need to define a new kind of spanning tree.We have already introduced spanning trees, as well asthe notion of a spanning tree Ti rooted at vertex i. Re-call that in a rooted spanning tree, every edge is directedtowards the root i (since a tree has no cycles, this direc-tion is defined unambiguously). Generalizing from this,we define a doubly-rooted spanning tree, schematicallydepicted in Fig. 12(a). We start with a spanning tree Sand two vertices i and j. We first note that all the edgesin the rooted trees Si and Sj are oriented in the samedirection except for those edges along the unique pathbetween i and j. This inspires us to pick a vertex k onthis path and define a doubly-rooted spanning tree Sij,kwith branch point k to be the spanning tree S with everyedge directed as it is in Si and Sj—when those direc-tions are the same—and otherwise directed towards i ifbetween k and i, and towards j if between k and j. Onecan view a (singly) rooted spanning tree Sj as a sort of“degenerate” doubly-rooted tree Sij,i with branch pointi.

Our mappings are then built from repeated applica-tions of the following operations on pairs of the form(Tb, Smn,b), where Tb is a spanning tree rooted at somevertex b, and Smn,b is a doubly-rooted spanning tree withroots m and n and branch point b:

• Cycle flip. Consider the unique edge e pointingout of b towards n in Smn,b. Reroot the tree Tbto the target t(e) to form Tt(e) and flip the edgee→ e∗ to form Smn,t(e) = (Smn,b \ e)∪ e∗. Output(Tt(e), Smn,t(e)).

• Edge swap. Consider the unique edge e pointingout of b towards n in Smn,b. Let f be first edge,along the unique directed path in Tb from t(e) to b,that reconnects Smn,b\e. Swap these edges to formT ′s(f) = (Tb \ f) ∪ e and S′mn,s(f) = (Smn,b \ e) ∪ f .

Output (T ′s(f), S′mn,s(f)).

The output of each of these operations is another pair(T ′b′ , S

′mn,b′) consisting of a tree rooted at b′ and a doubly-

rooted tree with branch point b′ (see Figure 12 for an il-lustration of this in the case of edge swap). Furthermore,no edges are reoriented in the edge swap, although edgesare exchanged between T and S. In a general cycle flip,no edges are exchanged, and the edges that are reoriented

Page 13: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

13

m n

b e e

f ?

f ?

m n

b

f (a) (b) (c)

m n

b

b0e0

FIG. 12. Effect of edge swap on Smn,b. (a) The structureof Smn,b, a doubly-rooted tree with roots m,n and branchpoint b. The edge e is the unique edge pointing out of s(e) =b towards n. (b) Removing the edge e disconnects Smn,b,since it is a tree. The edge f from Tb that reconnects it canpoint from the component containing n to any part of thecomponent containing m. It is also possible that f = e∗. (c)No matter where f points, removing e and adding f to Smn,b

yields a new doubly-rooted tree S′mn,b′ , with branch point

b′ = s(f).

form the single cycle obtained from the union of e withthe unique path in T from t(e) to s(e). In the degeneratecase where the path in T from t(e) to s(e) consists of thesingle edge e∗, cycle flip and edge flip are equivalent.

Notably, both of these operations are invertible, in thesense that given the output of either, and knowledge ofwhich was applied, we can uniquely recover the originalpair (Tb, Smn,b) from (T ′b′ , S

′mn,b′).

• To invert the cycle flip all we need is to identify theoriginal edge e—it is the reverse of the unique edgepointing out of b′ towards m. Note that s(e) = b,the original branch point of S and root of T .

• To invert the edge swap, we need to identify theoriginal edges e and f . The unique edge pointingout of b′ towards m is f . The original e is thefirst edge, going back along the path in T ′b′ froms(f) = b′ to t(f), that reconnects S′mn,b′ \ f .

Lemma 1 (“Tree surgery”). Let Emn be the set of span-ning trees of G containing the distinguished (undirected)edge m ↔ n. Then for any two distinct vertices i, j ofG, ∣∣∣∣∣

∑T∈Emn

∑S/∈Emn

w(Ti)w(Sj)∑T∈Emn

∑S/∈Emn

w(Tj)w(Si)

∣∣∣∣∣ ≤ exp(Fmax) (D2)

where Fmax is the magnitude of the cycle force that islargest in magnitude, among all those associated to cy-cles containing the distinguished edge m ↔ n (in eitherdirection).

Proof. To prove this result, it is sufficient to find a bi-jection between terms in the numerator and those in thedenominator, such that each term and its partner areequal or differ by a factor of exp(FC), where FC is thecycle force associated to a cycle C that contains the dis-tinguished edge.

Consider any term w(Ti)w(Sj) in the numerator. Wemap it to a term in the denominator as follows. Starting

with the pair (Ti, Sj), viewing Sj as a doubly-rooted treeSij,i, we repeatedly apply edge swap until the root of therooted tree (equivalently the branch point of the doubly-rooted tree) equals j, unless the edge f that would beremoved from Tb in the process is the distinguished edge(m→ n or n→ m). In that case, apply cycle flip in thatstep, so that the distinguished edge is not exchanged.

It is guaranteed that this iterative process will eventu-ally terminate, because at every step, the branch pointof the doubly-rooted tree Sij,b moves closer to j, and thepart of Sij,b rooted at i grows. Eventually, the branchpoint hits j, and the edge swap and cycle flip operationscannot be applied.

At the end of this iterative process the initial pair(Ti, Sj) has been transformed into a pair (T ′j , S

′i), whose

associated weight w(T ′j)w(S′i) appears in the denomina-tor of (D2). This defines a bijection between terms in thenumerator and terms in the denominator. To see that themap is invertible, we note that every step along the way(an application of edge swap or cycle flip) is invertible, aswe argued above. Therefore, as long as it is possible touniquely determine which was applied at each step, thewhole sequence of operations is invertible. But this ispossible, because when inverting a step, we can find theedge f that would have been removed from T by edgeswap in that step, and that determines whether or notedge swap or cycle flip was in fact applied in that step.Namely, cycle flip was applied if f was the distinguishededge, and edge swap was applied otherwise.

Having found this bijection between terms, it re-mains only for us to ask what is the ratio of the termsw(Ti)w(Sj) and w(T ′j)w(S′i)? The operation edge swaphas no effect on this product of weights, since it merelymoves edges between T and S. However, when cycle flipis applied, edges change the way they are directed, andthe weight w(Ti)w(Sj) changes by a factor of exp(FC)where C is the (directed) cycle that gets flipped. Sincewe only apply cycle flip if the path in Tb from t(e) toits root contains m ↔ n, C is always a cycle containingm ↔ n. Furthermore, in the iteration described above,cycle flip is applied at most once. To see this, note thatthe original tree Ti contains either m → n or n → m,never both. Furthermore, the edge f that comes up inedge swap always points from the part of S rooted at jto the part rooted i. Thus, if cycle flip flips the distin-guished edge to point the other way, it will never comeup as f in edge swap again, because the part of S rootedat i only ever grows during this algorithm.

So we have

w(Ti)w(Sj)

w(T ′j)w(S′i)= exp(FC) (D3)

for some cycle C that contains the edge m ↔ n, as de-sired.

To prove the inequality (D2), we now match up termsabove and below using this bijective “tree surgery”,putting them in an order such that each term above hasthe same position as its partner (e.g. image) below. The

Page 14: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

14

i jij

Edge swap

Cycle flip

Edge swap

Edge swap

T S

FIG. 13. Tree surgery. Illustration of the steps of the itera-tive “tree surgery” described in the proof of Lemma 1 appliedto particular pair of rooted spanning trees Ti and Sj of agraph G with 11 vertices. The sequence of edge swap andcycle flip operations applied has the effect of swapping theroots of the trees without swapping the distinguished edge(blue). At intermediate stages, S becomes a doubly-rootedtree whose branch point (labeled in green, also the root of T )moves between its roots i and j (labeled in red). The set ofdirected edges in the final pair of trees differs from the set inthe original pair by the edges in the cycle (which contains thedistinguished edge) flipped in the second step.

Lemma then follows from the inequality:∑ni=1 xi∑ni=1 yi

=

∑ni=1 (xi/yi) yi∑n

i=1 yi≤ max

i

(xiyi

). (D4)

Lemma 2 (“Cycle flip only”). For any spanning treesT, S and vertices i, j of G

w(Ti)w(Sj)

w(Tj)w(Si)≤ exp(Fi↔j) (D5)

where Fi↔j is largest possible value of ln[w(Pi→j ∪Pj→i)/w(P ∗i→j ∪ P ∗j→i)] where Pi→j is a (non-self-intersecting) path from i to j and Pj→i is a (non-self-intersecting) path from j to i, and the superscript ‘ ∗’denotes the reverse orientation.

Proof. As above, we consider the pair (Ti, Sj) but thistime just apply cycle flip to it repeatedly until it can

no longer be applied (because the branch point of S hasbecome j). The effect of these steps is to “swap theroots” of the two trees Ti and Sj , changing the directionsof edges without changing the underlying (undirected)spanning trees. Along any undirected spanning tree T ,there is a unique directed path Tv→w from any vertex vto any other vertex w. “Re-rooting” a tree changes itsweight as follows

w(Tv)w(Tv→w) = w(Tw)w(Tw→v) (D6)

which implies

w(Ti)w(Sj) = w(Tj)w(Si)w(Si→j)w(Tj→i)

w(Sj→i)w(Ti→j). (D7)

The fraction appearing here is of the form w(Pi→j ∪Pj→i)/w(P ∗i→j ∪ P ∗j→i), as required in the statement, es-tablishing the result.

It is important to note that neither Lemma 1 norLemma 2 implies the other, although their proofs canbe viewed as depending on a common technique.

Appendix E: Saturating the inequalities

We have established a number of thermodynamicbounds on steady-state response to edge perturbations.It remains an open question whether we can saturatethese inequalities. In this section, we exhibit one ex-ample where we can saturate our bounds—the case of asystem whose transition graph G consists of a single cy-cle C with cycle force FC = Fmax. While we are unableto prove this is the only way to saturate our inequalities,we do argue for its general relevance.

To keep the discussion as straightforward and preciseas possible, we focus on the ratio bound in (17), as thisturns out to be the simplest to investigate. We first spe-cialize to the case where we vary the edge parameterBmn associated to the edge emn, and ask for the responseof the ratio of steady-state probabilities of the adjacentstates m and n. In this case, the series of inequalitiesthat lead to our bound can be summarized as∣∣∣∣∂ ln (πm/πn)

∂Bmn

∣∣∣∣ =

∣∣∣∣ b0a1 − a0b1(b0 + b1)(a0 + a1)

∣∣∣∣≤

AM-GMtanh

(1

4

∣∣∣∣ln b0a1a0b1

∣∣∣∣)≤

“tree surgery”tanh(FC/4),

(E1)

where we use the notation

a1 =∑

T∈Emn

w(Tm) a0 =∑

S/∈Emn

w(Sm) (E2)

b1 =∑

T∈Emn

w(Tn) b0 =∑

S/∈Emn

w(Sn). (E3)

Page 15: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

15

The first inequality in Eq. (E1) is an application of theAM-GM inequality and the second comes about from our“tree surgery” argument of Lemma 1. We address eachin turn.

Let us begin with the tree surgery inequality, whichcomes about from analyzing the ratio

b0a1a0b1

=

∑T∈Emn

∑S/∈Emn

w(Tm)w(Sn)∑T∈Emn

∑S/∈Emn

w(Tn)w(Sm). (E4)

The tree surgery provides an invertible mapping betweenthe terms in the numerator and those in the denominator.For the case of a single cycle with vertices m,n adjacentto the distinguished edge emn, we have

w(Tm)w(Sn) = w(Tn)w(Sm)eFC (E5)

for all T ∈ Emn and S /∈ Emn. Thus, every term in thenumerator is proportional to a term in the denominatorwith the same proportionality constant:

b0a1a0b1

=

∑Tm∈Emn

∑Sn∈Emn

w(Tn)w(Sm)eFC∑Tn∈Emn

∑Sm∈Emn

w(Tn)w(Sm)= eFC .

(E6)Thus, the “tree surgery” inequality is exactly satisfied inthis case.

Equality in the AM-GM inequality is reached when

a0b0 = a1b1. (E7)

While there are numerous choices for the rates that causethis equality to be satisfied, we will just exhibit a par-ticular one to show that it is possible. To do so, we firstmake a simplifying observation: each term on both sidesof the equality is a product of the weight of a spanningtree rooted at m and one that is rooted at n. Therefore,each term has exactly the same dependence on the vertexparameters Ej , so we can cancel all the Ej on both sidesof (E7). Thus, all we need to do is fix the symmetricand asymmetric edge parameters. We first fix the asym-metric edge parameters by choosing all the weight of thecycle force to be on the perturbed edge emn,

Fkl = −Flk =

{FC k = m, l = n0 else

. (E8)

Solving Eq. (E7) for the symmetric edge parameters thenleads to the relation

eBmn =∑eij∈Gij 6=mn

eBij . (E9)

Thus, it is possible to saturate our inequality for the re-sponse of the ratio ln(πm/πn) to perturbations of Bmn.

This may seem like a rather special case, but we be-lieve the situation is more general than it first appears,since it is possible for the dynamics on more complicatedgraphs G (e.g. with multiple cycles) to effectively havethis “single-cycle” behavior. To see this, note that if the

rates of transitions in G are very small, apart from thosearound a single cycle containing the perturbed edge emn,then the graph is effectively composed of a single cycle,for the purposes of understanding response of the stateson the cycle. In addition, if we look at the response ofratios of arbitrary states on the cycle, such as ln(πi/πj),again the dynamics can effectively reproduce the situa-tion discussed above, where we focused on the verticesadjacent to emn. This is because if the rates along theunique paths from i to m and j to n on the cycle are ex-tremely fast, the states along these paths rapidly reach alocal steady-state distribution. The two paths then actas two “effective states” adjacent to the perturbed edgeemn.

These arguments suggest that, for a general graph G,there are limits of the rates that give rise to response ap-proaching arbitrarily closely the bound set by Corollary2.

Appendix F: Asymmetric edge perturbationinequality

Our asymmetric edge perturbation bound follows froma more general inequality for arbitrary perturbations ofa single rate:

Proposition 1. ∣∣∣∣ ∂πk∂ lnWij

∣∣∣∣ ≤ πk(1− πk) (F1)

Proof. By the matrix-tree theorem, we can write

πk =aWij + b

cWij + d. (F2)

where a, b, c and d are nonnegative quantities formedfrom sums of weights of rooted spanning trees that do notdepend on Wij . By normalization of probability πk ≤ 1,so we have c ≥ a, d ≥ b. Differentiating these expressionsyields

∂πk∂ lnWij

=(ad− bc)Wij

(cWij + d)2 (F3)

which after re-arranging gives∣∣∣∣ ∂πk∂ lnWij

∣∣∣∣= πk(1− πk)

∣∣∣∣ d− b(c− a)Wij + (d− b)

− b

aWij + b

∣∣∣∣(F4)

but the value of each of the two fractions on the righthand side is not smaller than zero or greater than one.This means their difference is not greater than one inmagnitude, implying the result.

Page 16: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

16

Corollary 6. ∣∣∣∣ ∂πk∂Fij

∣∣∣∣ ≤ πk(1− πk) (F5)

Proof. The asymmetric edge parameter Fij appears intwo rates, Wij and Wji. This implies, by the chain rule

∂πk∂Fij

=1

2

(Wij

∂πk∂Wij

−Wji∂πk∂Wji

), (F6)

which implies, by the triangle inequality,∣∣∣∣ ∂πk∂Fij

∣∣∣∣ ≤ 1

2

∣∣∣∣Wij∂πk∂Wij

∣∣∣∣+1

2

∣∣∣∣Wji∂πk∂Wji

∣∣∣∣ . (F7)

Now applying Proposition 1 establishes the desired re-sult.

Appendix G: Biochemical applications

So far, we have stated and proved equalities and in-equalities about the response to perturbations of phys-ical systems whose dynamics are well-modeled as con-tinuous in time and Markovian over a finite state space.In this section, we describe specializations of these gen-eral results to two well-known motifs found in biochem-ical networks. In each case, we find an inequality re-lating some figure of merit to a chemical potential differ-ence driving the network out of equilibrium (for example,∆µ = µATP − µADP − µPi for ATP hydrolysis).

There are several ways that studying a biochemicalnetwork might lead us to consider a linear time evolutionequation like (1),

pi(t) =

N∑j=1

Wijpj(t), (G1)

with∑iWij = 0 for all j. First, the chemical master

equation, which governs the evolution of the distributionover counts (nA, nB , . . . ) of chemical species A,B, . . . , isof this form. However, for chemical systems with manyparticles the number of states N in such a description isenormous.

However, for some chemical reaction networks, the lin-ear equation (1) arises as the rate equation governing thedeterministic evolution of the concentrations of chemicalspecies. As emphasized by Gunawardena [58, 59], this isa generic situation that can arise from strong time-scaleseparation. When the rate equation of a reaction net-work is of the form (1), we can equivalently view it asthe master equation of a continuous-time Markov chaindescribing the stochastic transitions of a single moleculesubject to a set of effectively monomolecular reactions[57, 60]. Whichever interpretation we take, the mathe-matics that arises is the same, and our results can be putto work.

k2[E1]W41 W14

W43

S E1S

E2S S*

k1[E1](a) (b)

1 2

4 3

W12

W21

W34

W23 W32

**

**

** *

*

*

**

**

*

*

*

FIG. 14. (a) The transition graph G arising from theGoldbeter-Koshland model in the low substrate limit, withproduct rebinding. Two transition rates (red) depend onthe (assumed constant) free enzyme concentration [E1] thatwe vary. Scaling [E1] is equivalent to a perturbation of twoedge parameters and one vertex parameter (blue). (b) Statenumbers and rate labels we use in this subsection. Keyequivalences are “1 = S”, “3 = S∗”, “W21 = k1[E1]”, and“W23 = k2[E1]”.

1. Covalent modification cycle

Goldbeter and Koshland [35] studied a model of the co-valent modification and demodification of a substrate bytwo enzymes, assuming the action of both enzymes obeysmass-action kinetics with a single intermediate complexand no product rebinding:

E1 + S ↔ E1S → S∗ + E1

E2 + S∗ ↔ E2S → S + E2. (G2)

The total substrate concentration Stot = [S] + [S∗] +[E1S] + [E2S] is conserved in these reactions, as are theenzyme totals E1,tot = [E1]+[E1S], E2,tot = [E2]+[E2S].In the limit of saturating substrate Stot � E1,tot, E2,tot,the kinetics are effectively Michaelis-Menten in form, andthe steady-state ratio [S∗]/[S] can exhibit unlimited sen-sitivity to changes in E1,tot and E2,tot [35].

Sensitivity of the steady state to changes in enzymeconcentrations is only possible out of equilibrium [40]. In(G2), the nonequilibrium nature of the system is reflectedin the combination of the irreversible product release re-actions with the overall reversibility of the modificationof S.

In the regime of low substrate Stot � E1,tot, E2,tot,we have that [E1] ≈ E1,tot and [E2] ≈ E2,tot, and thenonlinear mass-action dynamics implied by (G2) reduceto linear kinetics, with the enzyme concentrations “ab-sorbed” into the rate constants (see Fig. 14).

In this work, we consider the low-substrate limit,and study the relative probability πS∗/πS for a partic-ular substrate molecule to be modified. For thermody-namic consistency, all reactions must be reversible, sowe must include product rebinding. We further sup-pose that concentrations of other participants in thesereactions (e.g. ATP, ADP in the case of phosphoryla-tion/dephosphorylation) are held at fixed values. Thesechoices yield a system of the form we have studied inthe preceding sections, with linear dynamics of the form(1), held out of equilibrium by the cycle force FC =

Page 17: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

17

ln(W21W32W43W14

W12W23W34W41). For a system such as this one, driven

chemically, we can identify FC = ∆µ/kBT . Our resultsproved above then imply a bound, in terms of ∆µ, on the

sensitivity s of the steady-state ratio πS∗/πS to a changein [E1].

Perturbing [E1] is equivalent to perturbing two edgeparameters and a vertex parameter:

[E1]∂

∂[E1]= W21

∂W21+W23

∂W23(G3)

=

(W12

∂W12+W21

∂W21

)+

(W23

∂W23+W32

∂W32

)−(W12

∂W12+W32

∂W32

)= − ∂

∂B12− ∂

∂B23− ∂

∂E2.

Now we can apply Corollaries 1 and 4 to bound the sen-sitivity

s =

∣∣∣∣[E1]∂ ln(π3/π1)

∂[E1]

∣∣∣∣ (G4)

=

∣∣∣∣−( ∂

∂B12+

∂B23

)ln

(π3π1

)− ∂ ln(π3/π1)

∂E2

∣∣∣∣≤ tanh(F1↔3/4) = tanh(∆µ/4kBT ).

Remarkably, the form of the bound (G4) remains un-changed even if the assumption that catalysis proceedsvia a single intermediate complex is completely relaxed.In particular, following Gunawardena et al. [37, 38], weconsider an arbitrary reaction network built out of a col-lection of any number of reactions of the following form,which include an arbitrary number of intermediates andreactions between them:

E1 + S ↔ (E1S)i

E1 + S∗ ↔ (E1S)i

(E1S)i ↔ (E1S)j

E2 + S ↔ (E2S)i

E2 + S∗ ↔ (E2S)i

(E2S)i ↔ (E2S)j .

(G5)

A general network of this form is schematically rep-resented in Fig. 4(b). In any such network, considerthe subgraph whose vertices V are all the intermediates{(E1S)i} containing E1, together with S and S∗, andwhose edges E are all the edges between the vertices V .Scaling [E1] is equivalent to decreasing all the edge pa-rameters associated to edges in E , and the vertex param-eters associated to vertices in the set VI = V \ {S, S∗}.

This decomposition yields the result

s =

∣∣∣∣[E1]∂ ln(πS∗/πS)

∂[E1]

∣∣∣∣ (G6)

=

∣∣∣∣∣−(∑v∈VI

∂Ev−∑e∈E

∂Be

)ln

(πS∗

πS

)∣∣∣∣∣≤

∣∣∣∣∣(∑e∈E

∂Be

)ln

(πS∗

πS

)∣∣∣∣∣≤ tanh(FS∗↔S/4) = tanh(∆µ/4kBT )

where the last line follows from Corollary 5 with W ={S, S∗}, |W | = 2.

2. Biochemical sensing and the Govern–ten Woldetrade-off

Now we will show how our sensing bound (24) arises,and how it reduces to the results of Govern and ten Wolde[39] in the appropriate limits.

To arrive at (24), we first employ the approximation(23) (

σe[E1]

)2

≈ σ2s

(∂µs/∂[E1])21

[E1]2

=NπS∗(1− πS∗)

N2([E1]∂πS∗/∂[E1])2. (G7)

Now we recognize the derivative in the denominator asbeing an instance of the kinds we have considered al-ready. In particular, application of (20), in the form ofa ratio of general observables, together with our vertexperturbation results, yields∣∣∣∣[E1]

∂πS∗

∂[E1]+ πS∗πY

∣∣∣∣ ≤ πS∗(1− πS∗) tanh(∆µ/4kBT )

(G8)where πY is the fraction of the total substrate boundup in complexes involving E1. In [39], these enzymatic

Page 18: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

18

intermediates are neglected (e.g. equation [S22] of [39]),and to proceed we will do the same here, supposing thatπY is very small. We then get

(σe

[E1]

)2

&1

NπS∗(1− πS∗) tanh(∆µ/4kBT )2

≥ 4

N tanh(∆µ/4kBT )2(G9)

as desired. In the limit ∆µ� kBT ,

(σe

[E1]

)2

&4

N(G10)

whereas in the limit ∆µ� kBT , this yields

(σe

[E1]

)2

&64

N

(kBT

∆µ

)2

. (G11)

To make contact between our low potential limit ∆µ�kBT (G11) and the results of Govern and ten Wolde, wenow review the context of their results. In that paper, theauthors study the error δc/c in an estimate of the con-centration c of an extracellular ligand L, which binds toreceptor R, forming a complex: R+ L↔ RL. The com-plex RL then plays the role of E1 in our discussion above,participating in a covalent modification cycle. The con-centration of the ligand-bound receptor (in our notationE1, in their notation RL) is given by

[E1] = [RL] = RT p

p =c

c+K(G12)

where K is the dissociation constant and RT is the totalconcentration of receptors.

An estimate of c can be constructed, just as we describein the main text producing an estimate of [E1]. Usingthe approximation (23) together with the equation (G12)relates the error of these estimates,

(δc

c

)2

≈ 1

(1− p)2

(σe

[E1]

)2

(G13)

The authors of [39] compare the sensing error (δc/c)2

not to the chemical potential difference ∆µ, but to aquantity w, which is the dissipation rate of the wholesystem, normalized by the sum of all the rates (forwardand reverse) around the modification cycle (in the casestudied by the authors, consisting of only two states, inour notation, S and S∗, and no intermediates). We shall

call this quantity, which is the rate of relaxation to steadystate, R. So in our notation,

w =

(N × |J | × ∆µ

kBT

)/R. (G14)

where J is the net current around the cycle, per substratemolecule. The arguments of Govern and ten Wolde showthat in the limit ∆µ� kBT ,(

δc

c

)2

≥ 4

(1− p)2w. (G15)

This is slightly tighter than the inequality (S26) theywrite in [39], but also follows from their argument. As aconsequence of (G13) and (G15), we then get

(σe/[E1])2 ≥ 4/w. (G16)

To show that this bound coincides with our result (G11),we need to show that for fixed small force the smallestvalue achieved by 4/w is 64(kBT )2/N(∆µ)2. If this werenot so, it would imply that either our result was weakerthan that of Govern and ten Wolde, or vice versa.

To show that (G11) and (G16) do coincide for fixedsmall force, we use the inequality,

|J |R≤ πS∗(1− πS∗) tanh

(∆µ

4kBT

)≤ ∆µ

16kBT(G17)

which holds (for a two-state system), by the same algebra(C6)–(C8) that led to our symmetric edge perturbationresult. See also Malaguti and ten Wolde [61] (equationsS112 to S114), who give an explicit expression for |J |/R.

Plugging (G17) into the definition (G14), we get

4

w≥ 64

N

(kBT

∆µ

)2

. (G18)

Additionally, this inequality can be saturated in the limit∆µ� kBT , because there is a near equilibrium regime inwhich |J |/R ≈ ∆µ/16kBT . This establishes the desiredequivalence between our results.

3. Kinetic proofreading

In our presentation and analysis here, we follow closelythe papers of Murugan et al. [47, 48]. Our results gener-alize bounds on the discriminatory index ν found in thoseworks.

First, we consider the single-loop, three-state network(see Fig. 15) equivalent to the system studied by Hopfieldand Ninio [45, 46]. A perturbation of the binding energy∆ can be decomposed as a linear combination of vertexand symmetric edge parameter perturbations. In termsof the notation we introduce in Fig. 15(b), we have

Page 19: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

19

ES

ES

E

k2eΔ

k1eΔ

2

3

1W12

W13 W23

W21

W31W32

(a) (b)***

* *

*

**

*

FIG. 15. (a) The transition graph G for the single-loop kineticproofreading mechanism of Hopfield. The dissociation ratesof the complexes ES∗ and ES are the only rates that dependon the binding energy ∆. Perturbing ∆ is equivalent to vertexand edge perturbations (blue). (b) State numbers and ratelabels we use in this subsection. Key equivalences are “1 =E”, “3 = ES”, “W12 = k1e

∆”, and “W13 = k2e∆”.

∂∆= W12

∂W12+W13

∂W13(G19)

=

(W13

∂W13+W23

∂W23

)+

(W12

∂W12+W32

∂W32

)−(W23

∂W23+W32

∂W32

)=

∂E3+

∂E2+

∂B23.

Now we can apply Corollaries 1 and 2 to derive the bound

|ν − 1| =∣∣∣∣∂ ln(π1/π3)

∂∆− 1

∣∣∣∣ =

∣∣∣∣∂ ln(π1/π3)

∂E3+∂ ln(π1/π3)

∂E2+∂ ln(π1/π3)

∂B23− 1

∣∣∣∣ (G20)

≤ |1 + 0− 1|+∣∣∣∣∂ ln(π1/π3)

∂B23

∣∣∣∣≤ tanh (Fmax/4) = tanh (∆µ/4kBT ) .

In the case of the more general kinetic proofreadingscheme [47, 48] where m complexes can dissociate, de-scribed in Fig. 6, perturbing ∆ is equivalent to perturbingthe edge and vertex parameters associated to the edges E

and vertices V on one side of the “fence”. We then have

|ν − 1| =∣∣∣∣∂ ln(πE/πES)

∂∆− 1

∣∣∣∣ (G21)

=

∣∣∣∣∣(∑v∈V

∂Ev+∑e∈E

∂Be

)ln

(πEπES

)− 1

∣∣∣∣∣≤ |1− 1|+

∣∣∣∣∣(∑e∈E

∂Be

)ln

(πEπES

)∣∣∣∣∣≤ (m− 1) tanh(FE↔ES/4)

where in the last line we have applied Corollary 5.

[1] P. M. Chaiken and T. C. Lubensky, Principles of con-densed matter physics (Cambridge University Press,Cambridge, 1995).

[2] R. Kubo, M. Toda, and N. Hashitsume, Statisti-

cal Physics II: Nonequilibrium Statistical Mechanics(Springer-Verlag, Berlin, 1985).

[3] T. G. Mason and D. A. Weitz, Optical measuremetns offrequency-dependent lienar viscoelastic moduli of com-

Page 20: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

20

plex fluids, Phys. Rev. Lett. 74, 1250 (1995).[4] H. Qian, Coopertivity in cellular biochemical processes:

Noise-enhanced sensitivity, fluctuating enzyme, bistabil-ity with nonlinear feedback, and other mechanisms forsigmoidal responses, Ann. Rev. Biophys. 41, 179 (2012).

[5] C. C. Govern and P. R. ten Wolde, Energy dissipationand noise correlations in biochemical sensing, Physicalreview letters 113, 258102 (2014).

[6] L. Bintu, N. E. Buchler, H. G. Garcia, U. Gerlan, T. Hwa,J. Kondev, and R. Phillips, Transcriptional regulation bythe numbers: models, Curr. Opin. Genetics Dev. 15, 116(2005).

[7] A. Murugan, D. A. Huse, and S. Leibler, Speed, dissipa-tion, and error in kinetic proofreading, Proc. Natl. Acad.Sci. USA 109, 12034 (2011).

[8] J. Estrada, F. Wong, A. DePace, and J. Gunawardena,Information integration and energy expenditure in generegulation, Cell 166, 234 (2016).

[9] G. S. Agarwal, Fluctuation-dissipation theorems for sys-tems in non-thermal equilibrium and applications, Z.Phys. A 252, 25 (1972).

[10] H. Risken, The Fokker-Planck Equation: Methods ofSolution and Applications (Springer-Verlag, New York,1984).

[11] M. Baiesi, C. Maes, and B. Wynants, Fluctuations andresponse in nonequilibrium states, Phys. Rev. Lett. 103,010602 (2009).

[12] J. Prost, J.-F. Joanny, and J. M. R. Parrondo, General-ized fluctuation-dissipation theorem for steady-state sys-tems, Phys. Rev. Lett. 103, 090601 (2009).

[13] U. Seifert and T. Speck, Fluctuation-dissipation theo-rem in nonequilibrium steady states, Europhys. Lett. 89,10007 (2010).

[14] J. R. Gomez-Solano, A. Petrosyan, S. Ciliberto,R. Chetrite, and K. Gawedzki, Experimental verifica-tion of a modified fluctuation-dissipation relation for amicron-sized particle in a nonequilibrium steady state,Phys. Rev. Lett. 103, 040601 (2009).

[15] J. Mehl, V. Blickle, U. Seifert, and C. Bechinger, Experi-mental accessibility of generalized fluctaution-dissipationrelations for nonequilibrium steady states, Phys. Rev. E82, 032401 (2010).

[16] P. Bohec, F. Gallet, C. Maes, S. Safaverdi, P. Visco, andF. van Wijland, Probing active forces via a fluctuation-dissipation relation: application to living cells, Europhys.Lett. 102, 50005 (2013).

[17] B. Altaner, M. Polettini, and M. Esposito, Fluctuation-dissipation relations far from equilibrium, Phys. Rev.Lett. 117, 180601 (2016).

[18] T. Harada and S.-i. Sasa, Equality connecting energy dis-sipation with a violation of the fluctuation-response re-lation, Physical review letters 95, 130602 (2005).

[19] T. Harada and S.-i. Sasa, Energy dissipation and vio-lation of the fluctuation-response relation in nonequilib-rium langevin systems, Physical Review E 73, 026131(2006).

[20] S. Toyabe, T. Okamoto, H. Watanabe-Nakayama,T. Taketani, S. Kudo, and E. Muneyuki, Nonequilibriumenergetics of a single f1-atpase molecule, Phys. Rev. Lett.104, 198103 (2010).

[21] C.-C. S. Yan and C.-P. Hsu, The fluctuation-dissipationtheorem for stochastic kinetics - implications on geneticregulations, J. Chem. Phys. 139, 224109 (2013).

[22] K. Sato, Y. Ito, T. Yomo, and K. Kaneko, On the relation

between fluctaution and response in biological systems,Proc. Natl. Acad. Sci. USA 100 (2003).

[23] L. F. Cugliandolo, The effective temperature, J. Phys. A:Math. Theor. 44, 483001 (2011).

[24] E. Ben-Isaac, Y. Park, G. Popescu, F. L. H. Brown, N. S.Gov, and Y. Shokef, Effective temperature of red-blood-cell membrane fluctuations, Phys. Rev. Lett. 106, 238103(2011).

[25] E. Dieterich, J. Camunas-Soler, M. Ribezzi-Crivellari,U. Seifert, and F. Ritort, Single-molecule measurementof the effective temperature in non-equilibrium steadystates, Nat. Phys. 11, 971 (2015).

[26] A. C. Barato and U. Seifert, Thermodynamic uncertaintyrelation for biomolecular processes, Phys. Rev. Lett. 114,158101 (2015).

[27] T. R. Gingrich, J. M. Horowitz, N. Perunov, and J. L.England, Dissipation bounds all steady-state currentfluctuations, Phys. Rev. Lett. 116, 120601 (2016).

[28] U. Seifert, Stochastic thermodynamics, fluctuation the-orems, and moleculer machines, Rep. Prog. Phys. 75,126001 (2012).

[29] J. Schnakenberg, Network theory of microscopic andmacroscopic behavior of master equation systems, Rev.Mod. Phys. 48, 571 (1976).

[30] D. Andrieux and P. Gaspard, Fluctuation theorem fortransport in mesoscopic systems, J. Stat. Mech.: Theor.Exp. , P01011 (2006).

[31] R. Zia and B. Schmittmann, Probability currents as prin-cipal characteristics in the statistical mechanics of non-equilibrium steady states, J. Stat. Mech.: Theor. Exp.2007, P07012 (2007).

[32] T. L. Hill, Free Energy Transduction in Biology (Aca-demic Press, New York, 1977).

[33] D. Hartich, A. C. Barato, and U. Seifert, Nonequilibriumsensing and its analogy to kinetic proofreading, New J.Phys. 17, 055026 (2015).

[34] E. Thiede, B. Van Koten, and J. Weare, Sharp entrywiseperturbation bounds for markov chains, SIAM J. MatrixAnal. Appl. 36, 917 (2015).

[35] A. Goldbeter and D. E. Koshland, An amplified sensi-tivty arising from covalent modification in biological sys-tems, Proc. Nat. Ac. Sci. 78, 6840 (1981).

[36] H. Qian, Phosphorylation energy hypothesis: Openchemical systems and theire biological functions, Ann.Rev. Phys. Chem. 58, 113 (2007).

[37] Y. Xu and J. Gunawardena, Realistic enzymology forpost-translational modification: zero-order ultrasensitiv-ity revisited, J. Theor. Biol. 311, 139 (2012).

[38] T. Dasgupta, D. H. Croll, J. A. Owen, M. G. Vander Hei-den, J. W. Locasale, U. Alon, L. C. Cantley, and J. Gu-nawardena, A fundamental trade-off in covalent switch-ing and its circumvention by enzyme bifunctionality inglucose homeostasis, J. Biol. Chem. 289, 13010 (2014).

[39] C. C. Govern and P. R. ten Wolde, Optimal resourceallocation in cellular sensing systems, Proceedings of theNational Academy of Sciences 111, 17486 (2014).

[40] A. Goldbeter and D. E. Koshland, Energy expenditurein the control of biochemical systems by covalent modi-fication, J. Biol. Chem. 262, 4460 (1987).

[41] Y. Tu, The nonequilibrium mechanism for ultrasensitiv-ity in a biological switch: Sensing by maxwell’s demons,Proc. Natl. Acad. Sci. USA. 105, 11737 (2008).

[42] H. C. Berg and E. M. Purcell, Physics of chemoreception,Biophysical journal 20, 193 (1977).

Page 21: with biochemical applications - arXiv · Universal thermodynamic bounds on nonequilibrium response with biochemical applications Jeremy A. Owen, 1 Todd R. Gingrich, 2 and Jordan M.

21

[43] W. Bialek and S. Setayeshgar, Physical limits to biochem-ical signaling, Proceedings of the National Academy ofSciences 102, 10040 (2005).

[44] K. Kaizu, W. De Ronde, J. Paijmans, K. Takahashi,F. Tostevin, and P. R. Ten Wolde, The berg-purcell limitrevisited, Biophysical journal 106, 976 (2014).

[45] J. J. Hopfield, Kinetic proofreading: a new mechanismfor reducing errors in biosynthetic processes requiringhigh specificity, Proc. Natl. Acad. Sci. USA 71, 4135(1974).

[46] J. Ninio, Kinetic amplification of enzyme discrimination,Biochimie 57, 587 (1975).

[47] A. Murugan, D. A. Huse, and S. Leibler, Discriminatoryproofreading regimes in nonequilibrium systems, Phys.Rev. X 4, 021016 (2014).

[48] A. Murugan, D. A. Huse, and S. Leibler, Speed, dissi-pation, and error in kinetic proofreading, Proceedings ofthe National Academy of Sciences 109, 12034 (2012).

[49] T. E. Ouldridge, C. C. Govern, and P. R. ten Wolde,Thermodynamics of computational copying in biochemi-cal systems, Physical Review X 7, 021004 (2017).

[50] U. Seifert, Generalized einstein and green-kubo relationsfor active biomolecular transport, Phys. Rev. Lett. 104,138101 (2010).

[51] M. Baiesi, C. Maes, and B. Wynants, The modifiedsutherland-einstein realtion for diffusive non-equilibria,Proc. R. Soc. Lond. 467, 2792 (2011).

[52] A. Dechant and S. I. Sasa, Fluctuation-response inequal-ity out of equilibrium, arXiv:1804.08250.

[53] W. Tutte, The dissection of equilateral triangles intoequilateral triangles, in Mathematical Proceedings of theCambridge Philosophical Society, Vol. 44 (CambridgeUniversity Press, 1948) pp. 463–482.

[54] T. L. Hill, Studies in irreversible thermodynamics iv. di-agrammatic representation of steady state fluxes for uni-molecular systems, Journal of theoretical biology 10, 442(1966).

[55] B. O. Shubert, A flow-graph formula for the stationarydistribution of a markov chain, IEEE Transactions onSystems, Man, and Cybernetics , 565 (1975).

[56] F. Leighton and R. Rivest, Estimating a probability usingfinite memory, IEEE Transactions on Information Theory32, 733 (1986).

[57] I. Mirzaev and J. Gunawardena, Laplacian dynamics ongeneral graphs, Bull. Math. Biol. 75, 2118 (2013).

[58] J. Gunawardena, A linear framework for time-scale sep-aration in nonlinear biochemical systems, PloS one 7,e36321 (2012).

[59] J. Gunawardena, Time-scale separation–michaelis andmenten’s old idea, still bearing fruit, The FEBS journal281, 473 (2014).

[60] F. Wong, A. Amir, and J. Gunawardena, An energy-speed-accuracy relation in complex networks for bio-logical discrimination, arXiv preprint arXiv:1710.06038(2017).

[61] G. Malaguti and P. R. ten Wolde, Theory for the opti-mal detection of time-varying signals in cellular sensingsystems, arXiv preprint arXiv:1902.09332 (2019).


Recommended