arxiv: v2 [cond-mat.stat-mech] 3 Jun 2018

Similar documents
Maxwell's Demon in Biochemical Signal Transduction

J. Stat. Mech. (2011) P07008

Information Thermodynamics on Causal Networks

arxiv: v2 [cond-mat.stat-mech] 16 Mar 2012

Nonequilibrium Thermodynamics of Small Systems: Classical and Quantum Aspects. Massimiliano Esposito

Second law, entropy production, and reversibility in thermodynamics of information

Chapter 2 Review of Classical Information Theory

Beyond the Second Law of Thermodynamics

Entropy Production and Fluctuation Relations in NonMarkovian Systems

Thermodynamic Computing. Forward Through Backwards Time by RocketBoom

Emergent Fluctuation Theorem for Pure Quantum States

arxiv: v1 [cond-mat.stat-mech] 6 Jul 2015

Fluctuation theorem between non-equilibrium states in an RC circuit

Thermodynamic Cost Due to Changing the Initial Distribution Over States

Hardwiring Maxwell s Demon Tobias Brandes (Institut für Theoretische Physik, TU Berlin)

Nonequilibrium thermodynamics at the microscale

Information Landscape and Flux, Mutual Information Rate Decomposition and Connections to Entropy Production

PHYS 414 Problem Set 4: Demonic refrigerators and eternal sunshine

Physical description of nature from a system-internal viewpoint. Abstract

Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences

arxiv: v2 [cond-mat.stat-mech] 9 Jul 2012

arxiv: v2 [cond-mat.stat-mech] 23 Jul 2018

Transient Dissipation and Structural Costs of Physical Information Transduction

Introduction to Stochastic Thermodynamics: Application to Thermo- and Photo-electricity in small devices

Information in Biology

Information in Biology

Contrasting measures of irreversibility in stochastic and deterministic dynamics

5 Mutual Information and Channel Capacity

Maxwell's Demons and Quantum Heat Engines in Superconducting Circuits

LECTURE 4: Information-powered refrigerators; quantum engines and refrigerators

Stochastic thermodynamics

Mathematical Structures of Statistical Mechanics: from equilibrium to nonequilibrium and beyond Hao Ge

arxiv: v2 [cond-mat.stat-mech] 28 Mar 2013

Fluctuation theorems. Proseminar in theoretical physics Vincent Beaud ETH Zürich May 11th 2009

Fluctuation response inequality out of equilibrium

Information and Physics Landauer Principle and Beyond

Stochastic Thermodynamics of Langevin systems under time-delayed feedback control

Fluctuation relations and nonequilibrium thermodynamics II

Fluctuation theorems: where do we go from here?

Demon Dynamics: Deterministic Chaos, the Szilard Map, & the Intelligence of Thermodynamic Systems.

Statistical Complexity of Simple 1D Spin Systems

Thermodynamics for small devices: From fluctuation relations to stochastic efficiencies. Massimiliano Esposito

Stochastic thermodynamics of information

Optimal Thermodynamic Control and the Riemannian Geometry of Ising magnets

The physics of information: from Maxwell s demon to Landauer. Eric Lutz University of Erlangen-Nürnberg

arxiv: v1 [cond-mat.stat-mech] 8 Jul 2013

arxiv: v2 [cond-mat.stat-mech] 26 Jan 2012

arxiv: v2 [cond-mat.stat-mech] 18 Jun 2009

Chapter 16. Structured Probabilistic Models for Deep Learning

3. If a choice is broken down into two successive choices, the original H should be the weighted sum of the individual values of H.

Feedback Control and Counting Statistics in Quantum Transport Tobias Brandes (Institut für Theoretische Physik, TU Berlin)

Introduction to Fluctuation Theorems

Fluctuation Theorems of Work and Entropy in Hamiltonian Systems

LANGEVIN EQUATION AND THERMODYNAMICS

Stochastic Thermodynamics of Langevin systems under time-delayed feedback control

Statistical properties of entropy production derived from fluctuation theorems

Signal Processing - Lecture 7

Quantum Thermodynamics

arxiv: v4 [cond-mat.stat-mech] 3 Mar 2017

Optimum protocol for fast-switching free-energy calculations

Non equilibrium thermodynamic transformations. Giovanni Jona-Lasinio

Quantum-information thermodynamics

Information and thermodynamics: Experimental verification of Landauer s erasure principle

Entropy production and time asymmetry in nonequilibrium fluctuations

Thermodynamics of feedback controlled systems. Francisco J. Cao

Fluctuation, Dissipation and the Arrow of Time

Aspects of nonautonomous molecular dynamics

arxiv: v1 [cond-mat.stat-mech] 28 Nov 2016

Variational Inference (11/04/13)

Experimental Rectification of Entropy Production by Maxwell s Demon in a Quantum System

Multicanonical distribution and the origin of power laws. Abstract

A complex pattern of allosteric regulation is involved in its operation. The catalytic activity in a subunit is

arxiv: v1 [quant-ph] 30 Jul 2012

The Two-State Vector Formalism

arxiv: v1 [cond-mat.stat-mech] 6 Nov 2015

Microscopic Deterministic Dynamics and Persistence Exponent arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Sep 1999

Lecture 4: State Estimation in Hidden Markov Models (cont.)

Quantum Mechanical Foundations of Causal Entropic Forces

Introduction to Information Theory. Uncertainty. Entropy. Surprisal. Joint entropy. Conditional entropy. Mutual information.

Dept. of Linguistics, Indiana University Fall 2015

Undirected Graphical Models

arxiv: v2 [q-bio.pe] 27 May 2017

Chapter 2: Entropy and Mutual Information. University of Illinois at Chicago ECE 534, Natasha Devroye

Quantum measurement theory and micro-macro consistency in nonequilibrium statistical mechanics

Efficiency at Maximum Power in Weak Dissipation Regimes

Probabilistic Graphical Models

Fluctuation theorem in systems in contact with different heath baths: theory and experiments.

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013 Notes on the Microcanonical Ensemble

Thermodynamics of histories: application to systems with glassy dynamics

Stochastic equations for thermodynamics

Closed-loop analysis of a thermo-charged capacitor

How to Quantitate a Markov Chain? Stochostic project 1

arxiv: v1 [physics.class-ph] 14 Apr 2012

Bounds on thermal efficiency from inference

Chris Bishop s PRML Ch. 8: Graphical Models

arxiv: v2 [cond-mat.stat-mech] 26 Nov 2016

Shortcuts to Thermodynamic Computing: The Cost of Fast and Faithful Erasure

Thermodynamics of Information Processing in Small Systems )

Derivation of Bose-Einstein and Fermi-Dirac statistics from quantum mechanics: Gauge-theoretical structure

Even if you're not burning books, destroying information generates heat.

Transcription:

Marginal and Conditional Second Laws of Thermodynamics Gavin E. Crooks 1 and Susanne Still 2 1 Theoretical Institute for Theoretical Science 2 University of Hawai i at Mānoa Department of Information and Computer Sciences and Department of Physics and Astronomy Honolulu Hawai i 96822 USA ariv:1611.04628v2 [cond-mat.stat-mech] 3 Jun 2018 We consider the entropy production of a strongly coupled bipartite system. The total entropy production can be partitioned into various components which we use to define local versions of the Second Law that are valid without the usual idealization of weak coupling. The key insight is that causal intervention offers a way to identify those parts of the entropy production that result from feedback between the sub-systems. From this the central relations describing the thermodynamics of strongly coupled systems follow in a few lines. PACS numbers: 05.70.Ln05.40.a Rudolf Clausius famous statement of the second fundamental theorem in the mechanical theory of heat is that The entropy of the universe tends to a maximum. [1] Although this proclamation has withstood the test of time in practice measuring the entropy of the entire universe is difficult. As an alternative we can apply the Second Law to any system isolated from outside interactions (a Universe unto itself) as for example in Planck s statement of the Second Law: Every process occurringinnature...thesumoftheentropiesofallbodies taking part in the process is increased. [2] Of course perfectly isolating any system or collection of systems from outside influence is also difficult. Over the last 150 years thermodynamics has progressed by adopting various idealizations which allow us to isolate and measure that part of the total Universal entropy change that is relevant to the behavior of the system at hand. These idealizations include heat reservoirs work sources and measurement devices [3 4]. More recently information engines ( Maxwell demons [5 6]) have been added to the canon to represent idealized computational resources [4 7 10]. In this paper we demonstrate that we do not need in principle to resort to these idealizations. We show how the thermodynamics of strongly coupled systems follow in a straightforward manner from a causal decomposition of the dynamics. This unifying perspective greatly simplifies the treatment allowing us to assimilate the large recent literature on the thermodynamics of coupled systems in a few short pages. Looking at the problem in the rightwaythen makes it easy to showthat conditional and marginalized versions of the Second Law hold locally even when the system of interest is strongly coupled to other driven non-equilibrium systems. Partitions of entropy Before considering the partitioning of dissipation let us remember the partitioning of entropy in information theory. Suppose we have a pair of interacting systems and whose states are x and y respectively. The joint entropy of the total system is S = xy p(xy)lnp(xy). (1) The marginal entropy S of system is the entropy of the marginal distribution obtained by summing over the states of the other system S = x The conditional entropy p(x)lnp(x) p(x) = y S = S S = xy p(xy). (2) p(xy)lnp(x y) (3) is the average entropy of system given that we know the state of system. It is also useful to define the pointwise (specific) entropies of individual realizations of the system whose ensemble averages are the entropies defined previously. s(xy) = lnp(xy) = s(x y)+s(y) s(x) = lnp(x) s(x y) = lnp(x y) (4a) (4b) (4c) The negative log-probability Eq. (4b) is sometimes called surprisal in information theory. Dissipation Let us now consider dynamical trajectories of a bipartite system. We assume that each side of the system is coupled to idealized constant temperature heat reservoirs with reciprocal temperature β = 1/k B T and idealized work sources the controlled parameters of which we label u and v. Although the coupling to the heat baths and work sources are idealized the two subsystems may be strongly coupled to each other. A core tenet of non-equilibrium thermodynamics is the detailed fluctuation theorem which equates the entropy production (or dissipation) Σ to the log ratioof the probabilities of forward and time reversed trajectories[11 13]. Σ = ln p( x y ; u v) p( x y ; u v) (5)

2 Dissipation is a consequence of breaking time-reversal symmetry. Here x and y are trajectories of systems and respectively generated while the systems are driven by the external protocols u and v respectively. The trajectory x denotes x in reverse running time backwards. Consequently p( x y ; u v) is the probability of the forward time trajectories given the forward time dynamics and the forward time protocols whereas p( x y ; u v) is the probability of the conjugate time reversed trajectories given the time reversed driving. We use a semicolon before the controls to emphasize that the protocols are fixed parameters. But for notational simplicity we typically suppress the explicit dependance of the dynamics on the protocols writing p( x y) for p( x y ; u v) for example. Suppose we only observe the behavior of one of the subsystems. We can still define marginal trajectory probabilities and a marginal fluctuation theorem Σ = ln p( x) p( x) (6) and by similar reasoning the conditional entropy production x y) Σ = ln p( p( x y) = ln x y) p( p( y) p( x y) p( y) = Σ Σ. (7) Thus we can partition the total dissipation into local components. However in order to make these definitions of marginal and conditional dissipation concrete we will have to explore their physical significance. Dynamics To make the discussion unambiguous we adopt a specific model of the intersystem dynamics. We could opt for classical mechanics [4] or coupled Langevin dynamics [14] or a continuous time Markov process [15]. But we feel the discussion is most transparent when the dynamics are represented by coupled discrete time Markov chains. The dynamics of the joint system are assumed to be Markov and the dynamics of each subsystem is conditionally Markov given the state of the neighboring subsystems. The marginal dynamics are not Markov when we do not know the hidden dynamics of the other subsystem. We can use a causal diagram [16 18] to illustrate the time label conventions for the trajectories of the system and control parameters. First one subsystem updates then the other and so on. u 1 u 2 u 3 x 0 x 1 x 2 x 3 y 0 y 1 y 2 y 3 v 0 v 1 v 2 Horizontal arrows indicate time evolution and the other connections indicate causation where the dynamics of one sub-system are influenced by the external parameters and the current state of the other sub-system. For the corresponding time reversed trajectory the horizontal arrows flip but the vertical connections remain unchanged. ũ 1 ũ 2 ũ 3 x 0 x 1 x 2 x 3 ỹ 0 ỹ 1 ỹ 2 ỹ 3 ṽ 0 ṽ 1 ṽ 2 Here the tilde x τ labels the time reversed configurations of the time reversed trajectory. The probability for the joint trajectory (which is a product of individual transition probabilities) naturally splits into a product of two terms. p( x y x 0 y 0 ) =p(y 1 y 0 x 0 )p(x 1 x 0 y 1 )p(y 2 y 1 x 1 )p(x 2 x 1 y 2 ) =...p(y τ y x )p(x τ x y τ ) p(y t+1 y t x t ) p(x t+1 x t y t+1 ) =q( y ; xy 0 ) q( x ; yx 0 ) (8) Here p(x t+1 x t y t+1 ) is the probability of jumping from state x t to x t+1 given the current state of the other subsystem (and given knowledge of the driving protocol). The expressions q( y ; xy 0 ) and q( x ; yx 0 ) are the trajectory probabilities of one system given a fixed trajectory of the other system. Once again we set off parameters that are fixed (rather than observed and coevolving) with a semicolon. This can be represented by the following pair of diagrams: u 1 u 2 u 3 x 0 x 1 x 2 x 3 y 1 y 2 y 3 x 0 x 1 x 2 y 0 y 1 y 2 y 3 v 0 v 1 v 2 These expressions are not the same as the conditional distributions p( x yy 0 ) which describe the original process where both systems coevolve and influence each other. We ve chosen to use a different symbol (q) for these trajectory probabilities to make this distinction abundantly clear. The decomposition of the joint probability in Eq.(8) is refereedto asacausalintervention[17]. This decomposition is central to disentangling the direct and indirect effects of intersystem coupling. Detailed fluctuation theorem For the complete system the total path-wise entropy production consists of the change in the entropy of the environment (due to the flow of heat from the baths) and a boundary term s(xy) [11 13] Σ = s(xy) βq ( x y). (9)

3 This boundary term is the difference in pointwise entropy between the initial configurations of the forward and reverse trajectories s(xy) = lnp( x τ ỹ τ )+lnp(x 0 y 0 ). (10) Typically we either assume that the system is initially in thermodynamic equilibrium for both the forward and reversed processes (as we do for the Jarzynski equality [19]) or we assume that the final ensemble of the forward process is the same as the initial time reversed probabilities of the reversed process p( x τ ỹ τ ) = p(x τ y τ ) [13 20]. However in general the initial ensembles need not have any simple relationship: for instance we might be observing a short segment of a much longer driven process. We also assume that the energy E of the total system consists of the two subsystem Hamiltonians and an interaction term E (xy ; uv) = E (x ; u)+e (y ; v)+e int : (xy). (11) The external baths and control parameters couple to the internal states of each system separately and do not couple directly to the interaction energy which ensures that the source of energy flowing into the system is unambiguous. The heat is the flow of energy into the system due to interactions with the bath [11 19 21 23]. We can split the total heat into the heat flow for each of the two subsystems Q = Q +Q [ Q = E (x t+1 u t+1 )+E: int (x t+1y t+1 ) ] E (x t u t+1 ) E: int (x ty t+1 ) (12a) [ Q = E (y t+1 v t )+E:(x int t y t+1 ) ] E (y t v t ) E:(x int t y t ). (12b) Local detailed fluctuation theorems If the trajectory of system is fixed then its dynamics act as an idealized work source to system and we can write down a standard fluctuation theorem for system alone. ln q( x ; yx 0 ) p(x 0 ) q( x ; y x τ ) p( x τ ) = s(x) βq (13) This is the fluctuation theorem we would obtain were there no feedback from to. With feedback this quantity no longer correctly describes the entropy production of sub-system. Assuming that it does leads to apparent contradictions that would imply that the Second Law has to be modified [7]. What is the quantity that correctly describes the entropy production of sub-system while coupled to the co-evolving sub-system? The answer depends on what information the observer has at hand. If the observer knows the co-evolving state of sub-system at all times then the conditional entropy production Σ Eq. (7) best describes the dissipation encountered by system alone. In the absence of this knowledge the marginal Σ Eq. (6) describes the entropy production of subsystem. While we have no guarantee that the average of Eq. (13) is positive when there is feedback between the systems we do know that both the marginal and the conditional entropy production obey the Second Law because they can be written as Kullback-Leibler divergences which means that they are all non-negative quantities [24 38 40]. Therefore let us now consider a decomposition of the marginal dissipation: x) Σ = ln p( p( x) = ln x y) p( p( y x) p( x y) p( y x) (14a) x y x = ln p( 0 y 0 ) p(x 0 y 0 ) p( y xỹ τ ) p(ỹ τ x τ ) p( x y x τ ỹ τ ) p( x τ ỹ τ ) p( y xy 0 ) p(y 0 x 0 ) (14b) y ; xy = ln q( 0 ) q( x ; yx 0 ) p( y xỹ τ ) p(x 0 ) q( y ; xỹ τ ) q( x ; y x τ ) p( y xy 0 ) p( x τ ) = ln p(x 0) p( x τ ) +ln q( x ; yx 0 ) q( x ; y x τ ) (14c) (14d) y xy ln p( 0 ) q( y ; xỹ τ ) p( y xỹ τ ) q( y ; xy 0 ) = s βq Σ trn (14e) WewritedowninEq. (14a)thedefinitionofthemarginal dissipation (see Eq. (6)) and expand it using the definition of conditional probability p(a b) = p(a b)p(b). Then we split out the initial state probabilities in Eq. (14b) and in Eq. (14c) split the probability of the joint trajectory into components without feedback (see Eq. (8)). In Eq. (14d) we gather terms and in Eq. (14e) realize from this decomposition that the marginal entropy production of sub-system is comprised of the contribution without feedback (see Eq. 13) together with a term that arisesas aconsequenceofthe feedback. We call this term the transferred dissipation. If system did not influence the behavior of system then q( y ; xy 0 ) would equal p( y xy 0 ) (and similarly for the time reversed components) and the transferred dissipation would be zero. With this insight we can now appreciate the fact that causal intervention and decomposition of the joint probability of the system as a whole into disconnected parts (Eq. 8) is crucial in thinking about feedback systems because it allows us to decompose the entropy production revealing the contribution due to feedback. Σ trn

4 To summarize the menagerie of local detailed fluctuation theorems (Eqs. 5-7) can be expressed solely in terms of differences in local pointwise entropies Eq. (4) the heat flow Eq. (12) and the transferred dissipation. Σ = s βq βq Σ = s βq Σ trn Σ = s βq +Σ trn (15a) (15b) (15c) Transferred dissipation Transferred dissipation can be further decomposed into time-forward and timereversed components (analogous to the decomposition of the entropy production rate into the Shannon entropy rate and a time reversed entropy rate [24]). Σ trn = ln p( x yx 0 ) q( x ; y x τ ) p( x y x τ ) q( x ; yx 0 ) = ln p( x yx 0 ) q( y ; xy 0 ) p( y x x τ ỹ τ ) p( x y x τ ) q( y ; xỹ τ ) p( y x x 0 y 0 ) = ln q( y ; xy 0 ) p( y y 0 ) ln q( y ; xỹ τ ) p( y ỹ τ ) = T T (16) After some additional manipulation we recognize that the first term is the sum of the pointwise transfer entropies [25]. T = ln p(y t+1 y t x t ) ln p(y t+1 y 0:t ) = ln p(y t+1 y 0:t x 0:t ) p(y t+1 y 0:t ) = i(y t+1 : x 0:t y 0:t ) (17) Here i(a : b c) = lnp(ab c) lnp(a c)p(b c) is the pointwise conditional mutual information and the slice notation x a:b is shorthand for the sequence x a x a+1...x b. Thus T is the total pointwise transfer entropy from to for the forward time trajectory. Transfer entropy has been investigated as a measure of Granger statistical causality [25 27]. Recently transfer entropy has been recognized as a component of the thermodynamic dissipation[18 28 31]. But transfer entropy can only equal the total transferred dissipation if we construe a process with feedback only in the time-forward dynamics but no feedback in the time-reversed dynamics. Such time-reverse feedback-free reference systems have been studied e.g. [7 8]. However in general we must include the time-reversed component of the transferred dissipation in order to fully appreciate the thermodynamic costs associated with interactions between sub-systems [32]. Information Engines An interesting idealized limit to consider is when the interaction energy is zero E : = 0 but the dynamics are still coupled. Removing the energetic component to the interaction forces us to carefully consider the role of information flow and computation in thermodynamics and the relationship between information and entropy [7 10 28 30 33 37]. Although no energy flows between the systems the transferred dissipation can still be non-zero. From the point of view of system system becomes a purely computational resource (a Demon ). This resource has an irreducible thermodynamic cost which is captured by the transferred dissipation. Neglecting this cost leads to Maxwell Demon paradoxes where the Second Law appears to be violated. Local Second Law We are now in a position to express the averages of the total marginal and conditional dissipation by averages of the quantities we found in our decomposition (Eq. (15)). Remember that we can also write them as Kullback-Leibler divergences and hence each obeys a Second Law like inequality: Σ = S β Q β Q (18a) = x y) p( x y)ln p( p( x y) 0 x y Σ = S β Q Σ trn = x) p( x)ln p( p( x) 0 x Σ = S β Q + Σ trn = y p( y) x p( x y)ln p( x y) p( x y) 0 (18b) (18c) Since the total dissipation is the sum of a conditional and marginal dissipation it follows that on average all of the marginal and conditional dissipations are less than the total dissipation and we summarize: Σ { Σ Σ Σ Σ } 0. (19) Thus strongly coupled systems obey a Local Second Law of Thermodynamics. One has to either consider the dynamics of the system alone and study the marginal dissipation or account for the behavior of other systems directly coupled to the system of interest and study the conditional dissipation. In either case the system s average dissipation is non-negative and less than the total dissipation. When studying small parts of the entire Universe we are allowed to neglect the dissipation occurring elsewhere that is irrelevant to the behavior of the system at hand. Acknowledgments We would like to acknowledge enlightening discussions with David Sivak Rob Shaw Tony Bell Jordan Horowitz Sosuke Ito Takahiro Sagawa and Jim Crutchfield. This work was sup-

5 ported by the Foundational Questions Institute (Grant No. FQi-RFP3-1345). gec@threeplusone.com sstill@hawaii.edu [1] R. Clausius Philos. Mag. Ser. 4 35 405 (1868). [2] M. Planck Treatise on Thermodynamics (Longmans Green and Co. London 1897). [3] H. B. Callen Thermodynamics and an Introduction to Thermostatistics (Wiley New ork 1985) 2nd ed. [4] S. Deffner and C. Jarzynski Phys. Rev. 3 041003 (2013). [5] J. C. Maxwell Theory of heat (Appleton London 1871). [6] L. Szilard Z. Phys. 53 840 (1929). [7] T. Sagawa and M. Ueda Phys. Rev. Lett. 104 090602 (2010). [8] J. M. Horowitz and S. Vaikuntanathan Phys. Rev. E 82 061120 (2010). [9] J. M. R. Parrondo J. M. Horowitz and T. Sagawa Nat. Phys. 11 131 (2015). [10] A. B. Boyd D. Mandal and J. P. Crutchfield Phys. Rev. E 95 012152 (2017). [11] G. E. Crooks J. Stat. Phys. 90 1481 (1998). [12] C. Jarzynski J. Stat. Phys. 98 77 (2000). [13] U. Seifert Phys. Rev. Lett. 95 040602 (2005). [14] T. Munakata and M. L. Rosinberg Phys. Rev. Lett. 112 180601 (2014). [15] J. M. Horowitz and M. Esposito Phys. Rev. 4 031015 (2014). [16] J. A. Davis The Logic Of Causal Order (SAGE Publications London 1985). [17] J. Pearl Causality: Models Reasoning and Inference (Cambridge University Press New ork 2009) 2nd ed. [18] S. Ito and T. Sagawa Phys. Rev. Lett. 111 180603 (2013). [19] C. Jarzynski Phys. Rev. Lett. 78 2690 (1997). [20] G. E. Crooks Phys. Rev. E 60 2721 (1999). [21] J. E. Hunter III W. P. Reinhardt and T. F. Davis J. Chem. Phys. 99 6856 (1993). [22] K. Sekimoto Prog. Theor. Phys. Suppl. 130 17 (1998). [23] L. Peliti Phys. Rev. Lett. 101 098903 (2008). [24] P. Gaspard J. Chem. Phys. 120 8898 (2004). [25] T. Schreiber Phys. Rev. Lett. 85 461 (2000). [26] C. W. J. Granger Econometrica 37 424 (1969). [27] R. G. James N. Barnett and J. P. Crutchfield Phys. Rev. Lett. 116 238701 (2016). [28] T. Sagawa andm. Ueda Phys. Rev.E85 021104 (2012). [29] M. Prokopenko J. T. Lizier and D. C. Price Entropy 15 524 (2013). [30] D. Hartich A. C. Barato and U. Seifert J. Stat. Mech.: Theor. Exp. p. P02016 (2014). [31] J. M. Horowitz and H. Sandberg New J. Phys. 16 125007 (2014). [32] R. E. Spinney J. T. Lizier and M. Prokopenko Phys. Rev. E 94 022135 (2016). [33] S. Toyabe T. Sagawa M. Ueda E. Muneyuki and M. Sano Nat. Phys. 6 988 (2010). [34] T. Sagawa Prog. Theor. Phys. 127 1 (2012). [35] J. V. Koski V. F. Maisi T. Sagawa and J. P. Pekola Phys. Rev. Lett. 113 030601 (2014). [36] J. M. Horowitz J. Stat. Mech.: Theor. Exp. p. P03006 (2015). [37] A. Bartolotta S. M. Carroll S. Leichenauer and J. Pollack Phys. Rev. E 94 022102 (2016). [38] C. Jarzynski Phys. Rev. E 73 046105 (2006). [39] R. Kawai J. M. R. Parrondo and C. Van den Broeck Phys. Rev. Lett. 98 080602 (2007). [40] A. Gomez-Marin J. M. R. Parrondo and C. Van den Broeck Phys. Rev. E 78 011107 (2008).