42
Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut, University of Hamburg, 20144 Grindelberg 5, Hamburg, Germany. Department of Mathematics and Statistics, University of Reading, Reading, RG6 6AX, UK. Richard Blender, Salvatore Pascale, Jeroen Wouters Klimacampus, Meteorologisches Institut, University of Hamburg, 20144 Grindelberg 5, Hamburg, Germany. Corentin Herbert National Center for Atmospheric Research, P.O. Box 3000, Boulder, CO, 80307, USA. The climate is an excellent example of a forced, dissipative system dominated by nonlin- ear processes and featuring non-trivial dynamics of a vast range of spatial and temporal scales. The understanding of the climate’s structural and multiscale properties is cru- cial for the provision of a unifying picture of its dynamics and for the implementation of accurate and efficient numerical models. Many open questions remain when trying to put together such a complex picture. How to describe comprehensively the scale- scale interactions and couplings, and how to construct robust, seamless parametriza- tions in numerical models? How to study the climatic response to perturbations? How to construct a succinct picture of the climate able to summarize its energy fluxes, the energy pathways, and the entropy budget? How to take into account the fundamental symmetries of the system? There has always been a fruitful mutual exchange of ideas, stimulations, methods between climate science and many sectors of applied mathematics and theoretical physics, as clearly showed by examples such as chaos theory, stochastic dynamical systems, turbulence, time series analysis, partial differential equations, and extreme value theory. In this interdisciplinary review, we would like to point out some recent developments at the intersection between climate science, mathematics, and the- oretical physics which may prove extremely fruitful in the direction of constructing a more comprehensive account of climate dynamics. We are guided by our interest in exploring the nexus between climate and concepts such as energy, entropy, symmetry, response, multiscale interactions, and its potential relevance in terms of numerical mod- eling. We describe the powerful reformulation of fluid dynamics made possible by the adoption of the Nambu formalism, and the possible potential of such theory for the construction of more sophisticated numerical models of the geophysical fluids. Then we focus on the very promising results on the statistical mechanics of quasi-equilibrium geophysical flows in a rotating environment, which are extremely useful in the direction of constructing a robust theory of geophysical macro turbulence. The second half of the review is dedicated instead to ideas and methods suited to approaching directly the non-equilibrium nature of the climate system. First, we give an account of some recent findings showing how to use basic concepts of macroscopic non-equilibrium ther- modynamics for characterizing the energy and entropy budgets of the climate systems, with the ensuing protocols for intercomparing climate models and developing methods aimed at studying tipping points. These ideas can also create a link between climate science and the growing sector of astrophysics devoted to the investigation of exoplan- etary atmospheres. We conclude our review by focusing on non-equilibrium statistical mechanics, which allows for framing in a unified way problems as different as climate response to forcings of general nature, the effect of altering the boundary conditions or the coupling between geophysical flows, and the derivation of parametrizations for numerical models. I. INTRODUCTION The Earth’s Climate provides an outstanding exam- ple of a high-dimensional forced and dissipative complex system. The dynamics of such system is chaotic, so that there is only a limited time-horizon for skillful prediction,

Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

Mathematical and Physical Ideas for Climate Science

Valerio Lucarini

Klimacampus, Meteorologisches Institut,University of Hamburg,20144 Grindelberg 5, Hamburg,Germany.Department of Mathematics and Statistics,University of Reading,Reading, RG6 6AX,UK.

Richard Blender, Salvatore Pascale, Jeroen Wouters

Klimacampus, Meteorologisches Institut,University of Hamburg,20144 Grindelberg 5, Hamburg,Germany.

Corentin Herbert

National Center for Atmospheric Research,P.O. Box 3000, Boulder, CO, 80307,USA.

The climate is an excellent example of a forced, dissipative system dominated by nonlin-ear processes and featuring non-trivial dynamics of a vast range of spatial and temporalscales. The understanding of the climate’s structural and multiscale properties is cru-cial for the provision of a unifying picture of its dynamics and for the implementationof accurate and efficient numerical models. Many open questions remain when tryingto put together such a complex picture. How to describe comprehensively the scale-scale interactions and couplings, and how to construct robust, seamless parametriza-tions in numerical models? How to study the climatic response to perturbations? Howto construct a succinct picture of the climate able to summarize its energy fluxes, theenergy pathways, and the entropy budget? How to take into account the fundamentalsymmetries of the system? There has always been a fruitful mutual exchange of ideas,stimulations, methods between climate science and many sectors of applied mathematicsand theoretical physics, as clearly showed by examples such as chaos theory, stochasticdynamical systems, turbulence, time series analysis, partial differential equations, andextreme value theory. In this interdisciplinary review, we would like to point out somerecent developments at the intersection between climate science, mathematics, and the-oretical physics which may prove extremely fruitful in the direction of constructing amore comprehensive account of climate dynamics. We are guided by our interest inexploring the nexus between climate and concepts such as energy, entropy, symmetry,response, multiscale interactions, and its potential relevance in terms of numerical mod-eling. We describe the powerful reformulation of fluid dynamics made possible by theadoption of the Nambu formalism, and the possible potential of such theory for theconstruction of more sophisticated numerical models of the geophysical fluids. Thenwe focus on the very promising results on the statistical mechanics of quasi-equilibriumgeophysical flows in a rotating environment, which are extremely useful in the directionof constructing a robust theory of geophysical macro turbulence. The second half ofthe review is dedicated instead to ideas and methods suited to approaching directlythe non-equilibrium nature of the climate system. First, we give an account of somerecent findings showing how to use basic concepts of macroscopic non-equilibrium ther-modynamics for characterizing the energy and entropy budgets of the climate systems,with the ensuing protocols for intercomparing climate models and developing methodsaimed at studying tipping points. These ideas can also create a link between climatescience and the growing sector of astrophysics devoted to the investigation of exoplan-etary atmospheres. We conclude our review by focusing on non-equilibrium statisticalmechanics, which allows for framing in a unified way problems as different as climateresponse to forcings of general nature, the effect of altering the boundary conditionsor the coupling between geophysical flows, and the derivation of parametrizations fornumerical models.

I. INTRODUCTION

The Earth’s Climate provides an outstanding exam-ple of a high-dimensional forced and dissipative complex

system. The dynamics of such system is chaotic, so thatthere is only a limited time-horizon for skillful prediction,

Page 2: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

2

and is non-trivial on a vast range of spatial and temporalscales, as a result of the different physical and chemi-cal properties of the various components of the climatesystem and of their coupling mechanisms (Peixoto andOort , 1992).

Thus, it is extremely challenging to construct satis-factory theories of climate dynamics and it is virtuallyimpossible to develop numerical models able to describeaccurately climatic processes over all scales. Typically,different classes of models and different phenomenologi-cal theories have been and are still being developed byfocusing on specific scales of motion (Holton, 2004; Val-lis, 2006), and simplified parametrizations are developedfor taking into account at least approximately what can-not be directly represented (Palmer and Williams, 2009).

As a result of our limited understanding of and abilityto represent the dynamics of the climate system, it is hardto predict accurately its response to perturbations, werethey changes in the opacity of the atmosphere, in the so-lar irradiance, in the position of continents, in the orbitalparameters, which have been present for our planet dur-ing all epochs (Saltzman, 2001). The full understandingof slow- and fast-onset climatic extremes, such as droughtand flood events, respectively, and the assessment of theprocesses behind tipping points responsible for the multistability of the climate system are also far from beingaccomplished.

Such limitations are extremely relevant for problemsof paleoclimatological relevance such as the onset anddecay of ice ages or of snowball-conditions, for contingentissues like anthropogenic global warming, as well as in theperspective of developing a comprehensive knowledge onthe dynamics and thermodynamics of general planetaryatmospheres, which seems a major scientific challenge ofthe coming years, given the extraordinary developmentof our abilities to observe exoplanets (Dvorak , 2008).

Climate science at large has always been extremelyactive in taking advantage of advances in basic mathe-matical and physical sciences, and, in turn, in providingstimulations for addressing new fundamental problems.The most prominent cases of such interaction are relatedto the development of stochastic and chaotic dynamicalsystems, time series analysis, extreme value theory, andfluid dynamics, among others. At this regard, one mustnote that the year 2013 has seen a multitude of initiativesall around the world dedicated to the theme Mathematicsof Planet Earth, and, in this context, themes of climaticrelevance have been of outstanding relevance.

In this review we wish to present some recent re-search lines at the intersection between climate science,physics and mathematics. Such approaches seem ex-tremely promising for advancing, on one side, our abilityto understand and model climate dynamics, and repre-sent correctly climate variability and climate response toforcings. On the other side, the topics presented hereprovide examples of how problems of climatic relevance

may pave the way for new, wide-ranging investigationsof more general nature.

Of course, our selection is partial and non-exhaustive.We leave almost entirely out of this review very impor-tant topics such as extreme value theory (Ghil et al.,2011), multiscale techniques (Klein, 2010), adjoint meth-ods and data assimilation (Wunsch, 2012), partial differ-ent equations (Cullen, 2006), linear and nonlinear sta-bility analysis (Vallis, 2006), general circulation of theatmosphere (Schneider , 2006), macroturbulence (Love-joy and Schertzer , 2013), networks theory (Donges et al.,2009), and many relevant applications of dynamical sys-tems theory to geophysical fluid dynamical problems (Di-jkstra, 2013; Kalnay , 2003).

Our selection will focus, instead, on the concepts ofenergy, entropy, symmetry, coupling, fluctuations, andresponse, and also within this realm, we have to makepainful choices given the vastly of the field. We are specif-ically motivated by the desire of bridging the gap betweensome extremely relevant results in mathematical physics,statistical mechanics, and theoretical physics, and openproblems and issues of climate science, hoping to stimu-late further investigations and interdisciplinary activities.

We will first concentrate on the properties of invis-cid, unforced geophysical flows. In Sec. II, we providean overview of a very powerful variational formulationof hydrodynamics based on the formalism introduced by(Nambu, 1973) and present its applications in a geophys-ical context, suggesting how these ideas could lead tonew generation of numerical models. In Sec. III, start-ing from the classical investigation by (Onsager , 1949)of the dynamics of point vortices, we will show how todevelop a statistical mechanical theory of turbulenceforgeophysical flows in a rotating environment.

We will then move to the paradigm of out-of-equilibrium systems, i.e. we firmly set ourselves in therealm of dissipative chaotic flows. In Sec. IV, takinginspiration from the points of view of (Prigogine, 1961)and of (Lorenz , 1967), we explore how through classi-cal non-equilibrium thermodynamics one can constructtools for assessing the energy budgets and transport, theefficiency, and the irreversibility of the climate system,thus characterizing its large scale properties, and gatherinformation on tipping points. In Sec. V, we addressthe non-equilibrium statistical mechanics formulation ofclimate dynamics, and explore how the formalism of re-sponse theory allows for addressing in a rigorous frame-work the climatic response to perturbations, taking in-spiration from the work of (Ruelle, 1997). In Sec. VI,we present averaging and homogenization techniques, de-scribe how projector operator methods due to (Mori ,1965) and (Zwanzig , 1961) provide powerful tools for de-riving parametrizations and firm ground to the inclusionof stochastic terms and memory effects, and discuss howresponse theory can be used to derive similar results.

Finally, in Sec. VII we draw our conclusions and

Page 3: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

3

present some perspectives of future research.

II. BEYOND THE HAMILTONIANPARADIGM: NAMBU REPRESENTATIONOF GEOPHYSICAL FLUID DYNAMICS

A. Introduction

Hamiltonian formalism constitutes the backbone ofmost physical theories. In the case of a discrete sys-tem, the basic idea is to provide a full description ofthe degrees of freedom by defining a set of canonicalvariables q and of the related momenta p (q and p be-ing N -dimensional vectors), and by identifying the timeevolution to a flow in phase space such that the canon-ical Hamiltonian system H acts as a stream-function,q = ∂pH, p = −∂qH. H(q, p) corresponds to the en-ergy of the system. The flow is inherently divergencefree (solenoidal), so that the phase space does not con-tract nor expands, as implied by the Liouville Theorem.As suggested by Noether’s theorem, the presence of sym-metries in the system implies the existence of so-calledphysically conserved quantities Xi, such that Xi = 0. Ina system possessing time invariance the energy is con-stant, while in a system possessing translational invari-ance, the total momentum M is also constant. A systemcan possess many constant of motions, apart from energy,but the Hamiltonian plays a special role as it is the onlyfunction of phase space appearing explicitly in the defini-tion of the evolution of the system (Landau and Lifshits,1996).

Nambu (1973) presented a generalization of canonicalHamiltonian theory for discrete systems. The dynamicalequations are constructed in order to satisfy Liouville’sTheorem and are written in terms of two or more con-served quantities. The Nambu approach has been ex-tremely influential in various fields of mathematics andphysics and is viable to extension to the case of con-tinuum, so that it can be translated into a field the-ory. The construction of a Nambu field theory for geo-physical fluid dynamics went through two decisive steps.The first was the discovery of a Nambu representationof 2D and 3D incompressible hydrodynamics (Nevir andBlender , 1993). The second important step was the find-ing that the Nambu representation can be used to designconservative numerical algorithms in geophysical models,and that classical heuristic methods devised by Arakawafor constructing accurate numerical models actually re-flected deep symmetries coming from the Nambu struc-ture of the underlying dynamics of the flow (Salmon,2005).

The physical basis for the relevance of the Nambutheory for describing and simulating conservative geo-physical fluid dynamics, i.e. stratified and rotating flowscomes from the existence of relevant conservation prin-

ciples other than energy’s. The first one is the mate-rial conservation of potential vorticity, and the secondone is the particle relabeling symmetry. These proper-ties are independent from the particular approximatedmodel of the geophysical flow, and are valid for 2D and3D hydrodynamics, Rayleigh-Benard convection, quasi-geostrophy, shallow water model, and extends to the fullybaroclinic 3D atmosphere. In other terms, the Namburepresentation provides the natural description of geo-physical fluid dynamics and is superior to the more tra-ditional approaches based essentially on Navier-Stokesequations, just the action-angle representation of the dy-namics of a spring is superior to the simple descriptionprovided by the second Newton’s law of motion.

B. Hydrodynamics in 2D and 3D

In incompressible hydrodynamics enstrophy (in 2D)and helicity (3D) are known as integral conserved quan-tities besides energy (Kuroda, 1991). These two quan-tities originate in the particle relabeling symmetry inthe Lagrangian description of fluid dynamics. Nevir andBlender (1993) adapted Nambu’s formalism to incom-pressible nonviscous hydrodynamics by using enstrophyand helicity in the dynamical equations.

1. Two-dimensional hydrodynamics

In two dimensions incompressible hydrodynamics isgoverned by the vorticity equation

∂ζ

∂t= −u · ∇ζ (1)

with ζ = vx − uy and ∇ · u = 0. The Hamiltonian is thekinetic energy

H =1

2

∫u2 dA = −1

2

∫ζψ dA (2)

where ψ is the stream-function for u, u = k × ∇ψ (kdenotes the z-unit vector).

The Hamiltonian H is a functional of velocity. In gen-eral functionals are extensive functions in phase spaceand defined as maps assigning functions to numbers. Inthe dynamical equations the functional derivative δH/δζis needed which describes the change of the functionalH with respect to a changing dynamic variable ζ. Thedependency can be explicitly denoted as H[ζ]. The func-tional derivative of a functional G[f ] for a function f(x)is defined for a small variation δf in the linear expansion

G[f + δf ] = G[f ] +

∫δGδf(x)

δf(x)dx+ . . . (3)

Throughout this review we assume periodic boundaryconditions.

Page 4: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

4

The functional derivative δH/δζ for (2) is explicitlycalculated by

δH =

∫∇ψ · δ∇ψ dA

=

∫∇ · (∇δ∇ψ) dA−

∫ψδζ dA (4)

Since the first integral vanishes due to the boundary con-ditions, we obtain δH/δζ = −ψ.

Based on the material advection of the vorticity (1)any functional of the vorticity is conserved

C =

∫f(ζ) dA (5)

among these the most well-known is enstrophy

E =1

2

∫ζ2 dA (6)

The functional derivative of the enstrophy is simplyδE/δζ = ζ.

The 2D vorticity equation can be expressed in a Nambuform using the enstrophy E

∂ζ

∂t= −J

(δEδζ,δHδζ

)(7)

with the Jacobi operator

J (a, b) = ∂xa ∂yb− ∂ya ∂xb (8)

which is an anti-symetric bracket for the two arguments.Using the aforementioned functional derivatives the vor-ticity equation (1) is recovered as

∂ζ

∂t= J (ζ, ψ) (9)

The Nambu form (7) represents a field theoretic exten-sion of the dynamics introduced by Nambu (Nambu,1973) which uses additional conservation laws in the dy-namical equations besides the Hamiltonian.

In the following the relationships between Nambu me-chanics and Hamiltonian theory are briefly summarized.The time-evolution of an arbitrary functional of vorticityF = F [ζ] is determined by

∂F∂t

= −∫δF∂ζJ(δEδζ,δHδζ

)dA

= F , E ,H (10)

which defines a Nambu bracket for the three functionalsinvolved. The bracket is anti-symmetric in all arguments,E ,H,F = −H, E ,F, etc. In general a bracket is ananti-symmetric map in the space of functions. Using re-arrangements of these functionals and partial integrationit can be shown that the Nambu bracket is cyclic

F , E ,H = E ,H,F = H,F , E (11)

The cyclicity of this bracket is a main ingredient inSalmon’s application of Nambu mechanics (Salmon,2005) to construct conservative numerical codes (see Sec-tion II.C.2).

A system is Hamiltonian if its dynamics can be writtenas

∂F∂t

= F ,HP (12)

with an antisymmetric Poisson bracket. Additionally thebracket has to satisfy the Jacobi identity (see (Takhtajan,1994) for the extension to Nambu brackets).

The Poisson bracket for 2D hydrodynamics (Salmon,1988; Shepherd , 1990) is easily obtained from the Nambubracket if the dependency δE/δζ = ζ is evaluated

F ,HP = F , E ,H =

∫ζJ (Fζ ,Hζ) dA (13)

with the short cut Hζ for the functional derivative (herecyclicity is used (11)).

The Poisson bracket used in Eulerian hydrodynamicsis degenerate leading to the notion noncanonical Hamil-tonian mechanics. Degeneracy is equivalent to the exis-tence of a so-called Casimir function C[ζ] for which thePoisson bracket vanishes for any function f [ζ]

C, fP = 0 (14)

Casimir functions are automatically conserved due toC, HP = 0. In 2D hydrodynamics Casimirs are givenby integrals of arbitrary functions of the vorticity sincethis annuls (14). Noncanonical Hamiltonian dynamicshas been extensively used to investigate non-dissipativehydrodynamics and applications pertain to nonlinear sta-bility analysis, symmetries, and approximation theory(for reviews see (Salmon, 1988; Shepherd , 1990)).

The relationship (13) demonstrates that noncanonicalHamiltonian mechanics is embedded in Nambu mechan-ics. The main extension is that in Nambu mechanics twoHamiltonians, the enstrophy and the energy, are used (7),and that the Nambu bracket (10) is nondegenerate andvoid of Casimir functions.

2. Three-dimensional hydrodynamics

Incompressible inviscid fluid dynamics in 3D is deter-mined by the vorticity equation for ξ = ∇× u

∂ξ

∂t= ξ · ∇u− u · ∇ξ (15)

and ∇ · u = 0, with the velocity u. Total energy

H =1

2

∫u2 dV = −1

2

∫ξ ·AdV (16)

Page 5: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

5

and helicity

h =1

2

∫ξ · udV (17)

are conserved. A is the vector potential, u = −∇ ×A,with ∇ · u = 0. The functional derivative of the energywith respect to the vorticity is given by δH/δξ = −Aand for helicity δh/δξ = u (compare the 2D version (4)).

The Nambu form of the vorticity equation is

∂ξ

∂t= K

(δh

δξ,δH

δξ

)(18)

with

K(A,B) = −∇× [(∇×A)× (∇× B)] (19)

A functional F = F [ξ] evolves according to

∂F∂t

= −∫ (∇× δF

δξ

)×(∇× δh

δξ

)·(∇× δH

δξ

)dV = F, h,H (20)

The last equation defines the Nambu bracket for 3D hy-drodynamics based on the vorticity equation. Helicityis no longer a hidden conserved quantity but enters thedynamics on the same level as the Hamiltonian.

C. Geophysical fluid dynamics

In this section we summarize Nambu representationsof the most important models in geophysical fluid dy-namics: the quasi-geostrophic potential vorticity equa-tion (Nevir and Sommer , 2009), the shallow water model(Salmon, 2005; Sommer and Nevir , 2009), Rayleigh-Benard equations for two-dimensional convection (Bihlo,2008; Salazar and Kurgansky , 2010), and the baroclinicstratified atmosphere (Nevir and Sommer , 2009).

1. Quasi-geostrophic approximation

Quasi-geostrophic dynamics in absence of dissipativeprocesses and of forcings is determined by the materialconservation of the quasi-geostrophic potential vorticity

∂Q

∂t+

1

f0J(Φ, Q) = 0 (21)

where J is the Jacobian (8). Q is the quasi-geostrophicapproximation of Ertel’s potential vorticity

Q = ζg +f0

σ0

∂2Φ

∂p2+ f (22)

with the geostrophic vorticity ζg = 1/f0∇2hΦ, geopoten-

tial Φ, stability parameter σ0, and Coriolis parameter f .

The first conserved integral is the total energy

H =1

2

∫ [(∇hΦ

f0

)2

+

(1

N

∂Φ

∂z

)2]

dV (23)

where the first term is the density of kinetic energy andthe second term is the density of potential energy Holton(2004). The second conserved integral is potential en-strophy

E =1

2

∫Q2dV (24)

with the Brunt-Vaisala frequency N . The dynamics isdetermined by the Nambu bracket

∂F∂t

= −∫δFδQ

J

(δEδQ

,δHδQ

)dV = F , E ,H (25)

which is identical to (10). The Nambu representation ofthe quasi-geostrophic potential vorticity equation is

∂Q

∂t= −J

(δEδQ

,δHδQ

)(26)

Thus, the mathematical structure is analogous to thetwo-dimensional vorticity equation (10).

2. Shallow water model

The Nambu representation of the shallow water modelwas derived by Salmon (2005). Sommer and Nevir(2009) present a numerical simulation of these equationson a spherical grid, and Nevir and Sommer (2009) pub-lished the multilayer shallow water equations. Here thesingle layer model is summarized (Sommer and Nevir ,2009). The dynamics can be constructed on the evolu-tion equation for the vorticity for the divergence µ = ∇·v

∂tζ = −∇ · (ζav) (27)

∂tµ = k · ∇ × (ζav) +∇2(v2/2 + gh) (28)

The shallow water model possesses two conserved inte-grals, the total energy, given by the sum of kinetic andpotential energy

H =1

2

∫ (hv2 + gh2

)dV (29)

and potential enstrophy

E =1

2

∫hq2dV (30)

with the absolute potential vorticity q = ζa/h, ζa = ζ+f ,and the height of the fluid h.

Page 6: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

6

The dynamics of any functional F is determined bythe sum of three Nambu brackets

∂tF = F ,H, Eζ,ζ,ζ + F ,H, Eµ,µ,ζ + F ,H, Eζ,µ,h(31)

Here Fζ = δF/δζ, etc. The first bracket is

F , H, Eζ,ζ,ζ =

∫J(Fζ , Hζ)EζdA (32)

which is analogous to the 2D Nambu bracket (10). Forthe other brackets we refer to (Salmon, 2005; Sommerand Nevir , 2009). Salmon (2007) calculated the Nambubrackets based on the velocities instead of vorticity.

3. Rayleigh-Benard convection

Here the equations for the two-dimensional Rayleigh-Benard convection are summarized (see (Bihlo, 2008),Salazar and Kurgansky (2010) studied the three-dimensional problem). The fluid flow is incompressibleand obeys the Boussinesq approximation in a nondimen-sional form

∂ζ

∂t+ [ψ, ζ] =

∂T

∂x,

∂T

∂t+ [ψ, T ] =

∂ψ

∂x(33)

where [a, b] = ∂xa ∂zb − ∂za ∂xb denotes the Jacobianoperator. The equations preserve the Hamiltonian

H =

∫ (1

2(∇ψ)2 − Tz

)dxdz (34)

where, as usual, the term in the integral indicate thedensity of kinetic and potential energy, respectively, andthe Casimir

C =

∫ζ(T − z)dxdz (35)

This Casimir is a consequence of Kelvin’s circulation the-orem. The Nambu form of the coupled equations areexplicitly written as

∂ζ

∂t= −

[δCδT

,δHδζ

]−[δCδζ,δHδT

]= ζ, C,H (36)

∂T

∂t=

[δCδζ,δHδζ

]= T, C,H (37)

with the bracket

F ,G,H = −∫ (

δFδT

[δGδζ,δHδζ

]+δFδζ

[δGδT

,δHδζ

]+δFδζ

[δGδζ,δHδT

])dxdz (38)

Note that the Casimir (35), as helicity (17), has no defi-nite sign.

4. Baroclinic atmosphere

Nevir and Sommer (2009) published the equations de-termining the dynamics of a baroclinic atmosphere inNambu form (denoted as Energy-vorticity theory of idealfluid mechanics). The Nambu representation encom-passes the Eulerian equation of motion in a rotatingframe, the continuity equation, and the first law of ther-modynamics. The Nambu dynamics uses three brack-ets for energy helicity, energy-mass, and energy-entropy.Due to it’s special role in all three brackets the integral ofErtel’s potential enstrophy is coined as a super-Casimir.

The Nambu form shows an elegant structure wherefundamental processes are combined by additive terms.Incompressible, barotropic or baroclinic atmospheres areassociated to additive contributions. Thus approxima-tions are simply attained by the neglect of terms.

The momentum equation, the continuity equation andthe first law equation are

∂tv = v, ha, Hh + v,M,Hm + v, S,Hs (39)

∂tρ = ρ,M,Hm (40)

∂tσ = σ, S,Hs (41)

with σ = ρs, s being the entropy density.There are four conservation laws. The first is the total

energy

H =

∫ 1

2ρv2 + ρe+ ρΦ

dV (42)

e is the specific internal energy and Φ the potential ofthe gravitational and the centrifugal force. The absolutehelicity is

ha =

∫va · ξadV (43)

where the absolute velocity is va = v+Ω×r with angularvelocity Ω. Mass and entropy are

M =

∫ρdV, S =

∫ρsdV (44)

and potential enstrophy is with the potential vorticity Π

Eρ =

∫ρΠ2dV, Π =

ξa · ∇θρ

(45)

The three brackets are

F , ha,Hh = −∫ [

1

ρ

δFδv·(δhaδv× δHδv

)]dV (46)

F ,M,Hm = −∫ [

δMδρ

δFδv· ∇δH

δρ

+δFδρ∇ ·(δMδρ

δHδv

)]dV + cyc(F ,M,H) (47)

Page 7: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

7

F ,S,Hs = −∫ [

δSδρ

δFδv· ∇δH

δσ

+δFδσ∇ ·(δSδρ

δHδv

)]dV + cyc(F ,M,H) (48)

For a barotropic flow, the first law of thermodynamicsand therefore the energy-entropy bracket are omitted.The energy-entropy bracket has to be removed in the Eu-lerian equation of motion, leading to the energy-vorticityrepresentation of a compressible barotropic fluid. Notethe different brackets for velocity (46) and vorticity (20)in 3D hydrodynamics.

D. Conservative algorithms and numerical models

Salmon (2005, 2007) recognized that the existenceof a Nambu bracket with two conserved integrals al-lows the design of conservative numerical algorithms.The single necessary ingredient is that the discrete formof the Nambu bracket preserves antisymmetry. Forthe barotropic vorticity equation the Arakawa Jacobiancould be retrieved by equally weighting the cyclic permu-tations of the Nambu bracket. The approach is applicableto any kind of discretization, e.g. for finite differences,finite volumes, or spectral models. Furthermore, arbi-trary approximations of the conservation laws are pos-sible. The approach is extremely useful in GFD turbu-lence simulations because these flows are characterizedby the existence of conservation laws besides total en-ergy. In particular, the conservation of enstrophy inhibitsspurious accumulation of energy at small scales. Salmon(2007) presents the first numerical simulation of a shallowwater model by equations derived from Nambu brackets.The simulation is on a square rectangular grid and thedesign on an unstructured triangular mesh is outlined.

Sommer and Nevir (2009) report the first simulationof a shallow water atmosphere using Nambu brackets.The authors use an isosahedric grid (as in the ICONmodel, ICOsahedric Non-hydrostatic model, of the Ger-man Weather Service and the Max Planck Institute forMeteorology, Hamburg). The construction of the algo-rithm is as follows (Sommer and Nevir , 2009):

• First the continuous versions of the Nambu-brackets and conservation laws need to be given.

• On the grid the following expressions need to becalculated: functional derivatives, discrete opera-tors (div and curl), discretization of the Jacobianand the Nambu brackets.

• Finally the prognostic equations are obtained byinserting the variables in the brackets. The timestepping is arbitrary ((Sommer and Nevir , 2009)use a leap-frog with Robert-Asselin filter).

FIG. 1 Enstrophy tendencies in the enstrophy conservingICON shallow water model and the Nambu model of Sommerand Nevir (2009) (courtesy of Matthias Sommer, Universityof Munchen). Note that the tendency in the Nambu model isof the order of the numerical accuracy.

The authors find quasi constant enstrophy and energycompared to a standard numerical design (Fig. 1).

Gassmann and Herzog (2008) suggest a concept for aglobal numerical simulation of the non-hydrostatic atmo-sphere using the Nambu representation for the energy-helicity bracket F , ha,H (Nevir , 1998). Their sugges-tion incorporates a careful description of Reynolds aver-aged subscale processes and budgets. Gassmann (2013)describes a global non-hydrostatic dynamical core basedon an icosahedral nonhydrostatic model on a hexago-nal C-grid. The model conserves mass and energy in anoncanonical Hamiltonian framework. This study notesthat there are unsolved numerical problems which occurwhen the non-hydrostatic compressible equations are ina Nambu bracket form.

E. A phase space view: Lorenz-63 model

(Nambu, 1973) introduced his extension of classi-cal Hamiltonian dynamics on the basis of a dynami-cal system with three degrees of freedom and two con-served quantities. Here we consider the Lorenz-63 model(Lorenz , 1963) to present a geometric view of Nambudynamics in 3D space (for a 10 degree truncation see(Blender and Lucarini , 2013)). The conservative partsof the equations are extracted by eliminating the forcingand dissipation terms which violate Liouville’s Theorem(Nevir and Blender , 1994).

In order to distinguish the conservative and dissipa-tive contributions in the Lorenz-63 model a parameter mis introduced which controls the magnitude of the non-

Page 8: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

8

conservative terms (m = 1 is the standard case)

x = σy − σmx (49)

y = rx− xz −my (50)

z = xy −mbz (51)

The equations (49)-(51) with m = 0 have two con-served quantities, which can be found by integrating thetwo equations dy/dz = (r − z)/y and dz/dy = x/σ

H =y2

2+z2

2− rz, C =

x2

2− σz (52)

H represents available potential energy and C is totalenergy (see Fig. 2).

Using the two conserved quantities H and C in (52)we can write (49)-(51) in a Nambu representation forX = (x, y, z), ∇ = (∂x, ∂y, ∂z)

X = ∇C ×∇H (53)

The system satisfies the Liouville Theorem because theflow X is solenoidal in state space

∇ · X = ∇ · (∇C ×∇H) = 0 (54)

The flow is tangent to both conserved surfaces and paral-lel to the intersection. Thus, Nambu mechanics providesan illuminating global geometric view of the dynamics.Roupas (2012) visualizes the geometry of the conserva-tion laws H and C for different parameters.

FIG. 2 Conservation laws C and H in the Nambu representa-tion (Eq. 52) of the non-dissipative Lorenz equations (cour-tesy of Annette Muller, Freie Universitat Berlin). Parametersof the model: r = 1 and σ = 1.

F. Perspectives

Like Hamiltonian mechanics, the Nambu approach isa versatile tool for the analysis and simulation of dynam-ical systems. Here some possible research directions areoutlined.

Modular modeling and approximations: In sev-eral applications a Nambu representation can be found byadding brackets which conserve a particular Casimir, thisis already mentioned by Nambu (1973) (see the baroclinicatmosphere (Nevir and Sommer , 2009) and the classifi-cation in (Salazar and Kurgansky , 2010)). The dynam-ics is determined by these ’constitutive’ Casimirs (a no-tion coined in (Nevir and Sommer , 2009)) which are notconserved in the complete system. Decomposition leadsto subsystems where the constitutive Casimirs are con-served. An example is helicity which is constitutive inthe baroclinic atmosphere and only conserved in the 3Dincompressible flow. The decomposition is directly asso-ciated with approximations (Nevir and Sommer , 2009).Composition allows a process-oriented model design.

Statistical Mechanics: The statistical mechanics offluids is characterized by the existence of conservationlaws besides total energy (Bouchet and Venaille, 2012),see also Sect. III in this review. Thus these conserva-tion laws have a two-fold impact: They determine thedynamics in a Nambu bracket and the canonical proba-bility distribution in equilibrium.

Dynamics of Casimirs: Casimir-functions of a con-servative system are ideal observables to characterizethe dynamics in the presence of forcing and dissipation.Pelino and Maimone (2007) and Gianfelice et al. (2012)have determined recurrence maps of extremes of energyand a Casimir in a Lorenz-like map to assess predictabil-ity and properties of the invariant measure.

III. EQUILIBRIUM STATISTICAL MECHANICS FORGEOPHYSICAL FLOWS

A. Introduction

The basic idea of equilibrium statistical mechanics is tobuild a probability measure for the system under consid-eration, based only on the dynamical invariants. In par-ticular, this is an invariant measure for the original dy-namics, provided the Liouville theorem is satisfied (vol-ume in phase space is conserved by the dynamics). Thisprobability measure allows one to compute the statis-tics of any function of the dynamical variables, which isexpected to coincide with the macroscopic behavior ofthe system, without knowing the details of the micro-scopic dynamics. This approach, initiated at the end ofthe 19th century by Boltzmann and Gibbs, has provenvery fruitful in classical and quantum mechanics all alongthe 20th century. However, most of the standard ap-plications of equilibrium statistical mechanics deal withdynamics on a finite dimensional phase space (e.g. fi-nite number of molecules for ideal gas, or finite numberof spins for ferromagnetism models), with a finite num-ber of dynamical invariants. The equations describingthe dynamics of geophysical flows violate both these con-

Page 9: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

9

straints. Several solutions have thus been proposed: theyare reviewed briefly in the next sections, going from themain fundamental ideas1 to selected geophysical applica-tions2. The predictions of the theories take the form ofequilibrium states, that is to say the large-scale organi-zation of the flow, and equilibrium spectra for conservedquantities. Equilibrium states include a vast collection ofcoherent structures, which are indeed observed in natureand numerical experiments. Equilibrium spectra revealthe dominant directions of invariant quantities transfersacross the scales of motion. In both cases, the equilibriummethods indicate the natural tendency of the nonlinearevolution. In practice, the system may never reach theseequilibrium features, either because of non-ergodicity, orbecause it is subjected to forcing and dissipation whichare not accounted for in this approach. It is worthy ofnote that, in spite of this, the equilibrium approach stillyields useful results to understand the out-of-equilibriumsystem. Note that the response theory applied to equilib-rium states, as first suggested by (Kubo, 1957), providesa connection with the non-equilibrium approach, whichconstitutes a first step to understand departures fromequilibrium conditions.

In section III.B, we give a brief account of the pointvortex approach introduced by Onsager. Then we dis-cuss the geophysical ramifications of the pioneer work byKraichnan on Galerkin truncated inviscid flows (sectionIII.C), before introducing the theory developed by Miller,Robert and Sommeria, which treats continuous flows andtreats all the invariants (section III.D).

B. Point vortices

1. Negative temperature states and clustering of vortices

Onsager was the first to understand that the coherentstructures and persistent circulations that appear ubiq-uitously in planetary atmospheres and in the Earth’soceans could be explained on statistical grounds (On-sager , 1949). His work focused on 2D incompressible,inviscid fluids: ∂tu + u · ∇u = −∇P,∇ · u = 0, orequivalently in vorticity form, ∂tω + u · ∇ω = 0, withω = ∇ × u. To make the system tractable, he intro-duced an approximation of the vorticity field in termsof N point vortices with strength γi and position ri(t):

ω(r, t) =∑Ni=1 γiδ(ri(t) − r). Introducing the Hamilto-

1 Equilibrium statistical mechanics theories for geophysical fluidshave also led to progress in statistical mechanics itself, throughthe theoretical issues of ensemble equivalence, invariant mea-sures, ergodicity,... Due to length constraints, these theoreticaldevelopments are not reviewed here.

2 Although there are also interesting applications in magnetohy-drodynamics, potentially relevant for astrophysical flows, thistopic is left aside in the present review.

nian H = −∑i<j γiγjG(ri, rj), where G is the Green

function of the Laplacian, the dynamics reads simply

γidxidt

=∂H

∂yi, γi

dyidt

= −∂H∂xi

. (55)

This is a canonical Hamiltonian system with a finite num-ber of degrees of freedom, for which the standard meth-ods of statistical mechanics apply directly. In particu-lar, the microcanonical probability measure, associatinga uniform probability to all the configurations with thesame energy, is given by

ρE(ri1≤i≤N ) =δ(H(ri1≤i≤N )− E)

Ω(E), (56)

where Ω(E) is the structure function, which measuresthe volume in phase space occupied by configurationswith energy E. It is easily proved that, for a boundeddomain, and hence a finite volume phase space, this func-tion reaches a maximum for a given value of the energy.Hence, the thermodynamic entropy S(E) = kB ln Ω(E)decreases for a range of energies, and the statistical tem-perature 1/T = ∂S/∂E becomes negative. Negative tem-peratures, although counter-intuitive, have since beencommonly encountered in the study of systems with long-range interactions (Dauxois et al., 2002), and correspondto self-organized states. Here, the energy increases whentwo same-sign vortices move closer, while it decreases foropposite signs. When the temperature is negative, theentropy increases when energy increases, and there is noenergy-entropy competition, which results in the appear-ance of ordered states. Hence, equilibrium states exhibit-ing clusters of same-sign vortices are expected. This be-havior has been confirmed by numerical simulations withup to N = 6724 point vortices: see Fig. 3.

a small scale to a large scale [16]. In Fig. 5, the initial con-ditions are E ! 24:1"# E2 <Ec$ and I1 ! %1:17&10%13 ' 0 so that the temperature is positive. The vorticesdiffuse over the circular cross section, and the positiveand the negative vortices mix almost uniformly. This is acommon feature observed in systems with positive tem-perature and corresponds to a normal cascade.

Snapshots of the asymptotic states corresponding todifferent values of energy are shown in Fig. 6. The inversetemperatures !2, !3, and !4 associated with the systemenergy E2, E3, and E4, respectively, satisfy !2 > 0> !3 >!4. The small clumps (spots) are created as a result of thevortex condensation in the negative temperature cases inFigs. 6(b) and 6(c). It is noticeable that the size of theclump depends on the temperature. The size becomeslarger as the inverse temperature ! goes down at !< 0.

The other essential feature of the negative temperaturesystem is the energy concentration in a specific group ofthe vortices. There is no viscous term in the equation ofmotion (4), and the reduction of the total energy due to thenumerical dissipation is evaluated to be less than 0.1% inFig. 4. Thus, the total energy is conserved in the simulation.If the energy of a part of the system increases, the energy ofthe rest should decrease. In Fig. 4, ‘‘a part of the system’’ isthe vortices inside the clumps, and the rest is the back-ground vortices (outside the clumps). We evaluate the timeevolution of the energies belonging to the clumps and thebackground vortices for the case in Fig. 4. The result isshown in Fig. 7. The energy of the vortices that constitutes

the clumps certainly increases during the initial period upto " ! 20 when the formation of almost circular clumps isfinished. On the other hand, the energy belonging to thebackground vortices goes down. The time traces of theenergies belonging to the two groups are symmetric aboutthe line representing the ratio of 0:5. After " ! 20, theenergy of each group remains constant. It is noticeable thatthe exchange of the energy between the two groups of thevortices enables the vortex condensation. This observationindicates the common and essential role of backgroundvortices in supporting the condensation of two-sign vorti-ces as well as in assisting the generation of symmetricconfiguration of non-neutral plasma clumps [9,12–15].

Let us briefly examine the mechanism to determinethe asymptotic configuration of the two clumps by em-ploying a simple model. Suppose that the clumps areconfined uniformly within two circular patches of ra-dius 0:2R that are separated by the distance of 2r acrossthe center of the circular boundary of radius R. A calcu-lation indicates that the energy of the clump system ismaximized for r ' 0:5R. This maximum-energy configu-ration has a noticeable similarity to the distribution ob-tained at " ! 32 in Fig. 4. Though the contribution of thebackground vortices is neglected by assuming they aresufficiently neutralized, the model consideration suggeststhat the clumps tend to share the maximum energy avail-able under the constraint of conservation of the systemenergy and the angular impulse. It is consistent with thecommon feature of the negative temperature system.

In summary, we have numerically examined the dynam-ics of two-sign point vortices confined in the circularboundary. The state of the vortex system is characterized

FIG. 4 (color online). Time evolution of the negative tempera-ture system is plotted. The values of the parameters are N !6724, nc ! 4, Rc ! 0:5R, rc ! 0:2R, "c ! 15:6, and E !2:69& 104"! E1$.

FIG. 6 (color online). Snapshots of different energies(a) E2 ! 24:1, (b) E3 ! 142, and (c) E4 ! 315 at " ! 128 areshown.

FIG. 5 (color online). Time evolution of the positive tempera-ture system is plotted. The values of the parameters are N !6920, nc ! 20, Rc ! 0:2R, rc ! 0:07R, "c ! 9:3, and E !24:1"! E2$. Because the clumps overlap each other, we seeonly the outline of the clumps (hollow ring) at " ! 0.

FIG. 3. Critical energy Ec and width at half maximum areplotted as a function of N. In each case, the number of samplingis 107.

PRL 94, 054502 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending11 FEBRUARY 2005

054502-3

a small scale to a large scale [16]. In Fig. 5, the initial con-ditions are E ! 24:1"# E2 <Ec$ and I1 ! %1:17&10%13 ' 0 so that the temperature is positive. The vorticesdiffuse over the circular cross section, and the positiveand the negative vortices mix almost uniformly. This is acommon feature observed in systems with positive tem-perature and corresponds to a normal cascade.

Snapshots of the asymptotic states corresponding todifferent values of energy are shown in Fig. 6. The inversetemperatures !2, !3, and !4 associated with the systemenergy E2, E3, and E4, respectively, satisfy !2 > 0> !3 >!4. The small clumps (spots) are created as a result of thevortex condensation in the negative temperature cases inFigs. 6(b) and 6(c). It is noticeable that the size of theclump depends on the temperature. The size becomeslarger as the inverse temperature ! goes down at !< 0.

The other essential feature of the negative temperaturesystem is the energy concentration in a specific group ofthe vortices. There is no viscous term in the equation ofmotion (4), and the reduction of the total energy due to thenumerical dissipation is evaluated to be less than 0.1% inFig. 4. Thus, the total energy is conserved in the simulation.If the energy of a part of the system increases, the energy ofthe rest should decrease. In Fig. 4, ‘‘a part of the system’’ isthe vortices inside the clumps, and the rest is the back-ground vortices (outside the clumps). We evaluate the timeevolution of the energies belonging to the clumps and thebackground vortices for the case in Fig. 4. The result isshown in Fig. 7. The energy of the vortices that constitutes

the clumps certainly increases during the initial period upto " ! 20 when the formation of almost circular clumps isfinished. On the other hand, the energy belonging to thebackground vortices goes down. The time traces of theenergies belonging to the two groups are symmetric aboutthe line representing the ratio of 0:5. After " ! 20, theenergy of each group remains constant. It is noticeable thatthe exchange of the energy between the two groups of thevortices enables the vortex condensation. This observationindicates the common and essential role of backgroundvortices in supporting the condensation of two-sign vorti-ces as well as in assisting the generation of symmetricconfiguration of non-neutral plasma clumps [9,12–15].

Let us briefly examine the mechanism to determinethe asymptotic configuration of the two clumps by em-ploying a simple model. Suppose that the clumps areconfined uniformly within two circular patches of ra-dius 0:2R that are separated by the distance of 2r acrossthe center of the circular boundary of radius R. A calcu-lation indicates that the energy of the clump system ismaximized for r ' 0:5R. This maximum-energy configu-ration has a noticeable similarity to the distribution ob-tained at " ! 32 in Fig. 4. Though the contribution of thebackground vortices is neglected by assuming they aresufficiently neutralized, the model consideration suggeststhat the clumps tend to share the maximum energy avail-able under the constraint of conservation of the systemenergy and the angular impulse. It is consistent with thecommon feature of the negative temperature system.

In summary, we have numerically examined the dynam-ics of two-sign point vortices confined in the circularboundary. The state of the vortex system is characterized

FIG. 4 (color online). Time evolution of the negative tempera-ture system is plotted. The values of the parameters are N !6724, nc ! 4, Rc ! 0:5R, rc ! 0:2R, "c ! 15:6, and E !2:69& 104"! E1$.

FIG. 6 (color online). Snapshots of different energies(a) E2 ! 24:1, (b) E3 ! 142, and (c) E4 ! 315 at " ! 128 areshown.

FIG. 5 (color online). Time evolution of the positive tempera-ture system is plotted. The values of the parameters are N !6920, nc ! 20, Rc ! 0:2R, rc ! 0:07R, "c ! 9:3, and E !24:1"! E2$. Because the clumps overlap each other, we seeonly the outline of the clumps (hollow ring) at " ! 0.

FIG. 3. Critical energy Ec and width at half maximum areplotted as a function of N. In each case, the number of samplingis 107.

PRL 94, 054502 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending11 FEBRUARY 2005

054502-3

FIG. 3 Time evolution for a numerical simulation of two-sign point vortices, for positive temperatures (upper panel)and negative temperatures (lower panel) (Yatsuyanagi et al.,2005). For negative temperatures, we observe the clusteringof same-sign vortices.

Page 10: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

10

2. Mean-field equation

The above argument is qualitative; to characterize thecoherent structures which are expected to emerge fromthe clustering of same-sign vortices, we introduce themean field probability density for vorticity: ρ(σ, r, t) =(1/N)

∑j δ(σ − γj)δ(r − rj(t)). The probability den-

sity for a vortex with strength γi to be found at pointr at time t is just ρi(r, t) = ρ(γi, r, t)/nγi , where thenormalization factor nγ =

∑j δγjγ is needed so that∑

i

∫ρi(r, t)dr = 1. We define a coarse grained vor-

ticity field ω(r, t) =∫σρ(σ, r, t)dσ =

∑i γiρi(r, t).

The mean-field probability density is expected to con-verge towards its statistical equilibrium: the equilib-rium distribution maximizes the statistical entropy S =−∫ρ(σ, r) ln ρ(σ, r)dσdr = −

∑i

∫ρi(r) ln ρi(r)dr. The

solution of this variational problem is given by ρi(r) =

e−β(γiψ(r)−µi)/Z where β and βµi are the Lagrange pa-rameters associated with conservation of global energyand normalization of each ρi, respectively, and ψ =−∆−1ω is the coarse-grained stream function, while thenormalization factor Z is called the partition function.Averaging over this equilibrium distribution gives thecoarse-grained vorticity field, which satisfies the mean-field equation:

ω(r) =1

Z∑i

γie−β(γiψ(r)−µi). (57)

This is an equation of the form ω = F (ψ), characteris-tic of the steady-states of the 2D Euler equations. Awell-known particular case is that of N vortices withcirculation 1/N and N vortices with circulation −1/N .In that case, the mean-field equation can be recast asω = A sinh(βΨ), with Ψ = ψ+(µ+−µ−)/2 (Montgomeryand Joyce, 1974).

The theory can be generalized in a straightforwardmanner to quasi-geostrophic (QG) flows (Miyazaki et al.,2011). (DiBattista and Majda, 2001) have given solutionsof the mean-field equation for a two-layer model wherethe point vortices stand for hetons, introduced by (Hoggand Stommel , 1985) as a model of individual convectivetowers in the ocean. They have shown that a backgroundbarotropic current (the barotropic governor) confines po-tential vorticity and temperature anomalies, thereby sup-pressing the baroclinic instability, in accordance with nu-meric simulations (Legg and Marshall , 1993).

C. Galerkin truncated flows

1. 2D Turbulence

Rather than a discretization in physical space, onemay consider a finite number of modes in Fourier space,as proposed by (Lee, 1952) and (Kraichnan, 1967) inthe context of the Euler equations. For 2D flows —

for simplicity, we consider here a rectangular geome-try with periodic boundary conditions; the case of aspherical geometry can be found in (Frederiksen andSawford , 1980) — writing the vorticity field as a trun-cated Fourier series ω(x) =

∑k ω(k)eik·x, the evolution

in time of the Fourier coefficients follows an equationof the form ∂tω(k) =

∑p,qAkpqω(p)ω(q), where the

summation is restricted to a finite set of wave vectorsB = k ∈ 2π/LZ3, kmin ≤ k ≤ kmax. This dynam-ics preserves two quadratic quantities: the energy E =∑

k|ω(k)|2/(2k2) and the enstrophy Γ2 =∑

k|ω(k)|2.(Kraichnan, 1967) suggested to consider the canonicalprobability distribution for the Galerkin truncated sys-tem:

ρ(ω(k)k∈B) =e−βE−αΓ2

Z, (58)

In particular, the average energy at absolute equilibriumis given by

〈E〉 = −∂ lnZ∂β

=1

2

∑k∈B

1

β + 2αk2, (59)

which corresponds to an equipartition spectrum for thegeneral invariant βE + αΓ2: E(k) = πk/(β + 2αk2). In-viscid numerical runs indeed relax to this spectrum (Bas-devant and Sadourny , 1975; Fox and Orszag , 1973). Notethat the Lagrange parameters α and β cannot take ar-bitrary values; they are constrained by the realizabilitycondition — for the Gaussian integral defining Z to con-verge. Here, this condition reads: β + 2αk2

min > 0 andβ + 2αk2

max > 0. In particular, when α > 0, negativetemperatures can be attained. In this regime, which cor-responds to 〈Γ2〉/(2〈E〉) small enough (Kraichnan andMontgomery , 1980), the energy spectrum is a decreasingfunction of k. When β → −2αk2

min, a singularity appearsat k → kmin, which means that the energy is expected toconcentrate in the largest scales. Hence, statistical me-chanics for the truncated system predicts that when theenstrophy is small enough compared to the energy, weexpect the energy to be transferred to the large scales.(Kraichnan, 1967) gives other arguments to support andrefine this view; in particular he shows the existence oftwo inertial ranges, with a constant flux of energy andenstrophy, respectively, with the energy spectrum scal-ing as E(k) ∼ Cε2/3k−5/3 and E(k) ∼ C ′η2/3k−3 re-spectively, where ε and η are the energy and enstrophyfluxes. In particular, the equilibrium energy spectrum atlarge scales is shallower than the energy inertial rangespectrum. Assuming a tendency for the system to relaxto equilibrium — although the equilibrium is never at-tained in the presence of forcing and dissipation — wethus expect the flux of energy to be towards the largescales; a process referred to as the inverse cascade of2D turbulence (see Fig. 4). Similarly, the transfer ofenstrophy in the corresponding inertial range should be

Page 11: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

11

k-1

k-

53

k-3

10-1 1 10 102

10-3

1

103

k

EHk

L

k f

FIG. 4 Energy spectrum (solid blue curve) for 2D turbulenceat statistical equilibrium in the negative temperature regime;there is an infrared divergence at kmin, corresponding to con-densation of energy in the gravest mode. At large k, theequilibrium energy spectrum scales like k−1. We have super-imposed on this graph the slopes of the energy and enstrophyinertial ranges, k−5/3 and k−3, which build up in the forcedcase (see text for details).

towards the small scales. The dual cascade scenario hasbeen confirmed both by numerical simulations (Boffetta,2007) and laboratory experiments (Paret and Tabeling ,1997).

2. Topographic turbulence

The work of (Kraichnan, 1967) was extended to flowover a topography by (Salmon et al., 1976) and (Her-ring , 1977). If H(k) is the (prescribed) bottom to-pography spectrum, the energy spectrum now reads〈E(k)〉 = πk/(β + 2αk2) + 2α2k2/(β + 2αk2)2H(k).The second term arises because of a non-vanishing meanflow: straightforward computations show that 〈ω(k)〉 =

−2αk2h(k)/(β + 2αk2), so that the second term is theenergy spectrum of the mean flow, while the first termstems from the fluctuations. In the absence of topogra-phy, the ensemble average vorticity is zero, but in a par-ticular realization, the mean flow will not vanish; the sys-tem will spontaneously adopt a choice of phase for eachwave vector. Summing over the possible phase choicesyields the zero ensemble mean. Note that the meanflow at absolute equilibrium has vorticity proportionalto the stream function. This feature has been tested nu-merically, in connection with the emergence of Fofonoffflows (Griffa and Salmon, 1989; Wang and Vallis, 1994),see Fig. 5. The equilibrium energy spectrum and flow-topography correlation have also been confirmed numer-ically with very good agreement (Merryfield and Hol-loway , 1996).

3. Quasi-Geostrophic Turbulence

The dynamical equations of QG flow are very similarto the Euler equations, replacing vorticity by potentialvorticity. In particular, they conserve similar quadraticinvariants, and the theory can be extended in a straight-forward manner (Holloway , 1986; Salmon, 1998). Therole of bottom topography has already been investigatedin section III.C.2; we will discuss in this section the effectof stratification and beta effect.

Perhaps the simplest framework to consider the roleof stratification is the two-layer QG case. As in sectionIII.C.1, a canonical probability distribution can be con-structed, taking into account the three invariants: thetotal energy E and the enstrophies of each layer, Z1 andZ2. The corresponding partition function can be com-puted, and the spectrum studied in the various regimes,with similar results. In particular, negative temperaturestates are accessible, which correspond to condensationof the energy on the largest horizontal scales. Maybemore interestingly, although the various forms of energy(kinetic energy K1,K2 in each layer and potential energyP ) are not individually conserved, we can compute theiraverage value at equilibrium, as (Salmon et al., 1976)did. Alternatively, the standard decomposition in termsof the barotropic and baroclinic modes, with their ki-netic energies KT and KB , can be used. As (Salmonet al., 1976) highlighted, the Rossby deformation scalekD plays an important role. At scales smaller than thedeformation scale (k kD), the two layers behave es-sentially as two independent copies of 2D turbulence; theenergy spectrum in each layer is the same as in the 2Dcase, the correlation at statistical equilibrium is low, andthere is about as much energy in the barotropic mode andthe baroclinic mode: 〈KT (k)〉/〈KB(k)〉 ∼ 1. Besides,the potential energy is small compared to the kinetic en-ergy: 〈P (k)〉/〈KT (k)〉 = O(kD/k). At scales larger thanthe deformation radius (k kD), the system rather be-haves as a unique barotropic layer: the amount of energyin the two layers is about the same, but the energy isessentially in the barotropic mode, with negligible en-ergy in the baroclinic mode, and a statistical correla-tion between the two layers of order 1. This theoreticalanalysis goes in strong support of the standard pictureof two-layer QG turbulence, developed on phenomeno-logical grounds ((Rhines, 1979; Salmon, 1978), see also(Vallis, 2006, chap. 9)), and is in accordance with nu-merical simulations (Rhines, 1976). These results wereextended to an arbitrary number of layers and to continu-ously stratified flows by (Merryfield , 1998). Although theequilibrium mean, vertically integrated stream functionremains similar to the two-layer case, the distribution ofthe statistics on the vertical differs as higher-order mo-ments are considered. The ratio of potential to kineticenergy for instance, can become significantly underes-timated, especially in the limit of strong stratification

Page 12: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

12

100 Journal of Marine Research

a ----~:~:::~~~~~ ------------------------- izz:.-;- ________-----------------------

_.-- --.. _A .-

-____________-------------------------- ._ --._

/- ___---- _____--------------________ ..- --.

__.- --.. -.._ _--- ---

________----------------------- __--- --__ ---_ _____-- _________----------------------------------

b

1.0

.5

9 0

-.5

-1.0 - 1.:

d

2 -.6

Figure 8. Experiment using the smoothed average 5 field in ISRDl instead of random field as initial condition. Now the field has very little transient component. The fields are averaged from 10 T,, and then smoothed. Shown here are (a) relative vorticity 5, (b) absolute vorticity q, (c) streamfunction IJJ and (d) scatter-plot of q - I).

consistent with the theory. Roughly measuring from the scatter-plots, we obtain Apl CL- -0.37 where E is the average of p’s in the two experiments, A~J, is the increase in t,r, from 32 x 32 to 128 x 128 experiments. Therefore, we have:

AE 3Ak -- --- E - - 2 CL - 55-5%*

Both energies are the components contained in the time mean fields. Hence, when R, = 3.18 x 10e3 in the initial random flows, then as resolution is increased from

FIG. 5 Convergence towards the statistical equilibrium in inviscid Galerkin-truncated barotropic flow on a β-plane (Wang andVallis, 1994). Left: Stream function. Right: scatterplot of the q-ψ relation.

(kD → 0) where the two-layer model does not capturewell the possibility that an important fraction of the en-ergy may be trapped near the bottom.

The second dominant effect in geophysical flows, inaddition to stratification, is rotation. The Coriolis forceintroduces a linear term in the equations, which does notaffect directly the previous analysis of the nonlinear en-ergy transfers: the conserved quantities remain the sameand the statistical theory is easily extended by replacingrelative vorticity with absolute vorticity. However, thevariation of the Coriolis force with latitude is responsi-ble for the appearance of Rossby waves, which modify thephysical interpretation of the predicted cascade of energy.As anticipated by (Rhines, 1975) and verified numerically(e.g Vallis and Maltrud , 1993), the Rossby waves deflectthe inverse energy cascade. The ratio of the wave pe-riod to the eddy turnover time, ε(k) = τW (k)/τNL(k)gives an estimate of the validity of the above predictions;when ε 1, nonlinear effects dominate, while for ε 1,the waves are much faster than the nonlinear effects andthe statistical equilibrium may never be reached. Thecurve defined by ε(k) = 1 in k-space delimitates the tworegimes; it has a dumbell shape which shows that thewaves should prevent access to low wavenumbers alongone direction in k-space only. Hence the Rossby wavesdeflect the cascade and lead to the preferential formationof zonal flows.

4. Beyond balanced motion

Although the motions of the atmosphere and oceansof the Earth are nearly geostrophically balanced, non-balanced motions, like inertia-gravity waves, also playimportant roles. It is therefore important to understandhow energy is exchanged by nonlinear interactions be-tween the slow, balanced motions and the fast, wave mo-tions, and some insight has been gained through equilib-

rium statistical mechanics.

(Errico, 1984) first observed a tendency for unforcedinviscid flows described by hydrostatic primitive equa-tions to reach an energy equipartition state, in which theenergy in the fast wave modes is comparable to that inthe slow balanced modes. The study by (Warn, 1986),in the context of the shallow water equations, essentiallyconfirms that QG flows are not equilibrium states, andthat a substantial part of the energy may end up in thefast (surface) wave modes at statistical equilibrium, im-plying a direct cascade of energy to the small scales.(Bartello, 1995) has obtained analytically the equilib-rium energy spectrum for the Boussinesq equations (ne-glecting the nonlinear part of potential vorticity), in thepresence of rotation, confirming the direct cascade of en-ergy. In particular, there is no negative temperaturestates in this case, due to the presence of the inertia-gravity waves. Studying the resonant triadic interactions,he also discussed the possibility of a wave-vortical modedecoupling, and related the issue of geostrophic adjust-ment with the existence of an inverse cascade.

Without decoupling assumptions on the dynamics, in-verse cascades can be obtained by restricting the compu-tation of the partition function, as done generally whenstudying metastable states (Penrose and Lebowitz , 1979),to the subset of phase space made of balanced modes.This approach indicates that parity symmetry breaking(Beltrami) flows should display an inverse cascade of en-ergy (Herbert , 2013a). It has also been applied to ro-tating and/or stratified flows in the Boussinesq approx-imation (Herbert et al., 2013) to obtain the direction ofthe energy transfers in anisotropic Fourier space. Com-bined with an analysis of the deflecting effect of wavessimilar to the case of Rossby waves (see section III.C.3),this could explain why inverse cascades have been seen innumerical simulations of rotating (e.g. Mininni and Pou-quet , 2010) and rotating and stratified (e.g Marino et al.,

Page 13: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

13

2013) flows, but not in purely stratified flows (e.g. Waiteand Bartello, 2004).

D. The Robert-Miller-Sommeria theory

The RMS theory bears some resemblance with thepoint vortices theory; the main difference is that ratherthan following the (Lagrangian) motion of points withfixed vorticity, the theory is built as the continuous limitof lattice approximations in the Eulerian framework. Therationale behind this choice is, roughly speaking, that al-though Lagrangian motion is unpredictable, the statisticsof the Eulerian flow is (Robert and Rosier , 2001).

1. Mean-field theory

Above, we have considered finite-dimensional modelsconserving two quadratic quantities, the energy and theenstrophy. In fact, the majority of the flows consideredabove — and in particular 2D and QG flows — conservean infinite family of invariants, called Casimir invariants:for any function s,

∫s(ω)dr is conserved. The specific

case sn(x) = xn corresponds to the moments of the vor-ticity distribution, n = 1 is the circulation and n = 2 isthe enstrophy, while, sσ(x) = δ(x − σ) yields that thearea γ(σ) where the vorticity takes value σ is conserved.This is due to the absence of a vortex stretching term, incontrast with full 3D flows; here the vorticity (or poten-tial vorticity) patches are stirred in such a way that theirarea remains conserved, which implies that their bound-aries become more and more filamented. The theory de-veloped by (Miller , 1990) and (Robert and Sommeria,1991) introduces a coarse-grained vorticity field ω, whichcorresponds to the macroscopic state of the flow. Thiscoarse-grained vorticity field can be predicted based onthe invariants using statistical mechanics. To do so, weintroduce the probability distribution ρ(σ, r) for the vor-ticity field to take value σ at point r. The coarse-grainedvorticity field is given by ω(r) =

∫∞−∞ ρ(σ, r)dσ. The

invariants of the system are the energy

E [ρ] =

∫D2

drdr′∫R2

dσdσ′σσ′G(r, r′)ρ(σ, r)ρ(σ′, r′),

(60)

with G the Green function of the Laplacian, and theCasimir invariants

Gn[ρ] =

∫Ddr

∫Rdσσnρ(σ, r), (61)

or equivalently, the vorticity levels

Dσ[ρ] =

∫Ddrρ(σ, r). (62)

The idea of the theory is to select the probability dis-tribution ρ which maximizes a mixing entropy S [ρ] =−∫D∫R drdσρ(σ, r) ln ρ(σ, r), under the constraints of

conservation of the invariants, and point wise normal-ization N [ρ](r) =

∫R dσρ(σ, r) = 1. Hence, we are inter-

ested in the variational problem:

S(E, Γnn∈N) = maxρ,N [ρ](r)=1

S [ρ] | E [ρ] = E,

∀n ∈ N,Gn[ρ] = Γn, (63)

or equivalently,

S(E, γ(σ)) = maxρ,N [ρ](r)=1

S [ρ] | E [ρ] = E,

∀σ ∈ R,Dσ[ρ] = γ(σ). (64)

The expression for the mixing entropy S is justified bythe validity of the mean-field theory for long-range inter-actions (Bouchet and Venaille, 2012).

The critical points of the variational problem (64)are simply given by δS −

∫drζ(r)δN (r) − βδE −∫

dσα(σ)δDσ = 0, where β and α(σ) are the Lagrangemultiplier associated with the conservation constraints.Easy computations yield the solution

ρ(σ, r) =1

Ze−βσψ(r)−α(σ), (65)

so that the coarse-grained vorticity is given by

ω = F (ψ), with F (ψ) = − 1

β

δ lnZδψ

, (66)

and Z(ψ) =∫R dσe

−βσψ−α(σ). To compute the equilib-rium states of the system, one should solve the partialdifferential equation (66), referred to as the mean-fieldequation, and check afterwards that the obtained criticalpoints are indeed maxima of the constrained variationalproblem by considering the second derivatives. This willautomatically ensure that the equilibrium states are non-linearly stable (Chavanis, 2009).

2. Equilibrium states for 2Dand barotropic flows

The mean-field equation (66) is in general difficult tosolve; one issue is that the ω - ψ relation is in generalnonlinear. Most of the analytical solutions have beenobtained in the linear case, by decomposing the fields ona basis of eigenfunctions of the Laplacian on the domainD. This technique was first introduced in a rectangulardomain by (Chavanis and Sommeria, 1996), who showedthat the statistical equilibrium is either a monopole or adipole, depending on the aspect ratio (Fig. 6).

The same method was extended to the case ofbarotropic flows, replacing vorticity by potential vortic-ity. Taking into account the beta effect, Fofonoff flows are

Page 14: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

14

FIG. 6 Maximum entropy states as a function of the aspectratio for a rectangular domain, in the linear (strong mixing)ω-ψ limit (Chavanis and Sommeria, 1996). For τ < τc, theequilibrium is a monopole, while for τ > τc, it is a dipole.

obtained as statistical equilibria in a rectangular basin(Naso et al., 2011; Venaille and Bouchet , 2011a). On arotating sphere, the equilibria, in the linear limit, canbe either solid-body rotations or dipole flows (Herbertet al., 2012a,b). Note that in general, in the linear limit,all the energy is expected to condense into the gravestmodes in the equilibrium state (Bouchet and Corvellec,2010). The effect of the higher-order, fragile Casimir in-variants3 is to prevent complete condensation, leadingto a wider variety of flow topologies at equilibrium, andto non-linear ω - ψ relationships. The condensation canalso be arrested due to geometric effects, as in the case ofthe sphere, for which additional invariants appear (Her-bert , 2013b). (Bouchet and Simonnet , 2009) have alsoconsidered the role of a small nonlinearity in the ω - ψrelationship for a rectangular domain of aspect ratio closeto 1, thereby obtaining two topologies for the equilibriumstates: dipole and unidirectional flows. Adding a smallstochastic forcing generates transitions from one to theother equilibrium.

In the case of the equivalent barotropic model, forwhich the potential vorticity is given by q = −∆ψ +ψ/R2 + h, the Rossby deformation radius R introducesa new length scale. Another limit which is analyticallysolvable is that of a very small Rossby deformation ra-dius: R L, where L is the size of the domain. Thislimit leads to sharp interfaces separating phases of dif-ferent free energies, characterized by different, well ho-mogenized potential vorticities (Bouchet and Sommeria,

3 As the fragile invariants cannot be recovered from the predictedcoarse-grained field, it was suggested to treat them in a canoni-cal manner (Turkington, 1999). This introduces the conceptualdifficulty of defining a potential vorticity reservoir.

2002). Strong jets form at the interface between two suchhomogeneous potential vorticity regions; mid basin east-ward jets, such as the Gulf Stream and the Kuroshio,can be interpreted as such statistical equilibria when thebeta effect is neglected, but in the presence of beta effect,these solutions become metastable or unstable (Venailleand Bouchet , 2011b). In the jet regions, mesoscale ringsform and propagate westward (Chelton et al., 2007). (Ve-naille and Bouchet , 2011b) have shown that mesoscalerings could also be seen as local equilibrium states in achannel (see Fig. 7).

FIG. 7 Modulus of the velocity in the ring statistical equilib-rium of the QG model in the limit of small Rossby deforma-tion radius R (Venaille and Bouchet , 2011b). The width ofthe jet is of order R.

3. Stratified flows

In addition to the 2D and quasi-2D cases mentionedabove, the RMS theory has also been applied to strat-ified fluids (essentially in the quasi-geostrophic regime).(Herbert , 2013c) has obtained and classified the statis-tical equilibria of the two-layer QG model in the frame-work of the RMS theory, and updated the discussion ofthe vertical distribution of energy at statistical equilib-rium (see section III.C.3). In the context of continuouslystratified flows, (Venaille, 2012) has taken up the threadinitiated by (Merryfield , 1998) (see section III.C.3) andshown that bottom-trapped currents are indeed statisti-cal equilibria of the RMS theory. Still in the continuouscase, (Venaille et al., 2012) have also studied the verti-cal distribution of energy at statistical equilibrium, fo-cusing on the tendency to reach barotropic equilibriumstates; as also observed in the two-layer model, the con-straint of conservation of fine-grained enstrophy preventscomplete barotropization. As the beta effect increases,barotropization is facilitated, until we enter a regimedominated by waves.

Page 15: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

15

E. Subgrid scale parameterization

The equilibrium approach has found other practicalapplications with parameterization methods. A notableone is the Neptune effect parameterization: given the ten-dency of ideal Galerkin truncated flows to go towards thestate of maximum entropy (Zou and Holloway , 1994),(Holloway , 1992) suggested that instead of the usual sub-grid scale parameterizations in ocean models (e.g. eddyviscosity ν∗∆u), which drives the system towards rest,one should use a relaxation towards the statistical equi-librium state u∗ (e.g. of the form ν∗∆(u− u∗) for eddyviscosity). This parameterization has been used testedand commented in a number of studies (Cummins andHolloway , 1994, e.g.). For more perspective on this typeof subgrid-scale parameterizations, the reader is referredto (Holloway , 2004) and (Frederiksen and O’Kane, 2008).

Along similar lines, (Kazantsev et al., 1998) have pro-posed more generally to treat the subgrid scales so asto maximize the entropy production, inspired by the re-laxation equations formulated in the RMS theory (Cha-vanis and Sommeria, 1997). Note also that it has beenshown in direct numerical simulations of ideal 3D turbu-lence that the small scales thermalize progressively, andact as a sort of effective viscosity even in the ideal sys-tem, leading to the appearance of transient Kolmogorovscaling laws (Cichowlas et al., 2005). This seems to beconsistent with the above suggestions for subgrid scaleparameterizations.

IV. THERMODYNAMICS OF FORCED ANDDISSIPATIVE SYSTEMS

A. Introduction

The atmosphere and the oceans are out-of-equilibriumsystems, which exchange irreversibly matter and energyfrom their surrounding environment and re-export it ina more degraded form at higher entropy. For example,Earth absorbs short-wave radiation (lo entropy solar pho-tons at Tsun ≈ 6000 K) which is then re-emitted to spaceas infrared radiation (high entropy thermal photons atTearth ≈ 255 K). In addition to that, spatial gradientsin chemical concentrations and temperature as well astheir associated internal matter and energy fluxes can beestablished and maintained for long time within nonequi-librium systems (e.g. the temperature contrast betweenthe polar and equatorial regions and the associated large-scale, atmospheric and oceanic circulation). In contrast,closed, isolated systems cannot maintain disequilibriumand evolve towards structureless, homogenous thermo-dynamical equilibrium, as a result of the second law ofthermodynamics (Prigogine, 1961).

The basis of the physical theory of climate was estab-lished in a seminal paper by (Lorenz , 1955), who elu-

cidated how the mechanisms of energy forcing, conver-sion and dissipation are related to the general circula-tion of the atmosphere. Oceanic and atmospheric largescale flows results from the conversion of available poten-tial energy - coming from the differential heating due tothe inhomogeneity of the absorption of solar radiation-into kinetic energy. through different mechanisms of in-stability due to the presence of large temperature gra-dients. The understanding of the fundamentally ther-modynamical origin of such dynamical instabilities, asclarified by (Charney , 1947) and (Eady , 1949), is theother cornerstone of geophysical fluid dynamics. Suchinstabilities create a negative feedback, as they tends toreduce the temperature gradients they feed upon by fa-voring the mixing between masses of fluids at differenttemperatures. Attaining the closure of such a thermo-dynamical/dynamical problem would mean obtaining aself-consistent theory of climate able to connect instabili-ties and large scale stabilizing processes on longer spatialand temporal scales.

Furthermore, in a forced and dissipative system likethe Earth’s climate, entropy is continuously produced byirreversible processes and at steady state, the entropyproduction is balanced by a net outgoing flux of entropyat the boundary of the system, in our case, the top ofthe atmosphere. Therefore, on the average the entropybudget - just like the energy budget - vanishes (deGrootand Mazur , 1984; Prigogine, 1961). Besides the dissi-pation of kinetic energy due to viscous processes, manyother irreversible processes such as turbulent diffusion ofheat and chemical species, irreversible phase transitionsassociated to various processes relevant for the hydrolog-ical cycle, and chemical reactions relevant for the bio-geochemistry of the planet contribute to the total ma-terial entropy production(Goody , 2000; Kleidon, 2009).The study of the climatic entropy sources (Fraedrich andLunkeit , 2008; Goody , 2000; Kleidon, 2009; Kleidon andLorenz , 2005; Lucarini et al., 2011; Pascale et al., 2011;Pauluis and Held , 2002a,b) has been revitalized after(Paltridge, 1975, 1978) proposed a principle of maxi-mum entropy production (MEPP) as a constraint on theclimate system. Although there is still great confusionabout MEPP and other nonequilibrium variational prin-ciples (for an updated review see Dewar et al., 2013),this has lead the scientific community to refocus on theimportance of a thermodynamical approach – as com-plementary to the dynamical one – for studying classesof problems like those relevant for nonlinear geophysicalflows.

Page 16: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

16

B. Climatic energy budget and energy flows

1. Conservation of energy

A detailed knowledge of the energy flows and theirresponse to perturbations is fundamental for progress-ing understanding of the climate system. The right clo-sure of the energy budget and an accurate representa-tion of the transport of energy in all forms have to tobe properly accounted for in any climate model in or-der to have a good representation of the climate state,The total specific energy of a geophysical fluid is givenby the sum of internal, potential, kinetic and latent en-ergy, e = i + φ + k + l (e = cvT + gz + v2/2 + Lq forthe atmosphere and e = c0T + gz + v2/2 for the ocean,standard notation is used). The conservation of energyand mass require that (Peixoto and Oort , 1992):

∂ρe

∂t= −∇ · (Jh + FR + FS + FL)−∇(τ · v) (67)

where ρ is the density; t is time; v is the velocity vec-tor; FR, FS, and FL are the vectors of the radiative,turbulent sensible, and turbulent latent heat fluxes, re-spectively; p is the pressure; and τ is the stress tensor andthe total enthalpy transport Jh = (ρe + p)v = ρhv hasbeen introduced. By expressing Eq. (67) in spherical co-ordinates (r, λ, ϕ) and assuming the usual thin she’ll ap-proximation r = R+ z, z/R 1, where R is the Earth’sradius and z is the vertical coordinate of the fluid, it canbe shown that:

[E] = − 1

R cosϕ

∂TT∂ϕ

+ [FR]toa (68)

in which [E](ϕ, t) ≡∫E(λ, ϕ, t)dλ and the meridional

enthalpy transport has been defined as:

TT (ϕ, t) ≡∫∫

Jhϕ(ϕ, λ, z, t)R cosϕdzdλ. (69)

Equation (68) relates the rate of change of the verticallyand zonally integrated total energy to the divergence ofthe meridional transport by the atmosphere and oceansand the zonally integrated radiative budget at the top-of-the-atmosphere. Integrating along ϕ (X =

∫Xdϕ),

the expression for the time derivative of the net globalenergy balance is straightforwardly derived:

[FR]toa = [E]. (70)

Similar relationships can be written for the atmosphere,ocean and land provided that energy fluxes of sensible,latent heat as well as radiative fluxes are taken into ac-count at the surface (Lucarini and Ragone, 2011).

Under steady state conditions, the long term aver-

age E = 0 because otherwise, trends would be present.

Therefore from equation (70) the stationarity conditionimplies that

[FR]toa = 0. (71)

Equation (71) is a particular case of a general propertyof forced dissipative systems that, under generic bound-ary conditions, realize steady states obeying zero sumproperties for energy fluxes on the average(Prigogine,1961). A physically consistent climate model there-fore should feature a vanishing net energy balancewhen statistical stationarity is eventually obtained. Lu-carini and Ragone (2011) analyzed the behavior of morethan twenty atmosphere-ocean coupled climate models(PCMDI/CMIP3 intercomparison project, http://www-pcmdi.llnl.gov/) under steady state conditions (preindus-trial scenario) and found that models’ energy balancesare wildly different with global balances spanning be-tween −0.2 and 2 W m−2, with a few ones featuring im-balances larger than 3 W m−2. We have analyzed similarbudgets for the last generation of climate models (CMIP5intercomparison project, Taylor et al. (2012)) and havenot found a significant improvement (Fig. 8). Spuriousenergy biases may be associated with non-conservationof water in the atmospheric branch of the hydrologicalcycle (Liepert and Lo, 2013) and in the water surfacefluxes (Hasson et al., 2013; Lucarini et al., 2008), withthe fact that dissipated kinetic energy is not re-injectedin the system as thermal energy (Becker , 2003), as wellas with nonconservative numerical schemes (Gassmann,2013).

2. Meridional enthalpy transport

The meridional distribution of the radiative fields atthe top-of-the-atmosphere poses a strong constraint onthe meridional general circulation (Stone, 1978). As clearfrom equation (68), the stationarity condition (71) leadsto the following indirect relationship for TT :

TT (ϕ) = −2π

∫ π/2

ϕ

R2 cosϕ′(FR)toa(ϕ′). (72)

In other terms, the flux TT results from the fact that low-latitude zones feature a positive imbalance between thenet input of solar radiation – determined by planetaryalbedo, i.e. clouds (Donohoe and Battisti , 2012) – andthe output of longwave radiation, while the high latitudesfeature a negative imbalance. Atmospheric and oceaniccirculations act as responses needed to equilibrate suchan imbalance.

The climatic meridional enthalpy transport TT (ϕ) –and its atmospheric and oceanic components, TA andTO – is a fundamental measure of nonequilibrium andprovides a concise picture of the processes by which theclimatic fluid reduces the temperature difference between

Page 17: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

17

FIG. 8 Mean and standard deviation of globally averagedtop-of-the-atmosphere radiative budget (top), atmosphere en-ergy budget (top-middle), ocean (bottom-middle) and landenergy budget (bottom) for inter-comparable CMIP3 (red)and CMIP5 (blue) climate models control simulations.

the low and high latitude regions with respect to whatimposed by the radiativ-convective equilibrium picture.Stone (1978) showed that TT depends essentially on themean planetary albedo and on the equator-to-pole con-trast of the incoming solar forcing, while being mostlyindependent from dynamical details of atmospheric and

oceanic circulations. As emphasized by, Enderton andMarshall (2009), if one assumes drastic changes in themeridional distributions of planetary albedo differencesemerge with respect to Stone’s theory. A comprehen-sive thermodynamic theory of the climate system able topredict the peak location and strength of the meridionaltransport, the partition between atmosphere and ocean(Rose and Ferreira, 2013) and to accommodate the vari-ety of processes contributing to it, is still missing.

Besides theoretical difficulties, observational estima-tions of TT , TA and TO also poses non-trivial challenges.For simplicity, we here refer to TT . There is still notan accurate estimate of such a fundamental quantity fortesting the output of climate models, despite the efforts ofseveral authors (Fasullo and Trenberth, 2008; Mayer andHaimberger , 2012; Trenberth and Caron, 2001; Trenberthand Fasullo, 2010; Wunsch, 2005). The precision of theestimates relies on the knowledge of the boundary fluxesFR, FS, and FL and on the reanalysis datasets. Wun-sch (2005), by using measurements of the radiative fluxesat the top of the atmosphere and previous estimates ofthe oceanic enthalpy transport, gave a range of valuesof 3.0 − 5.2 PW (1 PW = 1015 W) for the maximumin the Northern Hemisphere (NH) and 4.0− 6.7 PW forthe maximum in the Southern Hemisphere (SH). Tren-berth and Fasullo (2010), by combining measurementsof top-of-the-atmosphere radiative fields with differentreanalyses and ocean datasets, found the range to be4.7− 5.1 PW for the SH maximum and 4.6− 5.6 PW forNH. Mayer and Haimberger (2012), using two reanalysisdatasets (ERA-40 and the more recent ECMWF reanaly-sis ERA-Interim), constrained the two peaks in narrowerconfidence intervals: 5.1 − 5.6 PW in the SH (4.4 − 4.9PW in the NH) for the ERA-40 data and 5.1 − 5.6 PWin the SH (4.4−4.9 PW in the NH) for the ERA-Interimdata. Unfortunately reanalysis datasets are affected bymass and energy conservation (e.g. +1.2 W m−2 at thetop-of-the-atmosphere and +6.8 W m−2 over oceans inERA-Interim, (Mayer and Haimberger , 2012)) problemsthat may potentially bias the transport estimates. Fur-thermore, these estimates are dependent on the analy-sis method and the model used – Trenberth and Caron(2001), using other reanalysis dataset (NCEP), found avalue of the maxima 0.6 PW larger in the NH than thosefound with the ECMWF reanalysis.

The use of numerical climate model does not help toreduce such uncertainties Lucarini and Ragone (2011)analyzed a large dataset of coupled climate mod-els (PCMDI/CMIP3, http://www-pcmdi.llnl.gov/) andfound a large spread in the meridional enthalpy trans-ports peaks with discrepancies of the order of 15-20 %around a typical value of about 5.5 PW. State-of-the-artclimate models (CMIP5 intercomparison project, Tayloret al. (2012)) show little improvement in terms of mu-tual agreement (Fig. 9). Donohoe and Battisti (2012)attributed such a large spread in TT to intermodel dif-

Page 18: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

18

FIG. 9 Value and position of the peak of the poleward merid-ional enthalpy transport in the pre-industrial scenario for thewhole climate (top), atmosphere (middle) and ocean (bottom)for the some of the CMIP3 (red) and CMIP5 (blue) generalcirculation models.

ferences in the meridional contrast of absorbed solar ra-diation, which, in turn, is mainly due to the inter-modeldifference in the shortwave optical properties of the at-mosphere (cloud distribution). Figure 9 also shows that,while the disagreement among models for the peak of theatmospheric transport is comparable to that for the peakof the total transport, enormous differences emerge whencomparing oceanic transports.

Interesting information emerge when looking at the po-sition of the peaks of the transport. Stone (1978) pre-dicts that the position of the maximum of TT is well con-strained by the geometry of the system and weakly de-pendent of longitudinal homogeneities, and, accordinglyin Fig. 9 both CMIP3 and CMIP5 models feature smallspread in the position of the peak of TT , with minute dif-ferences between the two hemispheres, except one outlier.Similarly, the spread among models is small with respectto the position of the peak of TA in both hemispheresand of TO in the Northern Hemispheres, while a largeruncertainty exists in the position of the peak of TO in theSouthern Hemisphere.

C. The maintenance of thermodynamicaldisequilibrium

The basic understanding of the maintenance of the at-mospheric general circulation was achieved nearly sixtyyears ago by E. N. Lorenz (Lorenz , 1955) through theconcepts of available potential energy and atmosphericenergy cycle. The concept of available potential energy,first introduced by Margules (1905) to study storms, isdefined as A =

∫cp(T −Tr)dV , where Tr is the tempera-

ture field of the reference state, obtained by an isentropicredistribution of the atmospheric mass so that the isen-tropic surfaces become horizontal and the mass betweenthe two isentropes remains the same. By its own def-inition, this state minimizes the total potential energyat constant entropy. Such a definition is somewhat arbi-trary and different definitions lead to different formula-tions of atmospheric energetics. For example, the choiceof a reference state maximizing entropy at constant en-ergy (Dutton, 1973) leads in a natural way to the conceptof exergy4, common in heat engines theory (Rant , 1956).A review on the various theories of available energeticsand their link to exergy, is given by (Tailleux , 2013). In-depth literature on the Lorenz’s energetics theory can befound in (Lorenz , 1967) and (Peixoto and Oort , 1992).Here a succinct review of the global thermodynamicalproperties of the atmosphere is presented and the mainimplications of the first and second law of thermodynam-ics are analyzed.

The total energy budget of the atmosphere can be writ-ten as E = P +K, where K = (1/2)

∫v2ρdV represents

the total kinetic energy and P =∫

(cpT +Lq)ρdV the to-tal potential energy5 and V is the atmospheric domain.It can be shown (Peixoto and Oort , 1992) that:

P = −W (P,K) + Ψ +D, (73)

K = −D +W (P,K), (74)

in which D =∫ρε2dV > 0 is the dissipation of ki-

netic energy due to friction, W (P,K) = −∫

v · ∇pρdVis the potential-to-kinetic energy conversion rate andΨ =

∫qnfdV is the non-frictional diabatic heating due to

the convergence of turbulent sensible heat fluxes, conden-sation/evaporation, and convergence of radiative fluxes.The conversion term W can be interpreted as the theinstantaneous work performed by the system. In this re-spect, equation (73) represents the statement of the firstlaw of thermodynamics for the atmosphere. Equations

4 Exergy is the part of the internal energy measuring the departureof the system from its thermodynamic and mechanical equilib-rium, that is a state of maximum entropy at constant energy.

5 The total potential energy is locally defined as cvT + gz + Lq.However under the hydrostatic approximation it is straightfor-ward to show that

∫(cvT + gz)ρdz =

∫cpTρdz (see e.g. Lorenz ,

1967).

Page 19: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

19

(73)-(74) imply that E = P + K = Ψ and therefore thefrictional heating D does not increase the total energysince it is just an internal conversion between kinetic andpotential energy. The terms involved in eqs. (73)-(74) areschematically shown in Fig. 10. Stationarity implies that

P = K = 0 and therefore D = W . A third independentmeasure of the strength of the Lorenz’s energy cycle notimplicit in equations (73)-(74) is the generation of avail-able potential energy, G =

∫qnf (1 − T/Tr)dV in which

Tr is the temperature field of the reference state and, fora steady state, G = W = D. Lorenz (1955) showed that

G ≈∫γT ′q′nfρdz, which emphasizes the dependence of

G on the gross static stability γ = −(Rθ/cppT )(∂θ/∂p)(vertical structure of the atmosphere) and on the covari-

ance of temperature and heating on pressure levels T ′q′nf(horizontal structure of the atmosphere).

The strength of the Lorenz energy cycle is a fundamen-tal nonequilibrium property of the atmosphere, whichjust as the meridional enthalpy transport (Sect. IV.B.2),is known with large uncertainty. Reanalysis (with allassociated problems, see Sect. IV.B.2) constrain D inthe range 1.5 − 2.9 W m−2 (Li et al., 2008). On theother hand, general circulation models feature values ofD from 2 to 3.5 W m−2 (Marques et al., 2011). Recentstudies suggest that small scales processes such as precip-itation (Pauluis and Dias, 2012) may significantly con-tribute to D, which might therefore be considerably un-derestimated. Furthermore, the knowledge of how theLorenz energy cycle will change in a warming scenariois an important metric to understand changes in futurelarge scale circulation. Numerical simulations show thata CO2 doubling causes a decrease of G of nearly 10% (Lu-carini et al., 2010). Warming patterns can alter G eitherby affecting the gross static stability γ or the meridionaltemperature/diabatic heating distribution. Hernandez-Deckers and von Storch (2012) show that the decrease inG is mostly associated with changes in the gross staticstability changes (i.e. in γ) rather than with meridionaltemperature gradient changes.

Another aspect to be considered is that usually the in-tensity of the Lorenz energy cycle is formulated assuminghydrostatic conditions and is computed for models thatdo not treat explicitly convection, but rather parametrizeit with some sort of adjustment scheme. Therefore,the Lorenz energy cycle in itself neglects any system-atic transfer of potential into kinetic energy occurringthrough non-hydrostatic, small scale motions. Obviously,the thermodynamics behind the set-up of organized con-vective motions can be framed in terms of an energy cyclesimilar to Lorenz’, and requires a separated treatment.See the comments in Steinheimer et al. (2008).

The oceans’ energy cycle is formally identical to thatintroduced for the atmosphere, with the only differencethat generation of kinetic energy due to wind stressesat the surface and dissipation of potential energy due

FIG. 10 Schematic figure showing atmospheric potential Pand kinetic K reservoirs and the transfer processes (ψ,W,D)

between them. Q = ψ +D is the total diabatic heating.

to molecular and small scale eddy mixing are not neg-ligible (Peixoto and Oort , 1992) and therefore must beaccounted for. As opposed to the atmospheric case, theoceanic energy cycle is powered mostly by an input of ki-netic energy, and not of available potential energy. Theoceanic Lorenz’s energy cycle is still poorly understoodfrom a theoretical point of view (Tailleux , 2013) and thevarious dissipation and generation terms are highly un-certain within the (1 − 2) × 10−2 W m−2 range (Oortet al., 1994; Storch et al., 2012).

1. Heat engine and efficiency

By defining the total diabatic heating q = qnf+ρε2 andsplitting the atmospheric domain V into the subdomainV + in which q = q+ > 0, and V −, where q = q− < 0, itcan be seen from equation (73) that:

W =

∫V +

q+ρdV +

∫−q−ρdV ≡ Φ+ + Φ−, (75)

with Φ+ > 0 and Φ− < 0 by definition. Therefore theatmosphere can be interpreted as a heat engine, in whichΦ+ and Φ− are the net heat gain and loss, and W themechanical work. The efficiency of the atmospheric heatengine, i.e. the capability of generating mechanical workgiven a certain heat input, can therefore be defined as:

η = (Φ+ + Φ−)/(Φ+) = W/Φ+. (76)

The analogy between the atmosphere and a (Carnot)heat engine can be pushed further if we introducethe total rate of entropy change of the system, S =∫q/TρdV = S+ + S−. In a steady state the following

expression holds:

S =Φ+

T++

Φ−

T−= 0, (77)

where T± ≡ Φ±/∫V ±

q±/TρdV from which it followsthat η = 1− T−/T+.

Page 20: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

20

Johnson’s approach generalizes to the global circu-lation similar heat engine theories previously used tocharacterize hurricane dynamics (Emanuel , 1991). InEmanuel’s theory a mature hurricane is depicted as anideal Carnot engine driven by the thermal disequilibriumbetween the sea-surface temperature Ts and the coolingtemperature T0 with an efficiency 1 − T0/Ts ≈ 1/3. Asimilar approach was extended also to moist convection(Emanuel and Bister , 1996; Renno and Ingersoll , 1996)for determining the the wind speed performed by the con-vective system for a certain rate of heat input Fin fromthe sea, W = Fin(1 − T0/Ts). Such an approach hasbeen used to study large scale, open systems (e.g. Hadleycell, Adams and Renno, 2005), and a grand goal seemsthe possibility of providing a thermodynamic theory ofmonsoonal circulation extending the results by Johnson(1989).

There are many, different definitions of efficiency usedto characterize global circulations(Ambaum, 2010; John-son, 2000; Perna et al., 2012; Schubert and Mitchell ,2013). As a future perspective, it would be desirable,from a theoretical point of view, to reconcile these defi-nitions of thermodynamic efficiency and understand theirdifferences.

2. Entropy production

More insight can be gained if we split S into the sumof two contributions, S = Se + Si, where Se is associatedwith the entropy fluxes with the surroundings and Si ≥ 0with internal irreversible processes (deGroot and Mazur ,1984). The term Se is associated with a certain amount ofexternal heat q added to the system by radiation (Goody ,2000), therefore the previous statement can be written,according to Clausius’ formulation, as:

S =

∫q

TρdV + Si, Si ≥ 0. (78)

For a steady state S = 0, hence Se ≤ 0 (which is theanalogue of Clausius’ inequality

∮δQ/T ≤ 0) implying

that non-equilibrium systems must export entropy to thesurroundings in order to maintain their patterns of flowsand associated gradients.

In the case of the atmosphere, Si = Sfric+Shyd+Sdiff ,that is the sum of contributions associated with fric-tional heating , hydrological cycle (diffusion of water andphase-changes) and heat diffusion respectively. Entropyproduction due to heat diffusion is generally small (≈ 2mW m−2 K−1, (Kleidon, 2009)) and associated mostlywith dry-convection occurring nearby the surface. Fric-tional heating is associated with turbulent energy cascadebringing kinetic energy from large scales down to scales(millimeters or less for geophysical flows) where viscos-

ity can act and Sfric =∫ε2/TρdV ≈ 10 mW m−2 K−1

(Fraedrich and Lunkeit , 2008; Pascale et al., 2011).Pauluis and Dias (2012) argued that another large sourceof frictional dissipation (comparable to Sfric) comes from

shears surrounding falling hydrometeors. Finally, Shydis due to irreversible processes associated with the hy-drological cycle – evaporation of liquid water in unsatu-rated air, condensation of water vapour in supersaturatedair and molecular diffusion of water vapour (Pauluis andHeld , 2002a,b) and requires the knowledge of relative hu-midity H and the molecular fluxes of water Jv:

Shyd =

∫(C − E)R(lnH+ Jv · ∇pw)dV

−∫z=0

Jv,zR lnH. (79)

An indirect estimate of (79) can be obtained from the en-tropy budget for the water substance Sw = Fl/T + Shyd(Pauluis and Held , 2002b), where Sw is the rate of changeof entropy of water substance and Fl the amount ofheat per time that the water substance receives fromits environment (i.e. through evaporation and conden-

sation ≈ 80 W m−2). For a steady state Sw = 0 and

so Shyd = Fl(1/Ts − 1/Ta) ≈ 37 mW m−2 K−1 (Pas-cale et al., 2011). Furthermore, in atmospheric mod-els aphysical entropy sources due to diffusive/dispersivenumerical advection schemes are also present(Johnson,1997). Detailed estimates of the entropy budget of theclimate system and of the material entropy production

(Si ≈ 50 mW m−2 K−1) can be found in (Goody , 2000;Pascale et al., 2011). Oceanic entropy production dueto small-scale mixing gives a small contribution (≈ 1mW m−2 K−1) to Si (Pascale et al., 2011).

Furthermore, since Si = −Se, an indirect estimationof the material entropy production may be obtained assoon as the outgoing entropy fluxes are known. In thecase of the climate systems, for example, this requires theknowledge of the radiative entropy fluxes. It is immedi-ately understood the direct method may be, in principle,applied also to other planets (for which radiative fluxesis the only thing we can know) in order to infer infor-mation about their dissipative processes in their interior(Schubert and Mitchell , 2013). Starting from eq. (78)(Lucarini et al., 2011) derived an approximate relationfor Si based only on 2-D radiative fluxes which permitsto decompose Si into the sum of contributions due tohorizontal and vertical processes (Pascale et al., 2012).

D. Applications and future perspectives

1. Bistabiliy and tipping points

Based on the evidence supported by Hoffman andSchrag (2002) and from numerical models (Budyko, 1969;

Page 21: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

21

FIG. 11 Material entropy production (mW m−2 K−1) as afunction of solar constant S∗ and the CO2 concentration. Thetransition SB→W and W→SB are marked with dashed arrowsstarting from the tipping point regions (courtesy of R. Boschi,Universitat Hamburg).

Ghil , 1976; Sellers, 1969), it is expected that the Earth ispotentially capable of supporting multiple steady statesfor the same values of some parameters such as, for ex-ample, the solar constant. Such states are the presentlyobserved warm state (W), and the entirely ice coveredSnowball Earth state (SB). This is due to the presenceof two disjoint strange (chaotic) attractors. The W→SBand SB→W transitions are due, mathematically, to thecatastrophic disappearance of one of the two strange at-tractors (Arnold , 1992) and, physically, to the positiveice-albedo feedback. The SB condition, which might bea common feature also of Earth-like planets, hardly al-lows for the presence of life, so this issue is of extremerelevance for defining habitability condition in extrater-restrial planets.

PLASIM (Fraedrich et al., 2005), a general circulationmodel of intermediate complexity, was used by Boschiet al. (2013); Lucarini et al. (2013) to reconstruct anextensive portion of the region of multistabilityin theplane described by the parameters (S∗, [CO2]). The sur-face temperature Ts(S

∗, [CO2]) is shown in Fig. 11. Theboundary of the domain in the parametric space wheretwo states are admissible correspond to the tipping pointsof the system.

The thermodynamical and dynamical properties of theW and SB states are largely different. In the W states,surface temperature are 40 − 60 K higher than in thecorresponding SB state and the hydrological cycle dom-inates the dynamics. This leads to a material entropyproduction (Fig. 12) larger by a factor of 4 – order of(40 − 60) × 10−3 W m−2 K−1 vs. (10 − 15) × 10−3 Wm−2 K−1 – with respect to the corresponding SB states(Boschi et al., 2013). The SB state is eminently a dryclimate, with entropy production mostly due to sensibleheat fluxes and dissipation of kinetic energy.

The response to increasing temperatures of the twostates is rather different: the W states feature a decreaseof the efficiency of the climate machine, as enhanced la-tent heat transports kill energy availability by reducingtemperature gradients, while in the SB states the effi-ciency is increased, because warmer states are associatedto lower static stability, which favors large scale atmo-spheric motions (Fig. IV.D.1). The entropy productionincreases for both states, but for different reasons: thesystem become more irreversible and less efficient in thecase of W states, while stronger atmospheric motions leadto stronger dissipation and stronger energy transports inthe case of SB states. A general property which has beenfound is that, in both regimes, the efficiency η increasesfor steady states getting closer to tipping points and dra-matically drops at the transition to the new state be-longing to the other attractor (Fig. IV.D.1). In a rathergeneral thermodynamical context, this can be framed asfollows: the efficiency gives a measure of how far fromequilibrium the system is. The negative feedbacks tendto counteract the differential heating due to the stellarinsolation pattern, thus leading the system closer to equi-librium. At the bifurcation point, the negative feedbacksare overcome by the positive feedbacks, so that the sys-tem makes a global transition to a new state, where, inturn, the negative feedbacks are more efficient in stabi-lizing the system (Boschi et al., 2013).

FIG. 12 Top: material entropy production (10−3 W m−2

K−1) vs. emission temperature TE(K) for Ω = Ωearth (ma-genta) and Ω = 0.5Ωearth (black). Bottom: as in top figurebut for efficiency (courtesy of R. Boschi).

Another interesting aspect is the determination of em-pirical functional relations between the main thermody-namical quantities and globally averaged emission tem-perature TE = (LWtoa/σ)1/4, as shown in Fig. 12. This

Page 22: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

22

would permit to express nonequilibrium thermodynami-cal properties of the system in terms of parameters whichare more directly accessible through measurements (Lu-carini et al., 2013).

2. Applications to planetary sciences

The discovery of hundreds of planets outside the solarsystem (exoplanets) (Seager and Deming , 2010) is ex-tending the scope of planetary sciences towards the studyof the so-called exoclimates (Heng , 2012a). A large num-ber of the exoplanets discovered so far are tidally lockedto their parental star, experiencing extreme stellar forc-ing on the dayside where temperature up to 2000 K canbe reached. Starlight energy, deposited within the at-mosphere at the planet’s dayside, is then transported byatmospheric circulation to the night side. Such a sys-tem, similarly to the Earth’s climate, works like a heatengine (Sect. IV.B.2, Sect. IV.C). The strength of theday-to-night enthalpy flux controls the ratio of outgo-ing longwave energy fluxes from the night and day sideξ = LWnight/LWday, called efficiency of heat redistribu-tion in the astrophysical literature. Observations throughinfrared light curves show that the hotter the planet, themore inefficient is the atmospheres at redistributing stel-lar energy leading to larger day-night temperature differ-ences. Numerical simulations (Perna et al., 2012) showthat ξ varies between 0.2 (low heat redistribution) and1 (full heat redistribution) and depends critically on theatmospheric optical properties and the intensity of thestellar irradiance. Relating this definition of efficiencywith what proposed in Sect. IV.C would provide a linkbetween energy conversion and energy transport in plan-etary atmospheres.

A thorough understanding of dissipative processesis fundamental for dealing with planetary atmospheres(Goodman, 2009; Pascale et al., 2013). Dissipative pro-cesses are poorly known on Solar System planets (Schu-bert and Mitchell , 2013) and on exoplanets where pro-cesses unusual on Earth as Ohmic dissipation6 and cas-cading mechanisms such as shock wave breaking are be-lieved to be common features (Batygin and Stevenson,2010; Heng , 2012b).

6 In hot Jupiters temperatures may be very high (≥ 1500 K), al-lowing for thermal ionization (governed by the Saha equation)and thus fast-moving (in hot Jupiters winds ∼ 1 km s−1) electriccharges. This induces an electric current towards the interior ofthe planet, where energy is then converted into heat by ohmicdissipation.

V. CLIMATE RESPONSE AND PREDICTION

A. Introduction

In the previous section, we have investigated the cli-mate as a non-equilibrium physical system and have em-phasized the intimate relation between forcing, dissipa-tion, energy conversion, and irreversibility. The sameapproach can be brought to a more theoretical level bytaking the point of view of non-equilibrium statisticalmechanics.

Non-equilibrium statistical mechanics provides thenatural setting for investigating the mathematical prop-erties of forced and dissipative chaotic systems, which livein a non-equilibrium steady state (NESS). In this state,typically, the phase space contracts, entropy is gener-ated, and the predictability horizon is finite. Deviationsfrom this behavior are possible, but extremely unlikely. Conceptually, non-equilibrium steady states are gen-erated when a system is put in contact with reservoirsat different temperatures or chemical potentials, and onedisregards the transient behaviors responsible for the re-laxation processes (Gallavotti , 2006).

The science behind non-equilibrium statistical me-chanical systems is still in its infancy, so that, as op-posed to the equilibrium case, we are not able to predictthe properties of a system given the parameters describ-ing its internal dynamics and the boundary conditions,except in special cases where the dynamics is trivial.

It is then important to choose a suitable mathemat-ical setting for being able to state some useful generalresults and compare numerical experiments with theory.It is advantageous to assume that the forced and dissi-pative systems under our consideration possess a uniqueSinai-Ruelle-Bowen (SRB) measure (Eckmann and Ru-elle, 1985; Ruelle, 1989; Young , 2002), which guaranteesthat, regardless of the initial state of the system, thesteady state statistical properties are well defined andstable with respect to adding some noise. The invari-ant measure is supported on the attractor on the sys-tem, which is smooth along the unstable directions, andsingular along the stable directions of the flow. In thevery intuitive language of Lorenz, the attractor locallylooks like the product of a manifold and a Cantor set.Moreover, the SRB measure can be considered as thezero-noise limit of the invariant measure of the originalsystem perturbed by (small) stochastic forcing, and isthereby referred to as physical measure. Note that al-most every time numerical simulations are performed,these hypotheses are implicitly assumed, because in theabsence of an SRB measure, for most cases of practicalinterests it would make little sense to compute any sta-tistical property of the system. In the case of multistablesystems, possessing a disjoint attractor, one can assumethat the statistical properties are well defined for all ini-tial conditions belonging to the basin of attraction of the

Page 23: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

23

same of the distinct piece of the attractor.(Ruelle, 1997, 1998a,b, 2009) recently proved that in

the case of an Axiom A system, its SRB measure, despiteits geometrical complexity, has also an extremely fasci-nating degree of regularity. In fact, there is a smoothdependence of the SRB measure to small perturbationsof the flow, and it is possible to derive corresponding ex-plicit formulas. Such response formulas boils down toa Kubo-like (Kubo, 1957; Kubo et al., 1988) perturba-tive expression where the terms describing the linear andnonlinear response of the system can be written as ex-pectation values of observables on the unperturbed SRBmeasure. This approach is especially useful for studyingthe impact of changes in the internal parameters of a sys-tem or of small modulations to the external forcing, andvarious studies have highlighted the practical relevanceof Ruelle theory for studying what we may call the sen-sitivity of the system to small perturbations. Among themany results, one can mention the actual implementa-tion of the response theory as an algorithm for predict-ing the finite and infinite time ensemble-averaged impactof forcings (Abramov and Majda, 2008), the provision ofa connection between response theory and ensemble ad-joint schemes (Eyink et al., 2004), the investigation of thefrequency-dependent linear response of the system (Re-ick , 2002). Finally, some efforts have been directed at ex-tending the analysis of the frequency dependent responseto the nonlinear case (Lucarini , 2008) and on testing therobustness of the theory with simple chaotic models (Lu-carini , 2009), and then applying the Ruelle’s formulasto simples cases of geophysical relevance (Lucarini andSarno, 2011). Recently the response theory has beenused to study the impact of stochastic perturbations ofchaotic systems (Lucarini , 2012) and derive rigorouslyparametrizations for reducing the complexity of multi-scale systems (Wouters and Lucarini , 2012) (see nextSection).

B. Response formulas

Let’s consider an Axiom A dynamical system whoseevolution equation can be written as x = F (x) andlet’s assume that it possesses an invariant SRB measureρ(0)(dx). Ruelle (1997, 1998a,b, 2009) has shown thatif the system is weakly perturbed so that its evolutionequation can be written as:

x = F (x) + Ψ(x)T (t) (80)

where Ψ(x) is a weak time-independent forcing and T (t)is its time modulation, it is possible to write the mod-ification to the expectation value of a general smoothobservable A as a perturbative series:

ρ(A)t =

∞∑n=0

ρ(n)(A)t, (81)

where ρ(0)(A)t = ρ(0)(A) is the expectation value of Aaccording to the unpertubed invariant measure ρ0, whileρ(n)(A)t with n ≥ 1 represents the contribution due tonth order processes and can be expressed as a n−upleconvolution product:

ρ(n)(A)t,=

∫ ∞−∞

dτ1 . . .

∫ ∞−∞

dτnG(n)A (τ1, . . . , τn)

× T (t− τ1) . . . T (t− τn). (82)

The integration kernel G(n)A (τ1, . . . , τn) is the nth order

Green function, which can be written as:

G(n)A (τ1, . . . , τn) =

∫ρ(0)(dx)Θ(τ1) . . .Θ(τn − τn−1)

× ΛΠ(τn − τn−1) . . .ΛΠ(τ1)A(x). (83)

where Λ(•) = Ψ·∇(•) describes the impact of the pertur-bation field and Π(σ) is the unperturbed time evolutionoperator such that Π(σ)K(x) = K(x(σ)). The Greenfunction obeys two fundamental properties

• its variables are time-ordered: if j > k, τj > τk →G

(n)A (τ1, . . . , τn) = 0;

• the function is causal: τ1 < 0→ G(n)A (τ1, . . . , τn) =

0.

It is important to note that many authors suggest that,at all practical purposes, the validity of the response the-ory extends well beyond the (rich but somewhat limited)mathematical world of Axiom A systems if one considerreasonable physical systems. See discussions in (Eyinket al., 2004; Lacorata and Vulpiani , 2007; Marconi et al.,2008). One way to rationalize this statement comes fromthe so-called Chaotic Hypothesis: Axiom A systems canbe considered as good effective models of actual systemswith many degrees of freedom (Gallavotti , 1996).

Limiting our attention to the linear case we have:

ρ(1)(A)t =

∫ +∞

−∞dτ1G

(1)A (τ1)T (t− τ1) (84)

where the first order Green function can be expressed asfollows:

G(1)A (τ1) =

∫ρ0(dx)Θ(τ1)Ψ(x) · ∇A(x(τ1)), (85)

In systems possessing a smooth invariant measure, likewhen equilibrium conditions apply or stochastic forcingis imposed, we can write ρ0(dx) = ρ0(x)dx, where ρ0(x)is the so-called density. In this case, we can rewrite Eq.85 as follows:

ρ(1)(A)t =

∫ +∞

−∞dτ1Θ(τ1)

∫dxρ0(x)

×B(x)A(x(τ1)T (t− τ1), (86)

Page 24: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

24

where B(x) = −∇ ·(ρ0(x)Ψ(x)

)/ρ0(x). In other terms,

one can predict the response at any time horizon t fromthe knowledge of the lagged correlation between the cho-sen observable A and the observable B, which dependson the invariant measure ρ0 and on the perturbation vec-tor field Ψ. Equation (86) provides a very general formof the Fluctuation-Dissipation Theorem (FDT) (Lacorataand Vulpiani , 2007; Ruelle, 1998a), which extends the re-sults by (Kubo, 1957) obtained considering the classicalcase of equilibrium system immersed in a heat bath. Re-cently, the FTD for system possessing a smooth invariantmeasure result has been extended to the nonlinear case(Lucarini and Colangeli , 2012).

The more common forms of the FDT can be obtainedby taking one or more of the following assumptions: :

• the perturbation flow is the form Ψ(x) = εxi;

• the observable is of the form A(x) = xj .

where xk is the kth component of the x vector and xk isthe corresponding unit vector. In this case, Eq. 86 takesthe form:

ρ(1)(xj)t = −ε∫ +∞

−∞dτ1Θ(τ1)

∫dxρ0(x)

× ∂i log[ρ0(x)]xj(τ1)T (t− τ1),(87)

If one takes the additional simplifying assumption thatunperturbed invariant measure has a Gaussian form, sothat ρ0(x) = 1/Z exp(−β

∑Nj=1 x

2j/2), where β > 0 and

Z is a normalizing factor, we obtain:

ρ(1)(xj)t = εβ

∫ +∞

−∞dτ1Θ(τ1)

∫dxρ0(x)

× xixj(τ1)T (t− τ1)

= εβ

∫ +∞

−∞dτ1Θ(τ1)Ci,j(τ1)T (t− τ1), (88)

where Ci,j is the lagged correlation between xi and xjin the unperturbed state. See also discussion in (Cooperand Haynes, 2011; Cooper et al., 2013).

Unfortunately, the link between linear response of thesystem to external perturbations and its internal fluctu-ations seems more elusive when the unperturbed state ,as in general in case of non-equilibrium deterministic sys-tems discussed above, has a singular invariant measure.In e.g. (Ruelle, 2009) it is shown that since the invariantmeasure is singular, the response of the system containstwo contributions, such that the first may be expressedin terms of a correlation function evaluated with respectto the unperturbed dynamics along the space tangent tothe attractor (unstable manifold) and is formally identi-cal to what given in Eq. 86. This part of the responsedecays rapidly due to mixing. On the other hand, thesecond term, which has no equilibrium counterpart, de-pends on the dynamics along the stable manifold, and,

hence, it may not be determined from the unperturbeddynamics and is also quite difficult to compute numer-ically. The response to forcings along the stable direc-tions can also be shown to converge as perturbations inthese directions get damped exponentially fast by defi-nition. These properties suggest the basic fact, alreadysuggested heuristically by (Lorenz , 1979), that in the caseof non-equilibrium systems internal and forced fluctua-tions of the system are not equivalent, the former beingrestricted to the unstable manifold only.

Despite such a serious mathematical difficulty, the ap-plication of FDT, even in extremely simplified, quasi-Gaussian, approximation, has enjoyed a good success inclimate science, even if it is clear that the ability of FDTin predicting the response to perturbation depends crit-ically on the choice of the observable of interest, on thelength of the integrations needed for constructing theapproximation of the invariant measure, and, of course,on the validity of the linear approximation (Cooper andHaynes, 2011; Cooper et al., 2013; Gritsun and Bransta-tor , 2007; Langen and Alexeev , 2005).

There are, in fact, various ways to circumvent the prob-lem of the rigorous non-equivalence between forced andfree fluctuations. First, one can consider the smooth-ing effect due to unavoidable physical or numerical noise.On a more basic level, one can observe that when con-sidering smooth, coarse-grained observables, one expectto see little influence of the fine structure of the invari-ant measure of chaotic deterministic systems, as projec-tions from high-dimensional spaces to lower dimensionalones are involved(Marconi et al., 2008). In the case ofclimate problems, one expects that, for a given lengthof the control run needed for accumulating the statistics,the FDT will perform better in predicting the response ofthe system to perturbations if we consider as observableA quantities like the globally averaged surface tempera-ture or the globally averaged emitted longwave radiationat the top of the atmosphere, than in predicting the re-sponse of, e.g. the surface temperature in an individualgrid point. Moreover, it is indeed not obvious a priory todetermine the range of forcings one which the responseof a given observable will be closely approximated by thelinear component of the response. Further comments canbe found at the end of Sec. VI

C. Computing the Response

1. Spectroscopic method

If we select T (t) = ε cos(ω0t) = ε/2(exp(−iω0t +exp(iω0t)) as modulating factor of the perturbation field

Page 25: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

25

Ψ(x), from equation Eq. 84 we derive:

ρ(1)(A)t = ε/2

∫ +∞

−∞dτ1G

(1)A (τ1) exp(−iω0(t− τ1))

+ ε/2

∫ +∞

−∞dτ1G

(1)A (τ1) exp(iω0(t− τ1))

= ε/2 exp(−iω0t)χ(1)A (ω) + c.c. (89)

where χ(1)A (ω0) is the Fourier Transform of G

(1)A (t), usu-

ally referred to as linear susceptibility, evaluated at fre-quency ω = ω0, and c.c. indicates complex conjugate.Therefore, under the hypothesis of linearity, by perform-ing an ensemble of experiments where the forcing is of theform T (t) = ε cos(ω0t), we can extract the linear suscep-tibility at frequency ω by selecting the ω0 component ofthe Fourier transform of the signal ρ(1)(A)t obtained bytaking the ensemble average of the difference between thetime series of A in the perturbed and unperturbed case.By changing systematically the frequency ω of the forc-

ing, one can reconstruct the susceptibility χ(1)A (ω) on a

chosen interval of frequencies. It is useful to recapitulatesome useful features of the susceptibility:

• In general the response of a system to a forcing canbe extremely different at for time scales of forcings.Therefore, one should use with great caution the

concept of scale analysis. Moreover, limω→0 χ(1)A (ω)

is the static response of the system and is a posi-tive number, which gives what is usually called thesensitivity.

• Resonances in the susceptibility function corre-spond to spectral ranges where the system is ex-tremely sensitive to forcings. In Fig. 13 we showthe real and imaginary part of the susceptibilityfor z variable of the (Lorenz , 1963) model - see Eqs.(49)-(51) - for the classical values of the param-eters (m = 1, σ = 10, r = 28, β = 8/3) anda given choice of the forcing (Ψ(x) = [0;x; 0]>,T (t) = 2ε cos(ωt)). We find that for ω ∼ 8.3,a very peaked spectral feature is apparent. Sucha resonance is due to the Unstable Periodic Or-bits (UPOs) (Cvitanovic, 1988) of the system withthe corresponding period (Eckhardt and Ott , 1994).UPOs populate densely the attractors of chaoticsystems and constitute the so-called skeleton ofthe dynamics. One can, in principle, reconstructthe whole statistics focussing on such special or-bits then on the usual trajectories of the flow. Inthe case geophysical flows, UPOs have been associ-ated to modes of low-frequency variability (Grit-sun, 2008). One can, more qualitatively, asso-ciate resonance to positive feedbacks acting on timescales corresponding to the resonant frequency.

• While |χ(1)A (ω)| measures the amplitude of the re-

sponse of the system to perturbation at frequency

ω, arctan(=χ(1)A (ω)/<χ(1)

A (ω)) gives the phasedelay between the forcing and the response, be-

cause the <χ(1)A (ω) (=χ(1)

A (ω)) gives the compo-nent of the response that is in phase (our of phase)with the forcing. Depending on the forcing, on thesystem, and on the observable, this angle can varysignificantly even in a relatively small range of fre-quencies, as a result of resonances. There is, ingeneral, nothing nonlinear in the presence of delaysbetween forcing and response.

• The two components =χ(1)A (ω) and <χ(1)

A (ω)are connected by integral equations, the so-calledKramers-Kronig relations (Lucarini , 2008, 2009;Lucarini et al., 2005; Ruelle, 2009). Such relationshave their foundation in the causality of the Greenfunction and establish a fundamental connectionbetween the response at different time scales:

<χ(1)A (ω) =

2

πP

∫dω′

ω′=χ(1)A (ω′)

ω′2 − ω2; (90)

=χ(1)A (ω) = −2ω

πP

∫dω′

ω′<χ(1)A (ω′)

ω′2 − ω2. (91)

where P indicates that the integral is taken in prin-cipal part. In particular one finds that:

<χ(1)A (0) =

2

π

∫dω′=χ(1)

A (ω′)ω′

, (92)

which provides a fundamental link between thestatic response - the sensitivity - and the out-of-phase response at all frequencies. An enormousliterature exists in optics, acoustics, condensedmatter physics, particle physics, signal processingon the theory and on the many applications ofKramer-Kronig relations and on the related sumrules, which provide integral constraints related tothe asymptotic behavior of the susceptibility (Lu-carini et al., 2005). This approach boils down toa spectroscopic investigation of the system and isthoroughly discussed, also in its numerical aspects,in (Lucarini , 2009; Lucarini and Sarno, 2011).

In Fig. 14 we present the real and imaginary part ofthe susceptibility of the mean energy e of the celebratedLorenz (1996) model:

dxidt

= xi−1(xi+1 − xi−2)− xi + F (93)

where i = 1, 2, ....., N , and the index i is cyclic so thatxi+N = xi−N = xi, and e = 1/N

∑Nj=1 x

2j/2. The

quadratic term in the equations simulates advection, thelinear one represents thermal or mechanical damping andthe constant one is an external forcing. The evolution

Page 26: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

26

100

101

102

−15

−10

−5

0

5

10

15

20

ω

Re[χ],Im

[χ]

FIG. 13 Measured real (blue line) and imaginary (red line)part of the susceptibility z variable of the Lorenz 63 model.Data from Lucarini (2009)

.

10−2

10−1

100

101

102

103

−0.5

0

0.5

1

1.5

2

ω

Re[χ],Im

[χ]

ω l ωh

FIG. 14 Measured real (blue line) and imaginary (red line)part of the susceptibility for the average energy of the Lorenz96 model. The rigorous extrapolation of the susceptibilityobtained via Kramers-Kronig analysis is reported (real part:black line; imaginary part: magenta line). Data from Lucariniand Sarno (2011)

.

equations are invariant under i → i + 1, so that the dy-namics is the same for all variables. Please find detailson the experiments in Lucarini and Sarno (2011), per-formed using N = 40 and F = 8. The system is per-turbed by the vector field Ψ(x) = [1; . . . ; 1]> modulatedby T (t) = 2ε cos(ωt). The resulting real and imaginary

part of χ(1)e (ω) are reported in Fig. 14, together with

the output of the data inversion performed via Kramers-Kronig relations. Once we obtain the susceptibility, asdiscussed in (Lucarini and Sarno, 2011), it is possible toderive the corresponding Green function by applying theinverse Fourier Transform. This is the first applicationof the Kramers-Kronig theory in a geophysical context.

2. Broadband forcing

If, instead, we select T (t) = δ(t), we derive from Eq.

85 that ρ(1)(A)t = G(1)A (t), i.e. the Green function corre-

sponds to the relaxation of an ensemble of trajectories ofthe system after a finite displacement along Ψ(x). Ob-

viously, we have that ρ(1)(A)ω = χ(1)A (ω), where the ˜

symbol, indicates, as customary, that a Fourier Trans-form has been applied, so that the Fourier Transform ofthe signal is the linear susceptibility. Therefore, usingjust one ensemble of experiments where the perturbationis described by an impulsive forcing, we can gather thesame information on the response of the system which,in the previous case required an accurate sampling of dif-ferent frequencies.

It is not always possible or advantageous to performeither of this class of experiments. Let’s look at the prob-lem from a slightly more general point of view., We applythe Fourier Transform to both sides of Eq. 84 and obtain:

ρ(1)(A)ω = χ(1)A (ω)T (ω) (94)

Choosing a sine or cosine function with argument ω0tfor the function T (t) amounts to selecting as T (ω) thesum of two δ’s centered in ω = ω0. Therefore, the input(forcing) allows only a small portion of the information toderived on the system from the output (response). Let’sassume that we choose the modulation T (t) such thatT (ω) is not vanishing for any ω. A good option is toconsider a broadband modulation, i.e. such that T (ω)is not vanishing for any ω and decreases slowly enough,e.g. like a power law. At this purpose, it is enoughto choose a function which is differentiable only a finitenumbers of times, including the case where there T (t) isnot continuous. If we perform an ensemble of simulationsof the forced system, measure ρ(1)(A)ω, we can readilyderive:

χ(1)A (ω) =

ρ(1)(A)ω

T (ω)(95)

Therefore, one single set of experiments is, in fact all weneed to do to learn about the linear response propertiesof the system for the observable A. If we want to predictthe response at finite and infinite time of the system toforcing with the same spatial pattern Ψ(x) but with dif-

ferent time modulation R(t), we can derive G(1)A (t) from

χ(1)A (ω) obtained via Eq. 95, and then plug it into Eq.

84. Alternatively, one can write:

ρ(1)(A)Rω = ρ(1)(A)TωR(ω)

T (ω)(96)

where the upper indices R and T have been inserted forclarity, and then compute the inverse Fourier transformto derive the response at all times.

Page 27: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

27

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.50

0.5

1

1.5

2

2.5

3

t

G(1

)e(t)

FIG. 15 Linear Green function G(1)e (t) for the average energy

e of the Lorenz (1996) model using the perturbation given inEq. 97. Compare with Fig. 4 in Lucarini and Sarno (2011)

.

D. Prediction via Response theory

The real test of the quality of an experimentally de-

rived linear Green function G(1)A is the assessment of

its ability to support predictions about the system’s re-sponse to any temporal pattern of forcing R(t). The realbenefit of the broadband approach described here relies

on exploiting linearity, and so deriving G(1)A from just one

ensemble of simulations, each performed with the same

modulation T (t). Computing the G(1)A per se might be,

in fact of little relevance 7.

At this regard, we have performed additional experi-ments on the (Lorenz , 1996) model mirroring what pre-sented in Sec. V.C.1. In this case, we have chosen astime modulation T (t) = εΘ(t), whose spectrum is indeedbroadband (T (ω)/ε = πδ(ω) + iP[1/ω]). In this case, wehave quite simply that:

G(1)e (t) =

d

dtρ(1)(e)t. (97)

Using about 1/100 of the computing time needed in (Lu-carini and Sarno, 2011), we have produced an estimateof the Green function of comparable quality; see Fig. 15.Additionally, we decided to check the predictive powerof the reconstructed Green function given in Fig. 15 bytesting its performance in predicting, through Eq. 84,the response of the system to a perturbation having tem-poral pattern given by T (t) = ε sin(2πt) (ε = 0.25). Theresults are presented in Fig. 16. The agreement betweenthe measured value of ρ(1)(e)t and the value predicted

using∫dτG

(1)e (τ)T (t − τ) is remarkable. One must em-

phasize that the agreement is comparable if one selectsε = 1, thus moving away from the linear regime.

7 The authors are thankful to F. Cooper for pointing out this issue.

0 1 2 3 4 5 6 7 8 9 10−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

t

ρ(1

) (e) t,∫dτG

(1)

e(t

−τ)T

(τ)

FIG. 16 Linear Green function G(1)e (t) for the average energy

e of the Lorenz (1996) model using the perturbation given inEq. 97. Compare with Fig. 4 in Lucarini and Sarno (2011)

.

0 50 100 150 2000

1

2

3

4

5

6

7

8

9

10

t (years)

∆Ts(K

)

FIG. 17 Comparison between the observed response in theglobally averaged surface temperature TS to an exponentialincrease at 1% per year of CO2 concentration from 360 to720 ppm, with subsequent stabilization (black line), with theprediction performed using linear response theory (blue line).The experiment has been carried out using an ensemble of200 climate simulations performed using the PLASIM model(Fraedrich et al., 2005)

.

E. Climate Response, Climate Change prediction

Let’s take inspiration from the previous example in or-der to get some results of stricter geophysical relevance:we want to perform predictions on the impact of increasesin the CO2 concentration on the globally averaged sur-face temperature as simulated by a climate model.

So, we consider that x = F (x) is the system of equa-tions describing the discretized version of a given modelof the continuum PDEs describing the evolution of theclimate in a baseline scenario with set boundary condi-tions - and in particular the the value of the CO2 con-centration and of the solar constant. We assume, forsimplicity, that system model does not feature daily orseasonal variations in the radiative input at the top of

Page 28: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

28

the atmosphere. Let’s choose for the observable A theglobally averaged surface temperature of the planet TS ,let’s consider as perturbation field Ψ(x) the convergenceof radiative fluxes due to change in the logarithm of theatmospheric CO2 concentration. We want to be able topredict at finite and infinite time the response of the thesystem to one of the standard CO2 forcing scenario givenby the IPCC by performing an independent set of per-turbed model integrations.

The test perturbation is modulated by the functionεΘ(t), where ε is such that we double the amount of CO2

concentration in the atmosphere. Our goal is to predictthe climate response to the customary 1% increase ofCO2 concentration from the baseline value to its double.We select as baseline concentration [CO2] = 360 ppm.We perform 200 simulations, each lasting 200 years forboth scenarios of CO2 forcing. Our experiments are per-formed using PLASIM (Fraedrich et al., 2005) with a T21spatial resolution, 10 vertical layers in the atmosphere,and swamp ocean having depth of 50 m.

From the time series of the ensemble mean of thechange of TS - ρ(1)(TS)t - resulting from the sudden in-

crease in the CO2, we derive the Green function G(1)TS

(t)using Eq. 97. Climate sensitivity is, in fact, defined byEq. 92. Given the chosen pattern of forcing, we canrewrite is as follows:

∆T = <χ(1)A (0) =

2

π

∫dω′<ρ(1)(TS)ω′, (98)

which relates climate response at all frequencies to itssensitivity.

We then convolute the Green function with the tem-poral pattern of forcing of the second set of experiments.Since our relevant control parameter is the logarithmof the CO2 concentration, the pattern of forcing corre-sponding to an exponential increase is, in fact, a straightline. The results are presented in Fig. 17, where we com-pare the predicted pattern of increase (blue line) with themeasure one (black line). The agreement is remarkable,both on the short and on the long time scales, while adiscrepancy lower than 10% exists between 50 and 100years lead time.

Apparently, despite all the nonlinear feedbacks of theclimate model, the response to changes in the logarithmof CO2 concentration can be accurately described by lin-ear response theory at all time scales, not only in thestatic case, corresponding to climate sensitivity, which isour model, is extremely high as a result of lack of theseasonal cycle. Nonlinearity in the underlying equationsand presence of strong positive and negative feedbacksdo not rule out the possibility of constructing accuratemethods for predicting the response. In fact, the meth-ods described here could be extended to the nonlinearcase by looking at the response in the frequency domain

(Lucarini , 2008, 2009), even if the data quality require-ment is obviously stricter.

The result presented here suggests that, in fact, manyof the scenarios of greenhouse gases concentration in-cluded in the IPCC reports (IPCC , 2001, 2007, 2013)may in fact be partly redundant, as for certain variablesmight be accurately described by linear response theorystarting from just one scenario. Equations 95 - 96 con-stitute the basis for predicting climate response at allscales.

Obviously, with a given set of forced experiments, it ispossible to derive the sensitivity to the the given forcingfor as many climatic observables as desired. It is im-portant to note that, for a given finite intensity ε of theforcing, the accuracy of the linear theory in describingthe full response depends also on the observable of inter-est: the performance of the methods described here maybe excellent for TS but probably not for the precipitablewater in the atmosphere, which, in first approximation,increases exponentially on the atmospheric temperature.Moreover, the signal to noise ratio and, consequently, thetime scales over which predictive skill is good may changea lot from variable to variable.

F. Noise and Prediction

Of course, in order to talk about predictability, we needto specify what are the time scales over which we expectto have satisfactory predictive skills. In fact, linear re-sponse theory allows for deriving some scaling laws foraddressing this matter. The main obstacle for achiev-ing a good degree of predictability is the uncertainty onthe estimate of response signal given in Eq. 94 from theoutcomes of the numerical experiments because of thefiniteness of the ensemble and of the duration of eachnumerical simulation.

The limits to predictability can be estimated as fol-lows. As a result of the presence of a finite ensembleof N simulations of finite length L, the estimates of theabsolute value of the response |ρ(1)(A)obsω | due to a forc-ing with spatial pattern Ψ(x) and temporal pattern T (t)differs from the true real and imaginary parts of the re-sponse |ρ(1)(A)ω| by a background function η(ω), whichis a white noise of standard deviation σ ∼ αN−1/2L−1/2,where α is a constant, because the system decor relatesexponentally fast thanks to its chaotic behavior. In thelimit of either N →∞ or L→∞, if our estimate - as weexpect - in unbiased, we obtain |ρ(1)(A)obsω | → |ρ(1)(A)ω|.

From Eq. 95 we derive, that, if |T (ω)| ∼ βω−ν

for large values of ω, then the absolute value of

|χ(1)A (ω)obs| is affected by a non stationary error due to

undersampling, whose standard deviation increases as

αβ−1N−1/2L−1/2ων . Let’ assume that |χ(1)A (ω)| ∼ κω−γ

as ω → ∞, with κ also a constant. Clearly, when ω islarge enough, the absolute value of the true susceptibil-

Page 29: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

29

ity |χ(1)A (ω)| s smaller than the frequency dependent noise

due to the unavoidable presence of η(ω). This happenswhen

ω & ωcrit =

βκ

)−1/(γ+ν)

N1/(2γ+2ν)L1/(2γ+2ν). (99)

In other terms, for frequencies larger than ωcrit, the esti-

mate of χ(1)A (ω) is of little utility. This translates into

the fact that the corresponding empirical G(1)e (t) will

perform poorly in predicting changes occurring on timescales smaller than

tcrit ∼(α

βκ

)1/(γ+ν)

N−1/(2γ+2ν)L−1/(2γ+2ν). (100)

Therefore, we are able to perform prediction on smallerand smaller scales if we have long integrations, manymembers of the ensemble, or a slow asymptotic decay ofthe susceptibility, and it is advantageous to choose, whenprobing the system, an observable which is as broadbandas possible. Equations 99-100 provide a guidance on howthe computational needs required to achieve predictiveskills on a given time scale. Similar results can be ob-tained by looking at the impact of noise due to under-sampling on the real and imaginary part of the responseand of the susceptibility.

VI. MULTISCALE SYSTEMS ANDPARAMETRIZATIONS

A. Introduction

The climate is a complex system featuring non-trivialbehavior on a large range of temporal and spatial scales(Lucarini , 2011; Peixoto and Oort , 1992; Vallis, 2006).It is clear that, no matter which are the available com-puting resources, we are able to simulate explicitly onlythe variable relevant for given ranges of spatial and tem-poral scales. Different choices of such ranges, in fact, cor-respond to different approximate theories of geophysicalfluid dynamics aimed at describing specific phenomenolo-gies, the most prominent case being, probably, the caseof quasi-geostrophic theory.

A manifestation of the inability to treat ultraslow vari-ability can be found in the usual practice in climate mod-eling of choosing fixed or externally driven boundary con-ditions, such as done when assuming a fixed extent forthe land-based glaciers, and, consequently, for the sea-level, or imposing a specific path of CO2 concentrationfor the atmosphere. See, at this regard, the distinctionbetween macroweather and climate proposed by (Lovejoyand Schertzer , 2013). Instead, the impossibility of treat-ing accurately fast processes requires the construction ofso-called parametrizations able to account, at least ap-proximately, the effect of the small scales on the large

scales, as a function of the properties of the large scalevariables. The process of deriving parametrizations isalso called model reduction or variable elimination.

Whereas deterministic and stochastic parametrizationsare by now common in geophysical fluid dynamical mod-els (Palmer and Williams, 2009), many of the approachesused up to now have been based on the existence of a timescale separation between microscopic and macroscopicprocesses. If one does assume a vast time-scale separationbetween the slow variables X and the fast variables Y ,averaging and homogenization methods (Arnold , 2001;Kifer , 2004; Pavliotis and Stuart , 2008) allow for deriv-ing an effective autonomous dynamics for the X vari-ables, able to encompass the impact of the dynamics ofthe Y variables. In this prototypical case, one speak ofa two-level system where the X and Y variables consti-tute the two levels, respectively. A typical example ofhow this kind of theory can be used in climate science isthe setting considered by Hasselmann, where fast weathersystems influence slow climate dynamics. These methodswill be reviewed in section VI.B.

Unfortunately, in many practical cases of interests geo-physical fluid dynamics, such a scale separation does notexist,so that there is no spectral gap able to support uni-vocally the identification of theX and Y variable. In fact,when the resolution of a numerical model is changed, allthe parametrizations have to be re-tuned, because theset of resolved variables has changed. The possibility ofconstructing a robust theory of parametrizations wouldallow for constructing schemes applicable seamlessly to amodel when its resolution is changed.

Here we will focus on analytical methods that allowone to derive reduced models from the dynamical equa-tions of the full model. Projector operator techniqueshave been introduced in statistical mechanics with thegoal of effectively removing the Y variables. In partic-ular, considerable interest has been raised by the Mori-Zwanzig approach, through which a formal - albeit prac-tically inaccessible - solution for the evolution of the Xvariables is derived (Zwanzig , 2001). These equationsin general contain both a correlated noise term and amemory term. Some attempts have been made to makeapproximation to the Mori-Zwanzig projected equationsto obtain practically useful equations. In applications ofstochastic mode reduction in climate science, the memoryterm is usually not taken into account. This term couldhowever be very relevant in systems without a time-scaleseparation, as for example in the parametrization of cloudformation in an atmospheric circulation model. The pres-ence of memory in such systems has been discussed in(Bengtsson et al., 2013; Davies et al., 2009; Piriou et al.,2007). The Mori-Zwanzig projection operator techniqueand related approaches are described in section VI.D.

Besides a limit of infinite time scale separation, thelimit of weak coupling between the different scales canalso be considered. In this limit, the dynamics retains the

Page 30: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

30

correlated noise and memory dependence that appearedin the Mori-Zwanzig reduced equations. The advantageof looking at this limit is however that the noise autocor-relation function and memory kernel can now be writtenas simple correlation and response functions of the un-resolved dynamics. This approach will be reviewed insection VI.E

Many other parametrization methods have been pro-posed from a more heuristic point of view. Particularlyrelevant with respect to the current study are the re-sults by (Kwasniok , 2012) indicating that non-Markovianparametrizations are superior to Markovian ones when alarge time scale separation is absent. A similar conclusioncan be reached from the work on empirical model reduc-tion by Chekroun et al. (2011a), where non-Markovianmodels are fitted to data without a clear time scale sep-aration.

B. Averaging and homogenization

When applying averaging and homogenization tech-niques, one considers dynamical systems where a smallparameter ε controls the time scale separation between aslow and fast evolution in the system. The prototypicalset of equations for such a problem is

x =f(x, y)

y =1

εg(x, y)

The principle behind averaging can be considered sim-ilar to the law of large numbers. If the time scale sep-aration becomes large, the fast system will go throughits entire attractor on the time scale of evolution of theslow system. In the limit of ε going to zero, the slowsystem x will therefore ’see’ all possible values of y. Asin the law of large numbers, the overall effect of all thesecontributions can be substituted by one single value. Itcan be shown that for finite time T , the trajectory x(t)converges to a solution of the equation

X = F (X)

where F (X) = ρ∞,X(f(X, y)) is the averaged value of thetendency, the average taken over the invariant measureρ∞,X of the dynamical system

y =g(X, y)

resulting for y when X is considered as a fixed forcingparameter. The convergence to the averaged solution de-pends on the type of the y dynamical system (stochasticor chaotic), on the coupling (full two-way or fast-to-slowonly), and on the initial conditions (Kifer , 2009). Exam-ples of dynamical systems can be constructed where for alarge set of initial conditions of y, the solution for x does

not converge to the averaged solution (Kifer , 2008). Fur-thermore, if the y system has long time correlations, suchas in a system with regime behaviour, the homogenizedsystem may converge badly and an extension based ona truncation of the transfer operator has been proposed(Schutte et al., 2004).

Let’s consider a simple example system.

x =xy

y =− λ

εy +

1√ε

dW

dt

The y system is an Ornstein-Uhlenbeck process, indepen-dent of x. The invariant measure of the fast y system isa Gaussian distribution with zero mean. Taking the av-erage of f(x, y) = xy, we see that the averaged equationin this case is the uninteresting equation X = 0.

This example immediately motivates the use of homog-enization methods. Here one scales the equation to alonger time scale θ = εt, the so called diffusive time scaleand then performs the asymptotic expansion. Similarlyto how correctly rescaling the sums of the law of largenumber leads to the more interesting central limit theo-rem instead, also in the setting of time scale separatedsystems, we get stochastic behavior on the diffusive timescale. For the example considered above, we get a weakconvergence to a reduced stochastic differential equationfor the X variable instead of the trivial dynamical systemobtained before.

(Abramov , 2012) has recently presented an applica-tion of this method to deriving a simplified dynamicsfor a system of geophysical relevance. A study of aver-aging and homogenization for idealized climate models,with a range of examples, can be found in (Monahan andCulina, 2011). Another rather successful attempt in thisdirection is given in (Majda et al., 2001).

A study of homogenization for geophysical flows wasperformed in Bouchet et al. (2013). The slow system isconsidered to be the evolution of zonal jets of a barotropicflow, which is forced by noise. The fast degrees of freedomare those representing the fast non-zonal turbulence. Ho-mogenization has also been applied in (Dolaptchiev et al.,2012) to the Burgers equation, where the slow variablesare taken to be averages over large grid boxes and thefast variables are the subgrid variables.

When one wants to consider very large time scales (forexamples times of the order of exp(1/ε)), one needs tolook beyond the central limit type theorems of homog-enization and consider so called large deviation results.These describe for example the transitions between dis-connected attractors of the averaged equations (Kifer ,2009).

Page 31: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

31

C. Dyson decomposition

The Dyson decomposition is an operator identity thatallows to decompose the time evolution of observables ofa dynamical system into a background evolution plus aterm representing the effect of a perturbative vector field.This decomposition forms the basis of the Mori-Zwanzigmethod and the weak coupling method described in sec-tions VI.E and VI.D.

Given a non-linear dynamical system x = F (x) ona manifold M, the analysis of perturbations can beformally simplified by turning the problem into a lin-ear equation for the evolution of observable functionsA : M → R. To this end, we define a linear differen-tial operator L(x) = (F (x).∇) called the Liouville op-erator, determining the time derivative of observables,dependent on time through a solution x(t) of the dy-namical system. We thus have that by the chain ruleA(x(t)) = L(x(t))A(x(t)). A solution of A over time isthen given by A(t) = Π(t)A(0) where Π(t) = exp(Lt).Although this equation does not tell us anything aboutthe actual value of the solution, it does allow us to ana-lyze perturbations, where the defining vector field F , andhence also L, are slightly altered.

As described e.g. in (Evans and Morriss, 2008), theseperturbative relations can be easily derived formally inthe resolvent formalism, by taking the Laplace transformof Π(t):

LΠ(s) =

∫ ∞0

dt exp(Lt) exp(−ts) = (s− L)−1 (101)

If L consists of a perturbation around an operator L0,i.e. L = L0 + L1, with L1 small in an appropriate sense,we can expand the Laplace transform using the equality

(A−B)−1 = A−1 +A−1B(A−B)−1 . (102)

In the case of the Laplace transform in Eq. 101, we takeA = s−L0 and B = L1, so that the A−1 and (A−B)−1

terms are themselves the Laplace transforms of Π0(t) =exp(L0t) and Π(t) respectively. Inverting the Laplacetransform, the product of transforms gives a convolutionterm, resulting in the following decomposition of Π(t):

Π(t) = Π0(t) +

∫ t

0

dτΠ0(t− τ)L1Π(τ) (103)

Another decomposition can be obtained when making useof the following equality for operator inverses:

(A−B)−1 = A−1 + (A−B)−1BA−1 . (104)

This gives rise to:

Π(t) = Π0(t) +

∫ t

0

dτΠ(t− τ)L1Π0(τ) (105)

These decompositions separate the evolution under thebackground evolution defined by L0 from the effect of theperturbation L1.

As the right hand side of Eq. (103) and (105) stillcontain Π(t), they can be iterated by to expand Π(t) −Π0(t) at different orders of the perturbation L1.

D. Projection operator technique

In the case of the Mori-Zwanzig approach (Mori et al.,1974; Zwanzig , 1960, 1961) a projection is carried out onthe level of the observables to remove unwanted, irrele-vant and usually fast degrees of freedom. Here the ex-pansion is performed around the evolution that involvesonly the relevant part of the phase space.

If a dynamical system is defined on a manifoldM, onedefines a projection P from the space of observable func-tions on the full phase spaceM to a space of observableswhich are considered to contain only the interesting dy-namics. Many different choices are possible; if the mani-foldM consists for example of a product of submanifoldsK of relevant and L of irrelevant variables, one can takea conditional expectation with respect to a measure onM, given the value of the relevant variables x ∈ K:

(PA)(x) =

∫N A(x, y)ρ(x, y)dy∫N ρ(x, y)dy

.

It is easily verified that this is a projection, i.e. P2 =P and that P is orthogonal w.r.t the product definedby ρ: 〈A,B〉 =

∫M ρAB. Another possible choice is a

projection onto a set of functions on M, such as linearfunctions of the coordinates in a Euclidean phase space.

The evolution operator L is now split into its projec-tion PL onto the relevant space of observables and thecomplement QL := (1−P)L. As described by (Zwanzig ,2001), a generalized Langevin equation can then be de-rived based on Eq. 105. We write the Liouville equationfor an observable A as

dA(t)

dt=LA(t) = etLLA

=etLPLA+ etLQLA

The factor exp(tL) in the second term can be furtherexpanded by making use of Eq. 105 with L0 = QL. Thisgives the following equation

dA(t)

dt=etLPLA

+ (etQL +

∫ t

0

ds e(t−s)LPLesQL)QLA

It is then argued (Zwanzig , 2001) that this equation is ageneralization of the Langevin equation, where the sec-ond term is a correlated noise term dependent on the ini-tial conditions of the irrelevant degrees of freedom and

Page 32: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

32

the third term represent the memory of the system dueto the presence of irrelevant variables that have inter-acted with the relevant ones in the past. Note that wehave done nothing more than manipulating the originalevolution equation A = LA. Correspondingly, the Mori-Zwanzig equation in itself does not simplify the problem.In order to derive a set of equations that are useful for nu-merical simulations, assumptions need to be made aboutthe dynamical system.

Several approximations to the Mori-Zwanzig equationshave been proposed in the literature. There are the shortand long memory approximations made in the method ofoptimal prediction (Bernstein, 2007; Chorin and Stinis,2006; Chorin and Hald , 2013; Chorin et al., 1998, 2000,2002, 2006; Defrasne, 2004; Hald and Kupferman, 2001;Park et al., 2007).

In the limit of an infinite time-scale separation betweenthe relevant and irrelevant variables, the stochastic com-ponent of the parametrization can be represented as awhite noise term, while the non-Markovian term van-ishes, as the irrelevant variables decorrelate quickly, thisbrings us back to the homogenization method of sec-tion VI.B. For a discussion of the applicability of Mori-Zwanzig in climate sciencess, see the paper by Gottwald(2010).

E. Weakly coupled systems

We now consider dynamical systems consisting of twosystems with a weak coupling. In this case an expansionof the dynamics can be made in orders of the coupling,giving insight into what properties of the coupled systemsdetermines the memory kernel and correlated noise thatappeared in the Mori-Zwanzig approach (Wouters andLucarini , 2012, 2013).

A possible application of this theory in climate sciencecan be found in the interaction between cloud forma-tion and large scale atmospheric flow, where there is nodistinct time scale separation, but instead the couplingcould be considered as weak.

In this setting the background vector field F consistsof a Cartesian product (FX , FY )> of the vector fields FXand FY defining the autonomous X and Y dynamics.The perturbing vector field δF is a coupling (ΨX ,ΨY )>

between the two systems. The full dynamical system isgiven by

dX

dt= FX(X) + ΨX(X,Y )

dY

dt= FY (Y ) + ΨY (X,Y ) (106)

For simplicity of presentation, for now we consider thecase where ΨX(X,Y ) = ΨX(Y ) and ΨY (X,Y ) =ΨY (X). We will come back to the general case later.

Writing (106) in terms of observables using the chainrule, we have

dA(X,Y )

dt=(LX(X,Y ) + LY (X,Y ))A(X,Y )

=(FX(X) + ΨX(X,Y )).∇XA(X,Y )

+ (FY (Y ) + ΨY (X,Y )).∇YA(X,Y )

where ∇X and ∇Y denote the gradients with respect tothe variables in X and in Y respectively, LX = (FX +ΨX)∇X and LY = (FY + ΨY )∇Y .

We now do a calculation in the style of the Mori-Zwanzig one in Section VI.D for the dynamical systemgiven in Eq. 106 and for observables AX that only dependon the relevant variables X. We start the dynamical sys-tem at a time −t in the past. The initial condition X0

of the resolved variables is assumed to be known exactly.The state of the Y variables is not known exactly. Assum-ing that the coupled system is its invariant distributionρ allows us to use the knowledge of X to further spec-ify the distribution of Y . The Y variable is distributedaccording to the conditional distribution of ρ given X0.Expectation values of observables of Y are given by∫

dY B(Y )ρ(X0, Y )∫dY ρ(X0, Y )

Since the coupling in the dynamical system correspond-ing to ρ is small, the invariant measure ρ is a perturba-tion around the uncoupled product measure ρ0,X ⊗ ρ0,Y ,where ρ0,X and ρ0,Y correspond to the unperturbed vec-tor fields FX and FY . Hence to zeroth order the condi-tional distribution of Y is simply ρ0,Y .

We now try to understand what the distribution oftendencies of X is, given these initial conditions. As inSection VI.D, we first perform a projection of the evolu-tion equation of AX , in order to separate the X and Yvariables. As projection, we choose the conditional ex-pectation given by ρ0,Y , i.e. PA =

∫dY A(X,Y )ρ0,Y (Y ).

Applying this projection to the time derivative of AX de-composes the coupling ΨX(Y ) into its average value andfluctuations around it. We then apply an expansion interms of Ψ to the evolution of the fluctuations.

The time derivative of AX at t = 0 is given by

d

dtAX(X,Y, t)|t=0 = (LXAX)(X,Y )

=((PLX +QLX)AX)(X,Y )

=(FX(X) + ρY (ΨX))∇XAX(X)

+ (ΨX(Y )− ρY (ΨX))∇XAX(X) (107)

We now want to find a formal solution for ΨX(Y ) −ρY (ΨX) that we can insert into the previous equation.

The evolution of ΨX is given by

ΨX(Y ) =et(LX+LY )ΨX(X0, Y0,−t)=et(LX+LY )ΨX(Y0)

Page 33: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

33

Making use of the decomposition of the Liouvillian L =LX + LY into L0(X0, Y0) = FX(X0)∇X + FY (Y0)∇Yand L1(X0, Y0) = ΨX(Y0)∇X + ΨY (X0)∇Y , we deriveby repeated use of Eq. 103 that

ΨX(Y ) =et(L0+L1)ΨX(Y0) = etL0ΨX(Y0)

+

∫ t

0

dτe(t−τ)L0L1eτL0ΨX(Y0) +O(L2

1)

(108)

Inserting this equation in (107), we get

(d

dtAX)(X,X0, Y0, t)|t=0

= (FX(X) + ρY (ΨX) + σ(t, Y0))∇XAX(X)

+

(∫ t

0

dτh(t, τ,X0, Y0)

)∇XAX(X))

where

σ(t, Y0) =etFY (Y0)∇Y ΨX(Y0)− ρY (ΨX)

h(t, τ,X0, Y0) =Π0(t− τ)L1(X0, Y0)Π0(τ)ΨX(Y0)

Due to the commutation of FX∇X and FY∇Y , we havethat

h =(e(t−τ)FX∇X ΨY (X0)

)× e(t−τ)FX∇X∇Y eτFY∇Y ΨX(Y0)

Now we can make use of the fact that Y is distributedaccording to ρY , the invariant measure under the flowgenerated by FY . The average of σ is then zero and theauto-correlation is given by the auto-correlation of thecoupling function in the uncoupled Y system:

ρY (σ(t, Y0)) = 0

ρY (σ(t, Y0)σ(t+ τ, Y0))

= ρY(ΨX(Y0)eτFY∇Y ΨX(Y0)

)and the averaged memory kernel is given by a responsefunction of the uncoupled Y system:

h = ρY (h) = ΨY (f t−τX (X0))ρY(∇Y eτFY∇Y ΨX(Y0)

)(109)

Proposing now a surrogate equation

dX(t)

dt=FX(X(t)) +M + σ(t)

+

∫ ∞0

dτh(τ, X(t− τ)) (110)

we have that the average tendency of X to third orderand its autocorrelation to second order in the couplingfunction correspond to those of the fully coupled system.

FIG. 18 Diagram describing the mean field effect of the Yvariables on the X variables. Term M in Eq. 112.

FIG. 19 Diagram describing the impact of fluctuations of theY variables on the X variables. Term σ in Eq. 112.

F. Response Theory

The response theory is a rather flexible tool as theformalism can be applied in a variety of situations, asthe perturbation flow Ψ(x) can be of very different na-ture. Considering Eq. 106 a special case of such a forc-ing is the internal coupling of degrees of freedom givenby (ΨX ,ΨY )> , to the unperturbed flow whose tendencyis given by (FX , FY )>. We can now use the formalismdescribed in Section V for computing at all orders thechange in the expectation value of A = A(X) due to thecoupling between the X and Y variables. After lengthycalculations, one obtain the explicit expression for

ρ(A)t = ρ0(A)t + ρ(1)(A)t + ρ(2)(A)t +O(Ψ3) (111)

As shown in (Wouters and Lucarini , 2012), if one col-lects these first and second order responses to the cou-pling Ψ, an identical change in expectation values fromthe unperturbed ρ0 up to third order in Ψ can be ob-tained by adding a Y -independent forcing to the ten-dency of the X variables as follows:

dX(t)

dt=FX(X(t)) +M + σ(t)

+

∫ ∞0

dτh(τ,X(t− τ)) (112)

where M = ρ0,Y (ΨX) is an averaged version of theY to X coupling, σ is a stochastic term, mimicking thetwo time correlation properties of the unresolved vari-ables and h is a memory kernel that introduces the non-Markovianity. A diagrammatic representation of pro-cesses responsible that these three additional terms areparametrizing is given in Figs. 18-20. The memory effect

Page 34: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

34

FIG. 20 Diagram describing the non-Markovian effect of theon the X variables on themselves, mediated by the Y vari-ables. Term h in Eq. 112.

is present due to the finite time scale difference betweenresolved and unresolved variables.

It should be noted that the choice of parametriza-tion is not unique. Any time-dependent forcing σ withthe correct two-point time-correlations will give the rightresponse up to second order. Also for the memory termthere is some freedom. If the coupling functions ΨX andΨY are allowed to be dependent on both X and Y , theabove analysis can still be carried out. For the case ofseparable couplings ΨX(X,Y ) = ΨX,1(X)ΨX,2(Y ) andΨY (X,Y ) = ΨY,1(X)ΨY,2(Y ) the average term becomesX dependent and the noise term becomes multiplicativeinstead of additive. An expression for more general cou-plings can be derived be decomposing the coupling func-tions into a basis of separable functions.

As obvious, Eqs. 110 and 112 are identical, which im-plies that, in terms of parametrization, it is equivalentto try to optimize the tendencies in an averaged senseor to try to minimize the bias in the statistical proper-ties of any observable. Equations 110-112 provide a verygeneral result concering direct methods for constructingparametrizations from the statistical properties of un-perturbed systems and the equivalence between the tworesults, obtained using rather distinct methods and con-cepts, suggest that there is a solid ground for exportingparametrizations developed for weather forecast into cli-mate models.

The results contained in Eqs. 110 and 112 imply that,up to second order, the dynamics of the X variables con-tains a stochastic term, so that one can expect that theprojection of the X variables of the invariant measure ofthe system given in Eq. 106 is, in fact to a good approx-imation, smooth, asa result of the regularizing effect dueto noise. Therefore, the FDT applies when we considerobservables A = A(X), i.e. functions of the resolved,slow variables only. This provides a solid reason why theempirical application of the FDT in a context of climatedynamics has proved partially successful, and of why thedegree of success depends critically on the climatic vari-able of interest.

VII. CONCLUSIONS

The main goal of this review paper is the provision ofan overview of some ideas emerging at the interface be-tween from theoretical physics, mathematics, and climatescience. The topics have been selected by the authorswith the goal of covering (at least partially) relevant as-pects of the deep symmetries of geophysical flows, of theprocesses by which they convert and transport energyand generate entropy, and of constructing relevant sta-tistical mechanical models able to address fundamentalissues like the response of the climate system to forcings,the representation of the interaction across scales, thedefinition of relevant physical quantities able to describesuccinctly but accurately the dynamics of the system.The themes covered in the review also inform the devel-opment and testing of climate models of various degreesof complexity, by analyzing their physical and mathemat-ical well-posedness and for constructing parametrizationsof unresolved processes, and by putting the basis for con-structing diagnostic tools able to capture the most rele-vant climate processes.

The Nambu formulation of geophysical fluid dynamicsexplored in Sec. II emphasizes the existence, in the in-viscid case, of non-trivial conserved quantities that areembedded in the equations of motion. Such quantities -which include potential vorticity in the three dimensionalcase - play a fundamental role, analogous to energy’s, inthe description of the state of the system and can be re-garded as observables of great relevance also in the casewhere dissipation and forcing are present. Moreover, theNambu formalism suggests us ways for devising very ac-curate numerical schemes, which do not have spuriousdiffusive behavior.

The symmetry properties of the flow in the inviscidlimit allow the construction of the ensembles describ-ing the equilibrium statistical mechanical properties ofthe geophysical flows (Sec. III), where the vorticity -in the two dimensional case - plays the role of the mostimportant physical quantity. Starting from the classi-cal construction due to Onsager of the gas of interact-ing vortices, the theory leads us to construct a theory ofbarotropic and baroclinic QG turbulence.

Taking the point of view of non-equilibrium system,the description of the climate system is enriched as thepresence of gradients of physical quantities like temper-ature and chemical concentrations - in first instance dueto the inhomogeneity of the incoming solar radiation, ofthe optical properties of the geophysical fluids, and of theboundary conditions- supports the possibility of work be-ing performed, resulting in organized fluid motions. InSec. IV the analysis of the energy and entropy budgets ofthe climate system is shown to provide a comprehensivepicture of climate dynamics, with promising outlook interms of investigation of the climate tipping points and asa method for studying the properties of general planetary

Page 35: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

35

atmospheres.

Section V introduces some basic concepts of non-equilibrium statistical mechanics, connecting the macro-scopic properties described in the previous section to thefeatures of the family of chaotic dynamical systems whichconstitute the backbone of the mathematical descriptionof non-equilibrium systems. For such systems, the re-lationship between internal fluctuations and response toforcings is studied with the goal of developing methodsfor predicting climate change. After clarifying the con-ditions under which the fluctuation-dissipation theoremis valid, we show for the first time examples of successfulclimate prediction for decadal and longer time scales. Inthis sense, we show that the problem of climate changeis mathematically well-posed.

Non-equilibrium statistical mechanics is also the sub-ject of Sec. VI, where we show how the formalism ofMori-Zwanzig projection operator supports the provisionof rigorous methods for constructing parametrizations ofunresolved processes. It is possible to derive a surrogatedynamics for the coarse grained variable of interest for cli-matic purposes, incorporating, as result of the couplingwith the small scale, fast variables, a deterministic, astochastic, and a non-Markovian contribution, which addto the unperturbed dynamics. The same results can beobtained using the response theory described in Sec. V,thus showing that the construction of parametrizationsfor weather and for climate models should have commonground.

Among the many topics and aspects left out of thisreview, we need to mention recent developments aimedat connecting the complementary, rather than oppos-ing (Lorenz , 1963) and (Hasselmann, 1976) perspectiveson complex dynamics dynamics, which focus on deter-ministic chaos and stochastic perturbations to dynam-ical systems, respectively. We refer in particular to theidea of constructing time-dependent measures for non au-tonomous dynamical systems (Chekroun et al., 2011b)through the introduction of the so-called pullback attrac-tor, which is the geometrical object the trajectories ini-tialized in a distant past tend to at time t with proba-bility 1 as a result of the contracting dynamics. Such anobject is not invariant with time, as a result of the time-dependent forcing, but, under suitable conditions on theproperties of the dynamical system, the supported mea-sure has at each instant properties similar to those of the(invariant) SRB measure one can construct for, e.g. au-tonomous Axiom A dynamical (Ruelle, 1989). Such anapproach allows for treating in a coherent way the pres-ence of modulations in the dynamics of the system, with-out the need of applying response formulas or of assumingtime-scale separations, and in particular allows for ana-lyzing the case where the forcing is stochastic, leading tothe concept of random attractor (Arnold , 1988). On adifferent line of research, it is instead possible to use Ru-elle response theory for computing the impact of adding

stochastic noise on chaotic dynamical systems (Lucarini ,2012). One finds the rate of convergence of the stochas-tically perturbed measure to the unperturbed one, anddiscovers the general result that adding noise enhancesthe power spectrum of any given observables at all fre-quencies. The difference between the power spectrum ofthe perturbed and unperturbed system can be used, mir-roring a fluctuation-dissipation result, for computing theresponse of the system to deterministic perturbations.

The methods, the ideas, the perspectives presented inthis paper are partially overlapping, partially comple-mentary, partly in contrast. In particular, it is not obvi-ous, as of today, whether it is more efficient to approachthe problem of constructing a theory of climate dynam-ics starting from the framework of hamiltonian mechan-ics and quasi-equilibrium statistical mechanics or takingthe point of view of dissipative chaotic dynamical sys-tems, and of non-equilibrium statistical mechanics, andeven the authors of this review disagree. The former ap-proach can rely on much more powerful mathematicaltools, while the latter is more realistic and epistemolog-ically more correct, because, obviously, the climate is,indeed, a non-equilibrium system. Nonetheless, the ex-perience accumulated in many other scientific branches(chemistry, acoustics, material science, optics, etc.) hasshown that by suitably applying perturbation theory toequilibrium systems one can provide an extremely ac-curate description of non-equilibrium properties. Such alack of unified perspective, of well-established paradigms,should be seen as sign of the vitality of many researchperspectives in climate dynamics.

ACKNOWLEDGMENTS

The authors wish to acknowledge fruitful interactionswith U. Achatz, A. Seifert, D. Ruelle, F. Cooper, P.Nevir, T. Kuna, J. Gregory, R. Boschi, M. Ambaum,M. Ghil, R. Tailleux, K. Heng, O. Pauluis, A. Muller,F. Lunkeit, F. Ragone, F. Sielmann, M. Sommer. VL,RB, SP, and JW wish to acknowledge the financial sup-port provided by the ERC-Starting Investigator GrantNAMASTE (Grant no. 257106). VL and JW acknowl-edge the kind hospitality of the Isaac Newton Institutefor Mathematical Sciences (Cambridge, UK) and of theMathematics for the Fluid Earth 2013 programme. Allauthors wish to acknowledge support by the Clusterof Excellence CLISAP. The National Center for Atmo-spheric Research is sponsored by the National ScienceFoundation.

REFERENCES

Abramov, R. V. (2012), Suppression of chaos at slow variablesby rapidly mixing fast dynamics through linear energy-

Page 36: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

36

preserving coupling, Communications in Mathematical Sci-ences, 10 (2), 595–624.

Abramov, R. V., and A. Majda (2008), New approximationsand tests of linear fluctuation-response for chaotic nonlin-ear forced-dissipative dynamical systems, Journal of Non-linear Science, 18, 303–341, 10.1007/s00332-007-9011-9.

Adams, D. K., and N. O. Renno (2005), Thermodynamic ef-ficiencies of an idealized global climate model, Clim. Dyn.,25, 801–813.

Ambaum, M. H. P. (2010), Thermal Physics of the Atmo-sphere, 256 pp., Wiley, New York.

Arnold, L. (1988), Random Dynamical Systems, Springer,New York.

Arnold, L. (2001), Hasselmann’s program revisited: Theanalysis of stochasticity in deterministic climate models,Stochastic climate models, 49, 141–158.

Arnold, V. (1992), Catastrophe theory, 3d edition, Springer,Berlin.

Bartello, P. (1995), Geostrophic adjustment and inverse cas-cades in rotating stratified turbulence., J. Atmos. Sci., 52,4410–4428.

Basdevant, C., and R. Sadourny (1975), Ergodic proper-ties of inviscid truncated models of two-dimensional in-compressible flows, J. Fluid Mech., 69, 673–688, doi:10.1017/S0022112075001620.

Batygin, K., and D. J. Stevenson (2010), Inflating hot jupiterswith ohmic dissipation, The Astrophysical Journal Letters,714 (2), L238–L24.

Becker, E. (2003), Frictional heating in global climate models,Mon. Weather Rev., 131, 508–520.

Bengtsson, L., M. Steinheimer, P. Bechtold, and J.-F. Ge-leyn (2013), A stochastic parametrization for deep con-vection using cellular automata, Quarterly Journal of theRoyal Meteorological Society, 139 (675), 15331543, doi:10.1002/qj.2108.

Bernstein, D. (2007), Optimal prediction of burgerss equa-tion, Multiscale Modeling & Simulation, 6 (1), 27–52, doi:10.1137/060651720.

Bihlo, A. (2008), Rayleigh-Benard convection as a Nambu-metriplectic problem, J. Phys. A, 41, 292,001.

Blender, R., and V. Lucarini (2013), Nambu representationof an extended Lorenz model with viscous heating, PhysicaD, 243, 86–91, doi:10.1016/j.physd.2012.09.007.

Boffetta, G. (2007), Energy and enstrophy fluxes in the doublecascade of two-dimensional turbulence, J. Fluid Mech., 589,253–260.

Boschi, R., V. Lucarini, and S. Pascale (2013),Bistability of the climate around the habitablezone: a thermodynamic investigation, Icarus, doi:http://dx.doi.org/10.1016/j.icarus.2013.03.017.

Bouchet, F., and M. Corvellec (2010), Invariant measures ofthe 2D Euler and Vlasov equations, J. Stat. Mech., 2010,P08,021, doi:10.1088/1742-5468/2010/08/P08021.

Bouchet, F., and E. Simonnet (2009), Random changesof flow topology in two-dimensional and geophysi-cal turbulence, Phys. Rev. Lett., 102, 94,504, doi:10.1103/PhysRevLett.102.094504.

Bouchet, F., and J. Sommeria (2002), Emergence of in-tense jets and Jupiter’s Great Red Spot as maximum-entropy structures, J. Fluid Mech., 464, 165–207, doi:10.1017/S0022112002008789.

Bouchet, F., and A. Venaille (2012), Statistical mechanics oftwo-dimensional and geophysical flows, Phys. Rep., 515,227, doi:10.1016/j.physrep.2012.02.001.

Bouchet, F., C. Nardini, and T. Tangarife (2013), Kinetictheory of jet dynamics in the stochastic barotropic and 2Dnavier-stokes equations, arXiv e-print 1305.0877.

Budyko, M. (1969), The effect of solar radiation variations onthe climate of the earth, Tellus, 21, 611–619.

Charney, J. G. (1947), The dynamics of longwaves in a baroclinic westerly current, Journalof Meteorology, 4 (5), 136–162, doi:10.1175/1520-0469(1947)004¡0136:TDOLWI¿2.0.CO;2.

Chavanis, P.-H. (2009), Dynamical and thermodynamical sta-bility of two-dimensional flows: variational principles andrelaxation equations, Eur. Phys. J. B, 70, 73–105, doi:10.1140/epjb/e2009-00196-1.

Chavanis, P.-H., and J. Sommeria (1996), Classification ofself-organized vortices in two-dimensional turbulence: thecase of a bounded domain, J. Fluid Mech., 314, 267–297,doi:10.1017/S0022112096000316.

Chavanis, P.-H., and J. Sommeria (1997), Thermody-namical approach for small-scale parametrization in2D turbulence, Phys. Rev. Lett., 78, 3302–3305, doi:10.1103/PhysRevLett.78.3302.

Chekroun, M. D., D. Kondrashov, and M. Ghil (2011a), Pre-dicting stochastic systems by noise sampling, and appli-cation to the el nio-southern oscillation, Proceedings of theNational Academy of Sciences, 108 (29), 11,766–11,771, doi:10.1073/pnas.1015753108, PMID: 21730171.

Chekroun, M. D., E. Simonnet, and M. Ghil (2011b),Stochastic climate dynamics: Random attractorsand time-dependent invariant measures, Physica D:Nonlinear Phenomena, 240 (21), 1685 – 1700, doi:http://dx.doi.org/10.1016/j.physd.2011.06.005.

Chelton, D. B., M. G. Schlax, R. M. Samelson, andR. A. de Szoeke (2007), Global observations of largeoceanic eddies, Geophys. Res. Lett., 34 (15), L15,606, doi:10.1029/2007GL030812.

Chorin, A., and P. Stinis (2006), Problem reduction, renor-malization, and memory, Communications in AppliedMathematics and Computational Science, 1 (1), 1–27, doi:10.2140/camcos.2006.1.1.

Chorin, A. J., and O. H. Hald (2013), Stochastic Tools inMathematics and Science, no. 58 in Texts in Applied Math-ematics, 3rd ed., Springer, New York.

Chorin, A. J., A. P. Kast, and R. Kupferman (1998), Optimalprediction of underresolved dynamics, Proceedings of theNational Academy of Sciences, 95 (8), 4094–4098, PMID:9539695.

Chorin, A. J., O. H. Hald, and R. Kupferman (2000), Opti-mal prediction and the MoriZwanzig representation of ir-reversible processes, Proceedings of the National Academyof Sciences, 97 (7), 2968–2973, doi:10.1073/pnas.97.7.2968,PMID: 10737778.

Chorin, A. J., O. H. Hald, and R. Kupferman (2002), Optimalprediction with memory, Physica D: Nonlinear Phenomena,166 (34), 239–257, doi:10.1016/S0167-2789(02)00446-3.

Chorin, A. J., O. H. Hald, and R. Kupferman (2006), Pre-diction from partial data, renormalization, and averaging,Journal of Scientific Computing, 28 (2-3), 245–261, doi:10.1007/s10915-006-9089-5.

Cichowlas, C., P. Bonaıti, F. Debbasch, and M. E. Brachet(2005), Effective Dissipation and Turbulence in SpectrallyTruncated Euler Flows, Phys. Rev. Lett., 95 (26), 264,502,doi:10.1103/PhysRevLett.95.264502.

Cooper, F. C., and P. H. Haynes (2011), Climate sensitivityvia a nonparametric fluctuation-dissipation theorem, Jour-

Page 37: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

37

nal of the Atmospheric Sciences, 68 (5), 937–953.Cooper, F. C., J. G. Esler, and P. H. Haynes (2013), Es-

timation of the local response to a forcing in a high di-mensional system using the fluctuation-dissipation theo-rem, Nonlinear Processes in Geophysics, 20 (2), 239–248,doi:10.5194/npg-20-239-2013.

Cullen, M. J. P. (2006), A Mathematical Theory of Large-ScaleAtmosphere/Ocean Flow, World Scientific Singapore.

Cummins, P. F., and G. Holloway (1994), On eddy-topographic stress representation, J. Phys. Oceanogr.,24 (3), 700–706.

Cvitanovic, P. (1988), Invariant measurement of strange setsin terms of cycles, Phys. Rev. Lett., 61, 2729–2732, doi:10.1103/PhysRevLett.61.2729.

Dauxois, T., S. Ruffo, E. Arimondo, and M. Wilkens (Eds.)(2002), Dynamics and Thermodynamics of Systems withLong Range Interactions, Lecture Notes in Physics, vol.602, Springer, New York.

Davies, L., R. S. Plant, and S. H. Derbyshire (2009), A sim-ple model of convection with memory, Journal of Geo-physical Research: Atmospheres, 114 (D17), n/an/a, doi:10.1029/2008JD011653.

Defrasne, S. (2004), Gute von Projektionsverfahren in der Dy-namik relevanter kollektiver Koordinaten in ausgewahltenphysikalischen Modellen.

deGroot, S., and P. Mazur (1984), Non-Equilibrium Thermo-dynamics, 510 pp., Dover.

Dewar, R., C. Lineweaver, R. Niven, and K. Regenauer-Lieb(2013), Beyond the second law: an overview, in Beyondthe Second Law: Entropy Production and Non-equilibriumSystems, edited by D. RC, L. CH, N. RK, and R.-L. K,Springer.

DiBattista, M., and A. J. Majda (2001), Equilibrium statisti-cal predictions for baroclinic vortices: The role of angularmomentum, Theor. Comput. Fluid Dyn., 14 (5), 293–322.

Dijkstra, H. (2013), Nonlinear Climate Dynamics, CambridgeUniversity Press, Cambridge.

Dolaptchiev, S., U. Achatz, and I. Timofeyev (2012),Stochastic closure for local averages in the finite-differencediscretization of the forced burgers equation, Theoreti-cal and Computational Fluid Dynamics, pp. 1–21, doi:10.1007/s00162-012-0270-1.

Donges, J. F., Y. Zou, N. Marwan, and J. Kurths (2009),Complex networks in climate dynamics, The EuropeanPhysical Journal Special Topics, 174 (1), 157–179, doi:10.1140/epjst/e2009-01098-2.

Donohoe, A., and D. Battisti (2012), What determines merid-ional heat transport in climate models?, J. Climate, 25,3832–3859.

Dutton, J. A. (1973), The global thermodynamics of atmo-spheric motions, Tellus, 25, 89–110.

Dvorak, R. (2008), Extrasolar Planets, Wiley.Eady, E. T. (1949), Long waves and cyclone waves, Tellus,

1 (3), 33–52, doi:10.1111/j.2153-3490.1949.tb01265.x.Eckhardt, B., and G. Ott (1994), Periodic orbit analysis of the

lorenz attractor, Zeitschrift fr Physik B Condensed Matter,93 (2), 259–266, doi:10.1007/BF01316970.

Eckmann, J. P., and D. Ruelle (1985), Ergodic theory of chaosand strange attractors, Rev. Mod. Phys., 57, 617–656, doi:10.1103/RevModPhys.57.617.

Emanuel, K. (1991), The theory of hurricanes, Ann. Rev.Fluid. Mech., 23, 179–196.

Emanuel, K., and M. Bister (1996), Moist convective velocityand buoyancy scales, J. Atmos. Sci., 53, 3276–3285.

Enderton, D., and J. Marshall (2009), Controls on the totaldynamical heat transport of the atmosphere and oceans, J.Atmos. Sci., 66, 1593–1611.

Errico, R. M. (1984), The statistical equilibrium solution of aprimitive-equation model, Tellus A, 36 (1), 42–51.

Evans, D., and G. Morriss (2008), Statistical Mechanics ofNonequilibrium Liquids, Cambridge University Press, Cam-bridge.

Eyink, G. L., T. W. N. Haine, and D. J. Lea (2004), Ruelle’slinear response formula, ensemble adjoint schemes and lvyflights, Nonlinearity, 17 (5), 1867.

Fasullo, J., and K. Trenberth (2008), The annual cycle of theenergy budget. part ii: Meridional structures and polewardtransports, J. Climate, 21, 23132325.

Fox, D. G., and S. A. Orszag (1973), Inviscid dynamics oftwo-dimensional turbulence, Phys. Fluids, 16, 169–171.

Fraedrich, K., and F. Lunkeit (2008), Diagnosing the entropybudget of a climate model, Tellus A, 60 (5), 299–304.

Fraedrich, K., H. Jansen, E. Kirk, U. Luksch, and F. Lunkeit(2005), The planet simulator: Towards a user friendlymodel, Meteorologische Zeitschrift, 14 (3), 299–304, doi:doi:10.1127/0941-2948/2005/0043.

Frederiksen, J. S., and T. J. O’Kane (2008), Entropy, Closuresand Subgrid Modeling, Entropy, 10, 635–683.

Frederiksen, J. S., and B. L. Sawford (1980), Statis-tical dynamics of two-dimensional inviscid flow on asphere., J. Atmos. Sci., 37, 717–732, doi:10.1175/1520-0469(1980)037¡0717:SDOTDI¿2.0.CO;2.

Gallavotti, G. (1996), Chaotic hypothesis: Onsager reci-procity and fluctuation-dissipation theorem, Journal ofStatistical Physics, 84 (5-6), 899–925.

Gallavotti, G. (2006), Stationary nonequilibrium statisticalmechanics, Encyclopedia of Mathematical Physics, ed. JPFrancoise, GL Naber, TS Tsun, 3, 530–539.

Gassmann, A. (2013), A global hexagonal C-grid non-hydrostatic dynamical core (ICON-IAP) designed for en-ergetic consistency , Quart. J. Roy. Meteorol. Soc., 139,152–175, doi:10.1002/qj.1960.

Gassmann, A. (2013), A global hexagonal C-grid nonhy-drostatic dynamical core (ICON-IAP) designed for ener-getic consistency, Quart. J. Roy. Meteorol. Soc., 139, doi:doi:10.1002/qj.1960.

Gassmann, A., and H.-J. Herzog (2008), Towards a consistentnumerical compressible non-hydrostatic model using gen-eralized Hamiltonian tools, Quart. J. Roy. Meteorol. Soc.,134 (635, B), 1597–1613, doi:10.1002/qj.297.

Ghil, M. (1976), Climate stability for a Sellers-type model,Journal of Atmospheric Sciences, 33, 3.

Ghil, M., P. Yiou, S. Hallegatte, B. D. Malamud, P. Naveau,A. Soloviev, P. Friederichs, V. Keilis-Borok, D. Kon-drashov, V. Kossobokov, O. Mestre, C. Nicolis, H. W.Rust, P. Shebalin, M. Vrac, A. Witt, and I. Zaliapin(2011), Extreme events: dynamics, statistics and predic-tion, Nonlinear Processes in Geophysics, 18 (3), 295–350,doi:10.5194/npg-18-295-2011.

Gianfelice, M., F. Maimone, V. Pelino, and S. Vaient (2012),On the recurrence and robust properties of lorenz?63model, Comm Math. Phys., 313, 745–779.

Goodman, J. (2009), Thermodynamics of atmospheric cir-culations on hot jupiters, The Astrophysical Journal, 693,1645, doi:10.1088/0004-637X/693/2/1645.

Goody, R. (2000), Sources and sinks of climate entropy, Q. J.R. Meteorol. Soc., 126, 1953–1970.

Gottwald, G. (2010), On recent trends in climate dynamics,

Page 38: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

38

AMS Gazette, 37 (5).Griffa, A., and R. Salmon (1989), Wind-driven ocean circu-

lation and equilibrium statistical mechanics, J. Mar. Res.,47, 457–492.

Gritsun, A., and G. Branstator (2007), Climate response us-ing a three-dimensional operator based on the Fluctua-tionDissipation theorem, Journal of the Atmospheric Sci-ences, 64 (7), 2558–2575.

Gritsun, A. S. (2008), Unstable periodic trajectories of abarotropic model of the atmosphere, Russ. J. Numer. Anal.Math. Modelling, 23, 345–367.

Hald, O. H., and R. Kupferman (2001), Convergence of opti-mal prediction for nonlinear hamiltonian systems, SIAMJournal on Numerical Analysis, 39 (3), 983–1000, doi:10.1137/S0036142900374482.

Hasselmann, K. (1976), Stochastic climate models, part I.theory, Tellus, 28 (6), 473–485.

Hasson, S., V. Lucarini, and S. Pascale (2013), Hydrologicalcycle over south and southeast asian river basins as simu-lated by pcmdi/cmip3 experiments, Earth System Dynam-ics, 4 (2), 199–217, doi:10.5194/esd-4-199-2013.

Heng, K. (2012a), The study of climates of alien worlds,American Scientist, 100 (4), 334–341.

Heng, K. (2012b), On the existence of shocks in irradiatedexoplanetary atmospheres, The Astrophysical Journal Let-ters, 761 (L1), 6.

Herbert, C. (2013a), Restricted Partition Functions and DualEnergy Cascades in parity symmetry breaking flows, Phys.Rev. E, submitted.

Herbert, C. (2013b), Additional invariants and statisticalequilibria for the 2D Euler equations on a spherical do-main, J. Stat. Phys., 152, 1084–1114.

Herbert, C. (2013c), Nonlinear energy transfers and phasediagrams for geostrophically balanced rotating-stratifiedflows, Phys. Rev. E, submitted.

Herbert, C., B. Dubrulle, P.-H. Chavanis, and D. Paillard(2012a), Phase transitions and marginal ensemble equiva-lence for freely evolving flows on a rotating sphere, Phys.Rev. E, 85, 056,304, doi:10.1103/PhysRevE.85.056304.

Herbert, C., B. Dubrulle, P.-H. Chavanis, and D. Paillard(2012b), Statistical mechanics of quasi-geostrophic flowson a rotating sphere, J. Stat. Mech., 2012, P05,023, doi:10.1088/1742-5468/2012/05/P05023.

Herbert, C., R. Marino, and A. Pouquet (2013), RestrictedEquilibrium, Waves, and Anisotropic Energy Cascades inRotating and Stratified Flows, J. Fluid Mech., in prepara-tion.

Hernandez-Deckers, D., and J.-S. von Storch (2012), Impactof the warming pattern on global energetics, J. Climate,25, 5223–5240.

Herring, J. R. (1977), On the statistical theory of two-dimensional topographic turbulence., J. Atmos. Sci., 34,1731–1750.

Hoffman, P., and D. Schrag (2002), The snowball earth hy-pothesis: testing the limits of global change, Terra Nova,14, 129.

Hogg, N. G., and H. M. Stommel (1985), Hetonic explosions:the breakup and spread of warm pools as explained by baro-clinic point vortices, J. Atmos. Sci., 42, 1465–1476.

Holloway, G. (1986), Eddies, waves, circulation, and mixing:Statistical geofluid mechanics, Ann. Rev. Fluid Mech., 18,91–147.

Holloway, G. (1992), Representing Topographic Stress forLarge-Scale Ocean Models, J. Phys. Oceanogr., 22, 1033–

1046.Holloway, G. (2004), From classical to statistical ocean dy-

namics, Surveys in Geophysics, 25, 203–219.Holton, J. (2004), An introduction to Dynamic Meteorology,

531 pp., Accademic Press.IPCC (2001), IPCC Third Assessment Report: Working

Group I Report ”The Physical Science Basis”, CambridgeUniversity Press.

IPCC (2007), IPCC Fourth Assessment Report: WorkingGroup I Report ”The Physical Science Basis”, CambridgeUniversity Press.

IPCC (2013), IPCC Fifth Assessment Report: WorkingGroup I Report ”The Physical Science Basis”, CambridgeUniversity Press.

Johnson, D. (1989), The forcing and maintenance of globalmonsoonal circulations: an isentropic analysis, in Advancesin Geophysics, vol. 31, edited by B. Saltzman, pp. 43–316.

Johnson, D. (1997), General coldness of climate models andthe second law: Implications for modeling the earth system,J. Climate, 10, 2826–2846.

Johnson, D. R. (2000), Entropy, the Lorenz energy cycle, andclimate, in General Circulation Model Development: Past,Present and Future, edited by D. A. Randall, pp. 659–720,Academic Press, New York.

Kalnay, E. (2003), Atmospheric Modeling, Data Assimila-tion and Predictability, Cambridge University Press, Cam-bridge.

Kazantsev, E., J. Sommeria, and J. Verron (1998), Subgrid-scale eddy parameterization by statistical mechanics in abarotropic ocean model, J. Phys. Oceanogr., 28, 1017–1042,doi:10.1175/1520-0485(1998)028¡1017:SSEPBS¿2.0.CO;2.

Kifer, Y. (2004), Some recent advances in averaging, Moderndynamical systems and applications: dedicated to AnatoleKatok on his 60th birthday, p. 385.

Kifer, Y. (2008), Convergence, nonconvergence and adiabatictransitions in fully coupled averaging, Nonlinearity, 21 (3),T27, doi:10.1088/0951-7715/21/3/T01.

Kifer, Y. (2009), Large deviations and adiabatic transitionsfor dynamical systems and markov processes in fully cou-pled averaging, Memoirs of the American Mathematical So-ciety, 201 (944), 0–0, doi:10.1090/memo/0944.

Kleidon, A. (2009), Non-equilibrium thermodynamics andmaximum entropy production, Naturwissenschaften, 96,653–677.

Kleidon, A., and R. Lorenz (2005), Non-equilibrium thermo-dynamics and the production of entropy, Springer, Berlin.

Klein, R. (2010), Scale-dependent models for atmosphericflows, Annual Review of Fluid Mechanics, 42 (1), 249–274,doi:10.1146/annurev-fluid-121108-145537.

Kraichnan, R. (1967), Inertial ranges in two-dimensional tur-bulence, Phys. Fluids, 10, 1417, doi:10.1063/1.1762301.

Kraichnan, R., and D. Montgomery (1980), Two-dimensionalturbulence, Rep. Prog. Phys., 43, 547, doi:10.1088/0034-4885/43/5/001.

Kubo, R. (1957), Statistical-mechanical theory of irreversibleprocesses. i. general theory and simple applications to mag-netic and conduction problems, Journal of the Physical So-ciety of Japan, 12 (6), 570–586, doi:10.1143/JPSJ.12.570.

Kubo, R., M. Toda, and N. Hashitsume (1988), Statisticalphysics II: nonequilibrium Statistical Mechanics, Springer,Heidelberg.

Kuroda, Y. (1991), On the Casimir invariant of Hamiltonianfluid mechanics, J. Phys. Soc. Japan, 60, 727–730.

Kwasniok, F. (2012), Data-based stochastic subgrid-scale

Page 39: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

39

parametrization: an approach using cluster-weighted mod-elling, Philosophical Transactions of the Royal Soci-ety A: Mathematical, Physical and Engineering Sciences,370 (1962), 1061–1086.

Lacorata, G., and A. Vulpiani (2007), Fluctuation-responserelation and modeling in systems with fast and slow dy-namics, Nonlin. Processes Geophys., 14, 681–694.

Landau, L. D., and E. M. Lifshits (1996), Mechanics,Butterworth-Heinemann, Oxford.

Langen, P. L., and V. A. Alexeev (2005), Estimating 2 xCO2 warming in an aquaplanet GCM using the fluctuation-dissipation theorem, Geophysical Research Letters, 32 (23),L23,708.

Lee, T. D. (1952), On some statistical properties of hydro-dynamical and magneto-hydrodynamical fields, Q. Appl.Math., 10, 69–74.

Legg, S., and J. Marshall (1993), A heton model of thespreading phase of open-ocean deep convection, J. Phys.Oceanogr., 23, 1040–1056.

Li, L., A. Ingersoll, X. Jiang, D. Feldmann, and Y. Yung(2008), Lorenz energy cycle of the global atmosphere basedon reanalysis datasets, Geophys. Res. Lett., 34, L16,813.

Liepert, B., and F. Lo (2013), CMIP5 update of inter-modelvariability and biases of the global water cycle in CMIP3coupled climate models, Environ. Res. Lett., 8, 029,401.

Lorenz, E. (1955), Available potential energy and the main-tenance of the general circulation, Tellus, 7, 157–167.

Lorenz, E. (1967), The nature and theory of the general cir-culation of the atmosphere, vol. 218.TP.115, World Mete-orological Organization.

Lorenz, E. (1979), Forced and free variations of weather andclimate, J. Atmos. Sci, 36, 1367–1376.

Lorenz, E. N. (1963), Deterministic nonperiodic flow, J. At-mos. Sci., 20, 130–141.

Lorenz, E. N. (1996), Predictability: A problem partly solved,in GARP Publication Series, vol. 16, pp. 132–136, WMO,Geneva, Switzerland.

Lovejoy, S., and D. Schertzer (2013), The Weather and Cli-mate: Emergent Laws and Multifractal Cascades, Cam-bridge University Press, Cambridge.

Lucarini, V. (2008), Response theory for equilibrium andnon-equilibrium statistical mechanics: Causality and gen-eralized kramers-kronig relations, Journal of StatisticalPhysics, 131, 543–558, 10.1007/s10955-008-9498-y.

Lucarini, V. (2009), Evidence of dispersion relations for thenonlinear response of the Lorenz 63 system, Journal of Sta-tistical Physics, 134, 381–400, 10.1007/s10955-008-9675-z.

Lucarini, V. (2011), Modelling complexity: the case of climatescience, arxiv:1106.1265.

Lucarini, V. (2012), Stochastic perturbations to dynamicalsystems: A response theory approach, Journal of StatisticalPhysics, 146 (4), 774–786, doi:10.1007/s10955-012-0422-0.

Lucarini, V., and M. Colangeli (2012), Beyond the lin-ear fluctuation-dissipation theorem: the role of causality,Journal of Statistical Mechanics: Theory and Experiment,2012 (05), P05,013.

Lucarini, V., and F. Ragone (2011), Energetics of cli-mate models: Net energy balance and meridional en-thalpy transport, Rev. Geophys., 49, RG1001, doi:10.1029/2009RG000323.

Lucarini, V., and S. Sarno (2011), A statistical mechanicalapproach for the computation of the climatic response togeneral forcings, Nonlin. Proc. Geophys., 18, 728.

Lucarini, V., J. J. Saarinen, K.-E. Peiponen, and E. M. Var-

tiainen (2005), Kramers-Kronig relations in Optical Mate-rials Research, Springer, New York.

Lucarini, V., R. Danihlik, I. Kriegerova, and A. Speranza(2008), Hydrological cycle in the danube basin in present-day and xxii century simulations by ipccar4 global climatemodels, Journal of Geophysical Research: Atmospheres,113 (D9), n/a–n/a, doi:10.1029/2007JD009167.

Lucarini, V., K. Fraedrich, and F. Lunkeit (2010), Thermody-namics of climate change: Generalized sensitivities, Atmos.Chem. Phys, 10, 9729–9737.

Lucarini, V., K. Fraedrich, and F. Ragone (2011), New resultson the thermodynamic properties of the climate system, J.Atmos. Sci., 68, 2438–2458.

Lucarini, V., S. Pascale, R. Boschi, E. Kirk, and N. Iro (2013),Habitability and multistablility in earth-like plantets, Astr.Nach., 334 (6), 576–588.

Majda, A., I. Timofeyev, and E. Vanden Eijnden (2001),A mathematical framework for stochastic climate models,Communications on Pure and Applied Mathematics, 54 (8),891–974.

Marconi, U. M. B., A. Puglisi, L. Rondoni, and A. Vulpiani(2008), Fluctuation-dissipation: Response theory in statis-tical physics, Phys. Rep., 461, 111.

Margules, M. (1905), On the energy of storms, transl. fromgerman by c. abbe, in Smithson. Misc. Collect., edited byD. A. Randall, pp. 533–595.

Marino, R., P. D. Mininni, D. Rosenberg, and A. Pouquet(2013), Inverse cascades in rotating stratified turbulence:fast growth of large scales, Europhys. Lett., 102, 44,006,doi:10.1209/0295-5075/102/44006.

Marques, C., A. Rocha, and J. Corte-Real (2011), Global di-agnostic energetics of five state-of-the-art climate models,Clim. Dyn., 36, 1767–1794.

Mayer, M., and L. Haimberger (2012), Poleward atmosphericenergy transports and their variability as evaluated fromECMWF reanalysis data, J. Climate, 25, 734–752.

Merryfield, W. J. (1998), Effects of stratification on quasi-geostrophic inviscid equilibria, J. Fluid Mech., 354, 345–356.

Merryfield, W. J., and G. Holloway (1996), Inviscid quasi-geostrophic flow over topography: testing statisticalmechanical theory, J. Fluid Mech., 309, 85–91, doi:10.1017/S0022112096001565.

Miller, J. (1990), Statistical mechanics of Euler equationsin two dimensions, Phys. Rev. Lett., 65, 2137–2140, doi:10.1103/PhysRevLett.65.2137.

Mininni, P. D., and A. Pouquet (2010), Rotating helical tur-bulence. I. Global evolution and spectral behavior, Phys.Fluids, 22, 035,105.

Miyazaki, T., T. Sato, H. Kimura, and N. Takahashi(2011), Influence of external flow field on the equi-librium state of quasi-geostrophic point vortices, Geo-phys. Astrophys. Fluid Dyn., 105 (4-5), 392–408, doi:10.1080/03091929.2010.502118.

Monahan, A. H., and J. Culina (2011), Stochastic av-eraging of idealized climate models, Journal of Cli-mate, 24 (12), 3068–3088, doi:10.1175/2011JCLI3641.1,WOS:000291585800010.

Montgomery, D., and G. Joyce (1974), Statistical mechanicsof “negative temperature” states, Phys. Fluids, 17, 1139,doi:10.1063/1.1694856.

Mori, H. (1965), Transport, collective motion, and Brownianmotion, Progress of Theoretical Physics, 33 (3), 423–455.

Mori, H., H. Fujisaka, and H. Shigematsu (1974), A new

Page 40: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

40

expansion of the master equation, Progress of TheoreticalPhysics, 51 (1), 109–122, doi:10.1143/PTP.51.109.

Nambu, Y. (1973), Generalized Hamiltonian dynamics, Phys.Rev. D, 7, 2403–2412.

Naso, A., P.-H. Chavanis, and B. Dubrulle (2011), Statisticalmechanics of Fofonoff flows in an oceanic basin, Eur. Phys.J. B, 80, 493–517, doi:10.1140/epjb/e2011-10440-8.

Nevir, P. (1998), Die Nambu-Felddarstellungen der Hydro-Thermodynamik und ihre Bedeutung fur die dynamischeMeteorologie, Habilitationsschrift.

Nevir, P., and R. Blender (1993), A Nambu representationof incompressible hydrodynamics using helicity and enstro-phy, J. Phys. A, 26, L1189–1193.

Nevir, P., and R. Blender (1994), Hamiltonian and Namburepresentation of the non-dissipative Lorenz equations,Beitr. Phys. Atmosph., 67, 133–144.

Nevir, P., and M. Sommer (2009), Energy-Vorticity Theory ofIdeal Fluid Mechanics, J. Atmos. Sci., 66 (7), 2073–2084,doi:10.1175/2008JAS2897.1.

Onsager, L. (1949), Statistical hydrodynamics, Il Nuovo Ci-mento, 6, 279–287, doi:10.1007/BF02780991.

Oort, A., L. Anderson, and J. Peixoto (1994), Estimates ofthe energy cycle of the oceans, J. Geophys. Res., 99, 7665–7688.

Palmer, T. N., and P. Williams (2009), Stochastic Physicsand Climate Modelling, Cambridge University Press, Cam-bridge.

Paltridge, G. (1975), Global dynamics and climate – a systemof minimum entropy exchange, Quarterly Journal of theRoyal Meteorological Society, 101, 475484.

Paltridge, G. (1978), The steady state format of global cli-mate, Quarterly Journal of Royal Meteorological Society,104, 927–945.

Paret, J., and P. Tabeling (1997), Experimental observationof the two-dimensional inverse energy cascade, Phys. Rev.Lett., 79, 4162–4165.

Park, J. H., N. S. Namachchivaya, and N. Neogi (2007),Stochastic averaging and optimal prediction, Journalof Vibration and Acoustics, 129 (6), 803–807, doi:10.1115/1.2748777.

Pascale, S., J. M. Gregory, M. Ambuam, and R. Tailleux(2011), Climate entropy budget of the hadcm3 atmosphere-ocean general circulation model and of FAMOUS, its low-resolution version, Clim. Dyn., 36, 1189–1206.

Pascale, S., J. M. Gregory, M. Ambuam, R. Tailleux, andV. Lucarini (2012), Vertical and horizontal processes in theglobal atmosphere and the maximum entropy productionconjecture, Earth System Dynamics, 3, 19–32.

Pascale, S., F. Ragone, V. Lucarini, Y. Wang, and R. Boschi(2013), Nonequilibrium thermodynamics of an opticallythin, dry atmosphere, Planetary and Space Science, 84, 48–65.

Pauluis, O., and I. Dias (2012), Satellite estimated ofprecipitation-induced dissipation in the atmosphere, Sci-ence, 335, 953–956.

Pauluis, O., and I. M. Held (2002a), Entropy budget of an at-mosphere in radiative-convective equilibrium. Part I: Max-imum work and frictional dissipation, J. Atmos. Sci., 59,125–139.

Pauluis, O., and I. M. Held (2002b), Entropy budget of an at-mosphere in radiative-convective equilibrium. Part II: La-tent heat transport and moist processes, J. Atmos. Sci.,59, 140–149.

Pavliotis, G., and A. Stuart (2008), Multiscale methods, Texts

in applied mathematics : TAM, Springer, New York, NY.Peixoto, J. P., and A. H. Oort (1992), Physics of climate,

American Institute of Physics, New York.Pelino, V., and F. Maimone (2007), Energetics, skele-

tal dynamics, and long-term predictions on kolmogorov-lorenz systems, Phys. Rev. E, 76, 046,214, doi:10.1103/PhysRevE.76.046214.

Penrose, O., and J. L. Lebowitz (1979), Towards a rigorousmolecular theory of metastability, in Fluctuation Phenom-ena, edited by E. W. Montroll and J. L. Lebowitz, chap. 5,p. 293, North-Holland, Amsterdam.

Perna, R., K. Heng, and F. Pont (2012), The effects of irra-diation on hot jovian atmospheres: heat redistribution andenergy dissipation, Astr. Journal, 751, 59–76.

Piriou, J.-M., J.-L. Redelsperger, J.-F. Geleyn, J.-P.Lafore, and F. Guichard (2007), An approach for con-vective parameterization with memory: Separating mi-crophysics and transport in grid-scale equations, Jour-nal of the Atmospheric Sciences, 64 (11), 4127–4139, doi:10.1175/2007JAS2144.1.

Prigogine, I. (1961), Thermodynamics of Irreversible Pro-cesses, Interscience, New York.

Rant, Z. (1956), Energie, ein neues wort fur technische ar-

beitfhigkeit, Forsch. Ing., 22, 36–37.Reick, C. H. (2002), Linear response of the Lorenz system,

Phys. Rev. E, 66, 036,103.Renno, N. O., and A. P. Ingersoll (1996), Natural convection

as a heat engine: a theory for CAPE, J. Atmos. Sci., 53,572–585.

Rhines, P. B. (1975), Waves and turbulence on a beta-plane,J. Fluid Mech., 69, 417–443.

Rhines, P. B. (1976), The dynamics of unsteady currents, inThe Sea, vol. VI, John Wiley & Sons, New-York, New York.

Rhines, P. B. (1979), Geostrophic turbulence, Ann. Rev. FluidMech., 11, 401–441.

Robert, R., and C. Rosier (2001), Long range predictabilityof atmospheric flows, Nonlin. Processes Geophys., 8, 55–67,doi:10.5194/npg-8-55-2001.

Robert, R., and J. Sommeria (1991), Statistical equilibriumstates for two-dimensional flows, J. Fluid Mech., 229, 291–310, doi:10.1017/S0022112091003038.

Rose, B., and D. Ferreira (2013), Ocean heat transport andwater vapor greenhouse in a warm equable climate: a newlook at the low gradient paradox, J. Climate, 26, 2127–2136.

Roupas, Z. (2012), Phase space geometry and chaotic attrac-tors in dissipative nambu mechanics, J. Phys. A: Mathe-matical and Theoretical, 45 (19), 195,101.

Ruelle, D. (1989), Chaotic Evolution and Strange Attractors,Cambridge University Press, Cambridge.

Ruelle, D. (1997), Differentiation of SRB states, Communica-tions in Mathematical Physics, 187 (1), 227–241.

Ruelle, D. (1998a), General linear response formula in sta-tistical mechanics, and the fluctuation-dissipation theoremfar from equilibrium, Phys. Lett. A, 245, 220–224.

Ruelle, D. (1998b), Nonequilibrium statistical mechanics nearequilibrium: computing higher-order terms, Nonlinearity,11 (1), 5–18.

Ruelle, D. (2009), A review of linear response theory for gen-eral differentiable dynamical systems, Nonlinearity, 22 (4),855–870.

Salazar, R., and M. V. Kurgansky (2010), Nambu brackets influid mechanics and magnetohydrodynamics, J. Phys. A,

Page 41: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

41

43, 305,501(1–8).Salmon, R. (1978), Two-layer quasi-geostrophic turbulence in

a simple special case, Geophys. Astrophys. Fluid Dyn., 10,25–52, doi:10.1080/03091927808242628.

Salmon, R. (1988), Hamiltonian fluid mechanics, Annu. Rev.of Fluid Mech., 20 (1), 225–256.

Salmon, R. (1998), Lectures on Geophysical Fluid Dynamics,Oxford University Press, Oxford.

Salmon, R. (2005), A general method for conservingquantities related to potential vorticity in numericalmodels, Nonlinearity, 18 (5), R1–R16, doi:10.1088/0951-7715/18/5/R01.

Salmon, R. (2007), A general method for conserving energyand potential enstrophy in shallow water models, J. Atmos.Sci., 64, 515–531.

Salmon, R., G. Holloway, and M. C. Hendershott (1976),The equilibrium statistical mechanics of simple quasi-geostrophic models, J. Fluid Mech., 75, 691–703, doi:10.1017/S0022112076000463.

Saltzman, B. (2001), Dynamical Paleoclimatology, AcademicPress New York.

Schneider, T. (2006), The general circulationof the atmosphere, Annual Review of Earthand Planetary Sciences, 34 (1), 655–688, doi:10.1146/annurev.earth.34.031405.125144.

Schubert, G., and J. Mitchell (2013), Planetary atmospheresas heat engines, in Comparative Climatology of TerrestrialPlanets, University of Arizona Press and Lunar and Plan-etary Institute.

Schutte, C., J. Walter, C. Hartmann, and W. Huisinga (2004),An averaging principle for fast degrees of freedom exhibit-ing long-term correlations, Multiscale Modeling & Simula-tion, 2 (3), 501–526, doi:10.1137/030600308.

Seager, S., and D. Deming (2010), Exoplanet atmospheres,Annual Review of Astronomy and Astrophysics, 48, 631–672.

Sellers, W. (1969), A global climatic model based on the en-ergy balance of the earthatmosphere system, J. Appl. Me-teorol., 8, 392–400.

Shepherd, T. (1990), Symmetries, conservation laws, andhamiltonian structure in geophysical fluid dynamics, Adv.in Geophysics, 32, 287–338.

Sommer, M., and P. Nevir (2009), A conservative schemefor the shallow-water system on a staggered geodesic gridbased on a Nambu representation, Quart. J. Roy. Meteorol.Soc., 135 (639), 485–494, doi:10.1002/qj.368.

Steinheimer, M., M. Hantel, and P. Bechtold (2008), Con-vection in lorenz’s global energy cycle with the ECMWFmodel, Tellus A, 60, 1001–1022, doi:10.1111/j.1600-0870.2008.00348.x.

Stone, P. H. (1978), Constraints on dynamical transports ofenergy on a spherical planet, Dyn. Atmos. Oceans, 2, 123–139.

Storch, J.-S. V., C. Eden, I. Fast, H. Haak, D. Hernndez-Deckers, E. Maier-Reimer, J. Marotzke, and D. Stammer(2012), An estimate of the Lorenz energy cycle for theWorld Ocean Based on the 1/10 STORM/NCEP simu-lation, J. Phys. Oceanogr., 42, 2185–2205.

Tailleux, R. (2013), Available potential energy and exergy instratified fluids, Ann. Rev. Fluid. Mech, 45, 35–58.

Takhtajan, L. (1994), On foundation of the generalizedNambu mechanics, Commun. Math. Phys., 160, 295–315.

Taylor, K. E., R. J. Stouffer, and G. A. Meehl (2012), Anoverview of CMIP5 and the experiment design, Bull. Amer.

Meteor. Soc., 93, 485–498.Trenberth, K. E., and J. M. Caron (2001), Estimates of merid-

ional atmosphere and ocean heat transports, J. Clim, 14,3433–3443.

Trenberth, K. E., and J. T. Fasullo (2010), Simulation ofpresent-day and twentyfirst-century energy budgets of thesouthern oceans, J. Climate, 23, 440–454.

Turkington, B. (1999), Statistical Equilibrium Measures andCoherent States in Two-Dimensional Turbulence, Comm.Pure Appl. Math., 52, 781–809, doi:10.1002/(SICI)1097-0312(199907)52:7¡781::AID-CPA1¿3.0.CO;2-C.

Vallis, G. K. (2006), Atmospheric and Oceanic Fluid Dynam-ics: Fundamentals and Large-scale Circulation, CambridgeUniversity Press, Cambridge.

Vallis, G. K., and M. E. Maltrud (1993), Generation of meanflows and jets on a beta-plane and over topography, J. Phys.Oceanogr., 23, 1346–1362.

Venaille, A. (2012), Bottom-trapped currents as statisticalequilibrium states above topographic anomalies, J. FluidMech., 699, 500, doi:10.1017/jfm.2012.146.

Venaille, A., and F. Bouchet (2011a), Solvable phase diagramsand ensemble inequivalence for two-dimensional and geo-physical turbulent flows, J. Stat. Phys., 143, 346–380, doi:10.1007/s10955-011-0168-0.

Venaille, A., and F. Bouchet (2011b), Oceanic rings and jetsas statistical equilibrium states, J. Phys. Oceanogr., 41,1860, doi:10.1175/2011JPO4583.1.

Venaille, A., G. K. Vallis, and S. M. Griffies (2012), The cat-alytic role of beta effect in barotropization processes, J.Fluid Mech., 709, 490–515.

Waite, M. L., and P. Bartello (2004), Stratified turbulencedominated by vortical motion, J. Fluid Mech., 517, 281–308, doi:10.1017/S0022112004000977.

Wang, J., and G. K. Vallis (1994), Emergence of Fofonoffstates in inviscid and viscous ocean circulation models, J.Mar. Res., 52, 83–127.

Warn, T. (1986), Statistical mechanical equilibria of the shal-low water equations, Tellus A, 38 (1), 1–11.

Wouters, J., and V. Lucarini (2012), Disentangling multi-levelsystems: averaging, correlations and memory, Journal ofStatistical Mechanics: Theory and Experiment, 2012 (03),P03,003.

Wouters, J., and V. Lucarini (2013), Multi-level dynami-cal systems: Connecting the ruelle response theory andthe mori-zwanzig approach, Journal of Statistical Physics,151 (5), 850860, doi:10.1007/s10955-013-0726-8.

Wunsch, C. (2005), The total meridional heat flux and itsoceanic and atmospheric partition, J. Clim., 18, 4374–4380.

Wunsch, C. (2012), Discrete Inverse and State Estimation -With Geophysical Fluid Applications, Cambridge Univer-sity Press, Cambridge.

Yatsuyanagi, Y., Y. Kiwamoto, H. Tomita, M. Sano,T. Yoshida, and T. Ebisuzaki (2005), Dynamics of Two-Sign Point Vortices in Positive and Negative Temper-ature States, Phys. Rev. Lett., 94 (5), 054,502, doi:10.1103/PhysRevLett.94.054502.

Young, L. (2002), What are SRB measures, and which dy-namical systems have them?, Journal of Statistical Physics,108, 733–754.

Zou, J., and G. Holloway (1994), Entropy maximization ten-dency in topographic turbulence, J. Fluid Mech., 263, 361,doi:10.1017/S0022112094004155.

Zwanzig, R. (1960), Ensemble method in the theory of irre-versibility, The Journal of Chemical Physics, 33 (5), 1338,

Page 42: Mathematical and Physical Ideas for Climate Science · 2016-03-03 · Mathematical and Physical Ideas for Climate Science Valerio Lucarini Klimacampus, Meteorologisches Institut,

42

doi:doi:10.1063/1.1731409.Zwanzig, R. (1961), Memory effects in irreversible thermody-

namics, Physical Review, 124 (4), 983–992.

Zwanzig, R. (2001), Nonequilibrium statistical mechanics, Ox-ford University Press, Oxford.