1. Introduction
Lorenz (1955), building on the work of Margules (1910), formulated the concept of available potential energy (APE) as that portion of the total potential energy, that is, the sum of the gravitational and internal energy in hydrostatic balance, that is available for conversion into kinetic energy. Since then the concept has been a cornerstone of large-scale dynamic meteorology and the theory of the general circulation (e.g., Lorenz 1967; Peixoto and Oort 1991). The Lorenz formulation is a Lagrangian one that determines the APE for an isentropic rearrangement of air parcels. For an atmosphere that is statically stable, the entropy increases with height. Thus, an isentropic vertical rearrangement of the parcels is excluded and the theory only assesses the available energy associated with the leveling of isentropic layers toward geopotential surfaces. This feature is an ideal means to assess the energy available for conversion to kinetic energy by baroclinic instability processes. However, this feature precludes the theory from evaluating the energy available for a purely vertical (e.g., convective) rearrangement. The theory requires the specification of a reference state. For example, if the APE is defined on isobaric coordinates, a reference (potential) temperature must be assigned. The specification of the reference temperature remains an unsolved problem in the theory (e.g., Dutton and Johnson 1967; Lorenz 1979; Pauluis 2007). [Randall and Wang (1992) advance a numerical, parcel-moving algorithm to find the temperature structure for their generalized convective available potential energy (GCAPE) that is applied to a single atmospheric sounding.]
Eulerian formulations of the available energy have appeared in several formulations: entropic energy (Dutton 1973; Livezey and Dutton 1976), static exergy (Karlsson 1990), extended exergy (Kucharski 1997), and available energy (Bannon 2005). These formulations differ slightly in their base state and their treatment of water vapor and hydrometeors. They all require the specification of a reference temperature. Dutton (1973) uses a reference temperature defined from the total potential energy. Karlsson (1990) uses a reference temperature that minimizes the entropy difference between the atmosphere and its reference atmosphere. Kucharski (1997) uses a reference temperature profile based on the horizontally averaged density field.
The present work reexamines and refines the formulation of the Eulerian atmospheric available energy (AE) of Bannon (2005) that defines the available energy as a generalized Gibbs function between the atmosphere and an isothermal reference atmosphere at the temperature Tr. The reference atmosphere is in thermal and hydrostatic equilibrium (Fig. 1) and, hence, is dynamically and convectively “dead.” It has the same mass per unit area of dry air and water as the atmosphere. The reference temperature is determined uniquely by minimizing this function. Section 2 presents a thermodynamic proof that the maximum kinetic energy that can be extracted from the total energy is specified by the minimum of the generalized Gibbs function. Minimization is defined to occur at the isothermal reference temperature T0. The outline of the proof follows but generalizes that in Reif (1965, section 8.3). Independently Karlsson (1990) suggested this minimization of his exergy formulation but did not pursue it. Section 3 describes the construct of the equilibrium atmosphere and the need for the careful treatment of the hydrometeors. Section 4 derives the governing equation for the evolution of the available energetics of a flow. The sources and sinks of AE are quantified as well as the partitioning of the AE into available potential and available elastic components. The positive definite nature of the AE is demonstrated along with its relation to the exergy approach of engineering thermodynamics (Bejan 1997). Section 5 presents some applications of AE to the standard atmosphere, idealized baroclinic zones, and observed moist soundings. It is demonstrated that the available energy shares properties with the Lorenz APE as well as convective available potential energy (CAPE). The available energy is partitioned into available baroclinic energy (ABE) and available convective energy (ACE) components. Section 6 estimates the available energetics of the general circulation.

Schematic diagram depicting an atmosphere A in thermal, mechanical, and diffusive contact with a hydrostatic, isothermal reference atmosphere. The reference surface vapor pressure is that for saturation at the temperature Tr for a reference atmosphere in contact with a water reservoir. The reference atmosphere is a dynamically and convectively “dead” state.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
2. Thermodynamic derivation of the atmospheric available energy





































The atmospheric Gibbs function δGA for a 25-km-deep standard atmosphere as a function of the reference temperature Tr. This function, denoted by the heavy solid curve, has a minimum at the equilibrium temperature, denoted by the asterisk on the abscissa, of T0 = 251.95 K. The solid and dashed curves denote the potential and elastic contributions to this function, respectively, defined in section 4b.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1







3. Specification of the equilibrium state











This asymmetry is physically appropriate for the equilibrium atmosphere. The criterion (Gibbs 1873, 1874, p. 146) for equilibrium of a system in a gravitational field is that the total chemical potential (i.e., the sum of the intrinsic chemical potential μ and the geopotential gz) is constant. The expression (3.7) for the dry air and water vapor satisfies this criterion. In contrast, the total potentials for the hydrometeors are not constant and these components are not in equilibrium. This disequilibrium is physically correct because all hydrometeors have nonzero terminal fall speeds and will eventually settle out.

4. Available energetics
This section uses a fluid dynamical approach to generalize the thermodynamic derivation of section 2 to include diabatic and frictional processes for an open atmosphere. It also quantifies the sources and sinks of the available energy and resolves issues related to the asymmetry in the chemical potentials discussed in section 3.
a. General derivation of the governing equation


























b. Partitioning and linearization of the available energy
The available energy (4.1) can be partitioned into the sum of the available energies of each component. Using relations of the form











The preceding formulas emphasize the enthalpic nature of the available energy. Their definitions in terms of normalized departures are computationally advantageous to ensure positive definite calculations as well as providing easy derivations of their quadratic nature.
5. Applications
The available energy formalism is applied to several atmospheres to determine their equilibrium temperature T0 and associated available energy.
a. Standard atmosphere
The case of the standard atmosphere in Fig. 2 is an example of the minimization of the atmospheric Gibbs function δGA. This hydrostatic atmosphere extends to a height of 25 km with a surface temperature and pressure of 288.15 K and 1013.25 hPa. The lapse rates are 6.5, 0, −1.0, and −2.8 K km−1 for z < 11, 20, 32, and 50 km, respectively. The potential and elastic contributions are defined by (4.17) and (4.18). The temperature profile is plotted in Fig. 3a, where it is compared to its equilibrium temperature T0 = 251.95 K. The total energy (TE) is slightly greater than that of its equilibrium atmosphere for both the 25- and 50-km-deep cases (Table 1). The internal energy (IE = U) differences between the atmosphere and its equilibrium are small and change sign between the two cases. For both, the potential energy (PE) is less for the equilibrium atmosphere. This reduction in PE is consistent with a lower center of gravity associated with the lower temperature of the equilibrium atmosphere in the lower troposphere (Fig. 3a). Because of the finite vertical extent of each atmosphere, the ratio of the potential to internal energy is less than that of R/cυa = 40%. Thus, the total energy is only approximately the enthalpy (cf. Lorenz 1955).

(a) Temperature and (b) available energy as a function of height z for a 25-km-deep standard atmosphere. In (a) the thick solid and solid curves are the soundings for the atmosphere and its equilibrium atmosphere, respectively, with T0 = 251.95 K. In (b) the solid, dashed, and thick solid curves denote the potential, elastic, and total contributions. The dotted vertical line is the zero line.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
Traditional energetics of two standard atmospheres of different depths. Values for the equilibrium atmosphere are in parentheses. The total energy is the sum of the internal and potential energies

The available energy (AE) for both atmospheres resides primarily (Table 2) in the available potential energy (APE) with about 10% in the available elastic energy (AEE). For the 25 km deep atmosphere, the available energy of 11.45 MJ m−2 is much less than its total energy (Table 1) of 2.4733 GJ m−2. Thus, only 0.46% is available for conversion into kinetic energy; 0.48% for the 50 km deep atmosphere. In contrast the available energy agrees well with the difference in total energy
Available energetics of the two standard atmospheres.

The vertical distribution of the available energy (Fig. 3b) is bimodal with a maximum at the surface, a secondary maximum at the tropopause, and a midtropospheric minimum near where the equilibrium temperature and the sounding (Fig. 3a) intersect. The secondary peak lies at the height of the change in the lapse rate. The elastic energy is significant in the stratosphere and its contribution dominates that of the potential energy above 20 km. This behavior holds for all of the cases analyzed and reflects that the elastic energy (4.18) is inversely proportional to the square of the pressure field.
b. Idealized baroclinic zones
The idealized baroclinic zones are based on a generalized compressible Eady base state (appendix C). The available energetics is summarized in Table 3 as a function of the total meridional temperature gradient
Available energetics of the idealized baroclinic zone as a function of horizontal temperature gradient

The spatial variation of the available energy (Fig. 4a) exhibits a bimodal variation with larger values where the difference between the temperature and the equilibrium temperature is larger. The dominance of the temperature variation, rather than pressure variation, reflects the dominance of the available potential energy contribution (4.17) to the total available energy. The efficiency factor

Cross section of the (a) specific available energy and (b) the efficiency factor N = (T − T0)/T for the idealized baroclinic zone with
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
The effect of a surface pressure gradient associated with a mean geostrophic wind is small (Table 4) with a slight increase in available energy as the wind transitions from westerly to easterly. Henceforth, the surface wind is set to zero.
Available energetics of the idealized baroclinic zone with

The effect of sloping topography on the available energy is readily included in this Eulerian formulation. Table 5 provides a comparison for topography sloping linearly upward toward the pole (henceforth poleward) or toward the equator (henceforth equatorward) to reach a maximum height of 3 km (appendix C). The poleward case contains more available energy. This result is consistent with the linear Eady model analysis of Mechoso (1980) that indicates greater growth rates for unstable baroclinic waves for terrain sloping with the isentropes.
Available energetics of the idealized baroclinic zones over topography sloping upward toward the pole or toward the equator. The horizontal temperature gradient is

The effect of the inclusion of a moist boundary layer (appendix C) on the available energy is summarized in Table 6. The vapor pressure has a surface relative humidity of 70% and decays vertically with a scale height of 3 km. Inclusion of moisture has increased the equilibrium temperature by ~5 K to 265.12 K and increased all available energy components of the dry air. The vapor contributes to the available energy primarily through its elastic component. The distribution of the available energy and the efficiency factor (not shown) is similar to that of Fig. 4 for the dry case. The major effect of the moisture is through the increased equilibrium temperature that moves the zero efficiency isopleth downward. Figure 5 illustrates the effects of water vapor on the minimization of the Gibbs function. For reference temperatures below about 270 K, the reference atmosphere contains an ice reservoir of tens of kilograms per square meter of water. The enthalpy change associated with this phase transformation increases the Gibbs function curve at those temperatures and moves the temperature minimum toward warmer temperatures. The equilibrium atmosphere is saturated with an ice reservoir that is 1.29 cm thick due to precipitation (3.3). The ice contributes the major amount of the difference in total energy between the atmosphere and its equilibrium by its enthalpy of deposition (Table 6). The vertical distributions of the available energies (not shown) are similar to those presented in section 5d, which analyzes observed moist atmospheric soundings.
Available energetics of the moist idealized baroclinic zone with


(a) The atmospheric Gibbs function δGA for the moist baroclinic zone as a function of the reference temperature Tr. The thick solid, solid, and dashed curves denote the total, dry air, and water vapor contributions. (b) The thick solid curve denotes the mass of water vapor Mυr in the reference atmosphere as a function of Tr. The horizontal dashed line is the mass of water vapor Mυ in the atmosphere. Their difference, Mυ − Mυr, denotes the amount of water in the atmosphere precipitated into the water reservoir in the reference state. The reference surface vapor pressure
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
c. Comparison with the available potential energy of Lorenz








Available baroclinic energetics of the idealized baroclinic zones with horizontal temperature gradient

d. Individual moist soundings
A real data case is afforded by the sounding from the Second Verification of the Origins of Rotation in Tornadoes Experiment [VORTEX2; 2155 UTC 5 June 2009, National Severe Storms Laboratory (NSSL1)] presented in Fig. 6. The sounding exhibits a large potential for deep convection for parcels ascending above about 700 hPa along the 70°C saturated adiabat. The convective available potential energy (CAPE) evaluated from 700 to 170 hPa is 2370 J kg−1. Assuming hydrostatic balance, the available energy may be analyzed in pressure coordinates. The minimization of the atmospheric Gibbs function (not shown) is similar to that depicted in Fig. 5. Table 8 and Fig. 7 present the results of an AE analysis of the profile. The dry air AE (Fig. 7a) exhibits a bimodal distribution with the potential dominating the elastic contribution. In contrast, the water vapor AE (Table 8) is dominated by the elastic contribution. This result implies that the vapor pressure perturbations dominate the potential temperature perturbations (section 4b). The water vapor AE (Fig. 7b) is also bimodal but with a thick 200-hPa layer of near-zero AE. This minimum reflects the impact of the low dewpoints between 700 and 500 hPa in the sounding (Fig. 6). The AE may be compared to the convective available potential energy by dividing by the total mass of the sounding M0 = 6.845 × 104 kg m−2, yielding a CAPE of 1541 J kg−1.

Skew T–logp diagram for the VORTEX2 sounding with the thick solid and dashed–dotted curves denoting the temperature and dewpoint temperature, respectively.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
Available energetics of the VORTEX2 sounding. The equilibrium temperature is T0 = 262.1 K, and a surface vapor pressure


Available energy for the (a) dry air and (b) water vapor as a function of pressure for the VORTEX2 sounding. The total, potential, and elastic contributions are denoted by the thick solid, solid, and dashed curves, respectively, for each component.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
A second case is phase III of the Global Atmospheric Research Program Atlantic Tropical Experiment (GATE) sounding analyzed by Randall and Wang (1992; see their Table 1 for the “given sounding”). This tropical sounding (Fig. 8) is relatively moist with the temperature profile close to the saturated adiabat of 70°C. This structure suggests a small value of CAPE. In contrast, the AE analysis of Table 9 and Fig. 9 indicates that the sounding is a potentially large source of kinetic energy. The general features of the AE profiles are consistent with those for the extratropical VORTEX2 sounding. These include a bimodal vertical distribution, a dominant potential contribution by the dry air, and a dominant elastic contribution by the water vapor. With a total mass of M0 = 9.135 × 104 kg m−2, the bulk CAPE is 2212 J kg−1. In contrast, Randall and Wang estimate their GCAPE to be 11 J kg−1.

As in fig. 6, but for the GATE III sounding.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
Available energetics of the GATE Phase III sounding. The equilibrium temperature is T0 = 271.4 K with


As in fig. 7, but for the GATE III sounding.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
6. Global available energy and the general circulation






Quantification requires the determination of the equilibrium temperature T0 for the atmosphere. Analysis of the 25-km-deep standard atmosphere with a 70% surface relative humidity and 3 km water vapor scale height yields an equilibrium temperature of 256 K with an available energy of 14 MJ m−2. A more definitive determination requires analysis of a global dataset. Such a calculation is beyond the scope of the current investigation. For definiteness in the following discussion, we take T0 = 255 K, the planetary/effective temperature.
a. Dissipation of kinetic energy
The dissipation term (6.3) includes three distinct processes: viscous dissipation of kinetic energy within the atmosphere and the surface flux of kinetic energy out of the atmosphere by the surface wind stresses and by the precipitation of hydrometeors.
1) Internal dissipation

2) Kinetic energy loss due to precipitation





3) Kinetic energy loss due to wind stress
With surface wind speeds exceeding that of the underlying surface, there is a downward transport of kinetic energy out of the atmosphere given by
Thus, the total dissipation rate is about 5–7 W m−2. Table 10 summarizes the discussion.
Dissipation of kinetic energy. In a steady state the total dissipation rate equals the rate of conversion of available energy to kinetic energy.

b. Generation of available energy
The generation rate (6.5) includes surface fluxes of exergy into the atmosphere and interior diabatic processes. The interior terms are weighted by the efficiency factor
1) Radiation
















The effect of clouds on the radiative generation could be significant. Low-level cloud bases, where the efficiency is positive generally, absorb longwave radiation emitted from the surface, leading to a positive AE generation. Emission of longwave radiation to space from high-level cloud tops, where the efficiency is negative, will also generate positive AE.
2) Dissipation

3) Precipitation exergy flux







4) Evaporation exergy flux





5) Dry air exergy flux







c. Irreversible entropy production
The last term in (6.4) represents irreversible entropy production processes (e.g., the subcloud evaporation of raindrops). It is positive definite and hence is a sink of available energy. The term
Table 11 summarizes the discussion on the available energy generation budget.
Generation of available energy. In a steady state, the total rate of generation of available energy less the lost work equals the rate of conversion of available energy to kinetic energy.

7. Conclusions
Minimization of an atmospheric Gibbs function has been utilized to determine the temperature
The analysis of the available energy cycle of the general circulation in section 6 is summarized in Fig. 10 and quantified in Table 12. A measure of the efficiency

Schematic diagram of the available energy cycle of the global atmosphere.
Citation: Journal of the Atmospheric Sciences 69, 12; 10.1175/JAS-D-12-059.1
Estimates of the available energy cycle of the global atmosphere.

The author benefited from suggestions from colleagues John A. Dutton, Dennis Lamb, Sukyoung Lee, Paul M. Markowski, and Raymond G. Najjar. Paul M. Markowski provided the VORTEX2 sounding. The comments of Olivier Pauluis and an anonymous reviewer also helped improve the manuscript.
APPENDIX A
Thermodynamics
The thermodynamic formulations are standard (e.g., Bohren and Albrecht 1998). The dry air and water vapor are treated as ideal gases with constant specific heats. The ice and liquid water are taken to be incompressible with constant specific heats. The enthalpies of phase change vary linearly with temperature following Kirchhoff’s law. This dependence is included in the Clausius–Clapeyron equations for the equilibrium vapor pressures. Table A1 summarizes the notation and values of various constants.
Physical constants.

APPENDIX B
Asymmetric Chemical Potentials










APPENDIX C
Idealized Baroclinic Zones
Idealized dry baroclinic zones are constructed with inspiration from the model of Eady (1949) but for a compressible atmosphere. The geometry is Cartesian with (y, z) indicating meridional and vertical coordinates. The central sounding is that for the standard atmosphere as are the lapse rates. The surface temperature varies linearly over the domain with a total variation of
The topography is
The effects of moisture on the baroclinic isolated jet are assessed by including a water vapor field that has a relative humidity of 70% at the surface with a scale height of 3 km. Then, the warmer lower latitudes will have greater moisture content than those at higher latitudes.
REFERENCES
Bannon, P. R., 2005: Eulerian available energetics in moist atmospheres. J. Atmos. Sci., 62, 4238–4252.
Bejan, A., 1997: Advanced Engineering Thermodynamics. 2nd ed. Wiley, 850 pp.
Bohren, C. F., , and B. A. Albrecht, 1998: Atmospheric Thermodynamics. Oxford, 402 pp.
Dutton, J. A., 1973: The global thermodynamics of atmospheric motion. Tellus, 25, 89–110.
Dutton, J. A., , and D. R. Johnson, 1967: The theory of available potential energy and a variational approach to atmospheric energetics. Advances in Geophysics, Vol. 12, Academic Press, 333–436.
Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1, 33–52.
Gibbs, J. W., 1873: On the equilibrium of heterogeneous substances. Trans. Conn. Acad. Arts Sci., 3, 108–248.
Gibbs, J. W., 1874: On the equilibrium of heterogeneous substances. Trans. Conn. Acad. Arts Sci., 3, 343–524.
Karlsson, S., 1990: Energy, entropy and exergy in the atmosphere. Ph.D. thesis, Institute of Physical Resource Theory, Chalmers University of Technology, 121 pp.
Kucharski, F., 1997: On the concept of exergy and available potential energy. Quart. J. Roy. Meteor. Soc., 123, 2141–2156.
Livezey, R. E., , and J. A. Dutton, 1976: The entropic energy of geophysical fluid systems. Tellus, 28, 138–157.
Lorenz, E. N., 1955: Available potential energy and the maintenance of the general circulation. Tellus, 7, 157–167.
Lorenz, E. N., 1967: The Nature and Theory of the General Circulation. WMO Press, 161 pp.
Lorenz, E. N., 1979: Numerical evaluation of moist available energy. Tellus, 31, 230–235.
Margules, M., 1910: On the energy of storms. Smithson. Misc. Collect., 51, 533–595.
Mechoso, C. R., 1980: Baroclinic instability of lows along sloping boundaries. J. Atmos. Sci., 37, 1393–1399.
Pauluis, O., 2007: Sources and sinks of available potential energy in a moist atmosphere. J. Atmos. Sci., 64, 2627–2641.
Pauluis, O., , and I. M. Held, 2002: Entropy budget of an atmosphere in radiative–convective equilibrium. Part I: Maximum work and frictional dissipation. J. Atmos. Sci., 59, 125–139.
Pauluis, O., , and J. Dias, 2012: Satellite estimates of precipitation-induced dissipation in the atmosphere. Science, 335, 953–956.
Pauluis, O., , V. Balaji, , and I. M. Held, 2000: Frictional dissipation in a precipitating atmosphere. J. Atmos. Sci., 57, 989–994.
Peixoto, J. P., , and A. H. Oort, 1991: Physics of Climate. American Institute of Physics, 564 pp.
Randall, D. A., , and J. Wang, 1992: The moist available energy of a conditionally unstable atmosphere. J. Atmos. Sci., 49, 240–255.
Reif, F., 1965: Fundamentals of Statistical and Thermal Physics. McGraw-Hill, 651 pp.
Trenberth, K. E., , J. T. Fasullo, , and J. Kiehl, 2009: Earth’s global energy budget. Bull. Amer. Meteor. Soc., 90, 311–323.