1. Introduction: Work, Heat, and Water
The second law of thermodynamics states that a process is thermodynamically spontaneous only if it yields work or at least is capable of yielding work (irrespective of whether it yields or costs heat): the maximum work that it can possibly yield is its free-energy yield. And that a process that costs work (irrespective of whether it yields or costs heat) is thermodynamically nonspontaneous: the minimum work required to enable it is its free-energy cost [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11]. These statements are true for all processes, isothermal or not. Two nonisothermal examples: (a) Neutralization of a temperature difference (clearly a nonisothermal process) is capable of yielding work, say, via a heat engine, and hence is thermodynamically spontaneous, but its reverse costs work and hence is thermodynamically nonspontaneous. (b) Adiabatic expansion of a gas (clearly a nonisothermal process) is capable of yielding work, say via a pressure engine, and hence is thermodynamically spontaneous if the external pressure is less than the pressure of the gas, but costs work and hence is thermodynamically nonspontaneous if the external pressure exceeds the pressure of the gas. Of course a thermodynamically spontaneous process can occur without yielding any work and hence with its ability to do work being wasted, e.g., process (a) above without intervention of a heat engine, and process (b) above without intervention of a pressure engine. However, any thermodynamically spontaneous process certainly has the ability to yield work, even though this ability can be wasted. The reverse of any thermodynamically spontaneous process has no ability to do work but costs work, and hence is thermodynamically nonspontaneous [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11]. (See: Reference [
1], Chapter 1; Reference [
2], Chapters 12–19, especially Chapters 13, 17, and 19 (in Chapter 17 most especially Section 17.2); Reference [
3], especially Chapters I–IV, VII, and VIII; Reference [
4], especially Chapters 1–5 and 7; Reference [
5], Chapter 8; Reference [
6], especially Chapters 1–3, 5, 7, and 10; Reference [
7], especially Chapters 1–5; Reference [
8], Chapters 2–5, especially Sections 2.1, 2.2, 4.2c, 4.5, and 4.6, and Chapter 5; and Reference [
9], Foundations (especially Part B) and Chapters 1–3, especially Part 2A and pp. 131–135. References [
8,
9,
10] base thermodynamics on the concept of work [
11], rather than on the concept of entropy. The advantage of this approach is that work is the most basic, simplest, most sharply defined, and most familiar concept in thermodynamics [
8,
9,
10,
11], and all more complex thermodynamic concepts, such as energy, heat, entropy, and temperature can be derived from the concept of work [
8,
9,
10,
11]. Concerning this point, in Reference [
8] see most especially Sections 2.2b and 2.2c and the second paragraph of Section 4.2c. Unfortunately, the most valuable of the discussions in Sections 2.1, 2.2, and 4.2c were omitted from, or at best stated less completely in, the 8th, 9th, and 10th Editions of Reference [
8]. But since the 10th Edition of Reference [
8] is the most recent one, we cite it as Reference [
9].)
Some clarification concerning the relation between heat and thermodynamic spontaneity may be required: I. A process that yields heat is thermodynamically spontaneous only if it also yields work or at least can yield work. Two examples: (i) A system can spontaneously yield heat to its surroundings only if it is hotter than its surroundings. However, if it is hotter than its surroundings, then a heat engine can employ this temperature difference to yield work. (ii) An exothermic (heat-yielding) chemical or nuclear reaction can occur spontaneously only if it is also exoergic (work-yielding). II. A process that costs heat is thermodynamically spontaneous only if it also yields work or at least can yield work. Two examples: (iii) A system can spontaneously absorb heat from its surroundings only if it is colder than its surroundings. However, if it is colder than its surroundings, then a heat engine can employ this temperature difference to yield work. (iv) An endothermic (heat-costing) chemical or nuclear reaction can occur spontaneously only if it is also exoergic (work-yielding). In example (i) above, the system is in its initial state sensibly hotter than its surroundings; in example (iii) above, sensibly colder. In example (ii) above, the chemical or nuclear reactants can be construed as latently hotter than their surroundings; in example (iv) above, latently colder. Thus for instance the warming of a cold room as per example (i) above via the physical process (sensibly hot object + sensibly cold room → sensibly lukewarm object + sensibly lukewarm room) is, in essence, thermodynamically equivalent to warming the room as per example (ii) above via the chemical process of burning coal (latently hot molecules (C + O) + sensibly cold room → latently lukewarm molecules (CO) + sensibly lukewarm room). (Note: “exergonic” and “endergonic” are sometimes used as synonyms for “exoergic” and “endoergic”, respectively, but mainly pertaining to biochemical reactions.)
A work-costing and hence nonspontaneous process can occur only if enabled by another, spontaneous, process that yields more work than the nonspontaneous one costs [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11]. It should be emphasized that a work-costing and hence nonspontaneous process would violate the second law of thermodynamics only if it could occur
without compensation. A work-costing and hence nonspontaneous process that occurs
with compensation, i.e., enabled by another, spontaneous, process that yields more work than the nonspontaneous one costs [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11], does
not violate the second law, because the
total process—the sum of both precesses combined—is a work-yielding process.
While the second law has been challenged [
12,
13,
14,
15,
16,
17], even if some, or even all, of these challenges are borne out it would still have a very wide range of validity: As stated in Reference [
14], (p. 13): “If the second law should be shown to be violable, it would nonetheless remain valid for the vast majority of natural and technological processes.” The range of validity of the second law would still encompass all processes considered in this paper. Hence for the purposes of this paper we need not consider challenges to the second law [
12,
13,
14,
15,
16,
17]. (The entireties of References [
12,
13,
14,
15,
16,
17] concern challenges to the second law of thermodynamics. Hence References [
12,
13,
14,
15,
16,
17] are cited in their entirety. This includes the Special Issues of
Entropy cited in References [
13,
17] and the Issue of
Foundations of Physics cited in Reference [
15].)
A thermodynamically neutral process neither yields nor costs work. Thus there is no tendency for such a process to occur, and hence it cannot proceed at a finite rate, in either the forward or reverse direction. An example is a work-costing Process B enabled or driven by an equally work-yielding Process A. The total Process A + B is thermodynamically neutral, and hence cannot proceed at a finite rate. Thus the limit of perfection or reversibility, i.e., Process B recovering all of the work yielded by Process A, is unattainable. In any real such total Process A + B, Process B can recover some—but not all—of the work yielded by Process A, so that the total Process A + B is a work-yielding process. In most—but not all—cases the limit of perfection or reversibility is approached more closely as a process proceeds more slowly. Counterexamples: (a) Adiabatic cooling during cosmological expansion of the Universe or adiabatic heating during its contraction of a sufficiently simple substance such as cosmic background radiation, dust, or possibly a monatomic ideal gas can occur reversibly at a finite rate [
18] (In Reference [
18] see Sections 130 and 169–171.) (b) Conversely, sliding and rolling friction do not vanish in the limit of vanishing speed, precluding perfection or reversibility from being approached even in this limit [
19,
20,
21,
22,
23]. (In Reference [
19] see
Section 6-1.) In fact, sliding friction is greater at vanishing speed than at moderate speed: (coefficient of static sliding friction) > (coefficient of kinetic sliding friction) [
19,
20,
21]. Hence sliding friction entails greater departure form perfection or reversibility at vanishing speed than at moderate speed [
19,
20,
21].
The generalized free energy associated with a process is the maximum work this process can yield if it is spontaneous, or the minimum work required to enable it if is nonspontaneous, under any circumstances. The immediately preceding sentence is true with respect to more specialized varieties of free energy (the most commonly employed ones being the Helmholtz and Gibbs free energies) only under restricted circumstances (usually taken to be isothermal, isochoric for the Helmholtz free energy and isothermal, isobaric for the Gibbs free energy). (See: Reference [
1], Section 1.27; Reference [
2], Chapter 17 (especially pp. 476–479) and Chapter 19; Reference [
3], Chapter VIII; Reference [
4], Chapters 5 and 7; Reference [
5], Sections 8.9–8.13; Reference [
6], Chapters 7 and 10, pp. 144–150, Section 11.7, pp. 270–282, and Section 12.10; Reference [
7], Chapter 5; Reference [
8], pp. 108–134; and Reference [
9], pp. 131–135.) However, we primarily consider free energy in the generalized sense indicated in the first sentence of this paragraph. Free energy and its relation to work will be discussed in more detail in
Section 3.
We show that in atmospheric thermodynamics condensation is an effect rather than a cause, i.e., a passenger rather than a driver: even though condensation always yields heat, it always costs work. Hence the second law of thermodynamics requires that condensation must be enabled by another, work-yielding process that yields more work than condensation costs. Condensation is most usually effected by cooling to below the dew point. In the case of most interest to us, convective weather systems and storms, the work-yielding process entails neutralization of super-moist-adiabatic vertical temperature gradients, hence tending to cool Earth’s (land or ocean) surface and warm the tropopause—in accordance with Nature’s abhorrence of gradients. The role of water vapor in convective-weather-system thermodynamics is to render available the super-moist-adiabatic portion of vertical temperature gradients and hence of the free energy associated therewith for driving convection. Convection, in turn, enables cooling of ascending air parcels to below the dew point and hence attainment of supersaturation in clouds, because water vapor cannot condense fast enough to maintain relative humidity in clouds exactly at saturation. Supersaturation, in turn, ultimately enables condensation and consequent release of heat of condensation: given supersaturation, condensation of water vapor and consequent release of heat of condensation is thermodynamically spontaneous. However, establishment of supersaturation, and maintenance of supersaturation in the face of condensation that would wipe it out, is itself never thermodynamically spontaneous in Earth’s atmosphere or anywhere else. It is driven by convection that is too fast to allow condensation to maintain relative humidity in clouds exactly at saturation, and this convection is in turn driven by super-moist-adiabatic lapse rates. Thus ultimately condensation and consequent release of heat of condensation is a thermodynamicaly nonspontaneous process forced by super-moist-adiabatic lapse rates. The super-moist-adiabatic portion of vertical temperature gradients—not water vapor—is the free energy source.
By contrast, in a dry atmosphere, only the super-dry-adiabatic portion of vertical temperature gradients is available to drive convection. At typical atmospheric temperatures in the tropics, where convective weather systems and storms are most frequent and active, the moist-adiabatic lapse rate is much smaller (thus much closer to isothermality) and hence represents much more extractable work than the dry—the thermodynamic advantage of water vapor. Thus the role of water vapor in atmospheric thermodynamics is to facilitate the extraction of work at the expense of a much larger fraction (the super-moist-adiabatic fraction) of vertical temperature lapse rates and hence of the free energy associated therewith than would be possible without it (the super-dry-adiabatic fraction). Moreover, the large heat of condensation (and to a lesser extent fusion) of water facilitates much faster heat transfer from Earth’s surface to the tropopause than is possible in a dry atmosphere, thereby facilitating much faster extraction of work, i.e., much greater power output, than is possible in a dry atmosphere—the kinetic advantage of water vapor. Thus the role of water vapor in convective-weather-system kinetics is to facilitate, via its large heat of condensation (and to a lesser extent fusion), much faster expenditure of this much larger—super-moist-adiabatic as opposed to super-dry adiabatic—fraction of vertical temperature gradients and hence of the free energy associated therewith than would be possible without it. Evaporation at Earth’s surface and condensation at high tropospheric altitudes greatly enhances the rate of heat flux from Earth’s surface to the tropopause over that attainable in a dry atmosphere.
What is more, the very properties of water help to maintain super-moist-adiabatic lapse rates in Earth’s troposphere. Water vapor, because it is a greenhouse gas, raises temperatures at Earth’s surface via reradiation of infrared frequencies from Earth’s atmosphere (mainly from the troposphere) back to Earth’s surface, and enhances radiation of infrared frequencies to space from tropospheric altitudes well above Earth’s surface thus lowering temperatures at these higher altitudes. Thus it helps to maintain larger average temperature lapse rates between Earth’s surface and the tropopause than would obtain without it. Condensed into clouds, water helps to do the same thing during the nighttime hours: Clouds act as a blanket trapping infrared radiation from Earth’s surface and reradiating it back to Earth’s surface, thereby raising nighttime temperatures at Earth’s surface, while infrared radiation to space from cloud tops, which are better radiators than air, helps to lower nighttime temperatures there. This helps to increase nighttime ambient vertical lapse rates in the air surrounding the clouds. The steepening of lapse rates with the help of water vapor at all times, and with the help of water vapor condensed into clouds at night, is greatest where there is the most water vapor—in the humid tropics—which is where convective weather systems and storms are most frequent and active. So it most assists convective weather systems and storms where they need it most. (Of course, during the daytime, clouds usually lower temperatures at Earth’s surface: reduced insolation usually more than offsets the heat-blanket effect.)
In
Section 2, we discuss condensation and consequent release of heat of condensation, and show that it is always a work-costing and hence nonspontaneous process in Earth’s atmosphere. The work-yielding process that pays for it is neutralization of super-moist-adiabatic vertical temperature gradients, i.e., cooling of Earth’s (land or ocean) surface and warming of the tropopause—Nature’s abhorrence of gradients. Super-moist-adiabatic vertical temperature gradients, indeed all (nongeothermal) temperature disequilibria on Earth, are, in turn, ultimately paid for by the even greater temperature disequilibrium between the
solar disk and the
cosmic background radiation [
24]. In
Section 3, the Helmholtz, Gibbs, and generalized free energies are discussed in some detail. Our preference for generalized free energy is explained. In
Section 4 the limitation upon convective weather systems and storms in a saturated atmosphere to extract only the work obtainable by partial neutralization to moist-adiabaticity of vertical temperature gradients with temperature decreasing upwards is discussed. It is contrasted with the more severe limitation upon those in an unsaturated atmosphere to extract only the work obtainable by partial neutralization to dry-adiabaticity of vertical temperature gradients with temperature decreasing upwards, and with the lack of any such limitation on the ability of heat engines in general to extract all of the work obtainable by total neutralization of any temperature gradient irrespective of direction to isothermality. Also in
Section 4 convective or hydrodynamic utilizability is contrasted with thermodynamic utilizability. The kinetic acceleration of extraction of work—the increased power—in a saturated atmosphere over that in a dry one is discussed in
Section 5. A synopsis is provided in
Section 6. Auxiliary topics are discussed in
Appendix A,
Appendix B, and
Appendix C.
We do not provide detailed calculations of convective-weather-system thermodynamics. These are provided in many other papers [
25,
26,
27,
28,
29,
30,
31,
32,
33,
34,
35,
36,
37,
38,
39]. (In Reference [
31] (a book) synopses are provided at pp. 12–16 and Chapter 10.) Our main goal is conceptual rather than computational—to show that condensation and consequent release of heat of condensation is always an effect rather than a cause, i.e., always a passenger rather than a driver, in Earth’s atmosphere, albeit an
important passenger.
2. Free Energy, Convection, and Condensation
Condensation is thermodynamically spontaneous and hence can yield work—not merely heat—but
only if there is supersaturation, i.e., relative humidity exceeding the saturation value, which is
given a plane surface of pure water. At any given temperature and pressure, the
energy of water vapor
always exceeds that of liquid water (or of ice), but the
Gibbs free energy of water vapor exceeds that of liquid water (or of ice)
only given
supersaturation [
27,
30,
40] (in Reference [
40] see Section 5.6). However, most of Earth’s atmosphere is
subsaturated. Nevertheless, supersaturation certainly exists in Earth’s atmosphere, especially in clouds. The point is that in all such cases the establishment of supersaturation, and its maintenance in the face of condensation that would wipe it out, is itself a work-costing and hence thermodynamically nonspontaneous process, and hence, ultimately, so is condensation and consequent release of heat of condensation that is enabled by supersaturation. Four examples of supersaturation that exist in Earth’s atmosphere: (a) Thermal radiation to space from cloud tops, which are better radiators than air, can lower cloud temperatures, especially at and near the tops of clouds, faster than condensation can maintain relative humidity in clouds exactly at saturation, thereby trapping some water vapor in metastable supersaturation. (b) Lapse rates exceeding the moist adiabatic can drive convection and thence condensation faster than condensation can maintain relative humidity in clouds exactly at saturation, thereby trapping some water vapor in metastable supersaturation. (c) At
relative humidity water vapor is at thermodynamic equilibrium with a plane surface of pure water. However, saturation or equilibrium relative humidity increases with decreasing water droplet size (owing to a small droplet’s curvature water molecules on its surface have fewer neighbors than on a plane surface and hence are less tightly bound and can evaporate more easily) and decreases with increasing nonvolatile solute concentration (solute molecules occupy part of the surface and if hygroscopic also bond with water molecules). Also saturation or equilibrium relative and absolute humidity with respect to metastable supercooled water droplets is higher than with respect to ice crystals at the same (subfreezing) temperature (because water molecules are less tightly bound in the liquid state than in the solid state and hence can evaporate more easily from the former). Thus water can spontaneously evaporate from small and/or low-solute-concentration droplets and condense onto large and/or high-solute-concentration droplets, and similarly spontaneously evaporate from supercooled water droplets and condense onto ice crystals—but never the reverse. (See: Reference [
40], Sections 5.8 and 5.10–5.11; Reference [
41], Sections 6.1 and 6.5; and Reference [
42], Chapter 5.) Note that any one, any two, or all three of these processes (a)–(c) can operate within any given cloud. (d) Supersaturation has been observed in the upper tropical troposphere and lower tropical stratosphere [
43]. However: (i) Supersaturation is a metastable state, in all four cases obtaining at the expense of tropospheric vertical super-moist-adiabatic temperature gradients ultimately paid for by insolation from the
solar disk heating Earth’s surface and the coldness of the
cosmic background radiation allowing radiative cooling of the tropopause [
24]. Thus, in all four cases, metastable supersaturation represents temporarily
stored free energy derived at the expense of vertical super-moist-adiabatic temperature gradients that are its
source. (ii) In all four cases the degree of supersaturation is typically small,
well below the maximum possible supersaturation of approximately
relative humidity that obtains in perfectly clean air, at which point water vapor molecules themselves become effective condensation nuclei [
44]: even the most pristine tropospheric and even lower stratospheric air almost always contains enough solid condensation nuclei so that condensation can begin at or at least very near the normal dew point [
44]. (In Reference [
44] see pp. 83–84.) (iii) Concerning Case (d): The upper tropical troposphere and lower tropical stratosphere are so cold that the total amount of water vapor trapped in metastable supersaturation even given its maximum possible value of approximately
relative humidity [
44] is bound to be very small. Moreover as noted in (ii) immediately above actual values of supersaturation probably are always
well below this maximum [
44]. (iv) Thus, in all four cases, supersaturation represents very limited free-energy
storage derived at the expense of super-moist-adiabatic vertical temperature gradients that are its
source, and ultimately at the expense of the
-solar-disk/
-cosmic-background-radiation disequilibrium that continually regenerates these gradients [
24]. A vertical temperature lapse rate exceeding the moist adiabatic in a saturated atmosphere between Earth’s hot surface and cold tropopause embodies the free energy that yields the work enabling convection and thence supersaturation, condensation, and consequent release of heat of condensation—not vice versa. (See: Reference [
7], Section 1.5 (especially Problem 1.40 on p. 27) and Section 5.3 (especially Problem 5.45 on pp. 177–178); Reference [
40], Sections 3.3–3.5 and 4.3, pp. 175–177, and Sections 6.5–6.9; Reference [
41], Sections 3.4–3.6; Reference [
42], Sections 1.2 and 3.5–3.7; and Reference [
45], Sections 2.7 and 2.9, and Appendix D.) It embodies
free energy rather than merely energy because it can yield
work. Once convection is enabled, no additional work is required to condense the water—condensation and consequent release of heat of condensation is a side result of the convective process. However, the work that it costs to enable convection is paid for by the work yielded by neutralization of super-moist-adiabatic temperature gradients, this neutralization tending to cool Earth’s (land or ocean) surface and warm the tropopause—by Nature’s abhorrence of gradients. Only if convection is thus enabled can the side result of supersaturation and thence condensation and consequent release of heat of condensation then also be enabled. It costs work—the work that it costs to enable convection—to
force supersaturation and thence condensation and consequent yielding of heat of condensation: supersaturation and thence condensation and consequent yielding of heat of condensation is thus a thermodynamically
nonspontaneous process that must be
forced.
Thus in all cases of supersaturation in convective weather systems and storms, the free-energy hierarchy is:
-solar-disk/
-cosmic-background-radiation disequilibrium → tropospheric vertical super-moist-adiabatic temperature gradients → convection → cooling of ascending air parcels to below the dew point → supersaturation → condensation and consequent release of heat of condensation [
24]. Condensation is a work-yielding and hence thermodynamically spontaneous given supersaturation. However, establishment of supersaturation, and its maintenance in the face of condensation that would wipe it out, is itself a work-costing and hence thermodynamically nonspontaneous process that can occur only because tropospheric vertical super-moist-adiabatic temperature gradients can yield more work than establishment and maintenance of supersaturation costs. Thus ultimately condensation and consequent release of heat of condensation can occur only because tropospheric vertical super-moist-adiabatic temperature gradients can yield more work than establishment and maintenance of supersaturation costs.
Perhaps this point is most clearly made evident by considering the
optimum circumstance for generation and maintenance of convective weather systems and storms with a
sub-moist-adiabatic vertical temperature gradient. This optimum circumstance with a sub-moist-adiabatic lapse rate obtains if the entire troposphere is saturated at
relative humidity and however high an absolute humidity, with however high an equivalent potential temperature
(the temperature of air if brought moist-adiabatically to a pressure of
). Release of heat of condensation limits the rate of cooling of saturated ascending air parcels to the moist adiabatic, resulting in positive buoyancy and thence in further convection—but only if the environmental lapse rate exceeds the moist adiabatic. Even given the
optimum circumstance with a
sub-moist-adiabatic environmental lapse rate—the entire troposphere saturated at
relative humidity and however high an absolute humidity, with however high an equivalent potential temperature
—buoyancy is negative even with the heat of condensation. Thus even given this optimum circumstance with a sub-moist-adiabatic lapse rate, the atmosphere is incapable of doing the work necessary to enable convection, thence supersaturation as per Item (b) of the first paragraph of this
Section 2, and finally the goal of condensation and consequent release of heat of condensation—despite the heat of condensation being equally large at any given temperature whether the lapse rate is sub- or super-moist-adiabatic. Thus even given this optimum circumstance with a sub-moist-adiabatic lapse rate, a jump from saturation to supersaturation—let alone a jump from subsaturation to supersaturation—cannot be made, and hence condensation and consequent release of heat of condensation cannot occur. Thus there can then be no convection, hence no supersaturation, and thence no condensation and no release of heat of condensation. Convective weather systems and storms—not even the smallest fair-weather cumulus clouds, let alone thunderstorms, tornadoes, and tropical cyclones—could not then exist. (See: Reference [
40], Chapters 5–6; Reference [
41], Sections 8.3 and 8.4; and Reference [
45], Sections 9.5–9.7.)
Thus supersaturation and thence condensation and consequent release of heat of condensation can occur only via being enabled, albeit as a side result of convection being enabled, by a process that yields
work. Neutralization of any temperature gradient is capable of yielding work. In the case currently under consideration, neutralization of a super-moist-adiabatic vertical temperature gradient—albeit only
partial neutralization to the moist adiabatic and not
total neutralization to isothermality because the sub-moist-adiabatic portion is convectively or hydrodynamically
unutilizable—yields the work required to enable convection, thence supersaturation, and ultimately condensation and consequent release of heat of condensation. (Convective or hydrodynamic utilizability will be discussed more thoroughly, and contrasted with thermodynamic utilizability, in
Section 4.) Owing to release of heat of condensation, heat transfer from Earth’s surface to the tropopause is more effective than would be possible without it, thereby facilitating this partial neutralization—but this partial neutralization drives condensation and consequent release of heat of condensation, and not vice versa.
The main point is that a super-moist-adiabatic lapse rate enables condensation and consequent release of heat of condensation, but condensation and consequent release of heat of condensation does not enable a super-moist-adiabatic lapse rate. A super-moist-adiabatic lapse rate is the cause and condensation is the effect, and not vice versa. This is what we mean by a super-moist-adiabatic lapse rate being the driver and condensation being the passenger, albeit an important passenger.
Of course it is well known that Earth’s atmosphere is always out of equilibrium, and that it is maintained out of equilibrium only at the expense of the even greater disequilibrium between the
solar disk and the
cosmic background radiation [
24]. The disequilibrium between the
solar disk and the
cosmic background radiation [
24] is at the top of the free-energy hierarchy
-solar-disk/
-cosmic-background-radiation disequilibrium → tropospheric vertical super-moist-adiabatic temperature gradients → convection → cooling of ascending air parcels to below the dew point → supersaturation → condensation and consequent release of heat of condensation [
24]. Hence the disequilibrium between the
solar disk and the
cosmic background radiation [
24] is the
ultimate driver of all organized (nongeothermally-driven and nontidal) activity—meteorological or otherwise—on Earth. But it is nonetheless of interest to inquire concerning, and to distinguish between,
immediate drivers (
immediate causes) and passengers (effects).
Wind drives (causes) swaying of tree branches and not vice versa. Thus wind is the immediate driver (immediate cause) and swaying of tree branches is the passenger (effect), and not vice versa. Likewise, super-moist-adiabatic lapse rate drives (causes) condensation and consequent release of heat of condensation in convective weather systems and storms, and not vice versa. Thus super-moist-adiabatic lapse rate is the immediate driver (immediate cause), and condensation and consequent release of heat of condensation is the passenger (effect), and not vice versa. The “tree theory” of wind, according to which wind is driven (caused) by trees swaying their branches, is wrong. Likewise, the “condensation theory” of convective weather systems and storms, according to which convective weather systems and storms are driven (caused) by condensation and consequent release of heat of condensation, is wrong. Both theories are wrong for the same reason: Setting of tree branches to swaying and condensation of water vapor are both work-costing and hence thermodynamically nonspontaneous processes in Earth’s atmosphere. A work-costing process can occur only if driven by a work-yielding process that yields more work than the work-costing one costs. Tree branches sway only because braking of wind yields more work than it costs to set them swaying. Condensation and consequent release of heat of condensation occurs in convective weather systems and storms only because neutralization of super-moist-adiabatic lapse rate towards the moist adiabatic yields more work than condensation and consequent release of heat of condensation costs.
Vertical temperature gradient exceeding the moist adiabatic is the major free-energy source that yields the work enabling convection and thence supersaturation, condensation, consequent release of heat of condensation, and convective weather systems and storms. While other, auxiliary, free-energy sources may yield some contribution, they cannot generate or maintain convective weather systems and storms in the face of a sub-moist-adiabatic lapse rate. For example, kinetic (wind) energy imported from the surrounding atmosphere is in many cases a significant auxiliary free-energy source for convective weather systems and storms, but still only an auxiliary one: it cannot generate or maintain convective weather systems and storms in the face of a sub-moist-adiabatic lapse rate.
3. Helmholtz, Gibbs, and Generalized Free Energy
The type of free energy that is expended by convective weather systems and storms should be specified. The two most commonly employed types are the Helmholtz and Gibbs free energies. Employing the standard notations
P for pressure,
V for volume,
E for internal energy,
S for entropy, and
for enthalpy, the differential changes in the Helmholtz function
and Gibbs function
(not yet specified as the Helmholtz and Gibbs free energies) of a system are, respectively,
and
In the second line of Equation (
1) we apply the first law of thermodynamics in the form
. Taking the limit of reversibility, in the third and fourth lines of Equation (
1)
and
, respectively. In the second line of Equation (
2) we apply
, and in the third line thereof the first law of thermodynamics in the form
. Taking the limit of reversibility, in the fourth and fifth lines of Equation (
2)
and
, respectively. Note that, strictly, in Equations (1) and (2)
P is the
external pressure acting
on the system of interest, but since we are considering the limit of reversibility
P is also the
internal pressure
of the system of interest. Since pressure, volume, temperature, and entropy are all state functions, the Helmholtz function
A and the Gibbs function
G are also state functions. (Note that the last lines of Equations (1) and (2), if taken per mole, are the Gibbs-Duhem relations for a single-component system pertinent to the Helmholtz and Gibbs free energies, respectively. For simplicity we consider single-component systems with fixed numbers of particles, i.e., with fixed
N. The single-component water-only system is of most interest to us (other components of air mixed with water vapor have negligible effects: see Reference 40, Section 5.7. If a single-component system can exchange like particles with its surroundings, then the term
must be added to Equations (1) and (2), where
is the chemical potential). For multi-component systems a
term for each component must be added to Equations (1) and (2).)
We are interested in special cases wherein the Helmholtz function A and the Gibbs function G are not merely state functions but useful state functions, i.e., state functions with a useful physical interpretation. Henceforth we will consider only special cases wherein A and G are useful state functions: the Helmholtz free energy and Gibbs free energy, respectively. It is usually asserted that the Helmholtz function A is a useful state function only in the special case of isothermal, isochoric conditions and that the Gibbs function G is a useful state function only in the special case of isothermal, isobaric conditions. For in the special case of isothermality and isochoricity, the Helmholtz function A is the Helmholtz free energy: In this special case the magnitude of the Helmholtz-free-energy change associated with a process is the maximum work this process can yield if it is spontaneous, or the minimum work required to enable it if it is nonspontaneous. Likewise, in the special case of isothermality and isobaricity, the Gibbs function G is the Gibbs free energy: In this special case the magnitude of the Gibbs-free-energy change associated with a process is the maximum work this process can yield if it is spontaneous, or the minimum work required to enable it if it is nonspontaneous.
To be precise, the restrictions with respect to temperature can be slightly relaxed, with
A and
G still being the
useful state functions described in the immediately preceding paragraph (see Reference [
2], pp. 476–479). For our special-case definition of
as the Helmholtz-free-energy change of a system to be valid, the volume of a system undergoing a process must be maintained
strictly constant (i.e., the process must be
strictly isochoric), but the temperature of the system can vary in intermediate states of a process so long as at the very least it is at the same temperature
T in the initial and final states. Likewise, for our special-case definition of
as the Gibbs-free-energy change of a system to be valid, a system undergoing a process must be in contact with a reservoir that maintains a
strictly constant external pressure (i.e., the process must be
strictly isobaric), but the temperature of the system can vary in intermediate states of a process so long as at the very least it is at the same temperature
T in the initial and final states.
Indeed, another relaxation of restrictions, with
A and
G still being these
useful state functions, seems possible. For, surely, since
T and
V are the natural independent variables for
A, by Equation (
1) under isothermal conditions in the limit of reversibility
. Likewise, equally surely, since
T and
P are the natural independent variables for
G, by Equation (
2) under isothermal conditions in the limit of reversibility
. (In the fifth line of Equation (
2) we equate
with
rather than with
so that in the last line of Equation (
2)
is expressed in terms of its natural independent variables
and
.) Thus, under isothermal conditions, the magnitude
is the maximum work
that an increase in the volume of a system can yield, or the minimum work
required to enable a decrease of its volume. Likewise, under isothermal conditions, the magnitude
is the maximum work
that can be yielded via a system being allowed to experience a decrease in pressure, or the minimum work
required to impose an increase in pressure upon it.
Yet, even with both of these relaxations of restrictions, the Helmholtz and Gibbs free energies
A and
G do not seem to be the most general useful state functions, physically interpretable in term of work. In atmospheric processes, temperature, pressure, and volume occupied per unit mass are typically
not constant, with even initial and final states typically of
unequal temperature,
unequal pressure, and
unequal volume occupied per unit mass. Hence neither the Gibbs free energy
G nor the Helmholtz free energy
A seem to always be the best, or at least the most general, choice for atmospheric processes. Nonetheless it should be noted that the Gibbs free energy has been employed in detailed calculations of convective-weather-system thermodynamics [
27,
30]. Also, the vapor pressure at which water vapor is at thermodynamic equilibrium with liquid water (or ice) at any given temperature, with no net tendency towards either condensation or evaporation, is that corresponding to the Gibbs free energy of water vapor being equal to that of liquid water (or ice) [
27,
30,
40]. Of course, this is true with respect to not only water, but all substances.
However, both the Helmholtz and Gibbs free energies are special cases of the generalized free energy
F. The magnitude
of the generalized-free-energy change
associated with a process is the maximum work this process can yield if it is spontaneous, or the minimum work required to enable it if it is nonspontaneous, under whatever conditions of temperature, pressure, and volume occupied per unit mass. The differential change
in the generalized free energy corresponding to a given differential process is
where
is the differential change in
total entropy
of the system of interest plus its surroundings combined including their interaction corresponding to this given differential process, and
is the temperature of the coldest available heat sink (the tropopause for convective weather systems and storms). The surroundings could be the heat sink itself but in any case must at least incorporate the heat sink. Interaction correlates a system with its surroundings and thus contributes
negative entropy (
negentropy)
to
. However, in all cases of interest to us,
is negligible and hence so is the consequent nonadditivity of entropy and of changes in entropy with respect to
and
exclusively, as per the second line of Equation (
3). (The dot-equal sign (≐) means “very nearly equal to”.)
Note that: (a) The concept of generalized free energy
F is equivalent to the concept of available work if the heat sink is construed as being included within the boundary of our system, so that by the first law of thermodynamics
(see Reference [
2], Section 17.2). (b) The location of the boundary of a system is arbitrary, and hence has no effect on thermodynamic behavior. For example, it is thermodynamically equivalent to construe that System A does
work on an external System B, or to construe that this
work is done internally within the isolated combined System A + B. (c) Our definition of
remains valid even if
and/or
not only vary in intermediate states, but are
not the same even in the initial and final states (The incorporation of
into
(or
) and hence into Equation (
3) will be explained in the paragraph containing Equations (6) and (7).) (d) In general entropy and temperature are state functions, and thus
,
,
,
, and
in particular are state functions (
being negligible in all cases of interest to us). Hence the generalized free energy
F is a state function. (e) Unlike
A and
G,
F is
always a
useful state function, i.e., always a
free energy: the magnitude
of the generalized-free-energy change
associated with a process
always is the maximum work this process can yield if it is spontaneous, or the minimum work required to enable it if it is nonspontaneous, under whatever conditions of temperature, pressure, and volume occupied per unit mass. Moreover, as we will now show, both the Helmholtz and Gibbs free energies are special cases of the generalized free energy.
The differential change in the Helmholtz free energy is
The first step in the first line of Equation (
4) implicitly, and the second step therein more explicitly, entails the assumption that
and hence the consequent nonadditivity of entropy and of changes in entropy with respect to
and
exclusively is negligible, as per the second line of Equation (
3). The second step in the first line of Equation (
4) is justified because, given the validity of the definition of
in terms of work if an
entire process is isochoric, i.e., with volume fixed and if the system of interest is at the same temperature
T at the very least in the initial and final states,
. Similarly, the differential change in the Gibbs free energy is
The first step in the first line of Equation (
5) implicitly, and the second step therein more explicitly, entails the assumption that
and hence the consequent nonadditivity of entropy and of changes in entropy with respect to
and
exclusively is negligible, as per the second line of Equation (
3). The second step in the first line of Equation (
5) is justified because, given the validity of the definition of
in terms of work if an
entire process is isobaric, i.e., with ambient pressure fixed, and if the system of interest is at the same temperature
T at the very least in the initial and final states,
. Thus Equation (
4) is a special isochoric case of Equation (
3), and Equation (
5) is a special isobaric case of Equation (
3), in both special cases with the system at the same temperature
T at the very least in the initial and final states. Hence both the Helmholtz and Gibbs free energies are special cases of the generalized free energy.
The differential form as employed in Equation (
3) is most applicable for our purposes, and is also employed in Equations (4) and (5) so that as special cases they are expressed in the same form as Equation (
3). In convective meteorological systems such as thunderstorms, tornadoes, and tropical cyclones, heat transfer from the heat source at Earth’s (land or ocean) surface at temperature
to the heat sink at the tropopause at temperature
may be sufficiently large and rapid that the decrease in
of Earth’s surface and/or the increase in
of the tropopause during the progress of a storm cannot be neglected: hence the differential form as employed in Equation (
3) is the most applicable form. Of course, any differential equation can be integrated if we wish to consider an entire process or any finite portion thereof. Note again that our definition of
(or
) in terms of work remains valid even if
and/or
not only vary in intermediate states, but are
not the same even in the initial and final states.
In accordance with the first and second laws of thermodynamics,
is incorporated into
(or
) and hence into Equation (
3) via
. Corresponding to a given differential heat increment
extracted from Earth’s (land or ocean) surface, a given differential heat increment
deposited at the tropopause, a given differential increment
of work
actually done, and the
maximum theoretical differential increment
of work allowed by the second law of thermodynamics corresponding to
in general and hence to Carnot efficiency
for heat engines operating in a cycle in particular, we have [
46] (in Reference [
46] see Sections 5.11 and 5.12 and Problem 5.22 on p. 199):
The first law of thermodynamics, in the form
, is applied in the second step of the first line of Equation (
6). The second law of thermodynamics requires
and hence
or equivalently
, where
is the differential loss of free energy
by the source thereof considered alone, not to be confused with the
net differential change in free energy
:
is lost by a super-moist-adiabatic gradient upon its differential neutralization, this loss being partially recovered as
performed by a convective weather system or storm. The equalities of the ≥ and ≤ signs in the last two lines of Equation (
6) obtain only in the limit of perfection or reversibility; only in this (unattainable) limit would
lost by the gradient be fully recovered as
performed by a convective weather system or storm. The truest thermodynamic measure of efficiency is not the Carnot efficiency
but the work efficiency
If as in tropical cyclones frictional dissipation of
(
in the (unattainable) limit of perfection or reversibility) is into the
hot reservoir at Earth’s (typically ocean [
31,
32,
33,
34,
35,
36,
37,
38,
39] but untypically desert [
38] (see also Reference [
39], Endnote 10) surface rather than into the cold reservoir at the tropopause, then the cost required of the hot reservoir is reduced from
to
(to
in the (unattainable) limit of perfection or reversibility). Tropical-cyclone performance is thereby enhanced. Dissipation into the hot reservoir also has been employed to enhance the performance of thermoelectric generators [
47].
The contribution to hurricane intensity by frictional dissipation and thence recycling of a hurricane’s work output (wind) into its hot reservoir has been investigated for
typical hurricanes, whose hot reservoir is the tropical ocean surface. Such recycling must also contribute, even if only slightly, to (re)intensification over land, rather than over water, as tropical or warm-core (not extratropical or cold-core) cyclones, of
untypical hurricanes in the tropical deserts of northern Australia, whose hot reservoir is the sandy desert soil [
38] (see also Reference [
39], Endnote 10). The major cause is the increased thermal conductivity of the sandy desert soil when moistened by the hurricane’s own rainfall, which more than compensates for the negative effects of cooling of the soil by this same rainfall. Recycling of a hurricane’s work output (wind) via friction into this hot reservoir must contribute, even if only slightly, towards partially counteracting this rain-cooling.
Contrasting energy sources versus
free-energy sources: A
free-energy source is a source
solely of
disequilibrium, or equivalently of
localization or
negentropy—not at all of
energy. Note well that by Equations (3)–(6) a differential change of
free energy is proportional
solely to a differential change of
total entropy (or total negentropy )—not at all to a differential change of
total energy. Indeed by the first law of thermodynamics even a differential or infinitessimal change of total energy is
impossible. Energy is incorporated into free energy only via multiplication of
by
. The dimensions of free energy are indeed those of energy, not those of entropy or negentropy, but these dimensions obtain only via multiplication of
by
— it is
and not
that provides the drive (except via contributing to
by being lower than
). A free-energy source can be the disequilibrium between ambient and either an energy source or an energy sink—burning fuel can run a heat engine but so can liquid nitrogen, ice, or evaporating water [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58] (Summaries of Reference [
52] are provided in References [
53,
54], and summaries of Reference [
56] are provided in References [
57,
58]. An auxiliary website is available online for Reference [
58].) Disequilibrium is localization: a heat engine running on burning fuel expends
localization of energy (
not energy itself) from the heat source (burning fuel), delocalizing this energy into the ambient environment; one running on liquid nitrogen, ice, or evaporating water [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58] expends
localization of energy (
not energy itself) in the ambient environment, delocalizing this energy into the heat sink (liquid nitrogen, ice, or evaporating water). Evaporation of water into an unsaturated atmosphere (relative humidity less than
) is a thermodynamically spontaneous process, yielding work even though it costs heat [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58]. Since most of Earth’s atmosphere is unsaturated, in most of Earth’s atmosphere evaporation—not condensation—is the work-yielding and hence thermodynamically spontaneous process, even though evaporation always costs heat and condensation always yields heat. This is even neglecting that establishment of supersaturation, and its maintenance in the face of condensation that would wipe it out, is always work-costing and hence thermodynamically nonspontaneous in Earth’s atmosphere or anywhere else, and hence that even in supersaturated regions of Earth’s or any other atmosphere condensation is ultimately always work-costing and hence thermodynamically nonspontaneous. The work yielded by evaporation of water sets into motion heat engines including but not limited to drinking birds and toy cars [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58]. It also sets thunderstorm downburst winds into motion (with the help of the weight of raindrops, and sometimes also with the help of melting of hail, which, like evaporation into an unsaturated atmosphere (relative humidity less than
), yields work even though it costs heat). (See: Reference [
41], pp. 350–352 and 361–362; and Reference [
42], Sections 5.7, 8.2, and 8.10, and pp. 268–270.) It also helps to lift water up trees [
48]. Research and development towards practical evaporation-powered engines for numerous applications has been undertaken [
52,
53,
54] and is currently underway [
55,
56,
57,
58]. Thus there are evaporation-driven engines [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58] but not condensation-driven engines: Convective weather systems and storms are lapse-rate-driven engines, not condensation-driven engines. Evaporation yields work in accordance with Equations (3) and (6): Evaporation costs heat and thus costs localization in momentum space and hence a decrease in momentum-space or thermal entropy—the width at half maximum of the Maxwellian velocity, or, better, momentum distribution decreases with decreasing temperature. However, in an unsaturated atmosphere (relative humidity less than
) delocalization of water molecules in position space increases enough upon evaporation to yield a more-than-compensating increase in position-space or configurational entropy. This delocalization in position space upon evaporation occurs, first, owing to breaking of correlations between water molecules, and second, after these correlations have been broken, owing to expansion in position space of the water vapor. Thus upon evaporation into an unsaturated atmosphere there occurs
net delocalization in the
total momentum-plus-position phase space, and hence the
total thermal-plus-configurational entropy
increases. (As already noted in the second sentence of
Section 2, at any given temperature and pressure, the
energy of water vapor
always exceeds that of liquid water (or ice), but the
Gibbs free energy of water vapor exceeds that of liquid water (or ice)
only given
supersaturation [
27,
30,
40].) To summarize this paragraph: In all cases the total energy of a system plus its surroundings combined is constant: Work is always obtained
solely at the expense of
delocalization of energy, i.e., at the expense of expenditure of
negentropy, never at the expense of
expenditure of energy. (One point: On p. 479 of Reference [
2] it is stated that in an adiabatic process all of the energy lost by a system can be converted to work, but that in a nonadiabatic process less than all of the energy lost by a system can be converted to work. However, if the entropy of a system undergoing a nonadiabatic process
increases, then
more than all of the energy lost by this system can be converted to work, because energy extracted from the surroundings can then also contribute to the work output. In some such cases positive work output can be obtained at the expense of the surroundings even if the change in the system’s energy is
zero, indeed even if the system
gains energy. Two examples: (a) Isothermal expansion of an ideal gas yields work even though the energy change of the ideal gas is
zero. (b) As per this paragraph and References [
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58], evaporation of liquid water (or ice) into an unsaturated atmosphere (relative humidity less than
) yields work even though liquid water (or ice)
gains energy in becoming water vapor. Both examples (a) and (b) immediately above are thermodynamically spontaneous processes; their reverses are nonspontaneous. They both entail delocalization of energy from the surroundings into the system. This emphasizes that delocalization of energy rather than energy per se enables work.)