Next Article in Journal
Regional Fluctuations in the Eastern Tropical North Pacific Oxygen Minimum Zone during the Late Holocene
Previous Article in Journal
Coupled Meteo–Hydrodynamic Approach in Semi-Enclosed Basins and Sensitivity Assessment of Wind-Driven Current
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TEOS-10 Equations for Determining the Lifted Condensation Level (LCL) and Climatic Feedback of Marine Clouds

1
Leibniz Institute for Baltic Sea Research (IOW), 18119 Warnemünde, Germany
2
Leibniz-Sozietät der Wissenschaften zu Berlin, 12587 Berlin, Germany
3
Leibniz Institute for Tropospheric Research (TROPOS), 04318 Leipzig, Germany
*
Author to whom correspondence should be addressed.
Oceans 2024, 5(2), 312-351; https://doi.org/10.3390/oceans5020020
Submission received: 19 March 2024 / Revised: 8 May 2024 / Accepted: 24 May 2024 / Published: 28 May 2024

Abstract

:
At an energy flux imbalance of about 1 W m−2, the ocean stores 90% of the heat accumulating by global warming. However, neither the causes of this nor the responsible geophysical processes are sufficiently well understood. More detailed investigations of the different phenomena contributing to the oceanic energy balance are warranted. Here, the role of low-level marine clouds in the air–sea interaction is analysed. TEOS-10, the International Thermodynamic Equation of State of Seawater—2010, is exploited for a rigorous thermodynamic description of the climatic trends in the lifted condensation level (LCL) of the marine troposphere. Rising sea surface temperature (SST) at a constant relative humidity (RH) is elevating marine clouds, cooling the cloud base, and reducing downward thermal radiation. This LCL feedback effect is negative and counteracts ocean warming. At the current global mean SST of about 292 K, the net radiative heat flux from the ocean surface to the LCL cloud base is estimated to be 24 W m−2. Per degree of SST increase, this net flux is expected to be enhanced by almost 0.5 W m−2. The climatic LCL feedback effect is relevant for the ocean’s energy balance and may be rigorously thermodynamically modelled in terms of TEOS-10 equations. LCL height may serve as a remotely measured, sensitive estimate for the sea surface’s relative fugacity, or conventional relative humidity.

1. Introduction

“Of the air, the part receiving heat is rising higher. So, evaporated water is lifted above the lower air”. (Leonardo da Vinci, ca. 1508. Paper sheets from the Arundel Codex, after 1508, in handwriting of the old Leonardo, discovered by Anna Maria Brizio in 1951 [1]). Translated from the Italian original by Marianne Schneider (1996) [2], p. 79. Published German text: “Der Teil der Luft steigt höher, der an der Wärme teilhat. Also dringt die Ausdünstung des Wassers hinauf über die niedrige Region der Luft, die sie berührt”.
In the title of this paper, emphasis is placed in particular on “TEOS-10 Equations”. TEOS-10, the Thermodynamic Equation of Seawater—2010 [3], was adopted in 2009 by UNESCO/IOC (IOC: Intergovernmental Oceanographic Commission, https://www.ioc.unesco.org/en, accessed on 10 March 2024) and in 2011 by IUGG (IUGG: International Union of Geodesy and Geophysics, https://iugg.org/, accessed on 10 March 2024) as an international geophysical standard providing the thermodynamic properties of seawater, ice, and humid air. TEOS-10 is designed to cover, with significantly improved quality and range of validity, not only the ocean’s interior [4], but also the properties of the ocean–atmosphere interface. In this paper, the related capabilities and rigor of TEOS-10 equations are demonstrated mathematically by developing and discussing conceptual models in the context of selected current problems from climate research. As with any other model, these are also only imperfect and simplified descriptions of a complex reality.
In typical global balance models, the energy exchange at the ocean–atmosphere interface includes quantitative estimates of five major contributions, namely, solar downward irradiation (≈165 W m−2), thermal upward radiation from the sea surface (≈−400 W m−2), thermal downward radiation from clouds and greenhouse gases (≈343 W m−2), the latent heat of evaporated water (≈−95 W m−2), and sensible heat exchange (≈−12 W m−2) by molecular conduction across the interface [5,6,7,8]. The figures given in brackets are respective rough estimates for the enthalpy fluxes per ocean surface area, which result in a net imbalance of about 1 W m−2 that is warming up the ocean. This is an amount 25 times as large as the world’s total energy consumption by humans in 2022 [9]. Similarly, the long-term global energy balance at the top of the atmosphere also amounts to about 1 W m−2 [10]. This way, the ocean contributes about 89% [11] or 90% [12,13] of the total to the globally accumulated energy of the warming climate, while the atmosphere takes only a minor share of just 1%. With the advent of the Thermodynamic Equation of State of Seawater—2010 (TEOS-10) as the current comprehensive international geophysical thermodynamic standard, previous methods for estimating the ocean heat content [12] can be replaced by more accurate modern formulas [14,15], as shown in Section 3.
Each of the energy fluxes mentioned above may, to some extent, be responsible for the observed ocean warming. However, identifying the causes of and processes behind the enhanced ocean heat imbalance is scientifically challenging. The magnitude of the air–sea flux imbalance is well below the typical uncertainty threshold of climate models of at least 10 W m−2 [6]. “The drivers of a larger [Earth energy imbalance] EEI in the 2000s than in the long-term period since 1971 are still unclear, and several mechanisms are discussed in literature. For example, Loeb et al. [16] argue for a decreased reflection of energy back into space by clouds (including aerosol cloud interactions) and sea ice and increases in well-mixed greenhouse gases (GHG) and water vapor to account for this increase in EEI. Kramer et al. (2021) [17] refer to a combination of rising concentrations of well-mixed GHG and recent reductions in aerosol emissions to be accounting for the increase, and Liu et al. (2020) [18] address changes in surface heat flux together with planetary heat redistribution and changes in ocean heat storage. Future studies are needed to further explain the drivers of this change, together with its implications for changes in the Earth system” [11].
“Low-altitude clouds with cloud top temperatures close to the ground environment generally lead to a weak greenhouse effect. In addition, the radiative cooling at the cloud top and the warming at the cloud base may enlarge the vertical temperature gradient and consequently intensify the turbulence” [19]. However, in some climate models “the low cloud response to [sea-surface temperature] SST change may be too weak” [20]. Developing climate models of marine low clouds and their sensitivity to SST variation is still a challenging task [21]. These phenomena are worth modelling using TEOS-10 equations from the Thermodynamic Equation of Seawater—2010 [3]. As a contributor to the marine heat balance, low-level clouds affect the upwelling thermal radiation of the surface area fraction they cover. By their altitude and due to their related cloud-base temperature, low-level clouds drive downward longwave radiation (Figure 1). As a function of the surface temperature and relative humidity (RH), an adiabatically ascending air parcel reaches its dew point at the so-called lifted condensation level (LCL). In this article, reference equations for the LCL are derived from the international standard TEOS-10. As a longwave radiation feedback effect, deviations in the downward black-body radiation of those clouds can be estimated this way from the observed trends related to the global warming of the ocean surface.
In Section 2 of this paper, selected details of ocean warming are briefly reviewed and discussed in relation to other climatic trends, such as cloudiness reductions, SST increases, or apparently constant marine RH and wind speeds. Section 3 provides a short introduction to the thermodynamic framework of TEOS-10. In Section 4, TEOS-10 equations are formulated for the LCL pressure and temperature as functions of surface pressure, temperature, and relative humidity. The rigorous results obtained using those equations may serve as reference values for other non-standard estimates. In addition to selected numerical results obtained from an iteration scheme, analytical approximations are suggested for the almost-saturated marine troposphere, as well as for the tutorial case of the simple crude Gibbs function, as explicitly defined in Appendix C. Section 5 investigates the climatic sensitivity of the LCL temperature to the SST at constant RH, with selected numerical solutions of TEOS-10 equations and approximate formulas for crude Gibbs functions. Section 6 concludes this paper with a discussion of the climatological context and derived solutions. Appendix A introduces the Jacobi method for a convenient manipulation of the partial derivatives used in TEOS-10. Appendix B calculates an equation from TEOS-10 for assessing the adiabatic lapse rate of the dew point, as required in Section 4. Appendix D mathematically explains Lambert’s W function, which is frequently encountered in atmospheric balance equations. Finally, Appendix E provides a list of symbols used in this paper.
Finally, it should explicitly be underlined at this point again that the very aim of this paper is an introduction to the mathematical apparatus of TEOS-10. For this purpose, selected current problems of climate research at the air–sea interface are briefly reviewed and exploited as tutorial examples, for which simple conceptual models are developed. Familiar such examples may serve to reduce the inevitable learning curve associated with actively making use of TEOS-10. For those models, TEOS-10 equations are presented and solved either numerically or via an approximate analytical method. Before TEOS-10, to the knowledge of the authors, geophysical and climate problems had never been described axiomatically in terms of non-trivial thermodynamic potentials which, however, are well known to be the most elegant, complete, consistent, and accurate tools available for this task. Moreover, TEOS-10 is the current international standard adopted in order to harmonise and overcome the widespread use of various collections of empirical property equations whose levels of mutual consistency are often unclear. An extended list of references regarding previous TEOS-10 work is given for readers who may be interested in the more detailed context of the particular approaches and equations presented here.

2. Application Context: Ocean Warming and Cloudiness

Through the last 70 years, the global mean sea surface temperature (SST) has increased from about 17.8 °C in 1955 to about 18.8 °C in 2023 [11], see the long-term trends graphically displayed later in this section, with an estimated linear trend of
t / ° C 18.4 + 0.015 × y r 2000 .
Here, y r is the number of the Gregorian calendar year. Accordingly, depending on the dataset used for the analysis, the heat content of the upper 2000 m of the world ocean is currently increasing at a rate of either about 0.9 W m−2 or 1.3 W m−2 per ocean surface area [9]. These values are consistent with a total Earth energy imbalance exceeding 1 W m−2 [20] and the fact that the ocean alone stores 90% of the globally accumulated heat [13].
Recent ocean warming is strongest along the global west-wind belts, as shown in Figure 2, and is spatially correlated with regions of high mean cloud coverage, as shown in Figure 3.
On the input side of the ocean’s energy imbalance, the systematic global reduction in cloudiness is estimated by the linear trend [23],
C 0.664 0.006 × y r 2000 ,
Here, C is defined by the fraction of the surface area covered by clouds. Evidently, this trend is negatively correlated with the tropospheric increase in temperature and specific humidity. Consistent with reduced cloudiness, between 1998 and 2017, Earth’s albedo declined by 0.5 W m−2 [24,25].
Decreased cloudiness increased the mean solar irradiation, J , passing through clear-sky regions,
J / W   m 2 0.41 + 0.037 × y r 2000 ,
by about 0.74 W m−2 in the past 20 years [26], see Figure 4 below. This phenomenon is known as the warming “shortwave cloud radiative effect” (SW CRE). This value may possibly explain the observed ocean warming if the ocean’s heat loss remains sufficiently below the SW CRE heat gain.
On the output side of the ocean’s energy imbalance, the sensible heat exchange of roughly 10 W m−2 between ocean and atmosphere is relatively small and does not exceed the uncertainty of flux estimates [6]. The ocean’s heat capacity exceeds that of the atmosphere by a factor of 1000, so that 1% of added ocean’s heat content corresponds to an atmospheric temperature increase 10 times larger that of the ocean. Sensible heat flux may be decreasing due to the faster rise in tropospheric temperatures compared to those of the sea. Estimates for such a trend are not available, though, and the quantitative contribution of sensible heat flow to ocean warming remains a pending problem.
On the output side of the ocean’s energy imbalance, the problem of ocean warming is closely related to a suspected change in the strength of the hydrological cycle. Cheng et al. [13] argue that the observed increase in salinity contrast in the upper 2000 m “is generally consistent with many atmosphere-based estimates and strengthens the evidence that the global water cycle has been amplified with global warming”. On the other hand, Held and Soden [27] have already emphasised that “it is important that the global-mean precipitation or evaporation, commonly referred to as the strength of the hydrological cycle, does not scale with Clausius–Clapeyron. We can, alternatively, speak of the mean residence time of water vapor in the troposphere as increasing with increasing temperature”. In contrast to some model predictions [28], there is no evidence yet that mean global marine evaporation has intensified, which would have cooled the ocean in contrast to the warming observed in reality [23,29,30].
Quantitatively, corresponding to about 1200 mm of annual evaporation, the latent heat exchange of roughly 100 W m−2 between ocean and atmosphere is the dominant means of oceanic heat loss and atmospheric energy supply [8,25,31]. According to the Dalton equation, the evaporation flux depends mainly on the RH and the wind speed [29,32,33]. Trends in marine surface RH are weak and are, if measurable at all, well below the level of measurement uncertainty [34,35,36]. “There is a tendency for the RH of the air to remain approximately constant as the climate changes” [31]. Similarly, there is no significant trend in marine wind speeds [37] so that relevant trends in global mean marine evaporation rates are neither expected nor observed [27]. However, the high sensitivity of latent heat flux with respect to RH may be estimated to be about 5 W m−2 per %rh [25,38,39], so that a small increase of 0.2 %rh may already suffice to explain the oceanic warming rate. Such a minor change in RH, however, could be disguised by the much larger observation uncertainty of 1–5 %rh [34]. “Relative humidity over oceans remained highly uncertain” [36]. Therefore, the relevance of latent heat with respect to ocean warming remains unclear.
On the output side of the ocean’s energy imbalance, the thermal radiation of the sea surface may easily be estimated from the Stefan–Boltzmann law of black-body radiation. Only a small fraction of this radiation may pass through the troposphere via the so-called atmospheric window at about 8–11 µm wavelength, as shown in Figure 5; most energy is blocked by the infrared opacity of clouds and greenhouse gases and sent back down to the ocean via long-wave radiation. This occurs according to their particular altitudes and temperatures. This “longwave cloud radiative effect” (LW CRE) of reduced cloudiness is cooling the ocean by letting more thermal radiation pass into space and emitting less downward radiation, as shown in Figure 6. It is estimated [26] that this cooling LW CRE almost completely cancels out the warming SW CRE of cloudiness. A consequence of this analysis is that ocean warming may only be attributed to decreasing cloudiness to a minor extent.
In addition to the lateral area covered by clouds, the mutual energy balance between SW and LW CRE is also influenced by other processes, namely:
(i)
In contrast to the SW CRE, the LW CRE strongly depends on the concentration of greenhouse gases present in the cloudless sky. If the clear sky absorbs infrared radiation to a similar extent as clouds do, the greenhouse effect will not decrease significantly due to the substitution of a cloud fraction by a cloudless fraction. The increasing concentration of water vapour in the marine troposphere [36] results in a stronger absorption of longwave radiation [25,40,41,42,43]. The vertically distributed opacity of the clear-sky troposphere results in an effective radiating height of roughly 5000 m at 500 hPa, where the temperature is around 255 K [44]. According to Figure 7, the infrared opacity of the troposphere is between 70% and 85%. A rule-of-thumb estimate [25] for the tropical marine infrared absorption coefficient of 71% is consistent with that range. These values are relatively close to those of the opacity of clouds. “The clear-sky infrared absorption/emission is very important, so ideally the assumed value for the clear sky is calculated using a radiative transfer model driven by reanalysis fields. Cloud radiative effect (as shown in Figure 6) is the difference between the all-sky observed and this modeled cloudless atmosphere” (Coda Phillips, priv. comm.).
(ii)
The SW CRE is relevant at daytime only, while the LW CRE acts all day and night. It is unclear whether the reported reduction in global cloudiness differs between day and night [45]. In that case, it may have distinct impacts on SW and LW CRE.
(iii)
Cloudiness is most pronounced in the tropics and the west-wind belts (Figure 3). Moreover, the SW CRE is most relevant at low latitudes, while the LW CRE acts all over the globe. It is unclear how the reported reduction in global cloudiness is correlated with latitude and it may have different impacts on SW and LW CRE.
(iv)
Through the LCL, the increasing ocean SST has an effect on the altitude of cloud formations. This changes the cloud base’s temperature and, in turn, its downward thermal radiation. This feedback effect is analysed thermodynamically in Section 4 and Section 5 of this paper. In addition, low-level cumulus cloud formation is highly correlated with the diurnal cycle of solar irradiation, latitude, and land–ocean distribution.
(v)
Through the LCL, a so far unnoticed minor increase in ocean surface RH may have an effect on the altitude of cloud formation. This could change the cloud base temperature and, in turn, cause its downward thermal radiation to be in the opposite direction compared to the SST trend. This negative feedback effect is briefly quantified thermodynamically in Section 4.2.

3. Mathematical Method: TEOS-10 Equations of State

The world’s ocean is a vast and dynamically changing store of heat. However, the theoretical description of its heat transport and conversion processes is thermodynamically challenging [12,14,15,46,47,48,49]. For example, the former 1980 Equation of State of Seawater (EOS-80) did not provide any equations for the calculation of seawater entropy or enthalpy [50,51]. In order to meet the growing demands of climate research and to support numerical circulation models of the ocean, starting in 2005, an improved comprehensive standard description of seawater thermodynamics has been developed by the SCOR/IAPSO Working Group 127 in cooperation with the International Association for the Properties of Water and Steam [52].
Proper axiomatic systems of mathematical statements possess the internal properties of independence, consistency, and completeness. In contrast to any earlier collections of empirical seawater property equations, the International Thermodynamic Equation of Seawater—2010 (TEOS-10) is designed in such an axiomatic manner and offers a comprehensive description of thermodynamic properties of water in its geophysical mixtures, phases, and mutual equilibria. Thermodynamics has a reputation of being mathematically difficult. For beginners, though, an encouraging anecdote quoted by Fink [53] is Arnold Sommerfeld’s experience with the use of the thermodynamic formalism: “Thermodynamics is a funny subject. The first time you go through it, you don’t understand it at all. The second time you go through it, you think you understand it, except for one or two points. The third time you go through it, you know you don’t understand it, but by that time you are so used to that subject, it doesn’t bother you anymore”. Appendix A provides a brief tutorial introduction to the mathematics used throughout TEOS-10. However, extensive software libraries are available from the web (www.teos-10.org, accessed on 10 March 2024), as outlined in the digital supplement of Wright et al. [54]. These conveniently provide direct access to numerical properties without detailed expertise being required regarding the sometimes-demanding algorithms behind these calculations [54,55].
More than 150 years ago, in 1873, Willard Gibbs discovered that a single mathematical function is sufficient to describe all thermodynamic properties of a given substance in its equilibrium state. Such functions are known as “thermodynamic potentials”. At its very core, TEOS-10 is mathematically defined [3] in terms of empirical correlation equations for just three thermodynamic potentials:
(i)
The specific Gibbs energy of seawater, g S W S , T , p , which is a function of absolute salinity, S , absolute temperature, T , and absolute pressure, p [56,57,58];
(ii)
The specific Gibbs energy of ambient hexagonal ice Ih, g I h T , p , which is a function of temperature and pressure [59,60];
(iii)
The specific Helmholtz energy of humid air, f A V A , T , ρ , which is a function of the dry-air mass fraction, A , and the mass density of humid air, ρ [61,62]. From these potential functions, all thermodynamic properties of seawater, ice, and humid air, as well as their mutual equilibria, can be derived mathematically in a perfectly consistent way via analytical or numerical means [39,63,64].
“Helmholtz energy” is sometimes also known as “free energy”, and “Gibbs energy” is sometimes referred to as “free enthalpy” in the literature. In TEOS-10, temperature is expressed on the 1990 International Temperature Scale [65]. In the vicinity of the triple point of water, ITS-90 deviates only insignificantly from the current Thermodynamic Temperature Scale [66] and will remain in practical use for the foreseeable future [67]. The isotopic composition of water is assumed to be that of Vienna Standard Mean Ocean Water (VSMOW) [68]. Salinity, S in IAPSO Standard Seawater, is defined by the Reference-Composition Salinity Scale [57] as the mass fraction of dissolved salt in seawater. The dry-air mass fraction A of humid air is related to the specific humidity q = 1 A , the mixing ratio r , and the water vapour mole fraction x by
1 r = 1 q 1 = 1 A 1 1 = M A M W 1 x 1 .
Here, M W = 18.015268   g m o l 1 and M A = 28.96546   g m o l 1 , respectively, are the molar masses of water and of dry air (IAPWS G08-10 2010) [69]. The freely adjustable constants for the absolute energies and entropies of water, sea salt, and dry air are consistently defined at reference states, for which the triple point of water and the standard ocean state are chosen [39,55,56,58,62,64,70].
The Gibbs function of freshwater, g S W 0 , T , p , equals the Gibbs function of liquid water derived from the specific Helmholtz energy, f F T , ρ , of fluid water [56]. This evaluated at high densities of liquid water via
g S W 0 , T , p g W T , p = f F T , ρ + p ρ ,
where pressure is given by
p = ρ 2 f F ρ T .

3.1. Ocean Heat Content

The specific entropy of seawater, η S W , is available from the Gibbs function by
η S W = g S W S , T , p T S , p ,
and the specific enthalpy of seawater, h S W , is available by
h S W S , η S W , p = g S W S , T , p + η S W T .
This way, the ocean heat content, O H C , is estimated from the integral [15] via
O H C = h S W S , η S W , p 0 h S O S W ρ S W S , η S W , p d V .
This is carried out using the ocean’s volume and its local in situ properties S , η S W S , T , p , p . By definition [3,58,64,70] the TEOS-10 enthalpy of the standard-ocean reference state is h S O S W 0 . The local in situ density ρ S W S , η S W , p is available from the relation
1 ρ S W = h S W p S , η S W .
Formula (9) describes a fictitious process by which each mass element, d m = ρ S W d V , in the volume V is isentropically lifted to the surface pressure, p 0 , where it transfers all its excess enthalpy, h S W S , η S W , p 0 h S O S W , to an external measurement device via sensible heat flux. Subsequently, the same amount of heat is reversibly conducted back, after which the parcel returns to its original spatial position. Then, O H C is the total amount of heat reversibly removed from the ocean across its surface by this process [39].
The approach of Equation (9) is therefore consistent with the fact that thermodynamically, heat is an exchange quantity rather than a state quantity, so that, rigorously speaking, any kind of stored “heat content” does not unambiguously exist, but is definite only as an amount of heat which is exchanged between a given system and its surroundings [71]. According to Clausius [72] and Gibbs [73],
d h S W = T * d η S W = c p S W d T *
is the proper thermodynamic expression for heat exchange at constant pressure and constant amounts of water and salt. Here, T * is the parcel’s changing surface temperature during the fictitious heat transfer process, and c p S W is its isobaric heat capacity. While the enthalpy difference d h S W is always an unambiguous state quantity, the total heat T * d η S W is ambiguous and depends on the particular heat exchange process performed along the integration path. Heat output, for example, may result from work input, or vice versa. “The obsolete hypothesis of heat being a substance is excluded” [74].

3.2. Humid Air

The Helmholtz function of dry air, f A V 1 , T , ρ , equals (up to modified reference state conditions) the equation of state of Lemmon et al. (2000) [61],
f A V 1 , T , ρ f A T , ρ .
The Helmholtz function of pure water vapour, f A V 0 , T , ρ , equals the specific Helmholtz energy, f F T , ρ , of fluid water [56], evaluated at low densities of water vapour
f A V 0 , T , ρ f F T , ρ .
For the following calculations of condensation levels, the specific entropy of humid air, η , is available from the related Gibbs function
g A V A , T , p = f A V A , T , ρ + ρ f A V ρ A , T
by
η = g A V A , T , p T A , p .
Note that, derived from Equation (14) for atmospheric applications, a simplified analytical low-pressure formulation for the Gibbs function g A V is available from Appendix B of the work of Feistel and Hellmuth (2023) [29].
The specific enthalpy of humid air is computed from the Gibbs function as
h A V A , η , p = g A V A , T , p + η T .

3.3. Relative Fugacity

For the computation of phase equilibria such as dew or frost points, TEOS-10 provides the chemical potentials of water in the different phases or mixtures.
The chemical potential of water vapour, μ V A V , in humid air is available from the Gibbs function by
μ V A V = g A V A g A V A T , p ,
or equivalently from the enthalpy by the following (also see Equation (A26))
μ V A V = h A V η h A V η A , p A h A V A η , p .
The chemical potential of liquid water equals its specific Gibbs energy,
μ W W = g W T , p ,
and, similarly, the chemical potential of ice is,
μ W I h = g I h T , p .
The dry-air mass fraction, A s a t T , p , of humid air saturated with respect to liquid water is defined by the equilibrium condition
μ V A V A s a t , T , p = μ W W T , p ,
and is defined with respect to ice by
μ V A V A s a t , T , p = μ W I h T , p .
To make the use of TEOS-10 equations more convenient, the numerical solutions of Equations (21) and (22) for either A s a t T , p , T A s a t , p , or p A s a t , T , together with numerous other properties (Feistel et al. 2010a; Wright et al. 2010), are implemented in the sea–ice–air (SIA) library of TEOS-10, freely available from the open source code and accessible at www.teos-10.org, accessed on 10 March 2024.
After the official adoption of TEOS-10, the IAPSO/SCOR/IAPWS Joint Committee on the Properties of Seawater (JCS) continued to address pending climate-related problems [75] that had not yet or only insufficiently been addressed by TEOS-10. Among those issues was the ambiguity and mutual inconsistency of different definitions of RH in practical use, such as the discrepancy between meteorology and climatology [34]. With respect to deviations from water phase equilibria such as saturation properties, the real-gas equivalent of conventional RH is the relative fugacity (RF) (Feistel et al. 2010b [62]; Equation (10) in Feistel and Lovell-Smith 2017 [76])
ψ f A , T , p exp μ V A V A , T , p μ V A V A s a t T , p , T , p R W T .
Here, R W = R / M W is the specific gas constant of water, and R is the universal molar gas constant. As a function of T , p , the chemical potential at saturation, μ V A V A s a t , T , p , may be expressed by means of either Equation (21) or (22). Note that Equation (23) is only valid for temperatures below the boiling point of water at a given pressure (Feistel and Lovell-Smith 2017) [76], a condition which is naturally fulfilled under ambient geophysical conditions.
In ideal gas approximation, RF is identical to conventional RH in terms of water vapour partial pressures. For air that is close to saturation, such as an amarine surface layer with about 80 %rh, the Clausius–Clapeyron formula is an excellent approximation of TEOS-10 values for RF, even if it refrains from the perfect gas assumption [77]. RF appears naturally in expressions for the Onsager driving force of non-equilibrium fluxes, such as evaporation from the ocean surface [29,33]. For the numerical computation of RF, a source code extension to the SIA library is available from Feistel et al. (2022) [77].
Mathematical transformation between different thermodynamic quantities and partial derivatives with respect to alternative independent variables, as is frequently exploited in TEOS-10 to derive the desired functional dependencies, may conveniently and error-free be executed by means of the formal Jacobi method developed by Norman Shaw (1935) [78], as briefly described in Appendix A.

4. Model Results: Isentropically Lifted Condensation Level (LCL)

“Synoptic weather observations from ships throughout the World Ocean have been analyzed to produce a climatology of total cloud cover and the amounts of nine cloud types. Among the cloud types, the most widespread and consistent relationship is found for the extensive marine stratus and stratocumulus clouds” [79]. Subtropical marine clouds show a pronounced “transition between unbroken sheets of stratocumulus and fields of scattered cumulus” [80]. The cumulus cloud base is typically observed below 500 m height, as shown in Figure 1. For the climatic feedback effect of marine clouds, the lifted condensation level (LCL) is a key parameter that describes the pressure (or altitude) at which an isentropically ascending parcel reaches its dew or frostpoint, starting from given values of temperature and RH, such as at the sea surface. A typical empirical equation for the LCL height, z L C L , is
z L C L z = γ L C L T T d p ,
where z , T and T d p , respectively, are the initial values of the parcel’s height, in situ temperature, and dew-point temperature. Commonly, the latter is related to the saturation vapour pressure and the RH via the Clausius–Clapeyron formula. Previous estimates for the coefficient γ L C L range widely between 123   m K 1 and 165   m K 1 [81], with an optimum of 125   m K 1 suggested by Lawrence (2005) [82]. Also, nonlinear mathematical expressions, more complicated than Equation (24), were suggested in the literature and derived from varying approximate thermodynamic relations for humid air, liquid water, and ice. Often, such relations are chosen according to the authors’ personal preferences and are scarcely supported by internationally agreed standard formulations. In Section 4.2 below, for typical values of the marine troposphere, the highly accurate TEOS-10 results selected for γ L C L range between 125 and 129 m K 1 .
Humid air at the sea surface has a typical climatological RH of about 80 %rh, a value that may vary seasonally or regionally between 70 %rh and 90 %rh, but appears likely to be only very weakly affected by global warming, if at all [8,30,31,83,84,85]. Assuming that an air parcel starts ascending at a constant specific humidity, A = A 0 = c o n s t , from the sea surface at a given air pressure, p 0 , and a certain sea surface temperature, T 0 , the value of the initial air fraction A 0 can be computed in implicit form from the sea surface relative fugacity, ψ f = ψ f A 0 , T 0 , p 0 , using Equation (23):
μ V A V A 0 , T 0 , p 0 = g W T 0 , p 0 + R W T 0 ln ψ f .
Here, the surface values of independent variables are indicated by the subscript 0 which may be given at constant A for simplicity. Similarly, the subscript LCL is used to assess properties at the lifted condensation state. Associated function values such as ψ f or g W are without state-dependent subscripts as long as there is no risk of confusion.
The tropospheric entropy, assumed to remain constant during the uplift, can be expressed by Equation (15),
η = η 0 = g A V A , T 0 , p 0 T A , p .
At the LCL, the entropy value,
η = η L C L = g A V A , T L C L , p L C L T A , p ,
needs to be the same as at the surface, Equation (26), and may be eliminated from the mathematical problem by equating Equation (26) with (27),
g A V A , T 0 , p 0 T A , p = g A V A , T L C L , p L C L T A , p .
The LCL pressure, p L C L , is approached when, at fixed values of A , η = A 0 , η 0 , the temperature drops to the dew point, T L C L . This is determined using Equation (21):
μ V A V A , T L C L , p L C L = g W T L C L , p L C L .
The chemical potential of water vapour in humid air, μ V A V , may be expressed by the related Gibbs function, Equation (17),
g A V A , T L C L , p L C L A g A V A , T L C L , p L C L A T , p = g W T L C L , p L C L .
If the initial value of RH is given, rather than the initial specific humidity, q = 1 A , then Equation (25) may be added to the system in the form of
g A V A , T 0 , p 0 A g A V A , T 0 , p 0 A T , p = g W T 0 , p 0 + R W T 0 ln ψ f .
Equations (28), (30) and (31) constitute a closed system of three nonlinear implicit equations for determining three unknowns A , T L C L , p L C L as functions of the given input values ψ f , T 0 , p 0 . These thermodynamic relations are the exact conditions for the LCL properties, with the only minor approximation performed because the dissolution of air into liquid water, influencing the saturation state, is neglected. This system may be solved iteratively using the TEOS-10 functions implemented in the SIA library, see Appendix B, or it may be exploited analytically after introducing quantitatively reasonable approximations.

4.1. Numerical Iterative Solution

The system of Equations (28) and (30), used for the calculation of T L C L , p L C L from A ; Equation (31), used for the determination of A from the given surface values of T 0 , p 0 ; and the relative fugacity can be solved iteratively, offering the best accuracy currently available. Expanding Equation (31) into a power series up to linear terms in the increment, A , of a starting estimate, A ,
g A V A + A , T 0 , p 0 A + A g A V A + A , T 0 , p 0 A T , p = g W T 0 , p 0 + R W T 0 ln ψ f
leads to the solution for the improvement, A i + 1 = A i + A i , of iteration step i ,
A i = g A V A i , T 0 , p 0 A i g A V A i , T 0 , p 0 A T , p g W T 0 , p 0 R W T 0 ln ψ f A i 2 g A V A i , T 0 , p 0 A 2 T , p .
Practically, this Newton–Cotes iteration may start from almost dry air, such as A 0 = 0.999999 .
In a similar manner, the other increments T L C L , p L C L follow on from Equations (28) and (30) and can be given in matrix notation as
a 11 a 12 a 21 a 22 T L C L p L C L = b 1 b 2 .
By Cramer’s rule, the solution is
T L C L = b 1 a 12 b 2 a 22 a 11 a 12 a 21 a 22 = b 1 a 22 b 2 a 12 a 11 a 22 a 21 a 12
and
p L C L = a 11 b 1 a 21 b 2 a 11 a 12 a 21 a 22 = a 11 b 2 a 21 b 1 a 11 a 22 a 21 a 12 .
The coefficients of this system are as follows:
a 11 = 2 g A V A , T L C L , p L C L T 2 A , p ,
a 12 = 2 g A V A , T L C L , p L C L T p A , p ,
a 21 = g A V A , T L C L , p L C L T A , p A 2 g A V A , T L C L , p L C L A T p g W T L C L , p L C L T p ,
a 22 = g A V A , T L C L , p L C L p A , T A 2 g A V A , T L C L , p L C L A p T g W T L C L , p L C L p T ,
b 1 = g A V A , T 0 , p 0 T A , p g A V A , T L C L , p L C L T A , p ,
b 2 = g W T L C L , p L C L g A V A , T L C L , p L C L + A g A V A , T L C L , p L C L A T , p .
All derivatives utilized here are numerically available from the TEOS-10 SIA library [54].
Table 1 reports selected LCL states computed from TEOS-10, iteratively solving Equations (34)–(36) at p 0 = 1013.25   h P a and ψ f = 80   % r h . The results are rounded to 6 digits. The climatic warming trend moves LCL to higher altitudes ( p L C L ), where the temperature difference between ocean ( T 0 ) and cloud base ( T L C L ) becomes larger and consequently, the downward radiative heat flux becomes reduced compared to (fictitious) lower clouds. This cooling effect is a negative, stabilising feedback of the cloud cover, slowing down ocean warming. The observed shrinking cloudiness, in turn, likely renders the cooling influence of clouds globally smaller.

4.2. Linear Analytical Approximation

If the initial temperature, T 0 , is only slightly higher than the dew point, T d p , as is typical at the ocean surface, the LCL pressure may be estimated from TEOS-10 equations that are linearised based on small differences as T L C L T 0 and p L C L p 0 . Performing this, the relative fugacity may be very accurately expressed by the Clausius–Clapeyron equation for the dew point (Feistel and Hellmuth 2022: Equation (14) therein) [77]
R W T 0 ln ψ f L L T d p , p 0 1 T 0 T d p .
Here, L L is the specific evaporation enthalpy of liquid water at an equilibrium with (saturated) humid air. Its value is available from the thermodynamically rigorous relations (Feistel et al. 2010b; Feistel and Hellmuth 2022: Equations (C.5) and (C.6) therein) [62,77],
L L T , p T = η A V A s a t T , p , T , p A s a t T , p η A V A T , p η W T , p ,
which is equivalent to
L L T , p T = A s a t T , p A s a t T p 2 g A V A 2 T , p .
Here, η A V and η W , respectively, are the specific entropies of humid air and liquid water.
Equation (43) represents the initial linear term of a truncated Taylor series with respect to powers of T 0 T d p , regardless of whether ideal gas assumptions have been applied or not.
A similar series expansion of the Gibbs function of liquid water in Equation (30) with respect to T T L C L T 0 and p p L C L p 0 gives
g W T L C L , p L C L g W T 0 , p 0 + g W T 0 , p 0 T p T + g W T 0 , p 0 p T p .
The expansion coefficients are the specific entropy, η W , and the specific volume, v W , of liquid water at T 0 , p 0 ,
g W g W T L C L , p L C L g W T 0 , p 0 η W T + v W p .
A similar expansion about T 0 , p 0 can be carried out for the Gibbs function of humid air,
g A V g A V A , T L C L , p L C L g A V A , T 0 , p 0 η A V T + v A V p .
Subtracting Equation (31) from (30) results in
g A V A g A V A T , p g W = R W T 0 ln ψ f ,
which may by expressed in terms of the linear offsets (47), (48), as
η A V A η A V A T , p η W T v A V A v A V A T , p v W p = R W T 0 ln ψ f .
Up to small linear terms in the subsaturation, A A s a t T 0 , p 0 , at the surface, the coefficient of T equals the specific evaporation entropy, Equation (44),
L L T d p , p 0 T d p = η L T d p , p 0 η A V A , T 0 , p 0 A η A V A T , p η W T 0 , p 0
and the term in front of p equals approximately the specific evaporation volume [62],
v L T d p , p 0 v A V A , T 0 , p 0 A v A V A T , p v W T 0 , p 0 .
Concerning linear approximation with respect to the LCL pressure and temperature offset, Equation (50) then reads
L L T d p , p 0 T d p T v L T d p , p 0 p = R W T 0 ln ψ f .
During adiabatic uplift, temperature and pressure change are mutually related by the moist lapse rate, Equation (A18),
T = Γ p .
Similar to Equation (24), the final formula for the LCL pressure, Formula (53), therefore also makes use of Equation (43),
p L C L p 0 = L L T d p , p 0 1 T 0 T d p L L T d p , p 0 T d p Γ v L T d p , p 0 = T 0 T d p Γ T d p , p 0 + Γ d p T d p , p 0 .
Here, the lapse rate of the dew point is negative, as shown in Appendix B, Equation (A31),
Γ d p = T d p p A = T d p v L L L T d p , p .
The hydrostatic balance equation (height z , gravity constant g E = 9.81   m   s 2 ),
v A V d p = g E d z ,
can be integrated along an isentropic vertical profile, due to the thermodynamic relation,
v A V = h A V p A , η ,
to give
z L C L z 0 = 1 g E h A V A , η , p L C L h A V A , η , p 0 .
This enthalpy difference may be used to estimate the coefficient of the LCL height, Equation (24), from the LCL pressure, through
γ L C L z L C L z 0 T 0 T d p .
Rigorous numerical TEOS-10 results for z L C L z 0 and γ L C L from Equations (59) and (60) at selected typical marine conditions, obtained through iteratively solving Equations (34)–(36), are reported in Table 2 and Table 3.
A linear approximation based on the LCL pressure, Equation (55), is performed as follows:
γ L C L v A V g E p L C L p T T d p v A V g E Γ + Γ d p .
In Table 3, in addition to the calculated LCL coefficients, the climatic feedback to possible RH trends is also reported as the thermal radiation flux, σ S B T L C L 4 , downward from the cloud base, which tends to warm the ocean if RH is rising.
If the specific volume of humid air in Equation (52) is much larger than the remaining terms of the evaporation volume, v L v A V , the simplified LCL coefficient given by Equation (61), expressing the lapse rate by Equation (A18), is Γ = α A V v A V T / c p A V . In terms of the thermal expansion coefficient, α A V , and the specific isobaric heat capacity, c p A V , of humid air, by virtue of Equation (56) and T T d p , we state that
γ L C L g E T d p α A V c p A V 1 L L 1 .
In ideal gas approximation, α A V 1 / T d p , this becomes
γ L C L g E 1 c p A V 1 η L 1 .
Here, η L = L L / T d p is the evaporation entropy of liquid water, as shown in Equation (51).
Analytical low-pressure approximations for the TEOS-10 Gibbs functions of liquid water and humid air are provided by Feistel and Hellmuth (2023: Appendix A therein) [29] and are given in an extremely simplified, crude form in Appendix C of this paper. These methods are successfully used in the following section.

4.3. Clausius–Clapeyron Expansion

In the context of global warming, the Clausius–Clapeyron equation for the saturated vapour pressure of water is a widely exploited and sufficiently accurate approximation method for estimating the increasing humidity of the troposphere. Here, an expression for the LCL is derived with a similar accuracy from the TEOS-10 equations.
Equations (28), (30) and (31) constitute a closed system of three nonlinear implicit equations for the three unknowns A , T L C L , p L C L as functions of the given input values ψ f , T 0 , p 0 . These relations may be formulated in terms of “crude” approximations, taking all heat capacities as constant, all liquids as incompressible, and all gases as perfect gases. The related crude Gibbs functions, g ~ W T , p and g ~ A V A , T , p , are indicated here by tildes and are provided in Appendix C.
Using those Gibbs functions, the conservation of entropy during adiabatic ascent, shown in Equation (28), is given by
g ~ A V A , T 0 , p 0 T A , p = g ~ A V A , T L C L , p L C L T A , p .
This takes the form of the conventional adiabatic equation of state,
c ~ p A V ln T 0 T L C L = R A V ln p 0 p L C L .
Here, c ~ p A V A 1 A c ~ p V + A c ~ p A and R A V A 1 A R W + A R A , respectively, are the heat capacity and the specific gas constant of humid air.
The saturation of air at the LCL, Equation (30), expresses the equation chemical potentials of water in the gas and the liquid phase,
g ~ A V A , T L C L , p L C L A g ~ A V A , T L C L , p L C L A T , p = g ~ W T L C L , p L C L .
In crude approximation, this condition reads
c ~ p W c ~ p V T L C L ln T L C L T t + R W T L C L ln x p L C L p t = g ~ 0 W + g ~ 1 W T L C L + p L C L ρ ~ W .
Here, T t and p t , respectively, are temperature and pressure at the triple point of water, and ρ ~ W is the density of liquid water. They appear in this context because the triple point is used as the reference state at which the arbitrary constants g ~ 0 W and g ~ 1 W are adjusted to the phase equilibrium properties of fluid water, computed from TEOS-10, as shown in Appendix C.
Finally, the relative fugacity (or RH) at the surface is specified by Equation (31) in terms of crude Gibbs functions
g ~ A V A , T 0 , p 0 A g ~ A V A , T 0 , p 0 A T , p = g ~ W T 0 , p 0 + R W T 0 ln ψ f .
Inserting the crude equations given in Appendix C, results in
c ~ p W c ~ p V T 0 ln T 0 T t + R W T 0 ln x p 0 ψ f p t = g ~ 0 W + g ~ 1 W T 0 + p 0 ρ ~ W .
From the latter equation, the vapour mole fraction at the surface, x ψ f , T 0 , p 0 , is obtained from the relative fugacity, ψ f , by
x ψ f , T 0 , p 0 = ψ f p t p 0 exp g ~ 0 W + g ~ 1 W T 0 R W T 0 + p 0 R W T 0 ρ ~ W c ~ p W c ~ p V R W ln T 0 T t .
From x , the dry-air mass fraction A is available through the use of Equation (4).
Subtracting Equation (67) from (69) and inserting entropy conservation (65) as
ln p L C L p t = ln p 0 p t + c ~ p A V R A V ln T L C L T 0 ,
the saturation Equation (68) can be written in the form of
Q ln T 0 T L C L ln ψ f = g ~ 0 W R W 1 T 0 1 T L C L + 1 ρ ~ W R W p 0 T 0 p L C L T L C L .
Here, the abbreviation
Q c ~ p W c ~ p V R W + c ~ p A V R A V
is introduced. The commonly used Clausius–Clapeyron equation for the saturation vapour pressure makes use of the fact that far from the critical point of water, the liquid density is much higher than the vapour density. This may be quantified by saying that the dimensionless parameter
ε = p 0 R W T 0 ρ ~ W 1
is a small number. The leading term of the perturbation expansion
T L C L = T L C L 0 + ε T L C L 1 + ,
applied to Equation (72) may be regarded as an Clausius–Clapeyron approximation,
Q ln T 0 T L C L 0 ln ψ f = g ~ 0 W R W 1 T 0 1 T L C L 0 .
Equation (76) may be rearranged in the form
B T 0 T L C L 0 exp B T 0 T L C L 0 = E
with the abbreviations,
B g ~ 0 W Q R W T 0
E B exp B + ln ψ f Q .
The solution of the nonlinear Equation (77) is
T L C L 0 = B w E T 0 .
Here, the function w y is the well-defined Lambert W function, as shown in Appendix D, which is one of the two solutions of the equation
w exp w = y .
Because y = E is negative, such as E = 0.174718 at T 0 = 292   K , solutions for the lower branch, w y < 1 , are chosen here,
T L C L 0 = 287.658   K .
Eventually, by means of Equation (71), the Clausius–Clapeyron approximation for the LCL pressure is
p L C L 0 = p 0 exp c ~ p A V R A V ln B w E .
In fact, these “crude” Clausius–Clapeyron approximations (80) and (83) already agree quite well with the rigorous iterative TEOS-10 results of Equation (34), see Table 4.

5. Model Results: Marine Climatic LCL Feedback

Refraining from discussing the complexity of spectrally resolved greenhouse gas attenuation, as shown in Figure 5 and in proper 3D radiation propagation models [42,86,87], low-level marine clouds as well as the ocean surface may approximately be described as planar black bodies, mutually interacting according to the Stefan–Boltzmann law of thermal radiation, as shown in Figure 8.
Denoting the Stefan–Boltzmann constant as σ S B = 5.670374419 × 10 8   W   m 2   K 4 , the net radiative heat loss, J E , of the cloud-covered ocean is given in a 1D model:
J E = σ S B T 0 4 T L C L 4 .
Its sensitivity with respect to ocean warming is determined as follows:
J E T 0 p 0 , ψ f = 4 σ S B T 0 3 T L C L 3 T L C L T 0 p 0 , ψ f ,
assuming that the marine surface RH remains unaffected by the SST change. For the LCL values listed in Table 1, the upward and downward radiative fluxes, J = σ S B T 0 4 and J = σ S B T L C L 4 , respectively, as well as estimated sensitivities, as in Equation (85), were calculated and are shown in Table 5.
The mathematical dependence of T L C L on T 0 is complicated, even in the case of crude approximations, as shown in Equation (80). In this section, a rigorous theoretical expression for this dependence is derived systematically from the TEOS-10 approach. As an aside, it should be noted that this longwave radiative interaction is an irreversible process with an entropy production rate [88,89,90] of
P = J E 1 T L C L 1 T 0 .
The latest assessments [30] have shown that at the ocean surface, current climate change is characterised by an increase in SST along with a constant surface RH, or similarly, it is characterised in the TEOS-10 formalism by a constant relative fugacity, ψ f . As a dimensionless quantity, a suitable thermodynamic expression for describing a related climatic trend of the LCL greenhouse effect is the sensitivity of the cloud base temperature, T L C L T 0 , p 0 , ψ f , with respect to an increase in the surface temperature, T 0 ,
β T L C L T 0 p 0 , ψ f .
This LCL temperature sensitivity controls the sensitivity of the ocean–cloud heat flux balance, Equation (85), with respect to rising SST. A rigorous formula for this sensitivity may be derived from the set of LCL equations that describe entropy conservation during uplift, as in Equation (28), where
f 1 g A V A , T 0 , p 0 T A , p g A V A , T L C L , p L C L T A , p = 0 .
Air saturation at the cloud base, as in Equation (30), is determined by
f 2 g A V A , T L C L , p L C L A g A V A , T L C L , p L C L A T , p g W T L C L , p L C L = 0 ,
and the given relative humidity at the sea surface, as in Equation (31), is determined by
f 3 g A V A , T 0 , p 0 A g A V A , T 0 , p 0 A T , p g W T 0 , p 0 R W T 0 ln ψ f = 0 .
During the climatic increase in the SST, T 0 , at a constant RH, ψ f , and surface pressure, p 0 , these equations always remain fulfilled for i = 1 , 2 , 3 ,
f i 0 , d f i d T 0 0 ,
so that other properties, namely, A , T L C L , p L C L , may vary along with T 0 . Taking the partial derivatives of the functions f 1 , f 2 and f 3 with respect to T 0 , while considering the dependence of A , T L C L , p L C L on T 0 , a set of three equations is obtained for their three unknown sensitivities with respect to ocean warming:
α A T 0 p 0 , ψ f ,   β ,   and   γ p L C L T 0 p 0 , ψ f .
These derivatives are mutually related by Equations (91),
f i A T 0 , T L C L , p L C L α + f i T L C L A , T 0 , p L C L β + f i p L C L A , T 0 , T L C L γ + f i T 0 A , T L C L , p L C L = 0 .
In matrix notation, this is
a 11 a 12 a 13 a 21 a 22 a 23 a 31 0 0 α β γ = b 1 0 b 3 .
By Cramer’s rule, the solution of this system for the LCL sensitivity β is straight forward:
β = a 11 b 1 a 13 a 21 0 a 23 a 31 b 3 0 a 11 a 12 a 13 a 21 a 22 a 23 a 31 0 0 = b 1 a 31 a 23 b 3 a 11 a 23 a 21 a 13 a 31 a 12 a 23 a 22 a 13 .
The coefficients of this system of linear equations are as follows
a 11 = 2 g A V A , T 0 , p 0 A T p 2 g A V A , T L C L , p L C L A T p ,
a 12 = 2 g A V A , T L C L , p L C L T 2 A , p = c p A V A , T L C L , p L C L T L C L ,
a 13 = 2 g A V A , T L C L , p L C L T p A ,
a 21 = A 2 g A V A , T L C L , p L C L A 2 T , p ,
a 22 = g A V A , T L C L , p L C L T A , p A 2 g A V A , T L C L , p L C L A T p g W T L C L , p L C L T p = L L T L C L , p L C L T L C L ,
a 23 = g A V A , T L C L , p L C L p A , T A 2 g A V A , T L C L , p L C L A p T g W T L C L , p L C L p T = v L ,
a 31 = A 2 g A V A , T 0 , p 0 A 2 T , p ,
b 1 = 2 g A V A , T 0 , p 0 T 2 A , p = c p A V A , T 0 , p 0 T 0 ,
b 3 = g A V A , T 0 , p 0 T A , p + A 2 g A V A , T 0 , p 0 A T p + g W T 0 , p 0 T p + R W ln ψ f .
All partial derivatives required here, in Equation (96) through to Equation (104), are explicitly numerically available from the TEOS-10 SIA library. Reported in Table 1, the LCL sensitivities computed from Equation (94) are the rigorous figures that are close to the previous estimates shown in Section 4. A value of β < 1 means that T L C L is rising slower than T 0 , so that the net radiation balance between sea surface and cloud base is distorted in favour of the upward heat flux, reducing the ocean’s warming rate.
The cloud base is cooling relative to global SST warming, as shown in Figure 7, and the net upward radiative heat flux of about 24 W m−2 is intensifying by 0.45 W m−2 per °C of SST as negative feedback to the warming ocean by currently about 1 W m−2 [9]. This feedback does not provide a possible explanation for the warming; on the contrary, it worsens the problem. While LCL feedback does not dominate the sea–air energy transfer, it is large enough anyway to not be ignored. The effect is about 100 times as large as the rate of atmospheric warming.
For tutorial reasons or first estimates, the coefficients of Equation (94) may be expressed as crude approximations of the Gibbs functions, shown in Appendix C. These are as follows:
a ~ 11 = c ~ p V c ~ p A ln T 0 T L C L + R A R W ln p 0 p L C L ,
a ~ 12 = c ~ p A V T L C L ,
a ~ 13 = R A V p L C L ,
a ~ 21 = R A R W 1 A R A V T L C L ,
a ~ 22 = L L T L C L ,
a ~ 23 = R W T L C L p L C L 1 ρ ~ W ,
a ~ 31 = R A R W 1 A R A V T 0 ,
b ~ 1 = c ~ p A V T 0 ,
b ~ 3 = g ~ 1 W + c ~ p V c ~ p W ln T 0 e T t R W ln x p 0 ψ f p t .
With either the TEOS-10 SIA library functions, or even with the simplified equations of Appendix C, the calculation of the LCL sensitivity, shown in Equation (95), is straightforward, but does likely need to be carried out numerically.

6. Discussion

Climate research relies on observation and causal prediction models. Causality itself cannot be observed, but is the most reliable prediction tool [91,92,93,94]. The comparison of predicted phenomena with future observations serves as a validity criterion for causal models [95,96]. In 2023, the global ocean heat content (OHC) increased by 1.5 × 10 22   J , a value 25 times as large as the world’s total human energy consumption [9], and even exceeded the atmospheric heat gain by at least 100 times [11]. This currently observed excessive OHC increment [13], rising at a rate of 1.3 W m−2, remains largely unexplained so far [16,30] and raises questions regarding the correctness, completeness, and internal consistency of climate models.
The dominating input–output processes of the ocean heat balance include solar shortwave and tropospheric longwave irradiation; thermal radiation from the ocean surface; latent heat exchange by phase transitions, including water vapour and ice; and sensible heat exchange with the atmosphere and lithosphere. The input and output processes each amount to a total of about 500 W m−2, and the observed imbalance arises from a small mutual mismatch by just 0.2% between those. Evidently, to reasonably explain a small difference of 1 W m−2 between two big numbers of about 500 W m−2, each of those must be known to within an uncertainty of below 1 W m−2. A key question is how this small amount may selectively be attributed to the various individual physical contributions involved. Unfortunately, at the ocean–atmosphere interface, current climate models exhibit uncertainties larger than 10 W m−2 and cannot reliably resolve the requisite heat flux differences.
TEOS-10, the Thermodynamic Equation of Seawater—2010, provides the most accurate and mutually consistent equations currently available for assessing the thermodynamic properties of seawater, humid air, and ice. It was adopted as an international geophysical standard by IOC/UNESCO in 2009 and by IUGG in 2011 to support climate research.
Current numerical climate models tend to underestimate ocean warming [28]. They typically implement Dalton equations to estimate evaporation rates [32] in a historic form, which understands increasing evaporation as a consequence of rising temperature and, in turn, of the vapour pressure of seawater. Quantitatively underpinned by TEOS-10, irreversible thermodynamics, however, suggests that constant marine relative humidity (RH) is the driving force of evaporation, rather than increasing vapour pressure [29,33,39]. A putative tiny increase in global mean RH, even below instrumental resolution, may suffice to explain the rate of ocean warming; this renders latent heat as the preferred candidate responsible for the observed OHC imbalance [25].
A second candidate possibly contributing to the OHC increase is the radiation balance, strongly affected by clouds and greenhouse gases. Global cloudiness has systematically been decreasing in the past decades, enhancing oceanic shortwave irradiation by decreasing shadowed areas. At the same time, the downward longwave irradiation from those clouds is also decreasing, with an estimated result that these two effects almost completely cancel one another [26]. However, global warming does not only reduce cloudiness; the trend of higher sea surface temperature (SST) also elevates marine clouds by an increasing the lifted condensation level (LCL), in turn cooling the cloud base and reducing their downward thermal radiation. In order to include such a cooling effect in the OHC balance investigations, in this paper the sensitivity of LCL and of the related ocean–cloud radiation balance is analysed with the TEOS-10 thermodynamic formalism. Rigorous equations are derived for estimating LCL effects in numerical climate models.
The 2023 global mean SST is about 292 K. At this temperature, the upward thermal radiation flux from the sea surface is 412 W m−2 (Table 5), while the downward flux from the LCL cloud base is 388 W m−2. Under LCL cloud cover, the remaining net upward radiation of 24 W m−2 grows by 0.45 W m−2 per one degree of further ocean warming. This LCL feedback effect is relevant in comparison to the 2023 OHC gain of 1.3 W m−2 and may not be ignored in balance investigations. As negative cooling feedback, this LCL effect does not help to explain observed ocean warming; rather, it enables the problematic heating rate to be explained by other contributing effects of the OHC balance, such as a putative uncertain minor climatic change in evaporation.
However, atmospheric RH can be observed to within an uncertainty in the range 1 %rh–5 %rh [34]. An increase in ocean surface RH of 1 %rh is estimated to reduce the latent heat flux of evaporation by about 5 W m−2 [25,29,38,39], thus warming up the ocean at this rate. In addition, a 1 %rh increase may lower the LCL by about 25 m, in turn intensifying the thermal downward radiation flux from the cloud bottom by more than 1 W m−2 (see Table 3). LCL height may serve as a remotely measured, sensitive estimate for sea surface relative fugacity or conventional RH. It appears to be a plausible working hypothesis that the excessive ocean warming of 2023 by 1.3 W m−2 may be caused to a large extent by slightly rising ocean surface RH, not exceeding the uncertainty of observation.

Author Contributions

R.F.: Manuscript idea and draft, elaboration of the theory; O.H.: independent verification of the theory and calculus, discussion and revision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data used are published in the cited literature, such as in open-access IAPWS and TEOS-10 documents.

Acknowledgments

Authors are grateful to Jessica Blunden and Coda Phillips for clarifying details of the published “State of the Climate in 2022”. The contribution of O. Hellmuth was provided within the framework of the research theme 2 “Aerosols and clouds, long-term processes and trends” of Leibniz Institute for Tropospheric Research (TROPOS), Leipzig and is part of the TROPOS activities within the framework of the EU project “Aerosol, Clouds and Trace gases Research InfraStructure” (ACTRIS). This work contributes to the tasks of the IAPSO/SCOR/IAPWS Joint Committee on the Properties of Seawater (JCS).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. The Jacobi Method

Consider N mathematical functions, y 1 , y 2 , y N , depending on N independent variables, x 1 , x 2 , x N . The matrix, A a i j , consisting of the pairwise partial derivatives, a i j = y i / x j , as its elements, is regarded as the Jacobian matrix of this set of functions. The determinant, J A = d e t a i j , of this matrix is commonly referred to as the functional determinant, or the Jacobian [97,98,99]. It is convenient to write J in the following form:
J = d e t y i x j = y 1 , y 2 , y N x 1 , x 2 , x N .
For N = 1 , the Jacobian equals the partial derivative,
J = d e t y i x j = y 1 x 1 = y 1 x 1 .
Determinants of higher dimensions, N , can be evaluated from their Laplace expansion with respect to their minors of dimension, N 1 . For N = 2 , the Jacobian can be reduced to two Jacobians of dimension 1,
J = y 1 , y 2 x 1 , x 2 = y 1 x 1 × y 2 x 2 y 1 x 2 × y 2 x 1 = y 1 x 1 y 2 x 2 y 1 x 2 y 2 x 1 .
For N = 3 , the Laplace expansion takes the form of a linear combination of three Jacobians of dimension 2,
J = y 1 , y 2 , y 3 x 1 , x 2 , x 3 = y 1 x 1 × y 2 , y 3 x 2 , x 3 y 1 x 2 × y 2 , y 3 x 1 , x 3 + y 1 x 3 × y 2 , y 3 x 1 , x 2 ,
in which the two-dimensional Jacobians are subsequently evaluated according to (A3).
Jacobians possess some helpful and convenient properties which simplify their formal manipulations. To bring the variables in a desired sequence, following from the general properties of determinants, the swapping of any pair of variables inverts the sign of the Jacobian, such that
y 1 , y i , y k , y N x 1 , x 2 , x N = y 1 , y k , y i , y N x 1 , x 2 , x N ,
and similarly,
y 1 , y N x 1 , x i , x k , x N = y 1 , y N x 1 , x k , x i , x N .
For N = 2 , this means
y 1 , y 2 x 1 , x 2 = y 2 , y 1 x 1 , x 2 = y 1 , y 2 x 2 , x 1 = y 2 , y 1 x 2 , x 1 .
As special cases for N = 3 , the backward sequence of variables inverts the sign,
y 1 , y 2 , y 3 x 1 , x 2 , x 3 = y 3 , y 2 , y 1 x 1 , x 2 , x 3 = y 1 , y 2 , y 3 x 3 , x 2 , x 1 = y 3 , y 2 , y 1 x 3 , x 2 , x 1 ,
while the rotation of variables preserves the sign,
y 1 , y 2 , y 3 x 1 , x 2 , x 3 = y 3 , y 1 , y 2 , x 1 , x 2 , x 3 = y 2 , y 3 , y 1 x 1 , x 2 , x 3 = y 1 , y 2 , y 3 x 3 , x 1 , x 2 = y 1 , y 2 , y 3 x 2 , x 3 , x 1 .
If one (or more) of the functions is an identity, say, y N x 1 , x 2 , x N x N , the Jacobian reduces by one (or more) dimension due to the Laplace expansion, so that
y 1 , y 2 , y N 1 , x N x 1 , x 2 , x N 1 , x N = y 1 , y 2 , y N 1 x 1 , x 2 , x N 1 .
In particular, the identical Jacobian equals unity,
x 1 , x 2 , x N 1 , x N x 1 , x 2 , x N 1 , x N = 1 .
In higher dimensions, the product rule of functional determinants is a generalisation of the chain rule for usual derivatives. If a set of functions z 1 , z 2 , z N depends on the variables y 1 , y 2 , y N , which in turn depend on the variables x 1 , x 2 , x N , then the relation between their Jacobians is (Bronstein and Semendjajew 1979) as follows
z 1 , z 2 , z N x 1 , x 2 , x N = z 1 , z 2 , z N y 1 , y 2 , y N × y 1 , y 2 , y N x 1 , x 2 , x N .
In particular, if the functions z 1 , z 2 , z N are chosen to be x 1 , x 2 , x N , then for the inverse functions there follows from (A12) the rule that
y 1 , y 2 , y N x 1 , x 2 , x N = x 1 , x 2 , x N y 1 , y 2 , y N 1 .
The Jacobi method developed by Shaw (1935) [78] is mathematically the most elegant way of transforming the various partial derivatives of different potential functions into each other, exploiting the convenient formal calculus of functional determinants [71,100].
Thermodynamic partial derivatives are commonly written in such a way that all variables are indicated that are kept constant when a particular derivative is carried out.

Appendix A.1. Example 1: Lapse Rate

As a first example for applying the Jacobi method, the adiabatic lapse rate Γ of humid air is defined as the change in a parcel’s temperature with pressure, while its entropy and specific humidity remain fixed, i.e.,
Γ = T p A , η .
Following the Jacobi method, making use of the property (A10), the constant variables may be included in the Jacobian so that this derivative is written as
Γ = T p A , η = T , A , η p , A , η .
If, say, the Gibbs function, g A V A , T , p , of humid air is given, it may be desired to express the lapse rate in terms of thermodynamic potential (similar rules apply to any other available such potential). To achieve this, the Jacobian (A15) may be transformed into the associated set of independent variables of this special case, A , T , p , making use of the “chain rule” (A12) and rearranging the quantities in the numerators by (A8) and (A9) in a sequence similar to those of the denominators:
Γ = T , A , η p , A , η = T , A , η A , T , p / p , A , η A , T , p = A , T , η A , T , p / A , η , p A , T , p .
Now, using (A10) again, the new Jacobians can be collapsed back to usual partial derivatives as follows:
Γ = A , T , η A , T , p / A , η , p A , T , p = η p A , T / η T A , p .
Entropy is the negative temperature derivative, Equation (26), of the Gibbs function, so that the final result of this formal procedure is the following lapse rate formula:
Γ = T p A , η = η p A , T / η T A , p = 2 g A V T p A / 2 g A V T 2 A , p = α v A V T c p A V .
Here, α = 1 v A V v A V T A , p , v A V = g A V p A , T and c p A V = T 2 g A V T 2 A , p are, respectively, the thermal expansion coefficient, the specific volume, and the specific isobaric heat capacity of humid air, available from the related TEOS-10 Gibbs function.
Alternatively, if enthalpy h A V A , η , p = g A V + T η is available as a thermodynamic potential, such as from the TEOS-10 SIA library, similar manipulations lead to the simple expression
Γ = T p A , η = 2 h A V η p A .

Appendix A.2. Example 2: Chemical Potential

Let a second example be the chemical potential of water in humid air, Equation (17),
μ V A V = g A V A , T , p A g A V A T , p ,
which is responsible for equilibria between humid air and condensed water phases, such as at saturation conditions. If adiabatic processes are studied, it is helpful to replace the Gibbs function, g A V A , T , p , by the enthalpy, h A V A , η , p ,
g A V = h A V A , η , p η h A V η A , p .
To achieve this, a transformation of the relative chemical potential,
μ = g A V A T , p = g A V , T , p A , T , p
to the new set of variables A , η , p is required, using (A3) and (A16),
g A V A T , p = g A V , T , p A , η , p / A , T , p A , η , p = g A V A η , p T η A , p g A V η A , p T A η , p T η A , p .
Now the derivatives of g A V may be replaced by those of h A V , using Equation (A21) so that the plain terms h A V / η A , p disappear,
g A V A T , p = h A V A η , p η 2 h A V A η A , p + η T A η , p T η A , p 2 h A V η 2 A , p .
Next, T can be expressed by enthalpy, T A , η , p = h A V / η A , p , to let several terms cancel out, as follows
μ = g A V A T , p = h A V A η , p .
Together with (A21), in terms of enthalpy and entropy, the final formula for the chemical potential of water in humid air is
μ V A V A , η , p = h A V A , η , p η h A V η A , p A h A V A η , p .

Appendix B. Adiabatic Lapse Rate of the Dew-Point Temperature

The adiabatic lapse rate of the in situ temperature of a humid air parcel is given by Equation (A19),
Γ = T p A , η = 2 h A V η p A .
The related lapse rate of the dew point is
Γ d p A , p = T d p p A .
At given composition, A , and pressure, p , the parcel’s dew point does not depend on its in situ temperature, T . The dew point T d p A , p is implicitly defined by the equilibrium condition for the chemical potential of water in humid air and in liquid water,
g A V A , T d p , p A g A V A , T d p , p A T , p = g W T d p , p .
Taken at constant A , the pressure derivative of this equation is as follows:
g A V T A , p 2 g A V A T p g W T p Γ d p = g A V p A , T 2 g A V A p T g W p T .
Up to a minus sign, the factor in front of Γ d p is the evaporation entropy, as in Equation (44), so that
L L T d p , p T d p = η A V A , T d p , p A η A V A T , p η W T d p , p ,
and the r.h.s. of (A30) equals the excess volume of evaporation at equilibrium, as in Equation (52), so that
v L T d p , p = v A V A , T d p , p A v A V A T , p v W T d p , p .
The final formula for the lapse rate of the dew point is as follows:
Γ d p = T d p p A = T d p v L L L T d p , p .
Due to the negative sign of this lapse rate, with falling pressure of an ascending parcel, the dew point rises, following the opposite trend to the in situ temperature.

Appendix C. Crude Gibbs Function Approximations

For practical rule-of-thumb estimates, as well as for tutorial purposes, it is often sufficient to employ crude approximations for the thermodynamic properties of liquid water and humid air. In this sense, gases may be considered to be perfect gases with constant heat capacities, and liquids can be incompressible and also have constant heat capacities. For liquid water, water vapour, dry air, and humid air, respectively, “crude” Gibbs functions g ~ W , g ~ V , g ~ A and g ~ A V are provided here, along with their 1st and 2nd partial derivatives. The tilde indicates such crudely approximated functions and variables. Those Gibbs functions may be defined by (see Feistel et al. 2010; Appendix H therein [62])
g ~ W T , p = g ~ 0 W + g ~ 1 W T c ~ p W T ln T T t + p ρ ~ W ,
g ~ V T , p = c ~ p V T ln T T t + R W T ln p p t ,
g ~ A T , p = c ~ p A T ln T T t + R A T ln p p t ,
g ~ A V A , T , p = A g ~ A + R A T ln 1 x + 1 A g ~ V + R W T ln x .
Here, c ~ p W , c ~ p V , and c ~ p A are the constant specific isobaric heat capacities of liquid water, water vapour, and dry air, respectively. The constant density of liquid water is ρ ~ W , and T t and p t are the temperature and pressure, respectively, at an arbitrary reference state, here chosen to be the triple point of water.
The absolute energy and absolute entropy of any given substance cannot be measured and are irrelevant in empirical thermodynamics; they may only be determined from theoretical models [88,101]. In TEOS-10, these coefficients are carefully specified using practically useful reference state conditions [58,64,70]. Here, g ~ 0 W and g ~ 1 W are coefficients that are adjusted below to the liquid–vapour equilibrium at the triple point, g ~ V T t , p t = g ~ W T t , p t , as well as to the related evaporation enthalpy. The related coefficients of water vapour and of dry air are set to zero here for the simplicity of the formulas, albeit without affecting any measurable properties. Note that such crude approximations of thermodynamic potentials need to be used with care; for example, g ~ W does not permit a reasonable computation of sound speed, in contrast to the original TEOS-10 Gibbs function.
In Table A1, the 1st and 2nd partial derivatives of the crude Gibbs functions are reported for their completeness and convenience. For quantitative estimates, the crude Gibbs functions may be adjusted to the properties of real substances at a suitably selected state. Here, the TEOS-10 triple-point properties of water at T t = 273.16   K and p t = 611.654   P a are used, rounded to 6 digits [56,58,62,70]:
c ~ p V = 1884.35   J   k g 1   K 1
c ~ p A = 1003.69   J   k g 1   K 1
c ~ p W = 4219.91   J   k g 1   K 1
ρ ~ W = 999.793   k g   m 3
Table A1. Analytical 1st and 2nd partial derivatives of the crude Gibbs functions (A34)–(A37). Note that e exp 1 , R A V A 1 A R W + A R A , and c ~ p A V A 1 A c ~ p V + A c ~ p A . The water vapour mole fraction is defined by Equation (4) as x A = 1 A R W R A V A .
Table A1. Analytical 1st and 2nd partial derivatives of the crude Gibbs functions (A34)–(A37). Note that e exp 1 , R A V A 1 A R W + A R A , and c ~ p A V A 1 A c ~ p V + A c ~ p A . The water vapour mole fraction is defined by Equation (4) as x A = 1 A R W R A V A .
g g ~ W g ~ V g ~ A g ~ A V
g A T , p 000 g ~ A + R A T ln 1 x
g ~ V R W T ln x
g T A , p g ~ 1 W c ~ p W ln e T T t c ~ p V ln e T T t
+ R W ln p p t
c ~ p A ln e T T t
+ R A ln p p t
c ~ p A V ln e T T t + R A V ln p p t
+ A R A ln 1 x + 1 A R W ln x
g p A , T 1 ρ ~ W R W T p R A T p R A V T p
2 g A 2 T , p 000 R A R W T A 1 A R A V
2 g A T p 000 c ~ p V c ~ p A ln e T T t + R A R W ln p p t
+ R A ln 1 x R W ln x
2 g T 2 A , p c ~ p W T c ~ p V T c ~ p A T c ~ p A V T
2 g T p A 0 R W p R A p R A V p
2 g p 2 A , T 0 R W T p 2 R A T p 2 R A V T p 2
2 g A p T 000 R A R W T p
Evidently, away from the triple point, the crude properties will deviate from those of TEOS-10.
The enthalpies of liquid water, h ~ W , and of water vapour, h ~ V , figured as
h ~ W T , p = g ~ W T , p T g ~ W T p = g ~ 0 W + p ρ ~ W + c ~ p W T ,
h ~ V T , p = g ~ V T , p T g ~ V T p = c ~ p V T ,
differ in terms of the latent heat of evaporation at the triple point,
L ~ L = h ~ V T t , p t h ~ W T t , p t = c ~ p V c ~ p W + g ~ 1 W T t .
The TEOS-10 value of the evaporation enthalpy of liquid water at the triple point is (Feistel et al., 2008 [70]) determined as follows:
L ~ L = 2500915   J k g 1
For the liquid–vapour equilibrium at the triple point, defined by equal chemical potentials,
g ~ W T t , p t = g ~ V T t , p t .
The adjustable constants take the values of
g ~ 1 W = L ~ L T t c ~ p V + c ~ p W = 11491.055   J   k g 1   K 1 ,
g ~ 0 W = g ~ 1 W T t p t ρ ~ W = 3138897   J   k g 1 .
Note that the latter figure is numerically crucial for the mutual consistency between the Gibbs functions of water, but that the value itself has no physical relevance and cannot be measured.
Up to 6 digits of the latest value, being exact by definition, the molar gas constant is (BIPM 2019) [6], given as
R = 8.31446   J m o l 1 K   1 ,
and the specific gas constant for water is
R W = R M W = 461.523   J   k g 1   K 1 ,
while for dry air, it is
R A = R M A = 287.047   J   k g 1   K 1 .
It should be noted that these crude Gibbs functions are not consistent with the TEOS-10 reference state conditions [58,64,70]. Consequently, attempts to compute phase-transition properties from combining proper TEOS-10 functions with these crude equations are strongly discouraged and will give unreasonable results.

Appendix D. Lambert’s W Function

The solution w y of the nonlinear equation
w exp w = y
is mathematically well studied and is known as “Lambert’s W function” [102,103,104].
For y 1 / e , the function w z has real values. Special values are
w 1 / e = 1 ,
w 0 = 0 ,
w 1 = 0.567143290
The value of w 1 is regarded as the mathematical “Omega constant”.
For arguments y > 0 , w y is a unique function. For 1 e < y < 0 , w y has two branches, as shown in Figure A1. The upper “principal” branch has values of 1 < w < 0 , and the lower branch has values of < w < 1 .
Numerically, the function w y can be computed by Newton–Cotes iteration of Equation (A52) or by various approximation expressions found in mathematical textbooks. For the principal branch, a simple formula is the series expansion, given by
w y = k = 1 k k 1 k ! y k = y y 2 + 3 2 y 3 8 3 y 4 +
Figure A1. Plot of Lambert’s W function. The blue upper curve is the principal branch. The red curve is the lower branch that is to be used in this paper. Modified public domain graphics from https://de.wikipedia.org/wiki/Lambertsche_W-Funktion, accessed on 10 March 2024.
Figure A1. Plot of Lambert’s W function. The blue upper curve is the principal branch. The red curve is the lower branch that is to be used in this paper. Modified public domain graphics from https://de.wikipedia.org/wiki/Lambertsche_W-Funktion, accessed on 10 March 2024.
Oceans 05 00020 g0a1

Appendix E. List of Symbols and Abbreviations

SymbolRemarkBasic Unit
A Dry-air mass fraction in humid air k g   k g 1
a 11 a 33 Matrix of coefficients
a ~ 11 a ~ 33 Matrix of coefficients, crude approximation
A s a t Dry-air mass fraction of saturated humid air k g   k g 1
B Abbreviation, B = g ~ 0 W / Q R W T 0 1
b 1 b 3 Vector of coefficients
C Cloudiness: cloud-covered surface fractionm−2/m−2
c p A V Specific isobaric heat capacity of humid air J   k g 1   K 1
c p S W Specific isobaric heat capacity of seawater J   k g 1   K 1
c ~ p A Crude specific isobaric heat capacity of dry air, c ~ p A = 1003.69   J   k g 1   K 1 J   k g 1   K 1
c ~ p A V Crude specific isobaric heat capacity of humid air J   k g 1   K 1
c ~ p V Crude specific isobaric heat capacity of water vapour,
c ~ p V = 1884.35   J   k g 1   K 1
J   k g 1   K 1
c ~ p W Crude specific isobaric heat capacity of liquid water,
c ~ p W = 4219.91   J   k g 1   K 1
J   k g 1   K 1
CRECloud radiative effectW m−2
e Euler number, e = exp 1 = 2.718281828  
E Abbreviation, E = B exp B + ln ψ f Q 1
EEIEarth energy imbalance
f A Specific Helmholtz energy of dry air J   k g 1
f A V Specific Helmholtz energy of humid air J   k g 1
f F Specific Helmholtz energy of fluid water J   k g 1
g A V Specific Gibbs energy of humid air J   k g 1
g ~ A V Crude specific Gibbs energy of humid air J   k g 1
g E Gravitational acceleration, g E = 9.81   m   s 2 m s 2
GHGGreenhouse gas
g I h Specific Gibbs energy of ambient hexagonal ice J   k g 1
g S W Specific Gibbs energy of seawater J   k g 1
g W Specific Gibbs energy of liquid water J   k g 1
g ~ W Crude specific Gibbs energy of liquid water J   k g 1
g ~ 0 W Adjustable constant, g ~ 0 W = 3271880   J   k g 1 J   k g 1
g ~ 1 W Adjustable constant, g ~ 1 W = 11977.9   J   k g 1   K 1 J   k g 1   K 1
h A V Specific enthalpy of humid air J   k g 1
h S W Specific enthalpy of seawater J   k g 1
h S O S W Specific enthalpy of the standard-ocean reference state J   k g 1
IAPSOInternational Association for the Physical Sciences of the Oceans
IAPWSInternational Association for the Properties of Water and Steam
ITS-901990 International Temperature Scale
JCSJoint Committee on the Properties of Seawater
J E Ocean–cloud radiative exchange fluxW m−2
J Upward thermal radiation fluxW m−2
J Downward thermal radiation fluxW m−2
LCLLifted condensation levelm
L L Specific evaporation enthalpy of liquid water J   k g 1
L ~ L Crude specific evaporation enthalpy of liquid water,
L ~ L = 2500915   J   k g 1
J   k g 1
LW CRELongwave cloud radiative effectW m−2
M A Molar mass of dry air, M A = 28.96546   g   m o l 1 g   m o l 1
M W Molar mass of water, M W = 18.015268   g   m o l 1 g   m o l 1
OHCOcean heat contentJ
p Pressure Pa
P Entropy production per surface area W m 2   K 1
p 0 Sea surface air pressurePa
p L C L Lifted condensation level pressurePa
p t Triple-point pressure of water, p t = 611.654   P a Pa
q Specific humidity k g   k g 1
Q Abbreviation, Q = c ~ p W c ~ p V R W + c ~ p A V R A V 1
r Mixing ratio k g   k g 1
R Molar gas constant, R = 8.31446 J   m o l 1   K 1 J   m o l 1   K 1
R A Specific gas constant of dry air, R A = 287.047   J   k g 1   K 1 J   k g 1   K 1
R A V Specific gas constant of humid air, R A V = 1 A R W + A R A J   k g 1   K 1
RFRelative fugacity%rh
RHRelative humidity%rh
R W Specific gas constant of water, 461.523   J   k g 1   K 1 J   k g 1   K 1
S Seawater salinity k g   k g 1
SCORScientific Committee on Oceanic Research
SSTSea surface temperatureK, °C
SW CREShortwave cloud radiative effectW m−2
t Celsius temperature°C
T Absolute temperature, ITS-90K
T 0 Sea surface temperatureK
T d p Dew-point temperature K
T L C L Lifted condensation level temperatureK
T t Triple point temperature of water, T t = 273.16   K K
TEOS-10Thermodynamic Equation of Seawater—2010
v A V Specific volume of humid air m 3   k g 1
VSMOWVienna Standard Mean Ocean Water
v W Specific volume of liquid water m 3   k g 1
w y Lambert’s W function1
x water vapour mole fraction m o l   m o l 1
y r Calendar year number (Common Era)
z Vertical coordinatem
z L C L Lifted condensation level heightm
α Humidity sensitivity K 1
α A V Thermal expansion coefficient of humid air K 1
β LCL temperature sensitivity1
Γ Adiabatic lapse rate of humid air K P a 1
Γ d p Adiabatic lapse rate of the humid-air dew point K P a 1
γ LCL pressure sensitivity
γ L C L LCL height coefficient m   K 1
A Increment of dry-air mass fraction k g   k g 1
J Solar irradiation increaseW m−2
J E Ocean–cloud exchange flux increaseW m−2
p Pressure increasePa
p L C L LCL pressure increasePa
T Temperature increase K
T 0 SST increaseK
T L C L LCL temperature increaseK
v L Specific evaporation volume of liquid water m 3   k g 1
η L Specific evaporation entropy of liquid water J   k g 1   K 1
ε Clausius–Clapeyron expansion parameter1
η Specific entropy J   k g 1   K 1
η A V Specific entropy of humid air J   k g 1   K 1
η S W Specific entropy of seawater J   k g 1   K 1
η W Specific entropy of liquid water J   k g 1   K 1
μ V A V Chemical potential of water vapour in humid air J   k g 1
μ W I h Chemical potential of ambient hexagonal ice J   k g 1
μ W W Chemical potential of liquid water J   k g 1
ρ Mass density k g   m 3
ρ S W Mass density of seawater k g   m 3
ρ ~ W Crude mass density of liquid water, ρ ~ W = 999.793   k g   m 3 k g   m 3
σ S B Stefan–Boltzmann constant, σ S B = 5.670374419   ×   10 8   W   m 2   K 4 W   m 2   K 4
ψ f Relative fugacity P a   P a 1

References

  1. Brizio, A.M. Primo Libro Delle Acque. Gheroni, Torino. 1951, pp. 159–266. Available online: https://en.wikipedia.org/wiki/Codex_Arundel (accessed on 10 March 2024).
  2. Schneider, M. Leonardo da Vinci—Das Wasserbuch; Schirmer/Model: München, Germany; Paris, France; London, UK, 1996. [Google Scholar]
  3. IOC; SCOR; IAPSO. The International Thermodynamic Equation of Seawater-2010: Calculation and Use of Thermodynamic Properties. Intergovernmental Oceanographic Commission, Manuals and Guides No. 56, UNESCO (English), Paris. 2010. Available online: http://www.TEOS-10.org (accessed on 10 March 2024).
  4. Almeida, L.; Lima de Azevedo, J.L.; Kerr, R.; Araujo, M.; Mata, M.M. Impact of the new equation of state of seawater (TEOS-10) on the estimates of water mass mixture and meridional transport in the Atlantic Ocean. Prog. Oceanogr. 2018, 162, 13–24. [Google Scholar] [CrossRef]
  5. Budyko, M.I. Der Wärmehaushalt der Erdoberfläche; Fachliche Mitteilungen der Inspektion Geophysikalischer Beratungsdienst der Bundeswehr im Luftwaffenamt: Köln, Germany, 1963; Volume 100, pp. 3–282. [Google Scholar]
  6. Josey, S.A.; Gulev, S.; Yu, L. Exchanges through the ocean surface. In Ocean Circulation and Climate. A 21st Century Perspective; Siedler, G., Griffies, S.M., Gould, J., Church, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 115–140. [Google Scholar] [CrossRef]
  7. Macdonald, A.M.; Baringer, M.O. Ocean Heat Transport. In Ocean Circulation and Climate. A 21st Century Perspective; Siedler, G., Griffies, S.M., Gould, J., Church, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 115–140. [Google Scholar] [CrossRef]
  8. Rapp, D. Assessing Climate Change in Temperatures, Solar Radiation, and Heat Balance; Springer: Cham, Switzerland, 2014. [Google Scholar]
  9. You, X. Oceans break heat records five years in a row. The heat stored in the world’s oceans increased by the greatest margin ever in 2023. Nature 2024, 625, 434–435. [Google Scholar] [CrossRef] [PubMed]
  10. Meehl, G.; Arblaster, J.; Fasullo, J.; Hu, A.; Trenbert, K.E. Model-based evidence of deep-ocean heat uptake during surface-temperature hiatus periods. Nat. Clim. Chang. 2011, 1, 360–364. [Google Scholar] [CrossRef]
  11. Von Schuckmann, K.; Minère, A.; Gues, F.; Cuesta-Valero, F.J.; Kirchengast, G.; Adusumilli, S.; Straneo, F.; Ablain, M.; Allan, R.P.; Barker, P.; et al. Heat stored in the Earth system 1960–2020: Where does the energy go? Earth Syst. Sci. Data 2023, 15, 1675–1709. [Google Scholar] [CrossRef]
  12. Abraham, J.P.; Baringer, M.; Bindoff, N.L.; Boyer, S.T.; Cheng, L.J.; Church, J.A.; Conroy, J.L.; Domingues, C.M.; Fasullo, J.T.; Gilson, J.; et al. A Review of Global Ocean Temperature Observations: Implications for Ocean Heat Content Estimates and Climate Change. Rev. Geophys. 2013, 51, 450–483. [Google Scholar] [CrossRef]
  13. Cheng, L.; Abraham, J.; Trenberth, K.E.; Boyer, T.; Mann, M.E.; Zhu, J.; Wang, F.; Yu, F.; Locarnini, R.; Fasullo, J.; et al. New record ocean temperatures and related climate indicators in 2023. Adv. Atmos. Sci. 2024, 41, 1068–1082. [Google Scholar] [CrossRef]
  14. McDougall, T.J. Potential enthalpy: A conservative oceanic variable for evaluating heat content and heat fluxes. J. Phys. Oceanogr. 2003, 33, 945–963. [Google Scholar] [CrossRef]
  15. McDougall, T.J.; Barker, P.M.; Holmes, R.M.; Pawlowicz, R.; Griffies, S.M.; Durack, P.J. The interpretation of temperature and salinity variables in numerical ocean model output and the calculation of heat fluxes and heat content. Geosci. Model Dev. 2021, 14, 6445–6466. [Google Scholar] [CrossRef]
  16. Loeb, N.G.; Johnson, G.C.; Thorsen, T.J.; Lyman, J.M.; Rose, F.G.; Kato, S. Satellite and Ocean Data Reveal Marked Increase in Earth’s Heating Rate. Geophys. Res. Lett. 2021, 48, e2021GL093047. [Google Scholar] [CrossRef]
  17. Kramer, R.J.; He, H.; Soden, B.J.; Oreopoulos, L.; Myhre, G.; Forster, P.M.; Smith, C.J. Observational Evidence of Increasing Global Radiative Forcing, Geophys. Res. Lett. 2021, 48, e2020GL091585. [Google Scholar] [CrossRef]
  18. Liu, C.; Allan, R.P.; Mayer, M.; Hyder, P.; Desbruyères, D.; Cheng, L.; Xu, J.; Xu, F.; Zhang, Y. Variability in the global energy budget and transports 1985–2017. Clim. Dynam. 2020, 55, 3381–3396. [Google Scholar] [CrossRef]
  19. Yang, X.; Ge, J.; Hu, X.; Wang, M.; Han, Z. Cloud-Top Height Comparison from Multi-Satellite Sensors and Ground-Based Cloud Radar over SACOL Site. Remote Sens. 2021, 13, 2715. [Google Scholar] [CrossRef]
  20. Loeb, N.G.; Mayer, M.; Kato, S.; Fasullo, J.T.; Zuo, H.; Senan, R.; Lyman, J.M.; Johnson, G.C.; Balmaseda, M. Evaluating twenty-year trends in Earth’s energy flows from observations and reanalyses. J. Geophys. Res. Atmos. 2022, 127, e2022JD036686. [Google Scholar] [CrossRef]
  21. Qu, X.; Hall, A.; Klein, S.A.; Caldwell, P.M. On the spread of changes in marine low cloud cover in climate model simulations of the 21st century. Clim. Dyn. 2014, 42, 2603–2626. [Google Scholar] [CrossRef]
  22. WMO Provisional State of the Global Climate 2023; World Meteorological Organization: Geneva, Switzerland, 2024. Available online: https://library.wmo.int/records/item/68835-state-of-the-global-climate-2023 (accessed on 10 March 2024).
  23. Foster, M.J.; Phillips, C.; Heidinger, A.K.; Borbas, E.E.; Li, Y.; Menzel, P.; Walther, A.; Weisz, E. PATMOS-x Version 6.0: 40 Years of Merged AVHRR and HIRS Global Cloud Data. J. Clim. 2023, 36, 1143–1160. [Google Scholar] [CrossRef]
  24. Goode, P.R.; Pallé, E.; Shoumko, A.; Shoumko, S.; Montañes-Rodriguez, P.; Koonin, S.E. Earth’s albedo 1998–2017 as measured from earthshine. Geophys. Res. Lett. 2021, 48, e2021GL094888. [Google Scholar] [CrossRef]
  25. Feistel, R.; Hellmuth, O. Relative Humidity: A Control Valve of the Steam Engine Climate. J. Hum. Earth Future 2021, 2, 140–182. [Google Scholar] [CrossRef]
  26. Phillips, C.; Foster, M.J. Cloudiness. In State of the Climate in 2022; Blunden, J., Boyer, T., Bartow-Gillies, E., Eds.; Bulletin of the American Meteorological Society (BAMS): Boston, MA, USA, 2023; Volume 104, pp. S60–S61. [Google Scholar]
  27. Held, I.M.; Soden, B.J. Robust Responses of the Hydrological Cycle to Global Warming. J. Clim. 2006, 19, 5686–5699. [Google Scholar] [CrossRef]
  28. Weller, R.A.; Lukas, R.; Potemra, J.; Plueddemann, A.J.; Fairall, C.; Bigorre, S. Ocean Reference Stations: Long-Term, Open-Ocean Observations of Surface Meteorology and Air–Sea Fluxes Are Essential Benchmarks. Cover. Bull. Am. Meteorol. Soc. 2022, 103, E1968–E1990. [Google Scholar] [CrossRef]
  29. Feistel, R.; Hellmuth, O. Thermodynamics of Evaporation from the Ocean Surface. Atmosphere 2023, 14, 560. [Google Scholar] [CrossRef]
  30. Blunden, J.; Boyer, T.; Bartow-Gillies, E. (Eds.) State of the Climate in 2022; Bulletin of the American Meteorological Society (BAMS): Boston, MA, USA, 2023; Volume 104, pp. S1–S516. [Google Scholar]
  31. Randall, D.A. Atmosphere, Clouds, and Climate; Princeton University Press: Princeton, NJ, USA, 2012. [Google Scholar]
  32. Stewart, R.H. Introduction to Physical Oceanography; Texas A & M University: College Station, TX, USA, 2008. [Google Scholar] [CrossRef]
  33. Feistel, R.; Hellmuth, O. Irreversible Thermodynamics of Seawater Evaporation. J. Mar. Sci. Eng. 2024, 12, 166. [Google Scholar] [CrossRef]
  34. Lovell-Smith, J.W.; Feistel, R.; Harvey, A.H.; Hellmuth, O.; Bell, S.A.; Heinonen, M.; Cooper, J.R. Metrological challenges for measurements of key climatological observables. Part 4: Atmospheric relative humidity. Metrologia 2016, 53, R40–R59. [Google Scholar] [CrossRef] [PubMed]
  35. Dunn, R.J.H.; Miller, J.B.; Willett, K.M.; Gobron, N. Global Climate. In State of the Climate in 2022; Blunden, J., Boyer, T., Bartow-Gillies, E., Eds.; Bulletin of the American Meteorological Society: Boston, MA, USA, 2023; Volume 104, pp. S20–S26. [Google Scholar]
  36. Willett, K.M.; Simmons, A.J.; Bosilovich, M.; Lavers, D.A. Surface Humidity. In State of the Climate in 2022; Blunden, J., Boyer, T., Bartow-Gillies, E., Eds.; Bulletin of the American Meteorological Society (BAMS): Boston, MA, USA, 2023; Volume 104, pp. S49–S52. [Google Scholar]
  37. Azorin-Molina, C.; Dunn, R.J.H.; Ricciardulli, L.; Mears, C.A.; Nicolas, J.P.; McVicar, T.R.; Zeng, Z.; Bosilovich, M.G. Land and Ocean Surface Winds. In State of the Climate in 2022; Blunden, J., Boyer, T., Bartow-Gillies, E., Eds.; Bulletin of the American Meteorological Society (BAMS): Boston, MA, USA, 2023; Volume 104, pp. S72–S74. [Google Scholar]
  38. Feistel, R. Salinity and relative humidity: Climatological relevance and metrological needs. Acta IMEKO 2015, 4, 57–61. [Google Scholar] [CrossRef]
  39. Feistel, R. TEOS-10 and the Climatic Relevance of Ocean-Atmosphere Interaction. EGUsphere 2024. preprint. [Google Scholar] [CrossRef]
  40. Goody, R.M. A statistical model for water-vapour absorption. Q. J. R. Meteorol. Soc. 1952, 78, 165–169. [Google Scholar] [CrossRef]
  41. Goody, R.M.; Robinson, G.D. Radiation in the troposphere and lower stratosphere. Rev. Mod. Meteorol. 1951, 77, 151–187. [Google Scholar] [CrossRef]
  42. Zhong, W.; Haigh, J.D. The greenhouse effect and carbon dioxide. Weather 2013, 68, 100–105. [Google Scholar] [CrossRef]
  43. Rothman, L.S.; Jacquemart, D.; Barbe, A.; Chris Benner, D.; Birk, M.; Brown, L.R.; Carleer, M.R.; Chackerian, C.; Chance, K.; Coudert, L.H.; et al. The HITRAN 2004 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 2005, 96, 139–204. [Google Scholar] [CrossRef]
  44. Colman, R.; Soden, B.L. Water vapor and lapse rate feedbacks in the climate system. Rev. Mod. Phys. 2021, 93, 045002. [Google Scholar] [CrossRef]
  45. Turbet, M.; Bolmont, E.; Chaverot, G.; Ehrenreich, D.; Leconte, J.; Marcq, E. Day–night cloud asymmetry prevents early oceans on Venus but not on Earth. Nature 2021, 598, 276–280. [Google Scholar] [CrossRef]
  46. McDougall, T.J.; Barker, P.M.; Feistel, R.; Roquet, F. A thermodynamic potential of seawater in terms of Absolute Salinity, Conservative Temperature, and in situ pressure. Ocean Sci. 2023, 19, 1719–1741. [Google Scholar] [CrossRef]
  47. McDougall, T.J.; Feistel, R.; Pawlowicz, R. Thermodynamics of Seawater. In Ocean Circulation and Climate, A 21st Century Perspective; Siedler, G., Griffies, S.M., Gould, J., Church, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 141–158. [Google Scholar] [CrossRef]
  48. Graham, F.S.; McDougall, T.J. Quantifying the Nonconservative Production of Conservative Temperature, Potential Temperature, and Entropy. J. Phys. Oceanogr. 2013, 43, 838–862. [Google Scholar] [CrossRef]
  49. Tailleux, R.; Dubos, T. A simple and transparent method for improving the energetics and thermodynamics of seawater approximations: Static energy asymptotics (SEA). Ocean. Model. 2024, 188, 102339. [Google Scholar] [CrossRef]
  50. Unesco Background Papers and Supporting Data on the International Equation of State of Sea Water 1980; Unesco Technical Paper Marine Science 38. UNESCO: Paris, France, 1981. Available online: https://unesdoc.unesco.org/ark:/48223/pf0000047363 (accessed on 10 March 2024).
  51. Fofonoff, N.P.; Millard, R.C. Algorithms for the Computation of Fundamental Properties of Seawater; Unesco Technical Papers in Marine Science 44; UNESCO: Paris, France, 1983. [Google Scholar] [CrossRef]
  52. Pawlowicz, R.; McDougall, T.J.; Feistel, R.; Tailleux, R. An historical perspective on the development of the Thermodynamic Equation of Seawater—2010. Ocean Sci. 2012, 8, 161–174. [Google Scholar] [CrossRef]
  53. Fink, J.K. Chapter 1: Mathematics of Thermodynamics. In Physical Chemistry in Depth; Springer: Berlin/Heidelberg, Germany, 2009. [Google Scholar] [CrossRef]
  54. Wright, D.G.; Feistel, R.; Reissmann, J.H.; Miyagawa, K.; Jackett, D.R.; Wagner, W.; Overhoff, U.; Guder, C.; Feistel, A.; Marion, G.M. Numerical implementation and oceanographic application of the thermodynamic potentials of liquid water, water vapour, ice, seawater and humid air—Part 2: The library routines. Ocean Sci. 2010, 6, 695–718. [Google Scholar] [CrossRef]
  55. Feistel, R.; Wright, D.G.; Jackett, D.R.; Miyagawa, K.; Reissmann, J.H.; Wagner, W.; Overhoff, U.; Guder, C.; Feistel, A.; Marion, G.M. Numerical implementation and oceanographic application of the thermodynamic potentials of liquid water, water vapour, ice, seawater and humid air—Part 1: Background and equations. Ocean. Sci. 2010, 6, 633–677. [Google Scholar] [CrossRef]
  56. Wagner, W.; Pruß, A. The IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water Substance for General and Scientific Use. J. Phys. Chem. Ref. Data 2002, 31, 387–535. [Google Scholar] [CrossRef]
  57. Millero, F.J.; Feistel, R.; Wright, D.G.; McDougall, T.J. The composition of Standard Seawater and the definition of the Reference-Composition Salinity Scale. Deep-Sea Res. 2008, 55 Pt I, 50–72. [Google Scholar] [CrossRef]
  58. Feistel, R. A Gibbs function for seawater thermodynamics for −6 to 80 °C and salinity up to 120 g kg−1. Deep-Sea Res. 2008, 55 Pt I, 1639–1671. [Google Scholar] [CrossRef]
  59. Feistel, R.; Wagner, W. High-pressure thermodynamic Gibbs functions of ice and sea ice. J. Mar. Res. 2005, 63, 95–139. [Google Scholar] [CrossRef]
  60. Feistel, R.; Wagner, W. A new equation of state for H2O ice Ih. J. Phys. Chem. Ref. Data 2006, 35, 1021–1047. [Google Scholar] [CrossRef]
  61. Lemmon, E.W.; Jacobsen, R.T.; Penoncello, S.G.; Friend, D.G. Thermodynamic Properties of Air and Mixtures of Nitrogen, Argon and Oxygen From 60 to 2000 K at Pressures to 2000 MPa. J. Phys. Chem. Ref. Data 2000, 29, 331–362. [Google Scholar] [CrossRef]
  62. Feistel, R.; Wright, D.G.; Kretzschmar, H.-J.; Hagen, E.; Herrmann, S.; Span, R. Thermodynamic Properties of Sea Air. Ocean. Sci. 2010, 6, 91–141. [Google Scholar] [CrossRef]
  63. IAPWS AN6-16 Advisory Note No. 6: Relationship between Various IAPWS Documents and the International Thermodynamic Equation of Seawater—2010 (TEOS-10); The International Association for the Properties of Water and Steam: Dresden, Germany, 2016; Available online: http://www.iapws.org (accessed on 10 March 2024).
  64. Feistel, R. Thermodynamic properties of seawater, ice and humid air: TEOS-10, before and beyond. Ocean Sci. 2018, 14, 471–502. [Google Scholar] [CrossRef]
  65. Preston-Thomas, H. The International Temperature Scale of 1990 (ITS-90). Metrologia 1990, 27, 3–10. [Google Scholar] [CrossRef]
  66. BIPM. The International System of Units (SI). 9th Edition of the SI Brochure. 2019. Available online: https://www.bipm.org/en/publications/si-brochure/ (accessed on 15 March 2023).
  67. Harvey, A.H.; Hrubý, J.; Meier, K. Improved and Always Improving: Reference Formulations for Thermophysical Properties of Water. J. Phys. Chem. Ref. Data 2023, 52, 011501. [Google Scholar] [CrossRef]
  68. IAPWS G5-01; Guideline on the Use of Fundamental Physical Constants and Basic Constants of Water. The International Association for the Properties of Water and Steam: Dresden, Germany, 2016. Available online: http://www.iapws.org (accessed on 10 March 2024).
  69. IAPWS G08-10; Guideline on an Equation of State for Humid Air in Contact with Seawater and Ice, Consistent with the IAPWS Formulation 2008 for the Thermodynamic Properties of Seawater. The International Association for the Properties of Water and Steam: Niagara Falls, ON, Canada, 2010. Available online: http://www.iapws.org (accessed on 10 March 2024).
  70. Feistel, R.; Wright, D.G.; Miyagawa, K.; Harvey, A.H.; Hruby, J.; Jackett, D.R.; McDougall, T.J.; Wagner, W. Mutually consistent thermodynamic potentials for fluid water, ice and seawater: A new standard for oceanography. Ocean Sci. 2008, 4, 275–291. [Google Scholar] [CrossRef]
  71. Landau, L.D.; Lifschitz, E.M. Statistische Physik; Akademie-Verlag: Berlin, Germany, 1966. [Google Scholar]
  72. Clausius, R. Die Mechanische Wärmetheorie; Friedrich Vieweg und Sohn: Braunschweig, Germany, 1876. [Google Scholar]
  73. Gibbs, J.W. Graphical methods in the thermodynamics of fluids. Trans. Conn. Acad. Arts Sci. 1873, 2, 309–342. Available online: https://www3.nd.edu/~powers/ame.20231/gibbs1873a.pdf (accessed on 10 March 2024).
  74. Sommerfeld, A. Thermodynamik und Statistik; Verlag Harri Deutsch: Thun, Switzerland; Frankfurt, Germany, 1988. [Google Scholar]
  75. Feistel, R.; Wielgosz, R.; Bell, S.A.; Camões, M.F.; Cooper, J.R.; Dexter, P.; Dickson, A.G.; Fisicaro, P.; Harvey, A.H.; Heinonen, M.; et al. Metrological challenges for measurements of key climatological observables: Oceanic salinity and pH, and atmospheric humidity. Part 1: Overview. Metrologia 2016, 53, R1–R11. [Google Scholar] [CrossRef]
  76. Feistel, R.; Lovell-Smith, J.W. Defining relative humidity in terms of water activity. Part 1: Definition. Metrologia 2017, 54, 566–576. [Google Scholar] [CrossRef]
  77. Feistel, R.; Hellmuth, O.; Lovell-Smith, J.W. Defining relative humidity in terms of water activity: III. Relations to dew-point and frost-point temperatures. Metrologia 2022, 59, 045013. [Google Scholar] [CrossRef]
  78. Shaw, A.N. The Derivation of Thermodynamical Relations for a Simple System. Philos. Trans. R. Soc. Lond. A 1935, 234, 299–328. [Google Scholar] [CrossRef]
  79. Eastman, R.; Warren, S.G.; Hahn, C.J. Variations in Cloud Cover and Cloud Types over the Ocean from Surface Observations, 1954–2008. J. Clim. 2011, 24, 5914–5934. [Google Scholar] [CrossRef]
  80. Sandu, I.; Stevens, B.; Pincus, R. On the transitions in marine boundary layer cloudiness. Atmos. Chem. Phys. 2010, 10, 2377–2391. [Google Scholar] [CrossRef]
  81. Romps, D.M. Exact Expression for the Lifting Condensation Level. J. Atmos. Sci. 2017, 74, 3891–3900. [Google Scholar] [CrossRef]
  82. Lawrence, M.G. The relationship between relative humidity and the dewpoint temperature in moist air: A simple conversion and applications. Bull. Amer. Meteor. Soc. 2005, 86, 225–233. [Google Scholar] [CrossRef]
  83. Thomas, G.E.; Stamnes, K. Radiative Transfer in the Atmosphere and Ocean; Cambridge University Press: Cambridge, UK, 1999. [Google Scholar]
  84. Liou, K.N. An Introduction to Atmospheric Radiation, 2nd ed.; International Geophysics Series; Academic Press: New York, NY, USA, 2002; Volume 84, Available online: https://www.sciencedirect.com/bookseries/international-geophysics/vol/84/suppl/C (accessed on 10 March 2024).
  85. Dai, A. Recent Climatology, Variability, and Trends in Global Surface Humidity. J. Clim. 2006, 19, 3589–3606. [Google Scholar] [CrossRef]
  86. Pelkowski, J. Of entropy production by radiative processes in a conceptual climate model. Meteorol. Z. 2012, 21, 439–457. [Google Scholar] [CrossRef]
  87. Pelkowski, J. On the Clausius-Duhem Inequality and Maximum Entropy Production in a Simple Radiating System. Entropy 2014, 16, 2291–2308. [Google Scholar] [CrossRef]
  88. Planck, M. Vorlesungen über die Theorie der Wärmestrahlung; Johann Ambrosius Barth: Leipzig, Germany, 1906. [Google Scholar]
  89. Kabelac, S. Thermodynamik der Strahlung; Vieweg, Braunschweig & Wiesbaden: Wiesbaden, Germany, 1994. [Google Scholar]
  90. Feistel, R. Entropy Flux and Entropy Production of Stationary Black-Body Radiation. J. Non-Equilib. Thermodyn. 2011, 36, 131–139. [Google Scholar] [CrossRef]
  91. Hume, D. Eine Untersuchung über den Menschlichen Verstand; Reclam: Ditzingen, Germany, 1967. [Google Scholar]
  92. Russell, B. Chapter IX: On the Notion of Cause. In Mysticism and Logic and Other Essays; Longmans, Green and Co.: London, UK, 1919; pp. 180–208. Available online: https://en.wikisource.org/wiki/Mysticism_and_Logic_and_Other_Essays (accessed on 10 March 2024).
  93. Feistel, R. On the Evolution of Symbols and Prediction Models. Biosemiotics 2023, 16, 311–371. [Google Scholar] [CrossRef]
  94. Feistel, R. Self-Organisation of Prediction Models. Entropy 2023, 25, 1596. [Google Scholar] [CrossRef] [PubMed]
  95. Hertz, H. Die Prinzipien der Mechanik; Johann Ambrosius Barth, Leipzig. Photocopy Reprint (1963); Wissenschaftliche Buchgesellschaft: Darmstadt, Germany, 1894. [Google Scholar]
  96. Stips, A.; Macias, D.; Coughlan, C.; Garcia-Gorriz, E.; Liang, X.S. On the causal structure between CO2 and global temperature. Sci. Rep. 2016, 6, 21691. [Google Scholar] [CrossRef] [PubMed]
  97. Bronstein, I.N.; Semendjajew, K.A. Taschenbuch der Mathematik; Nauka: Moscow, Russia; Teubner: Leipzig, Germany, 1979. [Google Scholar]
  98. Kaplan, W. Advanced Calculus; Addison-Wesley: Reading, MA, USA, 1984. [Google Scholar]
  99. Gradshteyn, I.S.; Ryzhik, I.M. Tables of Integrals, Series, and Products; Academic Press: San Diego, CA, USA, 2000. [Google Scholar]
  100. Margenau, H.; Murphy, G.M. The Mathematics of Physics and Chemistry; D. van Nostrand Company, Inc.: New York, NY, USA, 1943. [Google Scholar]
  101. Feistel, R. Distinguishing between Clausius, Boltzmann and Pauling Entropies of Frozen Non-Equilibrium States. Entropy 2019, 21, 799. [Google Scholar] [CrossRef] [PubMed]
  102. Corless, R.M.; Gonnet, G.H.; Hare, D.E.G.; Jeffrey, D.J.; Knuth, D.E. On the Lambert W Function. Adv. Comput. Math. 1996, 5, 329–359. [Google Scholar] [CrossRef]
  103. Kalugin, G.A.; Jeffrey, D.J.; Corless, R.M.; Borwein, P.B. Stieltjes and other integral representations for functions of Lambert W. Integral Transform. Spec. Funct. 2012, 23, 581–593. [Google Scholar] [CrossRef]
  104. Loczi, L. Guaranteed- and high-precision evaluation of the Lambert W function. Appl. Math. Comput. 2022, 433, 127406. [Google Scholar] [CrossRef]
Figure 1. Schematic of cloud radiation effects with respect to scattered subtropical marine cumulus clouds developing above a uniform lifted condensation level (LCL). Background photo taken at Las Brujas, Cuba, in February 2014.
Figure 1. Schematic of cloud radiation effects with respect to scattered subtropical marine cumulus clouds developing above a uniform lifted condensation level (LCL). Background photo taken at Las Brujas, Cuba, in February 2014.
Oceans 05 00020 g001
Figure 2. Observed upper 2000 m ocean heat content trend (i.e., energy difference divided by observation period) from 1958 to 2022 (WMO 2024) [22]. Image reproduction permitted by WMO Copyright. Image modified here by indicating regions of strongest cloudiness with 1, 2, and 3, as shown in Figure 3.
Figure 2. Observed upper 2000 m ocean heat content trend (i.e., energy difference divided by observation period) from 1958 to 2022 (WMO 2024) [22]. Image reproduction permitted by WMO Copyright. Image modified here by indicating regions of strongest cloudiness with 1, 2, and 3, as shown in Figure 3.
Oceans 05 00020 g002
Figure 3. Global distribution of cloudiness for July 2002–April 2015. NASA Earth Observatory image by Jesse Allen and Kevin Ward, using data provided by the MODIS Atmosphere Science Team, NASA Goddard Space Flight Center, https://earthobservatory.nasa.gov/images/85843/cloudy-earth, accessed on 10 March 2024. Image reproduction permitted by NASA Copyright. Image modified here by adding a legend and indicating regions of strongest cloudiness as 1, 2, and 3 for comparison with Figure 2.
Figure 3. Global distribution of cloudiness for July 2002–April 2015. NASA Earth Observatory image by Jesse Allen and Kevin Ward, using data provided by the MODIS Atmosphere Science Team, NASA Goddard Space Flight Center, https://earthobservatory.nasa.gov/images/85843/cloudy-earth, accessed on 10 March 2024. Image reproduction permitted by NASA Copyright. Image modified here by adding a legend and indicating regions of strongest cloudiness as 1, 2, and 3 for comparison with Figure 2.
Oceans 05 00020 g003
Figure 4. The shortwave cloud radiative effect (SW CRE, blue curve) on solar irradiation has an increasing annual trend of 0.037 W m−2 (dashed line). However, in the total cloud radiative effect (CRE, grey curve), this warming is almost completely cancelled against the cooling effect of reduced cloudiness on the ocean’s thermal radiation. This figure was kindly provided by Coda Phillips (priv. comm. 10 February 2024). Note that this figure is slightly corrected in comparison to the original publication (Phillips and Foster 2023) [26].
Figure 4. The shortwave cloud radiative effect (SW CRE, blue curve) on solar irradiation has an increasing annual trend of 0.037 W m−2 (dashed line). However, in the total cloud radiative effect (CRE, grey curve), this warming is almost completely cancelled against the cooling effect of reduced cloudiness on the ocean’s thermal radiation. This figure was kindly provided by Coda Phillips (priv. comm. 10 February 2024). Note that this figure is slightly corrected in comparison to the original publication (Phillips and Foster 2023) [26].
Oceans 05 00020 g004
Figure 5. Spectral distribution of radiative opacity between 70% and 85% of the troposphere. This figure was prepared by Robert A. Rohde for the Global Warming Art project. Reproduction permitted by Creative-Commons Copyright, https://de.m.wikipedia.org/wiki/Datei:Atmospheric_Transmission.png, accessed on 10 March 2024.
Figure 5. Spectral distribution of radiative opacity between 70% and 85% of the troposphere. This figure was prepared by Robert A. Rohde for the Global Warming Art project. Reproduction permitted by Creative-Commons Copyright, https://de.m.wikipedia.org/wiki/Datei:Atmospheric_Transmission.png, accessed on 10 March 2024.
Oceans 05 00020 g005
Figure 6. The longwave cloud radiative effect (LW CRE, orange curve) on thermal surface radiation has a decreasing annual trend of –0.038 W m−2 (dashed line). However, in terms of the total cloud radiative effect (CRE, grey curve), this cooling is almost completely cancelled by the warming effect of reduced cloudiness on solar irradiation. The figure was kindly provided by Coda Phillips (priv. comm. 10 February 2024). Note that this figure is slightly corrected in comparison to the original publication [26].
Figure 6. The longwave cloud radiative effect (LW CRE, orange curve) on thermal surface radiation has a decreasing annual trend of –0.038 W m−2 (dashed line). However, in terms of the total cloud radiative effect (CRE, grey curve), this cooling is almost completely cancelled by the warming effect of reduced cloudiness on solar irradiation. The figure was kindly provided by Coda Phillips (priv. comm. 10 February 2024). Note that this figure is slightly corrected in comparison to the original publication [26].
Oceans 05 00020 g006
Figure 7. Correlated, approximate linear long-term trends of observed global mean cloudiness (top, [23]), of solar irradiation (middle, from Figure 4) and of global mean oceanic sea surface temperature (bottom, [13]).
Figure 7. Correlated, approximate linear long-term trends of observed global mean cloudiness (top, [23]), of solar irradiation (middle, from Figure 4) and of global mean oceanic sea surface temperature (bottom, [13]).
Oceans 05 00020 g007
Figure 8. Radiative heat exchange between ocean surface and low-level cloud base, both modelled as planar black bodies in the longwave spectrum.
Figure 8. Radiative heat exchange between ocean surface and low-level cloud base, both modelled as planar black bodies in the longwave spectrum.
Oceans 05 00020 g008
Table 1. At constant surface RH of 80 %rh and pressure of 1013.25 hPa, as functions of the surface temperature T 0 and the surface dew-point temperature T d p , dry-air mass fraction A = 1 q , LCL pressure p L C L , the sensitivity of LCL pressure with respect to SST increases, LCL temperature T L C L , and the sensitivity of that temperature with respect to increasing SST are reported in the columns. The values result from iteratively solving TEOS-10 Equations (34)–(36). Differences p L C L , T 0 and T L C L are obtained from the lower minus the upper row of the related values, so that their signs reflect the climatic warming trend. Climatic sensitivities α, γ and β of A , p L C L and T L C L , respectively, are defined in Equations (87) and (92) and computed from Equation (94), as shown in Section 5. All values are calculated with full TEOS-10 accuracy and rounded only upon printing. The row printed in bold approximates the current global mean SST (Figure 7).
Table 1. At constant surface RH of 80 %rh and pressure of 1013.25 hPa, as functions of the surface temperature T 0 and the surface dew-point temperature T d p , dry-air mass fraction A = 1 q , LCL pressure p L C L , the sensitivity of LCL pressure with respect to SST increases, LCL temperature T L C L , and the sensitivity of that temperature with respect to increasing SST are reported in the columns. The values result from iteratively solving TEOS-10 Equations (34)–(36). Differences p L C L , T 0 and T L C L are obtained from the lower minus the upper row of the related values, so that their signs reflect the climatic warming trend. Climatic sensitivities α, γ and β of A , p L C L and T L C L , respectively, are defined in Equations (87) and (92) and computed from Equation (94), as shown in Section 5. All values are calculated with full TEOS-10 accuracy and rounded only upon printing. The row printed in bold approximates the current global mean SST (Figure 7).
T 0
K
T d p
K
A
%
α
% K−1
p L C L
hPa
p L C L T 0 γ
hPa K−1
T L C L
K
T L C L T 0 β
K K−1
286282.63399.2655−0.0483963.093−0.2757−0.2742281.8830.96320.9634
288284.58099.1631−0.0542962.542−0.2773283.8100.9629
290286.52699.0482−0.0608961.984−0.2823−0.2806285.7350.96210.9624
292288.47198.91960.0680961.4190.2841287.6590.9619
294290.41698.7758−0.0759960.847−0.2897−0.2878289.5830.96110.9614
296292.36198.6154−0.0846960.268−0.2917291.5050.9608
298294.30598.4368−0.0942959.680−0.2981−0.2959293.4260.96000.9603
300296.24898.2381−0.1047959.084−0.3004295.3460.9597
Table 2. Iterative solutions for the rigorous Equations (34)–(36), TEOS-10 values of LCL height, z L C L z 0 , computed from Equation (59), and the LCL coefficient, γ L C L = z L C L z 0 / T 0 T d p , computed from Equation (60) at a surface pressure of 1013.25 hPa and marine relative fugacity of ψ f = 80 % r h for selected SST values, T 0 . The row printed in bold approximates current global mean SST.
Table 2. Iterative solutions for the rigorous Equations (34)–(36), TEOS-10 values of LCL height, z L C L z 0 , computed from Equation (59), and the LCL coefficient, γ L C L = z L C L z 0 / T 0 T d p , computed from Equation (60) at a surface pressure of 1013.25 hPa and marine relative fugacity of ψ f = 80 % r h for selected SST values, T 0 . The row printed in bold approximates current global mean SST.
T 0
K
ψ f
%rh
T d p
K
A
%
z L C L z 0
m
γ L C L
m K−1
28680282.63399.2655423.468125.778
28880284.58099.1631431.481126.157
29080286.52699.0482439.660126.553
29280288.47198.9196448.017126.967
29480290.41698.7758456.561127.403
29680292.36198.6154465.305127.862
29880294.30598.4368474.263128.345
30080296.24898.2381483.449128.857
Table 3. Iteratively solving the rigorous Equations (34)–(36), TEOS-10 values of LCL height, z L C L z 0 , computed from Equation (59), and LCL coefficient, γ L C L = z L C L z 0 / T 0 T d p from Equation (60) at surface pressure of 1013.25 hPa and SST of T 0 = 292   K for selected relative fugacities ψ f . The row printed in bold approximates current global mean SST. Note the significant sensitivity of the LCL height on RH, affecting the LCL temperature, T L C L , as well as the related downward thermal radiation flux, σ S B T L C L 4 .
Table 3. Iteratively solving the rigorous Equations (34)–(36), TEOS-10 values of LCL height, z L C L z 0 , computed from Equation (59), and LCL coefficient, γ L C L = z L C L z 0 / T 0 T d p from Equation (60) at surface pressure of 1013.25 hPa and SST of T 0 = 292   K for selected relative fugacities ψ f . The row printed in bold approximates current global mean SST. Note the significant sensitivity of the LCL height on RH, affecting the LCL temperature, T L C L , as well as the related downward thermal radiation flux, σ S B T L C L 4 .
T 0
K
ψ f
%rh
T d p
K
A
%
z L C L z 0
m
γ L C L
m K−1
T L C L
K
σ S B T L C L 4
W m−2
29274287.26299.0012600.040126.632286.182380.348
29276287.67498.9740548.289126.745286.685383.030
29278288.07798.9468497.632126.857287.177385.668
29280288.47198.9196448.017126.967287.659388.263
29282288.85798.8923399.396127.076288.131390.818
29284289.23598.8650351.724127.184288.594393.334
29286289.60498.8378304.959127.291289.048395.812
29288289.96698.8105259.061127.396289.493398.255
Table 4. Comparison of LCL temperatures and pressures of the rigorous TEOS-10 solutions of Equation (34) with the Clausius–Clapeyron approximations, Equations (80) and (83). The row printed in bold approximates the current global mean SST.
Table 4. Comparison of LCL temperatures and pressures of the rigorous TEOS-10 solutions of Equation (34) with the Clausius–Clapeyron approximations, Equations (80) and (83). The row printed in bold approximates the current global mean SST.
T 0
K
T L C L
K
T L C L ( 0 )
K
p L C L
hPa
p L C L ( 0 )
hPa
286281.883281.883963.093963.066
288283.810283.810962.542962.525
290285.735285.735961.984961.972
292287.659287.658961.419961.395
294289.583289.584960.847960.858
296291.505291.506960.268960.276
298293.426293.428959.680959.712
300295.346295.348959.084959.118
Table 5. At a given surface RH of ψ f = 80 % r h and pressure of p 0 = 101325   P a , as functions of the SST and T 0 , the columns report upward thermal radiation flux, J = σ S B T 0 4 ; LCL temperature, T L C L ; downward thermal radiation flux, J = σ S B T L C L 4 ; upward net flux, J E = J J ; and the estimated sensitivity of the net flux with respect to increasing SST. The row printed in bold approximates the current global mean SST (Figure 7).
Table 5. At a given surface RH of ψ f = 80 % r h and pressure of p 0 = 101325   P a , as functions of the SST and T 0 , the columns report upward thermal radiation flux, J = σ S B T 0 4 ; LCL temperature, T L C L ; downward thermal radiation flux, J = σ S B T L C L 4 ; upward net flux, J E = J J ; and the estimated sensitivity of the net flux with respect to increasing SST. The row printed in bold approximates the current global mean SST (Figure 7).
T 0
K
J
W m−2
T L C L
K
J
W m−2
J E = J J
W m−2
J E T 0
286379.381281.883358.00621.3750.419
288390.105283.810367.89322.213
290401.055285.735377.97723.0780.447
292412.233287.659388.26323.971
294423.644289.583398.75124.8930.476
296435.290291.505409.44425.846
298447.174293.426420.34526.8290.508
300459.300295.346431.45527.845
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Feistel, R.; Hellmuth, O. TEOS-10 Equations for Determining the Lifted Condensation Level (LCL) and Climatic Feedback of Marine Clouds. Oceans 2024, 5, 312-351. https://doi.org/10.3390/oceans5020020

AMA Style

Feistel R, Hellmuth O. TEOS-10 Equations for Determining the Lifted Condensation Level (LCL) and Climatic Feedback of Marine Clouds. Oceans. 2024; 5(2):312-351. https://doi.org/10.3390/oceans5020020

Chicago/Turabian Style

Feistel, Rainer, and Olaf Hellmuth. 2024. "TEOS-10 Equations for Determining the Lifted Condensation Level (LCL) and Climatic Feedback of Marine Clouds" Oceans 5, no. 2: 312-351. https://doi.org/10.3390/oceans5020020

APA Style

Feistel, R., & Hellmuth, O. (2024). TEOS-10 Equations for Determining the Lifted Condensation Level (LCL) and Climatic Feedback of Marine Clouds. Oceans, 5(2), 312-351. https://doi.org/10.3390/oceans5020020

Article Metrics

Back to TopTop