You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272785359

Reversible and irreversible heat transfer by radiation

Article  in  European Journal of Physics · February 2015


DOI: 10.1088/0143-0807/36/3/035001

CITATION READS

1 572

2 authors:

Fernando Del Río Sara M T de la Selva


Universidad Autónoma Metropolitana, Mexico DF Metropolitan Autonomous University
103 PUBLICATIONS   1,258 CITATIONS    26 PUBLICATIONS   129 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Non-conformal soft-spheres View project

All content following this page was uploaded by Sara M T de la Selva on 03 March 2015.

The user has requested enhancement of the downloaded file.


Home Search Collections Journals About Contact us My IOPscience

Reversible and irreversible heat transfer by radiation

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2015 Eur. J. Phys. 36 035001

(http://iopscience.iop.org/0143-0807/36/3/035001)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 148.206.45.5
This content was downloaded on 20/02/2015 at 19:02

Please note that terms and conditions apply.


European Journal of Physics
Eur. J. Phys. 36 (2015) 035001 (20pp) doi:10.1088/0143-0807/36/3/035001

Reversible and irreversible heat transfer by


radiation
Fernando del Río1 and Sara María Teresa de la Selva
Departamento de Física Universidad Autónoma Metropolitana—Iztapalapa, Mexico
DF, Mexico

E-mail: fdr@xanum.uam.mx and tere@xanum.uam.mx

Received 22 July 2014, revised 17 December 2014


Accepted for publication 14 January 2015
Published 19 February 2015

Abstract
The theme of heat transfer by radiation is absent from most textbooks on
thermodynamics, and its treatment in the applied literature presents some basic
discrepancies concerning the validity of the Clausius relation between the
quantity of heat exchanged, δQ, and the accompanying entropy change, dS. We
review the reversible and irreversible heat transfers by radiation to clarify the
validity of the Clausius relation, and we show that in both cases, the Clausius
relation is obeyed, as it should be. We also deal with radiation diluted by the
presence of matter, introducing a dilution coefficient, ϕ, and an irreversibility
factor, χ (ϕ). This treatment requires the use of the correct relation between
energy and heat fluxes, the spectral fluxes of energy and entropy, and Planck’s
equation for the entropy of monochromatic radiation. For the irreversible case of
diluted radiation, we recover the ratio between the fluxes of heat and entropy
that agree with Clausius’ inequality, including an irreversibility factor,
(4 3) χ (ϕ). An improved modification for the explicit function χ (ϕ) is given.
As an illustration, the fluxes of energy and entropy from the Sun to the Earth are
obtained. We also calculate the fluxes re-emitted by the Earth, taking into
account the greenhouse effect. We find the value of 1.258 Wm−2K−1 for the re-
emitted entropy flux after the radiation has been thermalized, which is much
larger than the incident flux, in agreement with other authors.

Keywords: radiation, heat, entropy, reversible, irreversible, black body

1. Introduction

Radiation is universally considered a form of heat transfer with an associated energy flux. But
a body that radiates only heat, without any other change, will experience a decrease in
1
Author to whom any correspondence should be addressed.

0143-0807/15/035001+20$33.00 © 2015 IOP Publishing Ltd Printed in the UK 1


Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

entropy. Since according to the second law of thermodynamics, the entropy of the universe
cannot decrease, radiation must necessarily carry entropy in an amount at least equal to the
loss of entropy of the radiating body. As a matter of fact, the subject of the entropy of
radiation has a long history that can be traced back to its mention by Boltzmann in 1886, the
work of Wien in 1894, and, shortly after, the fundamental work of Planck [1–3].
Treatment of heat transfer by radiation is a topic almost absent in classical thermo-
dynamics texts, while it is an ever-present subject in the applied literature, mainly in engi-
neering and geophysics. However, not only do different authors use different relations
between the quotient formed by the transferred heat and absolute temperature, δQ T , and the
differential change in entropy, dS, but discrepancies are often present regarding the applic-
ability of Clausiusʼs relation,
δQ
dS ⩾ , (1)
T
where the equality sign applies to reversible processes and the strict inequality symbol applies
to irreversible ones, and where T is the absolute temperature of the source of heat [4, 5].
Confusion prevails regarding its validity in the case of radiation.
This confusion can be illustrated for a black body radiating freely into empty space. In
this case, the radiation process is irreversible, so that the ‘greater than’ symbol of Clausiusʼs
relation must be applied, and also the flux of heat must be recognized as equal to that of the
energy. Nevertheless, some authors consider that the fluxes of entropy and energy in this case
differ by a factor of 1 T only, where T is the black-body temperature [6–8]. This implies the
use of the equals sign in the Clausius relation, and therefore assumes that the process of
radiation emission by the black body is reversible. Other authors correctly consider a factor of
4 3T between those two fluxes, thus complying with Clausiusʼs strict inequality for an
irreversible process [9–11]. Some years ago, Wright published an article that partially made
this point [12]. We would like to present it here in a way that is more accessible to physics
students, and to also address the essential distinction between the reversible and irreversible
transfer of heat.
Since the Clausius relation embodies the second law of thermodynamics, we thought it
would be helpful to clarify matters, starting with the simple example of reversible heat
transfer by radiation from a reservoir into an quasistatically expanding empty cavity. In this
case, heat transfer complies with the Clausius equality. This is discussed in section 2.1.
In a second step, explored in section 2.2, we consider the irreversible heat transfer by the
free flow of radiation from the source into empty space, and show that it complies with the
Clausius strict inequality.
In a next step, developed in sections 2.3 and 2.4 , we include several possible sources of
irreversibility, coming from the interaction of radiation with matter within one global factor
for the spectral radiance, following the idea of Landsberg and Tonge [9], but with an
empirically calculated factor. These sources of irreversibility include scattering and absorp-
tion and give rise to what has been termed ‘diluted’ radiation. We find the relation among the
diluted fluxes of energy and entropy where the irreversibility factor appears as a modification
of the relation for empty space reinforcing the Clausius inequality.
Finally in section 3, we use the transfer of heat by radiation from Sun to Earth as an
example. The calculation is made of all the fluxes. We also take into account the radiation
reflected and re-emitted by the Earth. We consider the various sources of irreversibility
included in a global irreversibility factor. Our results agree completely with those recom-
mended by Wu and Liu [13], who have published a detailed critical review of the subject in
its application to calculate the energy and entropy balances of the Earthʼs system. We thus

2
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 1. Radiation heat transfer from a BB. (a) Reversible transfer into a cavity.
(b) Irreversible transfer into empty space.

find that the radiation emitted by the Earth carries away a much larger amount of entropy than
what it receives from the Sun.
The net entropy radiated by the Earth is sufficient to maintain a steady state, and therefore
to get rid of the excess entropy produced by all irreversible processes on Earth, including
those occurring within living beings. The subject of entropy and life has been a hot topic for
many years, but here we only recall the idea developed by Schrödinger years ago [14]. The
main point is that living organisms sustain their steady state by increasing the entropy of their
surroundings. The question of whether this process gives rise to measurable effects has been
considered by Ulanowicz and Hannon [15].
Our aim is to make this work as self-contained as possible, so a few references are made
to standard textbooks at the undergraduate level. Planck’s formula for the entropy of
monochromatic photons, which is necessary for this subject, is not usually treated at that
level; a brief derivation is thus given in the appendix.

2. Heat radiation

2.1. Reversible transfer of heat by radiation

We review first the calculation of the change in entropy when transferring an amount of heat
radiation in a reversible expansion.
For a heat reservoir, we consider a black body (BB), in equilibrium at temperature T. An
evacuated cavity, also filled with radiation, is in radiative contact with the BB through a
window perfectly transparent to radiation of all wavelengths. The cavity is cylindrical, with

3
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

one end being the window open to the BB and the other end closed by a piston that can be
moved in or out, as seen in figure 1(a).
The interaction between the cavity and its surroundings is determined by the nature of its
walls; the piston and the wall are assumed to be adiabatic and 100% reflecting. Radiation
within the cavity should be in equilibrium and isotropic, which is not true for radiation
leaving the BB. Hence, the cavity walls have to be diffusive (i.e., not specular).
The cavity is initially also at temperature T. For the radiation inside it, Maxwell equations
predict a pressure related to the energy density, e ≡ E V , as [16]
1
prad =e. (2)
3
From this equation, thermodynamics recovers Stefan’s experimental result:
prad = bT 4, (3)
where b = 4σSB 3c , c is the speed of light and the Stefan–Boltzmann constant is
σSB = π 2k 4 60 3c 2 . (4)
We shall use b or σSB alternatively, to keep the formulae simpler. From (3) and (4), we have
4σSB 4
e = 3bT 4 = T . (5)
c
We consider an isothermal expansion of the cavity with a change, dV, in its volume.
Since the expansion is adiabatic, the newly exposed walls do not emit new photons into the
cavity. While the change, dV, takes place, T is kept constant by the flux of energy from the
BB reservoir. In this way, the cavity reversibly absorbs a net quantity of heat, δQ, from the
BB, and the radiation in the cavity exerts an amount of work, δW .
Due to the first law, in expanding, the cavity would change its energy by
dE = δQ − δW . Thus, the heat transferred from the BB to the cavity is
δQ = dE + δW , (6)
and it becomes clear that δQ provides for both the energy needed to keep T and the work done
against the piston. Since T is constant, the energy density, dE dV , must be equal to e given in
(5),
dE = 3bT 4 dV . (7)
Now, given the pressure, prad, on the piston’s surface and the expanding volume dV, an
amount of work, δW = prad dV , is performed. That is, from (3)
δW = bT 4 dV . (8)
From equations (6)–(8), the heat transferred from the BB to the cavity is
δQ = 4bT 4 dV , (9)
or
dE = 3δQ 4; (10)
that is, only 3/4 of the entrant heat becomes internal energy since the remaining 1/4 of it is
used to do work.
We now calculate the amount of entropy, dS, gained by the cavity. Wien was the first to
study the entropy associated with radiation in 1894 [2], though Boltzmann had previously
mentioned the entropy in connection to radiation in 1886 [1]. This quantity is obtained from

4
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

the fundamental Gibbs equation


T dS = dE + prad dV , (11)
which, using (8) and (9), gives
dS = 4bT 3dV . (12)
Now, from (12) and (9), we recover the Clausius equality valid for reversible transfer:
dS = δQ T . (13)
Thus, in this respect, heat transfer by radiation is no different from heat transfer by
conduction! Actually, in his treatment of Wien’s work on the entropy of radiation, in a 1905
paper, Einstein explicitly assumes the validity of the Clausius equality (13) for radiation in the
case where the volume is kept constant and the heating is reversible [17].
We look now at the entropy balance. Taking into account the corresponding entropy
change of the BB source, due to the loss of heat − δQ, which, being a reservoir, equals
dSBB = −δQ T , we find the change of entropy in the Universe as dS + dSBB = 0 , as it should
be in this case.
Furthermore, with the help of (10) in (13), we find
4 dE
dS = . (14)
3 T
It is this relation that has given rise to some confusion in the literature, because if one
erroneously identifies the internal energy change, dE, as the heat δQ transferred from the BB,
one would find that in this reversible process, the transferred entropy would equal 4δQ 3T .
Instead, as we have seen by (10), the transferred heat is 4/3 of the internal energy increase,
thus leading back to the Clausius equality (13).
Notice that equation (14), integrated at constant T, is none other than the relation between
the energy and the entropy of the BB. Below, we shall need this relation written in terms of
densities,
4e
s= . (15)
3T

2.2. Irreversible transfer of heat by radiation

We now consider the irreversible transfer of heat from the BB in equilibrium at a temperature,
T, into its surroundings. We speak of the BB as the source of the radiation, and we assume
that the BB stays in equilibrium while the transfer occurs. This means that the amount of
energy radiated during the observation time is negligible with respect to the total energy of the
BB, or that there is a source of energy within the BB that keeps a steady state, as nuclear
reactions do in the Sun. In other words, this BB is considered to be a thermal reservoir.
Further on we shall use the Sun as an overall important example of such a BB, in which case
the boundary of the BB is held in place by gravity. Alternatively, we take the BB surrounded
by an adiabatic, perfectly reflecting wall except for a small window that is transparent to
radiation. We assume that the space surrounding the BB is a vacuum filled with equilibrium
radiation at a temperature, TV. Hence, the heat transfer from the BB in this case may be
considered to be a free expansion of radiation. Furthermore, we consider TV ≪ T , so that the
radiation from the surrounding space back into the BB is negligible. Actually, this condition
is not strictly necessary and is introduced here for simplicity; it will be removed in the
example in section 4.

5
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 2. Element of solid angle at the boundary of the source of radiation.

2.2.1. Fluxes of energy. We first consider the transfer of energy through the boundary of the
BB. From the definition of energy density, and given that the equilibrium radiation is
isotropic, the directional energy density (per unit solid angle at and around a fixed direction of
propagation, k) is just
ek = e 4π . (16)
Since e is taken at the inside of the boundary of the BB, it can be calculated with the
equilibrium result (5).
Then, because c is the speed of the radiation, the emitted energy per unit time, unit area,
and unit solid angle in a given direction, k , is simply given by
Ik = cek . (17)
Following the terminology of Nicodemus [18] and of Landsberg and Tonge [9], we call
this quantity, Ik , the radiance.
We want to obtain the energy crossing the boundary per unit area and unit time—that is,
the energy flux, Je (also called irradiance), in the direction normal to the boundary of the BB.
Let dΩ be the differential of the solid angle (i.e., the solid angle subtended by a narrow pencil
of radiation crossing the surface, as seen in figure 2). The normal component of the velocity
of the radiation with direction k at an angle θ is c cos θ . The total energy flux, Je, crossing the
surface is then obtained by integrating Ik cos θ dΩ over the solid angle. In the ideal BB case
we are considering, it is reasonable to assume that the radiation is isotropic, so that Ik is angle
independent. So we write the final answer as
Je = BIk, (18)
with

B= ∫ cos θ dΩ . (19)

We now calculate the integral in (19) for emission by the BB. The radiation with
direction k between zenithal angles θ and θ + dθ and all azimuthal angles subtends a solid
angle, dΩ = 2π sin θ dθ , as seen in figure 2. The integral can be calculated easily. At the BB
boundary, emitted radiation covers one whole hemisphere (0 ⩽ θ ⩽ π 2), in which case we
simply have B = π , and so the energy flux is given by

6
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Je0 = πIk. (20)


(We shall use the superindex 0 to identify quantities at the surface of the BB). From here,
using (5), (15), and (17), we recover the famous Stefan–Boltzmann law:

Je0 = σSB T 4. (21)


More generally, for the flux, Je, at a point away from the BB, we have to take into
account two main facts. The first is that now the solid angle in play is the one subtended by
the finite source at the point of observation, and with this caveat in regard to (19), we can still
use (18). The second fact is that the radiance, Ik , in (18) remains constant as the radiation
propagates in empty space. This constancy is a consequence of the conservation of energy
and was already considered by Planck. (For a modern derivation, see [18].) Thus, the energy
flux is given by (20), but with the factor, π, replaced in general by B. This result can be
expressed in terms of the flux at the boundary of the BB as
Je = (B π ) Je0. (22)
We notice that as the surface is moved away from the source, the solid angle diminishes and
Je decreases by the geometrical factor, B π . An explicit example of this calculation is given in
section 4 below.

2.2.2. Flux of entropy at the BB boundary. We now turn to look at the flux of entropy at the
boundary of the BB, J0s , where s can be taken as that at equilibrium (15). In analogy with the
procedure used for the energy, we now obtain a directional entropy density (per unit solid
angle at and around a fixed direction, k ): sk = s 4π . The entropy radiance along k at the BB
boundary is then given by ℓk0 = c sk = c s 4π , which can be written in terms of the energy
radiance, Ik , by using equations (15)–(17) as
ℓk0 = (4 3T ) Ik, (23)
where T is the temperature of the BB source. With the same arguments just used for the
energy flux, we now calculate the entropy flux by integrating ℓk cos θ dΩ over the solid angle
to find
Js0 = Bℓk0. (24)
At the boundary, we have B = π , so with (20) we arrive at
4 0
Js0 = Je . (25)
3T
This is none other than (15) in terms of fluxes.
Now comes a crucial point. Since no work is performed at the emergence of radiation out
of the BB into empty space, and taking into account the first law, all the energy leaving the
BB can be considered as heat. This means that the amount of heat, δQ, crossing an area, dA,
in time, dt, is simply Je0 dAdt , while the amount of entropy associated with the emitted
radiation is dS = Js0 dAdt . From (25), these quantities are then related by
4
dS = δQ , (26)
3T
which satisfies the Clausius strict inequality, dS > δQ T , which is valid for irreversible
processes. Considering the entropy balance of the original BB source together with the
radiation field during the time dt, we find that dSBB + dS = δQ 3T > 0. We have found a first
cause of irreversibility: the emergence of radiation from the BB into free space without the

7
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

performance of work. Radiation away from the boundary and other causes of irreversibility
are considered next.
A very different situation is the radiative heat transfer between two sources at somewhat
different temperatures, located a very short distance from each other [19, 20]. This interesting
point will be explained very briefly at the end of the article.

2.3. Diluted BB radiation

To examine other irreversible energy and entropy transfers, we consider the case where the
radiance is reduced due to the interaction of radiation with matter. This is usually referred to
in the literature as ‘diluted’ BB radiation. The cause of dilution can be either absorption,
imperfect emission, or scattering by a diffusive medium. Without specifying the precise cause
and without loss of generality, we shall assume a uniform dilution, which means that the
radiance is reduced by the same factor at all frequencies. To obtain the effect of the dilution
on the entropy of radiation, we must first look at its spectral (or frequency) components.
In a given state, x = (k, p), of definite wave vector k and polarization p , let the mean
occupation number of photons be given by n x . Focusing on the frequency interval
(ν, ν + dν ), we consider the density of states, ρ (x), per unit volume of radiation, with a wave
vector along a given direction (per unit solid angle) that is given by
ρ (x) = 2ν 2 c3; (27)
this is 1 4π the isotropic density of states [21]. Since a photon has energy eν = hν , then the
spectral directional energy density, eν, k , per unit volume, unit solid angle, and interval of
frequency, is given by hνρ (x) n x . Thus, using equation (27),
2h 3
eν, k =ν n x. (28)
c3
In equilibrium, the mean occupation number is given by the well-known Planck distribution
as [21]
1
n xeq = , (29)
ehν kT −1
where h and k are the Planck and Boltzmann constants, respectively.
The energy in (28) includes radiation with the two possible polarizations; integration over
all frequencies leads to the directional energy density, ek . In the equilibrium case, integration
over all directions of propagation and use of (29) again gives the classic BB result in (5).
The spectral radiance (energy transmitted per unit time, unit frequency interval, and unit
solid angle in a given direction), recalling that c is the speed of radiation, is simply given by
Iν, k = c eν, k , and from (28) we find
2h 3
Iν, k = ν n x. (30)
c2
The mean occupation number, n x , remains constant for free propagation in a vacuum; in
particular, this holds for the BB radiation away from its boundary where n x maintains its
equilibrium value (29). However, when radiation interacts dissipatively with matter, n x is no
longer conserved. This happens, for instance, in absorption, imperfect emission, or partial
reflection, when we can assume that n xeq is reduced by a factor, ϕ ⩽ 1 [9]. In the literature,
this type of radiation is called diluted BB radiation. For diluted radiation, the equilibrium
spectral radiance, Iνeq,k , becomes reduced by the same factor. For simplicity, we assume that ϕ
is the same constant for all frequencies, thus keeping the profile of the distribution. This

8
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

means that,
n xdil = ϕn xeq and Iνdil eq
,k = ϕIν,k, (31)
and the total diluted radiance is

Ikdil = ∫0 Iνdil
k dν . (32)

Of course, the total (diluted) radiance and the total (diluted) energy flux, in a direction normal
to the original BB surface, are then reduced by that same factor to become Ikdil = ϕIk and
Jedil = ϕJe , respectively. These two quantities are related by the equivalent of (18) that now
takes the form
Jedil = B ϕIk. (33)
From this equation follows a relation useful for emission by a so-called ‘gray body’, which is
a body similar to a BB but with an emissivity, ϵ < 1, playing the role of the dilution
coefficient, ϕ. Recalling (18) in (33) and setting B = π for the radiation at the boundary of the
gray body, we find the simple result
Je0,gray = ϵJe0. (34)
Next, we consider the fluxes of entropy away from the boundary of the BB.

2.4. Entropy of the diluted BB radiation

Radiation implies an entropy flux associated with a diluted energy flux. We seek an
expression for this diluted entropy flux in terms of the spectral radiance, Iν k .
The entropy of the photons in the state x can be obtained in quite a general way, even
away from equilibrium, by a simple procedure due to Landsberg [22]. Since this derivation is
not available in standard text books on statistical mechanics, we present it in the appendix.
The desired result is
sx = k ⎡⎣ ( 1 + n x ) ln ( 1 + n x ) − n x ln n x ⎤⎦ . (35)
This same formula was presented earlier by Slater [23]. Recently, Piña and de la Selva [24]
gave an equivalent equilibrium result for sx . Note sx is not a linear function of n x .
The spectral entropy radiance, ℓν k , is defined per unit volume of radiation with frequency
in (ν, ν + dν ), per unit frequency interval, and per unit solid angle along k . Using the density
of photon states, ρ (x), the entropy radiance is written cρ (x) sx , which becomes
2 2
ℓν k =ν sx . (36)
c2
Now, solving for n x in terms of Iν k in (30):
c 2Iν k
nx = . (37)
2hν 3
We reach the desired result by direct substitution of the last equation into equations (35) and
(36):

2k 2 ⎡⎢ ⎛ c 2Iν k ⎞ ⎛ c 2Iν k ⎞ c 2Iν k c 2Iν k ⎤



ℓν k = ν ⎜1 + ⎟ ln ⎜ 1 + ⎟− ln (38)
c 2 ⎢⎣ ⎝ 2hν 3 ⎠ ⎝ 2 hν 3 ⎠ 2 hν 3 2hν 3 ⎥⎦

9
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

This expression was first derived by Planck [3] for equilibrium as early as 1906. It
permits the calculation of the entropy radiance and the associated entropy flux for any
radiation field whenever the radiance, Iν, k , is known. The same result was re-derived by
Rosen [25] and by Ore [26], who discussed earlier derivations. A derivation valid for non-
diluted radiation in nonequilibrium situations was presented by Landsberg [27].
The relation between ℓν k and Iν, k is not linear; since for diluted radiation the spectral
radiance is ϕIνeqk , the corresponding spectral entropy radiance results in

2k 2 ⎡⎢ ⎛ c 2ϕIνeqk ⎞ ⎛ c 2ϕIνeqk ⎞ c 2ϕIνeqk c 2ϕIνeqk ⎤


⎥ . (39)
ℓνdil
k (ϕ ) = ν ⎜1 + ⎟ ln ⎜ 1 + ⎟− ln
c 2 ⎢⎣ ⎝ 2hν 3 ⎠ ⎝ 2hν 3 ⎠ 2hν 3 2hν 3 ⎥⎦
This is a central result: the relation between diluted entropy radiance, ℓνdil k , and diluted
energy radiance, ϕIνeqk , is the same as it is for nondiluted radiation (38). The bases of this result
are the validity of (36) and (37) in nonequilibrium situations.
Now in (39), we can use (29) and (30) to write ℓνdil k in terms of βhν and make one further
change of variable from ν to ζ = βhν to obtain
2k 4T 3 2 ⎡ ⎛ ϕ ⎞ ⎛ ϕ ⎞
ℓνdil
k (ϕ ) = ζ ⎢ ⎜1 + ζ ⎟ ln ⎜ 1 + ⎟
2
ch 3 ⎣ ⎝ e −1 ⎠ ⎝ ζ
e − 1⎠
ϕ ⎛ ϕ ⎞ ⎤
− ζ ln ⎜ ⎟ ⎥. (40)
e − 1 ⎝ eζ − 1 ⎠ ⎦
In similarity to the derivation of (24), the total entropy flux of the diluted radiation is now
obtained by integrating ℓνdil dil
k cos θ dΩ over all frequencies and solid angles. Since ℓν k is angle
independent, we again obtain a solid angle factor of B π ,
B ∞
Jsdil (ϕ) =
π
∫0 ℓνdil
k (ϕ)dν . (41)

Substitution of (40) in (41) allows us to use the quantity introduced in [9],

45 ⎡⎛ ∞ ϕ ⎞ ⎛ ϕ ⎞
χ (ϕ ) ≡ ζ2 ⎢ ⎜ 1 + ζ
∫0 ⎟ ln ⎜ 1 + ⎟
4π 4ϕ ⎣⎝ e −1 ⎠ ⎝ ζ
e − 1⎠
ϕ ⎛ ϕ ⎞ ⎤
− ζ ln ⎜ ⎟ ⎥dζ , (42)
e − 1 ⎝ eζ − 1 ⎠ ⎦

which, employing (4), leads to

4 ⎛⎜ B ⎞⎟
Jsdil = ϕ χ (ϕ) σSB T 3. (43)
3⎝ π⎠

A plot of χ (ϕ) is given in figure 3. It can be shown by numerical integration that


χ (ϕ = 1) = 1. Notice that the solid angle factor, B π , does not affect the dilution coefficient,
ϕ, in the argument of the function χ (ϕ). In terms of the corresponding diluted energy flux
(33), and using (21) and (20),
4 J dil
χ (ϕ ) e .
Jsdil = (44)
3 T
We first discuss the free propagation of radiation away from the boundary of the BB. In
this case ϕ = 1, so that from (44),

10
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 3. Entropy irreversibility factor, χ, as function of the dilution coefficient, ϕ. The


line represents the numerical integration in (42).

4
Js = Je. (45)
3T
We can see that the fluxes of energy and entropy away from the (finite) BB boundary keep the
same ratio as at that boundary, as seen in (25). This means that the entropy flux decreases by
the same geometrical factor, B π , as the energy flux. Therefore, no additional increase of
entropy occurs, and the process of free propagation once away from the BB is reversible, as
pointed out by Cailles and Herbert [28].
For diluted radiation, however, the ratio of the flux of entropy to that of energy does not
remain constant at 4 3T , but incorporates the extra factor χ (ϕ), as seen in (44). This form for
the entropy flux was first obtained by Landsberg and Tonge [9].
The irreversibility factor defined in (42) is shown as a function of ϕ in figure 3. It is
convenient to have this function, χ (ϕ), in closed form. Landsberg and Tonge [9] found an
explicit expression involving infinite series, which for small 0 < ϕ ⩽ 0.1 can be approxi-
mated by a closed formula. Alternative approximations have been proposed by Stephens and
O’Brien [29] and by Wright et al [30]. Here, we give the following empirical modification of
the formula proposed by Landsberg and Tonge [9]:

χ = 1 − 0.0348(1 − ϕ) − 0.277 ln (ϕ). (46)

One notices that the logarithmic term dominates the linear term, so that χ ⩾ 1 for ϕ ⩽ 1, as
shown in figure 3. The factor χ (ϕ) measures the increase in the entropy flux due to the
irreversibility of the dilution of the radiation, so we call it the irreversibility factor. Thus, the
ratio of entropy flux to energy flux is greater for diluted radiation than for undiluted radiation.
We should mention that in one of the earliest treatments of diluted radiation, Petela [31–33]
considers the case of a radiating ‘gray body’, which is a body with mean emissivity, ϵ, whose
energy radiance is ϵIν k . In his treatment, Petela assumes that the diluted entropy flux equals
the entropy flux of the BB multiplied by ϵ, which plays the role of the dilution coefficient, ϕ.
In terms of equations (43) and (44), this means that the dilution process is taken as reversible
—that is, with χ = 1.
Again, since no work has been done by the radiation in this process, from (44) we get the
relation that holds between the heat transferred, δQ, and the associated entropy change, dS:

11
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 4. Percent deviation, 100 δχ (ϕ) χ , of different approximations to the


irreversibility factor. Dotted line: Landsberg–Tonge formula [9]. Solid line:
equation (46). Short-dashed line: approximation of Stephens and O’Brien [29].
Dotted-dashed line and long-dashed line: first and second approximations of Wright
et al [30], equations (10) and (13) in [12].

4
dS = χ (ϕ ) δ Q , (47)
3T
satisfying again the Clausius inequality, dS > δQ T . We remind the reader that in these
equations, T is the temperature of the BB source.
In figure 4, we show the relative difference between the corresponding approximation or
model for χ (ϕ) and the result of the numerical integration in (42),
δχ χ = (χ (model) − χ (numerical)) χ , expressed as a percentage. The figure gives the plots
of various approximations. We can see that (46) provides the best approximation over the
range of ϕ of interest. The most accurate, except for small ϕ, are those by Wright et al [30]. In
contrast, (46) has the correct limiting behavior, χ → 1, when ϕ → 0 , and has a maximum
deviation of only 0.2% on the whole interval of interest.

3. Transfer of heat by radiation from Sun to Earth

3.1. Fluxes of energy

We discuss now, as a paradigmatic application of the above, the radiation the Earth receives
from the Sun and the global balances of energy and entropy thereof. Let us call RS the Sun’s
radius and rE the distance between Sun and Earth.
The radiation leaving the Sun’s surface has a BB spectral flux characteristic of the
effective temperature, TS, which we take as TS = 5778 K [34]. The energy flux is then found
from Stefan–Boltzmann’s law (21) to be

Je0 = 6.320 × 10 7 W m−2 .

The Planck distribution per unit interval of wavelength, eλ , can be obtained directly from
(30) by using ν = c λ :

12
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 5. Solid angle for radiation from the Sun on the Earth.

2 hc 2 1
Iλ, k = , (48)
λ5 ehc λkT − 1
and has its maximum at a wavelength given by Wien’s displacement law:
λ max = 2.898 × 10−3 m K T. With T = TS , we obtain λ max = 501.5 nm . The spectral flux
at this wavelength is found most easily from the spectral equivalent to (20), which is
Je, λ = πIλ, k , as Je,max −2
λ = 82 864 Wm nm .
−1

The flux of energy at a distance, r, from the Sun’s centre, Je(r), can be calculated very
easily from the conservation of energy. Since there is no absorption or scattering in empty
space, the total energy crossing a sphere of radius r ⩾ RS is constant for any r, and we have
simply that
R S2
Je (r ) = Je0, (49)
r2
where r ⩾ RS .
It is illustrative to obtain this same result directly from (22) by calculating the solid-angle
factor, B, (19), over the solid angle subtended by the Sun at a point on the Earth’s orbit. This
can be done with the help of figure 5. Take dΩ as the solid angle element subtended on the
Sun from a point on the Earth by incrementing the angle, α, by δα and the azimuthal angle
(not shown) by δφ . This solid angle equals the subtended area, Lδφ r δα , divided by the
square of the distance, r, to the vertex. Since L = r sin α , dΩ = sin α δα δφ. Hence, the
desired factor is
αS ⎛ sin2 α ⎞α S
B = 2π ∫0 cos α sin α dα = 2π ⎜
⎝ 2 ⎠0
⎟ = π sin2 αS . (50)

But from figure 5, sin αS = RS rE , and thus the solid-angle factor at the Earth’s orbit is
2
BE = π ( R S rE ) . (51)
It is worth noting that this value of B coincides with the solid angle subtended by a flat
solar disc as viewed from the Earth, but it is not the true solid angle subtended by a spherical
Sun. This is given simply by
αS
Ω 0 = 2π ∫0 sin α dα = 2π ( 1 − cos αS ). (52)

Both angles, B and Ω0 , tend to be equal when the ratio, RS rE , is very small, as is the case for
the Sun–Earth geometry where the two solid angles differ by only one part in 105. With this
value of BE, from (22) we recover (49) by making r = rE .

13
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

Figure 6. Spectral energy flux at the top of the atmosphere. The continuous line stands
for the satelliteʼs actual data, taken from G-173 [35]. The dashed line gives the
prediction by equation (53) from the Planck distribution at the mean Earth–Sun
distance.

From standard values2 of RS and rE and (51), we obtain BE = 6.807 × 10−5 sr, or
BE π = 2.167 × 10−5. With this value of BE π and the calculated J0e , we find from (22) the
energy flux at the Earth’s orbit as JeE = 1369 W m−2 . This value is within the range of the so-
called solar ‘constant’, which has a mean value between 1365 W m−2 and 1373 W m−2 , as
measured with satellites at the top of the Earth’s atmosphere3. Taking into account that the
Earth’s area is four times its cross section, the mean flux of energy incident per unit area of its
surface is 1/4 of the total, which is

Jeinc = 342 W m−2 ,


an often-quoted value.
The spectral energy flux, Jeλ , at the top of the atmosphere is shown in figure 6, where it is
compared with the predicted flux; this is obtained from the spectral equivalent of (20) and
(22) as
Je, λ = BE Iλ, k, (53)

where we put B = BE , as given by (51); Iλk maintains a constant value as the radiation crosses
the empty space between Sun and Earth, and thus is given by (48) with T = TS . The area
below the dashed curve in figure 6 that represents equation (53) equals 1369 W m−2 ; this value
corresponds to the magnitude of TS used in this work. The flux, Jeλ , has a maximum value of
1.795 Wm−2nm−1 at λ max . From figure 6, this value is close to that actually observed at the top
of the Earth’s atmosphere.
It is important to point out that, for equilibrium BB radiation, both the profile and
magnitude of the spectral energy flux are completely fixed once the temperature is given.
However, the spectrum in the figure is not really the BB equilibrium spectrum but one
reduced in magnitude by the solid-angle factor, BE.
2
Taken here as RS = 6.963 × 108 m and rE = 1.496 × 1011 m.
3
Actually, the effective temperature of the Sun is defined so as to obtain the measured value of the solar ‘constant’,
and thus it depends on the actual set of measurements selected. According to Wu and Liu (2010), TS = 5779 K ,
although their solar constant is only 1367 W m−2 .

14
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

3.1.1. Radiation reflected from the Earth. We now look at the flux of energy of the radiation
reflected from the Earth. We assume that this reflection is dispersive, with equal intensity over
all angles (so-called Lambertian) and characterized by an albedo, α. In this approximation, we
picture the Earth reflecting part of the radiation it receives uniformly in all directions. This
type of reflection produces a dilution of the radiation. Thus the incoming radiance, Ik , is
reduced due to two factors. The first is the proportion of the radiation reflected or albedo α,
which is a component of the dilution coefficient. The second is the angular dispersion from a
solid angle, Ω0 , of the incoming radiation given by (52), to a solid angle, 4π , of the reflected
radiation; this contributes to a dilution coefficient of Ω0 4π . Hence, the reflected radiance is
Ikref = ϕ ref Ik, (54)
where the dilution coefficient for reflection is obtained from the albedo coefficient that
determines the mean proportion of radiation reflected by the Earth. For this, we use the so-
called Bond albedo, α p ,4
ϕ ref = α p ( Ω 0 4π ). (55)
The flux of energy of the reflected radiation is obtained similarly to (33), by integrating
Ikref cos θ dΩ over the solid angle of the reflected radiation. Since by assumption the reflected
radiation is also isotropic, the integral factorizes to get
Jeref = ϕ ref Ik B ref , (56)
which, using (20), leads to

(
Jeref = ϕ ref B ref π Je0. ) (57)

Measuring the reflected radiation at the boundary of the reflecting body (top of the
atmosphere), the solid-angle factor is again B ref = π , because reflection from any given
surface element on the Earth is over a hemisphere, and we get
Jeref = ϕ ref Je0. (58)
This can be written in terms of the energy flux at the Earth’s orbit, Je. We first substitute ϕ ref
from (51) into (52), and then notice that Ω0 ≅ BE as calculated in (48), but from (22),
Je = (B π ) Je0 , arriving finally at
Jeref = α p Je 4. (59)
Taking an albedo of α p = 0.29 for the Earth5 and assuming that radiation is reflected by
the whole surface, we find that
Jeref = 99 W m−2 .
The rest of the energy, Jeabs = Jeinc − Jeref is absorbed by the Earth’s surface. This gives
Jeabs = 243 W m−2 .

3.1.2. Radiation re-emitted by the Earth. To complete the energy analysis, we consider the
radiation re-emitted by the Earth after it has been thermalized. Here, we assume that the Earth
is in a steady state, meaning that there is a balance between energy absorbed and energy
4
This type of albedo takes into account radiation reflected at all wavelengths; in contrast, the visual albedo involves
only visible light. See [7].
5
Value obtained from de Pater and Lissauer 2001 [36].

15
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

emitted. Why is the effective temperature of the Earth, TE, needed to produce the observed
outgoing flux of energy? For this, we can use the Stefan–Boltzmann law (21) in (34), with
given values of the flux, Je, and the dilution coefficient, ϕ; for emission, this coefficient equals
the Earth’s emissivity, ϵ. The temperature in this formula is the effective temperature of the
emitting body—that is, TE. At this point, it is illustrative to consider various assumptions. The
first and simplest is to take the Earth as a BB (albedo = 0, emissivity = 1) that absorbs all the
energy it receives from the Sun, meaning that Je = Jeinc ; this leads to TE = 278.7 K (5.6 °C).
As a second and more realistic assumption, we take into account the albedo, α p , and use
Je = Jeabs, but still with ϵ = 1, to find, of course, a much colder Earth: TE = 255.9 K ,
(−17.3 °C).
Finally, we take into account the greenhouse effect due to clouds and some gases in the
atmosphere, and we consider the mean effective emissivity of the Earth, taken here from
various sources as ϵE = 0.612. From (34), this gives TE = 289.0 K (16.1°C), which can be
considered a realistic value for the Earth’s effective temperature. Avoidance of empirical data
as albedo and emissivity is difficult because more realistic approaches involve specific models
for many geophysical processes. The maximum of the emitted radiation is at a wavelength,
λ max , well into the microwave range; with the last value of TE, we find λ max = 10 017 nm.

3.2. Fluxes of entropy

There has been some controversy in the literature about the correct expression for the flux of
entropy. Recently, Wu and Liu have made a thorough review of the subject [13]. Here we
only quote the main issues. Some authors, such as Peixoto et al [7] and Ozawa et al [8],
assume the validity of (13) for the emission by the BB. We have seen that this equation is
applicable to reversible processes only, and therefore does not hold for the radiation of the
Sun into interplanetary space. Other authors, like Petela [31–33], employ the correct
expression, (25) or (45), for the irreversible emission of light by a BB, but use it to treat
emission by a gray body (ϕ < 1), therefore omitting the crucial irreversibility factor, χ (ϕ).
Furthermore, there are several authors, such as Stephens and O’Brien [29] and Wright et al
[30], that use the correct Planck expression for the radiance of the radiation entropy (48) and
introduce different approximations for the irreversibility factor, χ (ϕ). Finally, Kabelac and
Drake discuss the important case of polarized radiation [37].
The different fluxes of entropy can be calculated very easily from the appropriate relation
between the fluxes of entropy and energy. The flux of entropy at the Sun’s surface given by
the BB emission formula (25), with T = TS , is found to be

JS0 = 14 580 Wm−2K−1.


Let us consider the incident entropy flux at a point on the Earth’s orbit. It is given by (43), but
with the dilution coefficient ϕ = 1, because the radiance stays constant in the travel of
radiation from the Sun to the Earth. Again, the solid-angle factor contributes BE π , so that we
obtain for the incident radiation a flux of entropy, JsE = 0.316 Wm−2K−1. This flux is across a
surface normal to the Sun–Earth centre-to-centre line. Per unit area of the Earth’s surface, the
incident flux is four times smaller than that amount, which is

Jsinc = 0.079 Wm−2K−1.


This value coincides with that of Wu and Liu [13] 6.
6
This is nominal. The process of distributing the energy over the entire Earth’s surface is just part of the irreversible
processes that eventually lead to a steady state.

16
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

The next piece to analyse is the radiation reflected by the Earth’s surface (oceans, clouds,
etc). The appropriate relation is now (44) with ϕ = ϕ ref , Jedil = Jeref and T = TS . Since
reflection is not specular and constitutes an irreversible process (diffusive scattering), the
entropy flux is enhanced by the irreversibility factor, which, using (46), is χ (ϕ ref ) = 4.68. All
this leads to a reflected entropy flux of

Jsref = 0.1072 Wm−2K−1.

This flux is larger than the incident flux, Jincs . Thus, the net flux of entropy into the Earth’s
system is negative: Jsinc − Jsref < 0 . The entropy carried away by the diluted (reflected)
radiation is larger than the input of entropy into the Earthʼs system. Nevertheless, the
phenomenon of reflection is in itself irreversible, and so this unbalance does not mean that the
reflected radiation is carrying away entropy otherwise generated within the Earthʼs system. (A
quantitatively different result is obtained if one assumes that the solar radiation is reflected
before it is distributed over the Earthʼs entire surface. A more plausible assumption would
consider reflection from only one hemisphere (the one actually illuminated).)
We finally consider entropy associated with the radiation re-emitted by the Earth after it
has been thermalized. This is not emission by a BB, but by a gray body due to the less-than-
unity emissivity of the Earth. Hence, we use (44) again, but now with ϕ = ϵ , Jedil = Jeabs, and
T = TE . Using the more realistic third temperature of 289.0 K one finds the Earth’s emitted
entropy flux to be

Jsem = 1.258 Wm−2K−1.

This value is much larger than that of the incident solar radiation. The irreversibility factor
here is χ (ϵ ) = 1.123 only, so that the relatively large value of Jem
s is mostly due to the low
effective temperature of the Earth, and not to the irreversible (diluted) character of the
radiation process. Taking into account the entropy of the radiation incident reflected and
emitted by the Earth, the net flux on entropy on the Earth is
Jsinc − Jsref − Jsem = −1.286 Wm−2 K−1, which means that the planet is getting rid of this
amount of entropy.

4. Conclusions

We have shown in detail that in all irreversible processes involving radiation that have been
considered here, the Clausius strict inequality relating entropy and heat is valid, as it should
be. The relation between heat transferred and entropy involving T, the constant temperature of
the source, is summarized in (47).
Irreversibility in the radiation processes has at least two possible origins. One is the
emergence of radiation from the source into cold empty space, which contributes with a factor
of 4/3 to the flux ratio (45) since no work is done. This process is similar to a transfer of heat
between two reservoirs of unequal temperatures situated at a macroscopic distance from one
another. The second is the dilution of radiation due to its interaction with matter, whose
contribution can be expressed by a single irreversibility factor, χ ⩾ 1, multiplying the
already-present factor of 4/3 in the ratio of fluxes. Dilution consists of the reduction of
radiance and may be due to the presence of absorption, imperfect emission, or scattering by a
diffusive medium. We also show that the mere spreading of radiation into empty space, once
out of the boundary of the source, which keeps the radiance constant, is a reversible process
and does not contribute to an increase in the ratio of the entropy to energy flux.

17
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

It is interesting to point out that when two bodies of unequal temperatures are at a
distance from each other that is smaller than the peak wavelength of the BB spectrum, the
heat transfer between the bodies can be several times more intense than that predicted by BB
radiation formulae [19]. This is the so-called near-field regime, which is present at nanoscale
distances and has been explored both experimentally and theoretically, including its ther-
modynamic efficiency. The reason for the phenomenon is the presence, in the interface, of
phonon-polariton excitations of quantum origin [20].

Acknowledgments

We gratefully acknowledge Professor Pablo Lonngi for helpful comments. FDR is thankful
for support from CONACYT (Mexico) project SEP-2008–105843.

Appendix A. Entropy of monochromatic radiation

For completeness we give a derivation of Planck’s formula (35) for the entropy of polarized
monochromatic photons. There are various derivations available in the specialized literature;
here we follow that of Landsberg [22] as presented by Kabelac and Drake [37].
We start with the well-known expression of Boltzmann for the entropy [38]:

S=− ∑Px ln Px, (60)


x

where Px denotes the probability, Px (N1, N2 , …), of finding N1 particles (photons) in the
single-particle quantum state x1, N2 particles in state x 2 , and so on. Since the photons are
noninteracting and follow Bose–Einstein statistics, we can safely assume that the probability
of finding Ni photons in state x j , written p j (N j ), is independent of the probabilities for other
states. Thus the joint probability Px factorizes,
Px ( N1, N2 , …) = p1 ( N1 ) p2 ( N2 ) · · · (61)
where the p j (N j ) are normalized so that
∑ p j ( N j ) = 1. (62)
N j=0

Now, we can write for the entropy of the photons in state x j


sx = −k ∑ p j ( N j ) ln p j ( N j ). (63)
N j=0

For photons (bosons) we now assume that the probability of finding an extra photon in state
x j is independent of any photons already occupying that state. This means that the
probability, 0 ⩽ q j ⩽ 1, of finding any one photon in x j only depends on the state, and thus
( )
the probability for Nj also factorizes: p j N j ∝ q jN j . The proportionality constant is obtained
from the normalization condition and the expression for the sum of a geometric series as
(1 − q j ), so that

( ) (
p j N j = 1 − q j q j j. ) N
(64)

Substituting this expression in (62), and using again the geometric series sum result, we find
that

18
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

⎡ qj ⎤
sx = −k ⎢ ln 1 − q j +
( ) ln q j ⎥ . (65)
⎢⎣ 1 − qj ⎥⎦

Equation (63) allows us to calculate the mean occupation number of photons in x j ,


nj = ∑ N j p j ( N j ),
N j=0

which, using the geometric series sum result for a third time, is shown to be

n j = qj ( 1 − q ).
j

This equation can be used to rewrite (64) as

sx = k ⎡⎣ 1 + n j ln 1 + n j − n j ln n j ⎤⎦ ,
( ) ( )
which is the result (35).

References

[1] Broda E 1983 Ludwig Boltzmann. Man, Physicist, Philosopher (Woodbridge, CN: Ox Bow Press)
pp 79–80
[2] Wien W 1894 Temperatur and Entropie der Strahlung Ann. Phys. Chem. 52 132–65
[3] Planck M 1991 The Theory of Heat Radiation (New York: Dover) pp 97–169
[4] Fermi E 1956 Thermodynamics (New York: Dover) pp 54–55
[5] Pippard A B 1974 The Elements of Classical Thermodynamics (London: Cambridge University
Press) p 94
[6] Noda A and Tokioka T 1983 Climates at minima of the entropy exchange rate J. Meteorol. Soc.
Jpn. 61 894–908
[7] Peixoto J, Oort A, Almeida M D and Tomé A 1991 Entropy budget of the atmosphere J. Geophys.
Res. 96 981–8
[8] Ozawa H, Ohmura A, Lorenz R D and Pujol T 2003 The second law of thermodynamics and the
global climate system: A review of the maximum entropy production principle Rev. Geophys.
41 10181
[9] Landsberg P T and Tonge G 1979 Thermodynamics of the conversion of diluted radiation J. Phys.
A: Math. Gen. 12 551–62
[10] Edgerton R H 1980 Second law and radiation Energy 5 693–707
[11] Aoki I 1983 Entropy productions on the Earth and other planets of the solar system J. Phys. Soc.
Jpn. 52 1075–8
[12] Wright S 2007 Comparative analysis of the entropy of radiative heat transfer and heat conduction
Int. J. Thermodynamics 10 27–35
[13] Wu W and Liu Y 2010 Radiation entropy flux and entropy production of the Earth system Rev.
Geophys. 48 2008RG000275
[14] Schrödinger E 1967 What is Life? (Cambridge: Cambridge University Press) 29–40
[15] Ulanowicz R E and Hannon B M 1987 Life and the production of entropy Proc. R. Soc. London
232 181–92
[16] Heald M A and Marion J B 1995 Classical Electromagnetic Radiation 3rd edn (New York:
Brooks/Cole Thompson Learning) p 524
[17] Einstein A 1905 Concerning an heuristic point of view toward the emission and transformations of
light Ann. Phys. 17 132–48
[18] Nicodemus F E 1963 Radiance Am. J. Phys. 31 368–77
[19] Shen S, Narayanaswamy A and Chen G 2009 Surface phonon polaritons mediated energy transfer
between nanoscale gaps Nano Lett. 9 2909
[20] Latella I, Pérez-Madrid A, Lapas L C and Rubí J M 2014 Near-field thermodynamics: useful work,
efficiency, and energy harvesting J. Appl. Phys. 115 124307
[21] Mandl F 1988 Statistical Physics 2nd edn (New York et al: Wiley) p 250

19
Eur. J. Phys. 36 (2015) 035001 F del Río and S M T de la Selva

[22] Landsberg P T 1959 The entropy of a non-equilibrium ideal quantum gas Proc. Phys. Soc. 74
486–8
[23] Slater J C 1939 Introduction to Chemical Physics 1st edn (New York: McGraw-Hill) p 72
[24] Piña E and de la Selva S M T 2010 Thermodynamics of radiation modes Eur. J. Phys. 31 393–400
[25] Rosen P 1954 Entropy of radiation Phys. Rev. 96 555
[26] Ore A 1955 Entropy of radiation Phys. Rev. 98 887–8
[27] Landsberg P T 1961 Thermodynamics with Quantum Statistical Illustrations (New York:
Interscience) p 296
[28] Cailles U and Herbert F 1988 Radiative processes and non-equilibrium thermodynamics J. App.
Math. Phys. (ZAMP) 39 242–66
[29] Stephens G L and O’Brien D M 1993 Q. J. R. Meteorol. Soc. 119 121–52
[30] Wright S E, Scott D S, Haddow J B and Rosen M A 2001 On the entropy of radiative transfer in
engineering thermodynamics Int. J. Eng 39 1691–706
[31] Petela R 1961 Exergy of heat radiation of a perfect gray body (in Polish) Zesz. Nauk. Politech.
Slask. Energ. 5 33–45
[32] Petela R 1964 Exergy of heat radiation J. Heat Transfer 86 187–92
[33] Petela R 2003 Exergy of undiluted thermal radiation Sol. Energy 74 469–88
[34] Williams D R 2013 Sun fact sheet, nssdcgsfc.nasa.gov/planetary/factsheet/sunfact.html (retrieved
30 September 2013)
[35] Renewable Resource Data Center 2014 ASTM G173-03 reference spectra, available from http://
rredc.nrel.gov./solar/spectra/am1.5/ASTMG173/ASTMG173.html
[36] de Pater I and Lissauer J 2001 Planetary Science (Cambridge: Cambridge University Press) p 66
table 4.1
[37] Kabelac S and Drake F D 1992 The entropy of terrestrial solar radiation Sol. Energy 48 239–48
[38] See [21] p 61
[39] See [21] p 249

20

View publication stats

You might also like