The Stefan-Boltzmann Law: Two Classical Laws Give A Quantum One

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Physica Scripta

PAPER Related content


- Einstein's part in the development of
The Stefan–Boltzmann law: two classical laws give quantum concepts
M A El'yashevich
a quantum one - Scaling, scattering, and blackbody
radiation in classical physics
Timothy H Boyer
To cite this article: H Paul et al 2015 Phys. Scr. 2015 014027
- Irreducible decomposition of Gaussian
distributions and the spectrum of black-
body radiation
Sándor Varró
View the article online for updates and enhancements.

Recent citations
- The Riemann hypothesis illuminated by
the Newton flow of
J W Neuberger et al

This content was downloaded from IP address 141.216.78.40 on 23/10/2017 at 03:21


| Royal Swedish Academy of Sciences Physica Scripta

Phys. Scr. T165 (2015) 014027 (7pp) doi:10.1088/0031-8949/2015/T165/014027

The Stefan–Boltzmann law: two classical


laws give a quantum one
H Paul1, D M Greenberger2, S T Stenholm3 and W P Schleich4,5
1
Institut für Physik der Humboldt Universität, D-12489 Berlin, Germany
2
City College of New York, New York, New York 10031, USA
3
Physics Department, Royal Institute of Technology, KTH, Stockholm, Sweden and Laboratory of
Computational Engineering, HUT, Espoo, Finland
4
Institut für Quantenphysik and Center for Integrated Quantum Science and Technology (IQST ),
Universität Ulm, D-89069 Ulm, Germany
5
Texas A&M University Institute for Advanced Study (TIAS), Institute for Quantum Science and
Engineering (IQSE) and Department of Physics and Astronomy, Texas A&M University, College Station,
Texas 77843-4242, USA

E-mail: harry.paul@physik.hu-berlin.de

Received 12 December 2014


Accepted for publication 3 March 2015
Published 7 October 2015

Abstract
Due to the universality of blackbody radiation the constant in the Stefan–Boltzmann law
connecting the energy density and temperature of blackbody radiation is either a universal
constant, or built out of several universal constants. Since the Stefan–Boltzmann law follows
from thermodynamics and classical electrodynamics this constant must involve the speed of light
and the Boltzmann constant. However, a dimensional analysis points to the existence of an
additional universal constant not present in the two classical theories giving birth to the Stefan–
Boltzmann law. In the most elementary version this constant has the dimension of an action and
is thereby proportional to Planck’s constant. We point out this unusual phenomenon of the
combination of two classical laws creating a quantum law and speculate about its deeper origin.

Keywords: blackbody radiation, Stefar–Boltzmann law, classical–quantum border

1. Introduction: dimensional analysis predicts new [9] in 1879, and derived in 1884 by Boltzmann [10] from
physics thermodynamics with elements of classical electrodynamics
predicts the existence of a universal constant heretofore not
During its long life, thermodynamics [1] has played an present in physics. In the most elementary version of
extremely fruitful role in fostering quantum physics. Indeed, dimensional analysis this new constant has the units of action
this theory has emerged from Max Planck’s ingenious and is proportional to Planck’s constant.
application [2] of thermodynamics to blackbody radiation.
Moreover, laser physics [3] has profited from novel concepts
such as negative temperature [4] to describe an inverted 1.1. Two routes to the Stefan–Boltzmann law
medium. Finally quantum thermodynamics [5] which ori-
ginally was intended to derive thermodynamics from quantum In figure 1 we highlight this surprising phenomenon and
mechanics has led to a wealth of studies of quantum ther- compare and contrast two approaches towards the Stefan–
modynamics machines [6, 7] and speculations that quantum Boltzmann law. The most frequently presented derivation is
coherence might act as a resource giving rise, for example, to summarized for the sake of completeness in appendix A and
solar cells [8] with improved efficiencies. proceeds by first establishing the Planck law, and by then
However, due to today’s dominance of quantum theory integrating it over all frequencies. Hence, this approach
we often underestimate the predictive power of thermo- employs a combination of a classical and a quantum law to
dynamics. Indeed, in this article we point out that the Stefan– obtain a new quantum law. Indeed, the mode density of the
Boltzmann radiation law obtained experimentally by Stefan classical electromagnetic field together with the quantum

0031-8949/15/014027+07$33.00 1 © 2015 The Royal Swedish Academy of Sciences Printed in the UK


Phys. Scr. T165 (2015) 014027 H Paul et al

theories provide directly an expression for the integrated


energy density with the help of a differential equation. Not
only is the resulting expression finite but it is also of quantum
nature. In this sense the finiteness of the integrated energy
density is another indication that the Stefan–Boltzmann law is
a quantum law.

1.2. Speed of light is a universal constant

How can the combination of two classical laws lead to a


quantum law? This question arises from a dimensional ana-
lysis of the Stefan–Boltzmann law derived from classical
electrodynamics and thermodynamics.
However, this is not the only example where a phe-
nomenon embedded in a well-established theory already
Figure 1. The amazing power of the combination of classical points to a deeper concept. Indeed, we now show that the
electrodynamics and thermodynamics in deriving a quantum law, Coulomb and Ampère laws suggest a new universal constant,
that is the Stefan–Boltzmann law, illustrated by comparing and that is the speed of a light. In this way we get a glimpse of
contrasting two approaches: (i) a mixed classical–quantum, and (ii) a special relativity.
purely classical one. The mode density of the classical electro-
magnetic field together with the quantum mechanical average energy We consider two equally equipped laboratories, S and S′
of the field oscillator yields the Planck law. Integration over all which differ from each other only by the fact that they are
frequencies leads us to the Stefan–Boltzmann law. In this approach moving at a uniform velocity v relative to each other.
we use one element (mode density) of a classical theory and one Experimenters in these labs set out to test the laws of elec-
(average energy of oscillator) from quantum mechanics to arrive at a tricity. Lab S explores first the Coulomb law, and concludes
quantum law. In contrast, in the Boltzmann derivation [10] two
classical elements (the pressure-energy equivalence and a thermo- that the force Fe between two equal charges e is given by the
dynamic relation) suffice to cross the classical–quantum border and formula
arrive at the Stefan–Boltzmann law.
e2
Fe = k1,
r2
mechanical average of the energy of the field oscillator yields where r denotes their separation. Here k1 depends on the units
the Planck radiation formula. used, and also determines the numerical value of the force.
In contrast, the derivation [10, 11] of Boltzmann starts Lab S′ performs the same experiment, and presumably
from two classical laws and crosses somehow magically the obtains the same result. Provided it uses the same units it will
classical–quantum border to arrive at a quantum law. Indeed, get an identical value for k1.
from a connection between the pressure and the energy Next lab S explores the Ampère law, by measuring the
density of the radiation established in classical electro- force Fm between two parallel wires of length L, each carrying
dynamics, and a thermodynamic relation based on the first a current I. In this case the experimenter finds
and the second law of thermodynamics we find the Stefan–
L2
Boltzmann law. We emphasize that this approach does not Fm = k 2 I 2
.
pass through the Planck formula but goes straight to its fre- r2
quency-integrated form, that is to the Stefan–Boltzmann law. Again S′ performs the same experiment, and presumably
It is well-known that the Rayleigh–Jeans expression for arrives at the same result. Assuming the use of identical units
the energy density valid for small frequencies is independent the experimenter will get the same value for k 2 . Both labs
of Planck’s constant and therefore a classical result. More- interprete a current as a flow of charge with units
over, the quadratic increase with frequency leads to a diver- charge time = [e] [t ]. Here square brackets denote the
gent integral of the energy density. Nevertheless, the Stefan– dimensions of physical quantities.
Boltzmann law, which is an expression for the integrated Lab S knows that all forces have the same dimension and
energy density, and derived solely from classical concepts thereby establishes based on dimensional analysis the identity
provides us with a finite result.
⎡ e2 ⎤ ⎡ e2 ⎤
This apparent contradiction resolves itself when we rea- ⎣ ⎦ ⎣ ⎦
lize that the Stefan–Boltzmann law is indeed a quantum law. [ Fe] = [ k1] ⎡ r 2⎤
= [ Fm] = [ k 2 ]
⎡ t 2⎤
,
⎣ ⎦ ⎣ ⎦
Since Planck’s constant h appears to the third power in the
proportionality constant the energy density diverges as h which reduces to
approaches zero.
Moreover, this observation demonstrates in a striking ⎡ r 2⎤
[ k1] ⎣ ⎦
way the power of the two classical theories. Rather than = . (1)
deriving first an expression for the energy density in its fre- [ k2 ] ⎡ t 2⎤
⎣ ⎦
quency dependence and subsequently integrating it, the two

2
Phys. Scr. T165 (2015) 014027 H Paul et al

Hence, the ratio of the two constants k1 and k2 has the units of Stefan–Boltzmann law, and not the appearance of Planck’s
a square of a velocity, regardless of the electrical units they constant. This insight is in complete agreement with our
were measured in. Moreover, the experimenter in S′ finds the dimensional analysis in the main body of our article. In order
same constant. It is remarkable that all measurements were to keep the article self-contained, we summarize in
static measurements. Nothing was moving in either labora- appendix B the thermodynamic relation which is at the very
tory. Hence, this velocity cannot be the velocity of anything heart of the derivation of the Stefan–Boltzmann law.
in particular. It is a universal constant in the sense that it can
be determined by performing static measurements in the
individual laboratories. 2. Basic elements of Stefan–Boltzmann Law
Next the two labs explore the Faraday law and the
Maxwell law of displacement. Obviously both will find that We start by briefly reviewing the basic elements of the Ste-
E ∝ dB dt , and B ∝ dE dt , which leads them to conclude fan–Boltzmann law. In this way we bring out most clearly the
that electromagnetism is a wave phenomenon that travels with classical underpinning of this early radiation formula.
a characteristic velocity. But what is that velocity? According to the Stefan–Boltzmann law [11]
It is built right into the equations they had determined by
static measurements. Therefore, they both conclude that these u (T ) = a T 4 (2)
waves travel with the characteristic velocity c given by the the total energy density
ratio of k1 and k2, equation (1). Hence, if we accept the ∞
validity of the Maxwell equations in both frames, we are led u (T ) ≡ ∫0 d ν u (ν , T ) (3)
inexorably by a dimensional argument to the universality of
this velocity. One does not need the Michelson–Morley of the blackbody radiation integrated over all frequencies ν is
experiment. Rather now, it is that result that must be proportional7 to the fourth power of the temperature T of the
explained. body. Here u denotes the monochromatic energy density, that
is the density corresponding to the frequencies between ν and
1.3. Outline of the article ν + dν divided by dν . The proportionality constant a is an
integration constant which arises from a differential equation
Our article is organized as follows: in order to establish the of first order obtained from thermodynamics with elements of
fact that the Stefan–Boltzmann is solely based on classical classical electrodynamics.
theories we briefly recall in section 2 the key elements of its Indeed, it follows from two ingredients: (i) the connec-
derivation. Here we follow closely the original approach tion
[10, 11] by Boltzmann. We then dedicate section 3 to a
1
dimensional analysis [12] 6 of the Stefan–Boltzmann law and p= u (4)
demonstrate that in its most elementary version the new 3
universal constant has the units of an action. Therefore, it is between the energy density u of the radiation and its pressure
up to a numerical factor proportional to Planck’s constant. In p on the walls of a cavity established in classical electro-
section 4 we show that the sole contribution of quantum dynamics, and (ii) the relation
mechanics in the form of the Planck radiation formula to the
⎛ ∂E ⎞ ∂ ⎛⎜ p ⎞⎟
Stefan–Boltzmann law is to fix this numerical constant. We ⎜ ⎟ = T2 (5)
conclude in section 5 by summarizing our results and by ⎝ ∂V ⎠T ∂T ⎝ T ⎠
providing a brief outlook. for the energy E in a volume V which is a consequence [11] of
In appendix A we present yet another example where the the first and the second law of thermodynamics as shown in
dimensional analysis of a formula derived in a well-estab- appendix B.
lished theory provides us with a hint of a deeper theory. Since When we substitute the identity
here we consider the Wien radiation law which can also be
obtained from thermodynamics, this discussion provides E = Vu (6)
additional support for the main theme of our article—classical together with the expression equation (4) for the radiation
equations can point to the existence of the quantum world. pressure into the thermodynamic relation equation (5) we
Moreover, we obtain the Stefan–Boltzmann law starting from arrive at
the Wien one which contains a free function f. We compare
and contrast two examples for f : (i) first we consider the du
4u = T , (7)
Planck expression and recall the basic elements of its deri- dT
vation. (ii) Then we analyze the high-frequency limit. This which after integration yields the Stefan–Boltzmann law
approach suggests that the main achievement of quantum equation (2).
mechanics is the determination of the numerical factor in the
7
Apart from the version of the Stefan–Boltzmann law given by equation (2)
6
Both authors in [12] perform a dimensional analysis of the Kirchhoff law there exists another one which connects the power radiated by a surface
given by equation (A.1). When they use soley the constants c and kB the element of the blackbody with the fourth power of the temperature. The
energy density integrated over all frequencies is infinite. However, a corresponding proportionality constant is called σ and differs from a by one
dimensional analysis including an additional constant yields a finite result. power of the velocity of light and a factor 4, that is σ = ac 4 .

3
Phys. Scr. T165 (2015) 014027 H Paul et al

3. Emergence of a new constant 4. Origin of numerical factor

Due to the universality of blackbody radiation, that is the We emphasize that a measurement of the energy density u of
radiation is independent of the specific properties of the the blackbody radiation as a function of the temperature T
radiating body as long as it is in thermal equilibrium, the together with the Stefan–Boltzmann law, equation (2), and the
constant a of integration emerging from equation (7) must be knowledge of the universal constants c and kB provides us
either a universal constant, or built out of several universal with the value of Planck’s constant h only up to a numerical
constants. Since the two theories giving rise to the result factor N. Nevertheless, it is surprising that the value of h
equation (7) are classical electrodynamics and thermo- obtained in this way agrees within an order of magnitude with
dynamics the two universal constants that appear naturally are the correct one.
the speed of light c and the Boltzmann constant kB. Although In order to bring this fact out most clearly we now
compare the thermodynamic result, equation (9), for a to the
we cannot predict the value of a a dimensional analysis
one following from the Planck radiation law [1]. In
immediately shows us that a must contain a universal constant
appendix A we recall the relation
alien to these classical theories.
Indeed, from the definition equation (6) of the energy E
8π 5 k B4 k B4
in terms of V and u, and the Stefan–Boltzmann law a= = , (10)
15 h3c3 (Nh)3c3
equation (2), we find with the fact that the product k B T
represents an energy the relation where N ≡ [15 (8π 5)]1 3 ≅ 0.183.
Indeed, this comparison yields the identification
[a ] =
[u ]
=
[E ] 1
=
[E ] [ k B ]4 =
[ k B ]4 .
⎡ T 4⎤ ⎡ ⎤ J = N · h,
⎣ ⎦ [V ] ⎣ T ⎦
4 [V ] [ k B T ]4 [V ][E ]3
which shows that the thermodynamic result for the universal
constant J differs from Planck’s constant by a factor 0.183.
Moreover, the speed of light c which has the dimensions It is the numerical factor 8π 5 15 in equation (10) which
of length/time ≡ [V 1 3] [t ] allows us to eliminate [V ] and we reflects the correct quantum mechanical radiation law. As
arrive at shown in appendix A, this fact stands out most clearly in the
Wien approximation of the Planck law which does not alter
[a ] =
[ k B ]4 . (8)
the appearance of the universal constants in the so-obtained
[c]3 [Et ]3 Stefan–Boltzmann law but brings in a slightly different
numerical factor.
This feature confirms that the appearance of h in a is
Hence, a must contain, apart from the fourth power of kB and solely due to dimensional reasons. Only the numerical factor
the inverse of the cube of c, the inverse of the cube of a 8π 5 15 in a is a consequence of quantum mechanics, in a
constant J with the dimensions of action, that is complete agreement with the thermodynamics approach.
k B4
a= . (9)
J 3c3 5. Summary and outlook

Moreover, since a can only consist of universal constants and Dimensional arguments have always been used not only to
c and kB are elements of the underlying classical theories the ensure consistent use of units and concepts, but occasionally
quantity J must be a new universal constant. to suggest surprising connections between variables. In the
We note that we can multiply the numerator and present article we have pointed out such a connection. Indeed,
denominator of the expression for [a] in equation (8) by an our dimensional analysis of the Stefan–Boltzmann radiation
arbitrary power s of [c]. In this case, the new constant J has formula shows that two classical laws can lead to a quantum
the dimension of action × (velocity)s . However, we empha- law. This surprising ability of two theories providing us with
size that this non-uniqueness does not change the fact that a a glimpse of a new one also appears in electro and magneto
statics represented by the Coulomb and the Ampère law
must contain a new universal constant which in its most
which suggest special relativity, and in the combination of
elementary case, that is for s = 0, has the units of an action.
classical electrodynamics and thermodynamics giving rise to
We conclude this section by noting that there exist many
the Wien law which alludes to quantum mechanics. But how
examples where dimensional analysis suggests the existence can it be that we obtain a quantum law from classical ones?
of Planck’s constant, for example in defining an elementary Unfortunately, at this point we cannot present an answer but
phase space volume. However, the present argument is fun- can only speculate.
damentally different since in addition to the information An elementary explanation may start from the fact that
provided by dimensional analysis we have the knowledge that thermodynamics is characterized by an unique contrast of an
the integration constant a consists of the product of universal austerity of assumptions and a richness of applications. One
constants. This fact is unique to the Stefan–Boltzmann law. might suspect that it is exactly this simplicity and generality

4
Phys. Scr. T165 (2015) 014027 H Paul et al

of the foundations of thermodynamics which gives this theory obtaining the Stefan–Boltzmann law. We then turn to a brief
the enormous predictive power. discussion of the essential ingredients of the Planck law,
Another insight into this miraculous revelation of a new analyze various limits of the corresponding function fP and
constant of nature might come from the form of the differ- compare and contrast the numerical constants resulting from
ential equation equation (7) leading to the fourth power of the the integration of the respective energy densities over
temperature. Here the variable T appears in front of the frequency.
highest derivative, that is of the first derivative of the energy
density with respect to T. Therefore, the behavior of the A.1. Planck’s constant from the Wien law
solution when T is extended into the complex plane is gov-
erned by a branch point leading to a logarithmic phase sin- Even before James Clark Maxwell unified the laws of elec-
gularity [13]. tromagnetism, Gustav Kirchhoff showed in 1859 that the
This feature also manifests itself in the quantum radiation in a cavity which is in equilibrium at temperature T
mechanical problem of the inverted harmonic oscillator [14] possesses a monochromatic energy density u , that is an
where the kinetic energy is still positive but the quadratic energy per unit volume and frequency ν which is a universal
potential is repulsive rather than attractive. The differential function of T and ν, that is
equation corresponding to an energy eigenstate, when viewed u = u (ν , T ). (A.1)
[14] in rotated quadrature variables is of the same form as the
differential equation equation (7) for the Stefan–Boltzmann In 1896 Wilhelm Wien was able to show that u takes the
law. In the case of the inverted oscillator it gives rise to a form
logarithmic phase singularity which is the origin of the Unruh
u (ν , T ) = ν 3f (ν T ), (A.2)
effect [15], the Hawking radiation [16] or optical analogues
[17] of event horizons of black holes. Moreover, it is the where the function f cannot be determined within the
origin of the tunneling through the parabolic barrier whose framework used to derive this result. Indeed, Wien obtained
rate [18] is given by an expression reminiscent of the Planck [11] his law by adiabatically expanding the walls of the
radiation formula. cavity, and showing that this procedure induces a Doppler
A recent article [19], which revisits tunneling using the shift in the reflected light, from which he was able to derive a
Wigner function [20] emphasizes the non-analytic behavior of partial differential equation, whose solution is the functional
the Wigner function across the separatrix in phase space relation equation (A.2).
distinguishing in- and outgoing waves. It is this feature which This is as far as we can go by using purely thermo-
gives rise to the logarithmic phase singularity and to the dynamic arguments. But equation (A.2) carries important
various forms of radiation including the familiar transmission dimensional information that points to new physics when we
formula [18]. cast the function ν 3f (ν T ) into the appropriate dimen-
To connect these aspects with the Stefan–Boltzmann law, sional form.
and provide with the help of the analytic properties of the The relevant ingredients from electrodynamics and
underlying differential equation extended into complex space thermodynamics are the speed c of light and k B T . Here
a bridge between the quantum and the classical world is a Boltzmann’s constant k B acts as a conversion factor to express
fascinating topic. Unfortunately it goes beyond the scope of the temperature T in units of energy E, that is ⎡⎣ k B T ⎤⎦ = [E ].
the present article and has to be postponed to a future
Since udν has units energy volume = [E ] [V ], and ν 3 c3 has
publication.
units of 1 [V ] we find
k B T ν 3 ⎛ bν ⎞
Acknowledgments u (ν , T )dν = α f⎜ ⎟ dν . (A.3)
ν c3 ⎝ k B T ⎠

We thank M Freyberger, Th W Hänsch, W Hüttner, and WH Here α is a dimensionless numerical constant.


Zurek for fruitful discussions on this topic. WPS gratefully We have inserted in equation (A.2) an energy factor k B T
acknowledges the support by a Texas A&M University but have preserved the form while giving it the correct
Institute for Advanced Study (TIAS) Faculty Fellowship. dimensions. In this process we had to introduce a new con-
stant b, so that bν is an energy, giving b the dimension of
action. This approach brings in a new universal constant
Appendix A. Blackbody radiation laws which is determined by Planck’s constant h.
The only way to avoid the appearance of b is if f is equal
In this appendix we first recall that a dimensional analysis of to unity corresponding to the Rayleigh–Jeans law discussed
the Wien radiation law also reveals the need for an additional later in this appendix. Unfortunately, in this case we are
universal constant in blackbody radiation. Although this confronted with the ‘ultraviolet catastrophe’, since the integral
argument can be found in standard textbooks [11] we repeat it over all frequencies diverges.
here in a slightly different form because it is highly relevant Hence, the permissible functions f belong to a restricted
for the present discussion. In particular, we emphasize the class. In the next sections we discuss two such examples of f
crucial role of the dimensionless argument of the function f in corresponding to the Planck law and the Wien approximation.

5
Phys. Scr. T165 (2015) 014027 H Paul et al

A.2. Stefan–Boltzmann law When we substitute equations (A.8) and (A.9) into
equation (A.7) we obtain the Planck radiation formula
When we now integrate equation (A.3) over frequency we
arrive with the dimensionless integration variable 8πν 2 hν
u P (ν , T ) ≡ . (A.10)
c3 ⎛ hν ⎞
bν exp ⎜ ⎟−1
ξ≡ (A.4) ⎝ kB T ⎠
kB T
A comparison of this expression for the energy density to the
at the Stefan–Boltzmann law one originating from the Wien formula, equation (A.2), and
brought into a more concrete form by dimensional analysis ,
∞ k B4 equation (A.3), allows us to identify the constant b = h
u (T ) ≡ ∫0 d ν u (ν , T ) = α ′
b3c3
· T4 (A.5)
together with the function
ξ
for the total energy density u. fP (ξ ) ≡ , (A.11)
ξ
e −1
We emphasize that the numerical constant α of
equation (A.3) is modified by this integration and reads and the numerical constant


α ≡ 8π . (A.12)
α′ ≡ α ∫0 2
dξ ξ f (ξ ). (A.6) With the integral relation [21]
∞ ξ3 π4
Hence, the form of the Stefan–Boltzmann law is independent ∫0 dξ
eξ − 1
=
15
(A.13)
of the so-far unknown function f. In contrast, the constant α′
strongly depends on it. Moreover, the fact that the argument we find the explicit form
of f has to be dimensionless which leads to the introduction of
8π 5
the new universal constant b, together with the cubic α P′ = (A.14)
dependence of the frequency ν as dictated by the Wien law, 15
equation (A.2), determines the appearance of b in the of the numerical factor α′, equation (A.6), in the Stefan–
denominator of equation (A.5) with the third power. Boltzmann law equation (A.5) corresponding to the
Planck law.
A.3. Planck law
A.4. Wien approximation
Next we discuss our first example for f given by the Planck
radiation formula. For this purpose we first briefly summarize Next we consider the high-frequency limit of fP which cor-
the underlying principles and then focus on the mixed clas- responds to neglecting the term unity in the denominator of fP,
sical–quantum approach towards the Stefan–Boltzmann law equation (A.11), and leads us to the function
illustrated in figure 1.
The energy density u (ν, T )dν is the product fW (ξ ) ≡ ξ e−ξ (A.15)

u = n (ν) E¯ (ν , T ) (A.7) representing the Wien approximation.


With the integral relation
of the average energy E¯ (ν, T ) of the mode oscillator with ∞
frequency ν, and the mode density ∫0 dξ ξ 3e−ξ = 6 (A.16)
8π 2
n (ν)dν = 2 · 4πk 2dk = ν dν (A.8) we find the corresponding numerical constant
c3
αW′ ≡ 6 · 8π , (A.17)
following from the dispersion relation k ≡ ∣ k ∣ ≡ ν c of light.
The factors 2 and 4π arise from the two polarization and arrive at the ratio
directions and the integration over the two angles in the α P′ π4 1
three-dimensional wave vector space k , respectively. = = 1.11, (A.18)
αW′ 15 6
Quantum mechanics enters into u only through the
average energy that is the two constants from the Planck formula and the
Wien approximation are rather close.

E¯ P (ν , T ) ≡ (A.9)
⎛ hν ⎞
exp ⎜ ⎟−1 A.5. Rayleigh–Jeans law
⎝ kB T ⎠
We conclude by briefly considering the Rayleigh–Jeans law
which emerges when we perform the average over all where the function fRJ , resulting from the expansion of fP for
energies of the field oscillator with a sum rather than an ξ ≪ 1, takes the form
integral. This substitution reflects the fact that the energies are
fRJ (ξ ) ≡ 1. (A.19)
discrete rather than continuous.

6
Phys. Scr. T165 (2015) 014027 H Paul et al

Obviously now the expression for α′ given by equation (A.6) differentiations, which leads us to the formula
displays a cubic divergence which gives rise to the ultraviolet ⎛ ∂ ⎡ 1 ⎛ ∂E ⎞ ⎤ ⎞ ⎛ ∂ ⎡ 1 ⎛ ∂E ⎞ p ⎤⎞
catastrophe. ⎜⎜ ⎢ ⎜ ⎟ ⎥ ⎟⎟ = ⎜⎜ ⎢ ⎜ ⎟ + ⎥ ⎟⎟ .
⎝ ∂V ⎣ T ⎝ ∂T ⎠V ⎦ ⎠T ⎝ ∂T ⎣ T ⎝ ∂V ⎠T T ⎦⎠
V

With the relation


Appendix B. Thermodynamic relation
∂ 2E ∂ 2E
=
In this appendix we rederive for the sake of completeness the ∂T ∂V ∂V ∂T
thermodynamic relation equation (5) which is a consequence
this procedure yields
of the first and the second law of thermodynamics. Our
derivation follows closely the one in [11]. 1 ⎛ ∂E ⎞ ∂ ⎜⎛ p ⎟⎞
0=− ⎜ ⎟ + ,
We start from the first law of thermodynamics which T 2 ⎝ ∂V ⎠
T ∂T ⎝ T ⎠
reflects conservation of energy and connects the change dE in
which is the thermodynamic relation equation (5).
the internal energy E to the added heat δQ, and the work
− pdV due to the change dV of the volume V, that is
dE = δQ − pdV . (B.1) References
Next we recall the change
[1] Planck M 1945 Treatise on Thermodynamics (New York:
δQ Dover)
dS =
T [2] See for example, Gasiorowicz S 1974 Quantum Physics (New
York: Wiley)
of the entropy S at the temperature T due to δQ, that is the [3] See for example, Sargent M, Scully M O and Lamb W E 1974
second law of thermodynamics, which casts the first law, Laser Physics (Reading: Addison-Wesley)
equation (B.1), into the form [4] Ramsey N F 1956 Phys. Rev. 103 20
[5] See for example, Gemmer J, Michel M and Mahler G 2009
1 p Quantum Thermodynamics (New York: Springer)
dS =dE + dV . (B.2)
T T [6] Linden N, Popescu S and Skrzypczyk P 2010 Phys. Rev. Lett.
105 130401
When we express E and S in the thermodynamic vari- [7] Levy A and Kosloff R 2012 Phys. Rev. Lett. 108 070604
ables T and V, that is E = E (T , V ) and S = S (T , V ) we [8] Scully M O 2010 Phys. Rev. Lett. 104 207701
obtain with [9] Stefan J 1879 Sitz.ber. Math.-Nat.wiss. Cl. kaiserlichen Akad.
Wiss. 79 391
⎛ ∂E ⎞ ⎛ ∂E ⎞ [10] Boltzmann L 1884 Ann. Phys. Chem. 22 291
dE = ⎜ ⎟ dT + ⎜ ⎟ dV
⎝ ∂T ⎠V ⎝ ∂V ⎠T [11] Planck M 1959 The Theory of Heat Radiation (New York:
Dover)
from equation (B.2) the relation Hund F 1966 Theoretische Physik: Eine Einführung
(Wärmelehre und Quantentheorie vol 3) (Stuttgart: B G
1 ⎛ ∂E ⎞ ⎡ 1 ⎛ ∂E ⎞ p⎤ Teubner)
dS = ⎜ ⎟ dT + ⎢ ⎜ ⎟ + ⎥ dV ,
T ⎝ ∂T ⎠V ⎣ T ⎝ ∂V ⎠T T⎦ [12] See for example, Sommerfeld A 1959 Thermodynamics and
Statistical Mechanics: Lectures on Theoretical Physics 5
and by direct comparison with (New York: Academic)
Epstein P S 1954 Am. J. Phys. 22 402
⎛ ∂S ⎞ ⎛ ∂S ⎞ [13] Berry M V 1992 Huygens Principle 1690–1990: Theory and
dS = ⎜ ⎟ dT + ⎜ ⎟ dV
⎝ ∂T ⎠V ⎝ ∂V ⎠T Applications ed H Blok, H A Ferwerda and H K Kniken
(Amsterdam: Elsevier)
[14] Barton G 1986 Ann. Phys. 166 322
the identities
Balazs N L and Voros A 1990 Ann. Phys. 199 123
⎛ ∂S ⎞ 1 ⎛ ∂E ⎞ [15] Unruh W G 1976 Phys. Rev. D 14 870
⎜ ⎟ = ⎜ ⎟ , (B.3) [16] Kiefer C, Morto J and Moniz P V 2009 Ann. Phys. 18 722
⎝ ∂T ⎠V T ⎝ ∂T ⎠V [17] Leonhardt U 2002 Nature 415 406
[18] Kemble E C 1935 Phys. Rev. 48 549
and Hill D L and Wheeler J A 1953 Phys. Rev. 89 1102
⎛ ∂S ⎞ 1 ⎛ ∂E ⎞ p [19] Heim D M, Schleich W P, Alsing P M, Dahl J P and Varro S
⎜ ⎟ = ⎜ ⎟ + . (B.4) 2013 Phys. Lett. A 377 1822
⎝ ∂V ⎠T T ⎝ ∂V ⎠T T [20] See for example Schleich W P 2001 Quantum Optics in Phase
Space (Weinheim: Wiley-VCH)
Next we differentiate equations (B.3) and (B.4) by V and [21] Gradstein I S and Ryzhik I M 1994 Table of Integrals, Series
T, respectively and recall that we can interchange the two and Products (Boston: Academic Press)

You might also like