Working PaperPDF Available

Planck's law revisited

Authors:

Figures

Content may be subject to copyright.
Planck’s law revisited
Sergei Viznyuk
Abstract
I review a textbook derivation of Planck’s formula for spatial density of radiation
energy. I point out at one inconsistency, and a couple of factitious assumptions used in
derivation. I propose a derivation more aligned with quantum mechanical principles. I
show the de-coherence of oscillator modes is the major factor in Planck’s law.
By nearly universal consent, the day of Dec.14, 1900 when Planck’s formula [1, 2] for spatial
density of radiation energy has been published, is considered [3] the birthday of quantum theory:




(1)
Since then, the formula became a staple piece of every textbook on statistical physics [4, 5].
Its close match to experimental results is touted as one of the greatest achievements of quantum
theory. For instance, the spectral intensity of cosmic microwave background (CMB) deviates from
Planck’s formula less than 0.03% [6, 7]. The Planck’s derivation [1, 2] of (1) follows from:


(2)
, where is the average energy of a resonator of frequency in a thermodynamic ensemble of
resonators, with fixed total energy of the ensemble; 
is, supposedly, the number of all
hypothetically possible resonator modes per unit volume of space per unit frequency range. The
formula for the average energy is derived from the energy levels of a quantum harmonic
oscillator:


(3)
, and Boltzmann’s postulate that probability to find a member of thermodynamic ensemble in a
state with energy is proportional to 
 :



 

 
 
(4)
The reason the so-called zero-point energy term  
in (4) is not included into Planck’s formula
(1) is the topic of a hundred-year controversy [8, 9]. A short argument follows here. A radiation
mode manifests itself via electromagnetic interaction with the matter. As a result of this interaction,
the energy quantum  is transferred from the matter to the radiation mode, or from radiation mode
to the matter. If radiation mode is in its ground state with  
, it cannot lose another 
quantum to the matter. That in itself does not prove the zero-point energy does not exist. However,
the gravity which would have been exerted by zero-point energy of all hypothetically possible
radiation modes is over 58 orders of magnitude bigger than the gravity which could be derived
from empirical evidence [8]. I suggest this discrepancy is because, in textbooks, the radiation is
[implicitly] considered a classical object, which exists “out there” in the open space, and possesses
properties independent of the measurement context. Such view is contrary to the base principles
of quantum mechanics [in Copenhagen interpretation]. I argue that the radiation modes which are
present, are the ones which were actually emitted by the matter, not all the hypothetically possible
radiation modes. Since the number of oscillators in the matter, which interact with radiation is
limited, so is the number of radiation modes. A radiation mode and the corresponding matter
oscillator should be viewed as one and the same [entangled] system. Planck may have had similar
view, albeit not clearly stated. The word entanglement was not in physicist’s vocabulary at the
time. Planck’s reasoning was [2]:
Let us consider a large number of monochromatically vibrating resonator of frequency
(per second), of frequency ,  of frequency , ..., with all large number which are
at large distances apart and are enclosed in a diathermic medium with light velocity and
bounded by reflecting walls. Let the system contain a certain amount of energy, the total energy
(erg) which is present partly in the medium as travelling radiation and partly in the resonators
as vibrational energy. The question is how in a stationary state this energy is distributed over
the vibrations of the resonator and over the various of the radiation present in the medium, and
what will be the temperature of the total system
…we first of all consider the vibrations of the resonators and assign to them arbitrary definite
energies, for instance, an energy to the resonators , to the resonators ,...
Dividing by , by ,... we obtain the stationary value of the energy , , ... of a
single resonator of each group, and thus also the spatial density of the corresponding radiation
energy
As it sounds, Planck implied the average energy (4) of a resonator in a given resonator group
is one and the same as the “corresponding radiation energy”. Thus, an actually present radiation
mode should have a corresponding matter resonator. The number of radiation modes is the same
as the number of matter resonators.
On the other hand, (1) still gives the correct result, even as the factor 
is considered
as the number of all hypothetically possible radiation modes per unit volume of space per unit
frequency range, not just actually present modes. This contradiction stems from the way the factor

is derived in textbooks. Note, that (4) is obtained from fundamental quantum mechanical
expression (3) in thermodynamic limit of a large number of oscillators present in a given mode.
However, the expression 
for the number of modes per unit volume per unit frequency
range is obtained with purely classical approach, by treating each mode as a standing wave
enclosed in a limited volume with ideally conducting walls, so as to nullify the wave amplitude at
the boundary. To combine an expression obtained from fundamentally quantum mechanical
principles, with an expression obtained from purely classical prospective, into a single formula (1)
is the inconsistency I wish to call out. Furthermore, the way the expression 
is derived in
textbooks [4, 5] ought to raise eyebrows, since it is based on completely improbable assumptions:
1. that the radiation, e.g. cosmic background, can be considered as enclosed in a cavity
2. that the enclosing cavity has ideal conductor walls so the wave amplitude is a mathematical
zero at the boundary
Both of these assumptions are crucial for considering the available phase space discrete, which is
necessary for the textbook derivation of 
factor. Even a miniscule deviation from ideal
conductor walls of the cavity immediately breaks the discreteness of phase space, and effectively
makes the number of hypothetically possible radiation modes of the same frequency infinite. Thus,
the factor 
in (2) would have to be derived from a different context, as I do below.
A measurement of radiation intensity with e.g. a bolometer, is effectively the measurement of
the energy contained in matter oscillators inside the sensitive element of the device. Consider an
oscillator immersed into radiation field. If the system oscillator+radiation is in a state at ,
the probability to find it in the same state at , from Schrödinger equation, is:
 

(5)
, where , are eigenstates and eigenvalues of -matrix. From (5), it follows,  
,
i.e. the transition rate is zero. This result is referred to as quantum Zeno effect [10, 11]. It is the
result of a coherent coupling (entanglement) between the oscillator and radiation modes,
manifested by the phase relationship between superposed eigenstates of -matrix in (5):
(6)
From (5), the transition rate is also zero in a more general case, if phase difference
between -states is any analytic function of time, i.e. if phases of -states predictably relate to
each other. In order for the transition to happen, the superposed eigenstates of -matrix must
undergo de-coherence, i.e. the phase relation (6) must be broken. There are various mechanisms
which may cause de-coherence of -states, such as:
1. Rayleigh scattering [12, 13]
2. Brownian motion [14, 15]
3. Dispersive media [16, 17]
4. Recombination of electron-hole pairs in semiconductors [18]
It is not in the scope of this paper to consider de-coherence mechanisms in detail. Rather, I shall
pursue a generic approach. I write (5) in a more general form, given (6) may no longer hold:
 
(7)
The probability distribution in (7) for an oscillator (e.g. a dipole) over radiation modes
is well known (see e.g. dipole radiation). For this discussion it is only important that the number
of radiation modes within a given spectral width  is large, so I can later take the limit .
If the matter is in equilibrium with radiation, a transition changes oscillator energy by 
with equal probability in either direction. In between transitions, the phase difference
evolves according to (6). The case of , where , is equivalent to consecutive
transitions in the same direction. In time , the phase of each of the -states in (7) undergoes a
total number  
of positive and negative increments with equal probability  Here, has a
meaning of mean free time between transitions, i.e. the de-coherence time. The resultant
increments are binomially distributed around mean , with variance of binomial distribution
 
. The variance in phase
 
. The
variance in phase difference is 

. Here  is the angular frequency.
Figure 1 shows numeric simulation of (7), with binomially distributed phases . The
calculation established the following:

(8)
The result (8) is interesting in a couple of ways. First, if , then , i.e. no transition
can occur if oscillator couples into a single radiation mode. This result can be obtained directly
from (5), since means the initial state is also the eigenstate of -matrix. Second, (8)
shows exponential decay over time, which is the characteristic feature of classical behavior, the
result of the de-coherence of radiation modes. In the limit ,  
, i.e. the probability
spreads equally among radiation modes. That is the consequence of simplification  
.
A more accurate calculation may require a dipole radiation distribution to be used for in (7), a
subject for a separate exercise.
From (8), in the limit of a large number of modes (), the transition rate is:

  
(9)
To evaluate the de-coherence time in (9), I consider a matter oscillator, formed when a set of
elements (e.g. electron-hole pairs) on the surface of the detector entangle through some medium
(e.g. electromagnetic field). An analog of such entanglement is a Cooper pair in superconductor,
mediated by phonon interaction. The surface area which encloses a set of elements in entangled
Figure 1
Blue line is a calculation of (7), with binomially distributed phases . The
red line is a plot of formula (8). The graphs were calculated using GNU Octave
code http://phystech.com/download/ph2.m with the following parameters:


 

The blue curve smooths out and becomes identical with red curve as .
state is limited by the coherence radius , where is the speed of light, is the
refractive index of the material, and  [rad/s] is the spread in internal transition frequencies of
the entangled elements. If is the number of entangled elements per unit surface area of the
detector; the dimensionless scattering rate;  the scattering frequency within  spectral
width, then, the de-coherence time of the oscillator is:



(10)
, and the transition rate:

 

(11)
In equilibrium, the loss of a number of oscillators in a particular mode is compensated by the
radiation-stimulated induction into the mode of the same number of oscillators. The energy balance
equation is


 
(12)
, where is the spectral radiance of incident radiation; is the efficiency of the conversion of
the incident radiation into the oscillator energy; is the number of oscillators per unit surface area
of the detector. I have to subtract zero-point energy term  
from the ensemble-average energy
in (12), because an oscillator cannot lose energy in the ground state. Combining (12) with (11):



(13)
The term in square brackets can be considered as pertaining to the incident radiation, and
parameters outside the brackets as properties of the detector. Then, (13) can be split into formula
for the spectral radiance, and formula for the detector efficiency:



(14)
(15)
, where  
can be interpreted as the number of entangled elements making up one oscillator.
The Planck’s formula (1) for the spectral energy density readily follows from (14):




(16)



(17)
I have shown the Planck’s law follows from consideration of radiation and matter oscillators as
parts of the same quantum system. I have argued against considering radiation as an entity which
exists and possesses properties independent of the matter it interacts with.
References
[1]
M. Planck, "Zur Theorie des Gesetzes der Energieverteilung im Normalspectrum,"
Verhandlungen der Deutschen Physikalischen Gesellschaft, vol. 2, p. 237245, 1900.
[2]
M. Planck, "On the Theory of the Energy Distribution Law of the Normal Spectrum," [Online].
Available: www.lawebdefisica.com/arts/distributionlaw.pdf.
[3]
L. Boya, "The Thermal Radiation Formula of Planck," 12 2 2004. [Online]. Available:
https://arxiv.org/pdf/physics/0402064.pdf.
[4]
L. Landau and E. Lifshitz, Statistical Physics, vol. 5, Elmsford: Pergamon Press, Ltd., 1980.
[5]
C. Kittel and H. Kroemer, Themal physics, W. H. Freeman, 1980.
[6]
J. Mather, E. Cheng, D. Cottingham, R. Eplee, D. Fixsen, T. Hewagama, R. Isaacman, K.
Jensen, S. Meyer, P. Noerdlinger, S. Read, L. Rosen, R. Shafer, E. Wright, C. Bennett, N.
Boggess, M. Hauser and e. al, "Measurement of the Cosmic Microwave Background Spectrum
by the COBE FIRAS instrument," The Astrophysical Journal, vol. 420, pp. 439-444, 10
January 1994.
[7]
"CMB measured intensity vs frequency," NASA Goddard Space Flight Center, 2017. [Online].
Available: https://asd.gsfc.nasa.gov/archive/arcade/cmb_intensity.html. [Accessed 2017].
[8]
G. Grundler, "The zero-point energy of elementary quantum fields," Novermber 2017.
[Online]. Available: https://arxiv.org/abs/1711.03877.
[9]
P. Jordan and W. Pauli, "Zur Quantenelektrodynamik ladungsfreier Felder," Zeitschrift für
Physik, vol. 47, no. 3, pp. 151-173, 1928.
[10]
S. E. Misra B., "The Zeno's paradox in quantum theory," Journal of Mathematical Physics,
vol. 18, no. 4, pp. 756-763, 1977.
[11]
S. A. Koshino K., "Quantum Zeno effect by general measurements," 3 2005. [Online].
Available: http://arxiv.org/abs/quant-ph/0411145.
[12]
H. Uys, M. Biercuk, A. VanDevender, C. Ospelkaus, D. Meiser, R. Ozeri and J. Bollinger,
"Decoherence due to Elastic Rayleigh Scattering," Phys.Rev.Letters, vol. 105, p. 200401, 2010.
[13]
M. Schlosshauer, Decoherence and the quantum-to-classical transition, Springer, 2007.
[14]
J. Paz, "Decoherence in Quantum Brownian Motion," 02 1994. [Online]. Available:
http://arxiv.org/pdf/gr-qc/9402007v1.pdf.
[15]
K. Hornberger, "Introduction to Decoherence Theory," Lect. Notes Phys., pp. 221-276, 2009.
[16]
C. Antonelli and e. al, "Sudden Death of Entanglement Induced by Polarization Mode
Dispersion," Phys.Rev.Letters, vol. 106, no. 8, p. 080404, 02 2011.
[17]
S. Salemian and S. Mohammadnejad, "Analysis of Polarisation Mode Dispersion Effect on
Quantum State Decoherence in Fiber-based Optical Quantum Communication," in
Proceedings of the 2011 11th International Conference on Telecommunications (ConTEL),
Graz, 2011.
[18]
Y. Jho, X. Wang, D. Reitze, J. Kono, A. Belyanin, V. Kocharovsky and G. Solomon,
"Cooperative recombination of electron-hole pairs in semiconductor quantum wells under
quantizing magnetic fields," PHYSICAL REVIEW B, vol. 81, pp. 155314-1, 2010.

Supplementary resource (1)

Data
December 2017
Sergei Viznyuk
... , where Ω (( ); ) = ln Γ( + 1) − ∑ ln Γ( + 1) ∈ (24) Hence, for elementary eigenstates, energy equals difference between entropy Ω ( , ( )) of equilibrium, and entropy Ω ({ }; ) of the sample, i.e. energy equals knowledge [about object state]. As entropies (5), (24) are in units of nats, so is the energy (6,16,22,23). Figure 5 demonstrates function √ℰ/ calculated for two sets of parameters ( ) using exact expression (6) and approximations (16), and (22). I plotted √ℰ/ instead of ℰ to show asymptotic behavior of (6) and (16) in comparison with quadratic form (22). Using (9,15,22) I obtain multivariate normal approximation [17] to multinomial distribution (1) as Figure 6 shows graphs of ln (( ); , ( )) as a function of 1 calculated for = 1000 and four sets of probabilities ( ), using exact formula (1), and multivariate normal approximation (25). ...
... From (35) the characteristic difference between adjacent energy levels is: Figure 11 shows the numeric calculation of the difference Δℇ between adjacent energy levels averaged over distinct event samples with the given value of , and = 3. I can use (46), with degeneracy of each level =6, as an approximation for combined energy levels (49) in expression for partition function (63), and obtain mean energy of modes with given as [23]: ...
... The conventional theory views radiation as existing "out there", decoupled from the matter. Such view leads to the infinite energy density due to the infinite number of hypothetically possible decoupled radiation modes, each multiplied by ℏ 2 ⁄ [23]. The conventional theory has no upper limit on ℏ , short of an artificial cut-off, usually assumed at Planck energy. ...
Preprint
Full-text available
Understanding of physical reality is rooted in the knowledge obtained from observations. The knowledge is encoded in variety of forms, from sequence of letters in a book, to neural circuits in a brain. At the core, any encoded knowledge is a sample of correlated events (symbols). I show the event samples bear attributes of a physical reality: energy, temperature, momentum, mass. I show that treating measurement as event sampling is consistent with predictions of quantum mechanics (QM). I discuss QM basics: wave function, Born rule, and Schrödinger equation, emphasizing their true meaning, which is rarely, if ever, mentioned in textbooks. I derive similar expressions using event sample as base construct, demonstrating the connection between QM and the presented model. I explain the mechanics of observation, and the role of observer. I show how model extends to include dispersion, decoherence, transition from quantum to classical state. I prove decoherence is a key factor in Fermi's golden rule, in Planck's radiation law, and in emergence of time. The controversial aspects of QM, such as wave function collapse, and measurement problem, do not appear in presented framework, which I call the knowledge mechanics (KM)
... (2) 1 It was shown, e.g., that the assumption about radiation existing "out there", in the open space, does not allow selfconsistent derivation of Planck's radiation formula [31]. It also leads to zero-point energy paradox [33]: the gravity from all zero-point energy modes would exceed the observed gravity by at least 58 orders of magnitude [32] 2 The term "quantum information" is widely used [35,15,3], but with no clear definition in sight. ...
... It was shown, e.g., that the assumption about radiation existing "out there", in the open space, does not allow selfconsistent derivation of Planck's radiation formula[31]. It also leads to zero-point energy paradox[33]: the gravity from all zero-point energy modes would exceed the observed gravity by at least 58 orders of magnitude[32] 2 The term "quantum information" is widely used[35,15,3], but with no clear definition in sight. ...
Preprint
Full-text available
I show any theory assuming state of an object, or even object’s existence, will be at odds with empirical evidence. I discuss QM relation with special relativity (SR). I argue all paradoxes are artifacts of factitious assumptions
... , where ( • ) is a characteristic decoherence time; ℰ is in joules ( ), and ℎ is in ( • ). The expression (10) for transition rate, containing ℰ 2 in one form or another, is present in a number of QM artifacts, from Fermi's golden rule [9], to Planck's radiation law [10], to the predicted gravityinduced decay [11]. ...
Preprint
Full-text available
I show the unitarily equivalent solutions of Schrödinger equation predict different measurement results in different observation bases. I identify the criterion for choosing an observation basis in which the predicted measurement results are consistent with observed macroscopic behavior
Article
Full-text available
We present results of detailed investigations of light emission from semiconductor multiple quantum wells at low temperatures and high magnetic fields excited by intense femtosecond laser pulses. The intensity and linewidth as well as the directional and statistical properties of photoemission strongly depended on the magnetic field strength and pump laser fluence. We also investigated the effects of spot size, temperature, excitation geometry, and excitation pulse width on the emission properties. The results suggest that the initially incoherent photoexcited electron-hole pairs spontaneously form a macroscopic coherent state upon relaxation into the low-lying magnetoexcitonic states, followed by the emission of a superfluorescent burst of radiation. We have developed a theoretical model for superfluorescent emission from semiconductor quantum wells, which successfully explained the observed characteristics.
Article
Full-text available
We study the decoherence of polarization-entangled photon pairs subject to the effects of polarization mode dispersion, the chief polarization decoherence mechanism in optical fibers. We show that fiber propagation reveals an intriguing interplay between the concepts of entanglement sudden death, decoherence-free subspaces, and nonlocality. We define the boundaries in which entanglement-based quantum communications protocols relying on fiber propagation can be applied.
Article
Since 1925, exactly four arguments have been forwarded for the assumption of a diverging (respectively --- after regularization --- very huge) zero-point energy of elementary quantum fields. And exactly three arguments have been forwarded against this assumption. In this article, all seven arguments are reviewed and assessed. It turns out that the three CONTRA arguments against the assumption of that zero-point energy are overwhelmingly stronger than the four PRO arguments.
Conference Paper
The implementation of quantum communication protocols in long optical fibers is limited by several decoherence mechanisms. In this contribution we review this mechanisms and analysis their effect on the quantum states. Because of asymmetry in the real fibers, the two orthogonal polarization modes are propagated at different phase and group velocities. The difference in group velocities results in Polarization Mode Dispersion (PMD). It is created by random fluctuations of the residual birefringence in optical fibers, such that the State of Polarization (SOP) of an optical signal will turn randomly over time, in an unpredictable way. In optical quantum communication, quantum state measurement is necessary. The bottleneck for communication between far apart nodes is the increasing of the error probability with the length of the channel connecting the nodes. For an optical fiber, the value of absorption and depolarization of a photon (i.e., the qubit) increases exponentially with the length of the fiber.
Article
In Weiterfhrung der Diracschen Theorie, bei der die elektrodynamischen Feldgren als nicht vertauschbare Zahlen (q-Zahlen) angesehen werden, werden hier, wenigstens im Spezialfall des Fehlens von geladenen Teilchen (reines Strahlungsfeld), Vertauschungsrelationen zwischen den Feldgren aufgestellt, die eine relativistisch invariante Form haben. Es wird gezeigt, da diese Relationen auch ohne Benutzung der Fourierzerlegung des Feldes formuliert werden knnen. Ferner wird eine allgemeine mathematische Methode angegeben, die gestattet, Relationen zwischen q-Zahlen, die von Raum- und Zeitkoordinaten stetig abhngen (q-Funktionen), umzudeuten in Relationen zwischen geeignet gewhlten Operatoren, die auf verallgemeinerte, vom ganzen Feldverlauf abhngige -Funktionen (Funktionale) anzuwenden sind.
Article
We present theoretical and experimental studies of the decoherence of hyperfine ground-state superpositions due to elastic Rayleigh scattering of light off resonant with higher lying excited states. We demonstrate that under appropriate conditions, elastic Rayleigh scattering can be the dominant source of decoherence, contrary to previous discussions in the literature. We show that the elastic-scattering decoherence rate of a two-level system is given by the square of the difference between the elastic-scattering amplitudes for the two levels, and that for certain detunings of the light, the amplitudes can interfere constructively even when the elastic-scattering rates from the two levels are equal. We confirm this prediction through calculations and measurements of the total decoherence rate for a superposition of the valence electron spin levels in the ground state of 9Be+ in a 4.5 T magnetic field.
Article
Thesis (Ph. D.)--University of Washington, 2005. The main theme of this dissertation is the emergence of the "classical" world of our experience from quantum mechanics. The underlying key ingredient to be discussed is decoherence, i.e., the dislocalization of local phase relations through entanglement with environmental degrees of freedom. The analysis of a decoherence-based quantum-to-classical transition is approached from four different points of view.First, key features of the decoherence program are clarified, and implications of decoherence for the quantum measurement problem and for the existing interpretations of quantum mechanics are extensively discussed.Second, a critical analysis of the "self-induced decoherence" approach is presented. This approach has claimed to provide an alternative viewpoint on decoherence and to bypass some of the conceptual difficulties often associated with environmental decoherence. It is shown that this approach not only falls short of describing a physically relevant decoherence mechanism, but also fails in a large class of model systems.Third, a derivation of the quantum-probability concept and the Born rule from symmetries in quantum entanglement is analyzed. In particular, important key assumptions not mentioned in the original exposition of the derivation are identified.Fourth, recent experiments are discussed that have demonstrated the existence of superpositions of macroscopically distinct states. The crucial role of decoherence in these experiments is analyzed, and the implications for minimal non-collapse quantum mechanics and physical collapse models are investigated.Based on these studies, it is concluded that pure quantum features, most notably, quantum entanglement and the resulting decoherence effects, are well capable of explaining the emergence of classicality in unitary quantum mechanics.