Black-Body Radiation in an Accelerated Frame
Abstract
We derive Planck’s radiation law in a uniformly accelerated frame expressed in Rindler coordinates. The black-body spectrum is time-dependent by its temperature and Planckian at each instantaneous time, but it is scaled by an emissivity factor that depends on the Rindler spatial coordinate and the acceleration magnitude. The observer in an accelerated frame will perceive the black-body as black, hyperblack, or grey, depending on its position with respect to the source (moving away or towards), the acceleration magnitude, and the case of whether it is accelerated or decelerated. For an observer accelerating away from the source, there exists a threshold on the acceleration magnitude beyond which it stops receiving radiation from the black-body. Since the frequency and the number of modes in Planck’s law evolve over time, the spectrum is continuously red or blue-shifted towards lower (or higher) frequencies as time progresses, and the radiation modes (photons) could be created or annihilated, depending on the observer’s position and its acceleration or deceleration relative to the source of radiation.
I Introduction
The status of temperature in relativistic thermodynamics remained a subject of ongoing debate, marked by the lack of a universal agreement in the physics community regarding the transformation of temperature for moving bodies (Farias, ; Derakshani, ). In their works, both Planck and Einstein independently derived that the temperature of a moving body would transform as , with representing the Lorentz factor associated with the velocity of the moving body (or moving observer, if the body is at rest) (Planck, ; Einstein, ; Einstein2, ). However, under different assumptions, Ott and Arzelis derived the transformation of temperature to be (Ott, ; Arzelies, ). Furthermore, Landsberg proposed the invariance of temperature under transformations between inertial frames, namely, (Landsberg, ; Lansberg2, ). Each of these different formulation for temperature transformations in inertial frames is supported by compelling arguments, hence, although the most widely-accepted theory is the one by Einstein-Planck (Einstein, ; Planck, ), the attempt to reconcile these different approaches remains an ongoing research. A recent advancement in the field of special relativistic thermodynamics was achieved by Nakamura in (Nakamura, ; Nakamura2, ) who revised the Israel-van Kampfen covariant, inverse, 4-temperature (vanKampen, ; Israel, ). This revision allows the derivation of the three distinct temperature transformations within the Israel-van Kampen formulation, depending on the definition of the 3-dimensional volume and the chosen decomposition of the 4-momentum (Nakamura, ; Nakamura2, ).
Moreover, Tolman pioneered the study of general relativistic thermodynamics in (Tolman, ; Tolman2, ), by proposing a law that in a gravitational field, the temperature of a system in thermal equilibrium is inversely proportional to the square root of the gravitational potential at that location. This law can be expressed as , where represents the purely-time component of the metric tensor . In general relativity, the metric tensor describes the gravitational field as the curvature of spacetime, and the component corresponds to the gravitational potential in the Newtonian limit (Einstein-1, ; Carrol, ).
On the other hand, the concept of temperature can be defined from several theoretical framework. In thermodynamics, temperature is defined as the parameter shared by two bodies in thermal equilibrium, a state characterized by the absence of net heat transfer between the bodies (Salinger, ); this definition is widely accepted among -century physicists including Clausius, Carnot, Kelvin, and many others. From the statistical mechanics approach, it is defined as the average of kinetic energy of a many-body system, and this is described by the relation , with is the Boltzmann constant (Boltzmann, ). Statistical mechanics provides a procedure for bridging macroscopic observables of a system, such as temperature and entropy, to its microscopic properties, i.e., the position, velocity, and kinetic energy of the individual degrees of freedom (Liboff, ). Combining these two perspectives, temperature can be defined as an observable that emerge from the statistical properties of individual degrees of freedom (i.e., the kinetic energy) of a many-body system that is in a thermal equilibrium with its surrounding. As far as we are concern, this is the current status of the temperature conceptualization widely-accepted by physics community at the moment.
In our study, we adopt a more pragmatic approach to the concept of temperature -specifically, through the use of Planck’s law on black-body radiation, originally proposed in 1900 (Planck2, ). This law initiated the birth of quantum mechanics and remains valid through experiments to the present day, and it founds wide applications in pyrometric measurements. As explained in the preceding paragraphs, the thermodynamic definition of temperature -a parameter indicating two systems are in thermodynamic equilibrium- is particularly applicable for temperature measurement in a rest frame. The measurement apparatus must establish direct contact with the system and afford sufficient time for thermalization until a (thermal) equilibrium is reached. In this case, the temperature is well-defined. However, for these two systems to be in thermal equilibrium, in prior, they need to be in a mechanical equilibrium, while maintaining a direct contact to undergo thermalization. This requirement is challenging to fulfill if one of the systems is not at rest relative to the other (vanKampen, ; vanKampen3, ). In principle, at least by the definition originating from thermodynamical approach, the notion of temperature can only emerge as a consequence of local measurement (Mares, ). Furthermore, a quantum-field theory calculation, using the Unruh-DeWitt detector as a thermometer moving through a thermal bath, gives a result that the particle distribution is non-Planckian, which made it difficult to define temperature in this context (Costa, ). Based on this work, then Landsberg and Matsas proposed an argument on the impossibility of a universal relativistic temperature transformation, due to the fact that there is no continuous function that could map the non-Planckian distribution from (Costa, ) to a Planckian one (Matsas, ; Matsas2, ). These reasons supports the argument that the concept of temperature of a moving body is not well-defined (Mares, ).
However, using the Planckian spectrum, one can predict the temperature of a distant object without necessitating direct contact, thereby skipping the need to reach thermal equilibrium with the object under observation. This methodology, treating temperature as a derived parameter, finds widespread application – from the infrared thermometers to the calculation of the temperatures of celestial bodies such as stars and the cosmic microwave background (Bbracewell, ). The process involves the collection of the complete radiation spectrum emitted by the object and, -under the assumption that the object behaves as a black-body-, the comparation of the spectrum data with the theoretical black-body spectrum. Temperature prediction can be carried out using either Wien’s displacement law (Wien, ; Wien2, ) or the Stefan-Boltzmann law (Stefan, ). In the case of a black-body source, these two distinct temperature calculation methods coincide. However, it is important to note that assuming all objects behave as perfect black-bodies is overly restrictive. For a more realistic approach, the collected data should be compared to a calibrated grey-body spectrum derived from the black-body spectrum, but with emissivity or absorptivity values set below 1. This method, to the best of our knowledge, represents the sole viable approach for ’measuring’ the temperature of a moving object. Furthermore, experiments based on this measurement technique are relatively straightforward to conduct. These facts emphasize the importance of Planck’s law in the determination of temperatures of moving bodies.
We derived Planck’s law and calculated the black-body spectrum in a uniformly accelerated frame expressed in Rindler coordinates. The reason behind our quest on this subject can be explained as follows: In the framework of general relativity, an inertial frame is an idealization, hence one needs to move to a more general reference of frame. As a first realistic step towards achieving this objective, we consider a uniformly accelerated frame, expressed in Rindler coordinates. The calculation of the black-body spectrum in the inertial frame had been done in (Peebles, ; Heer, ; Henry, ). These works demonstrated that the Planckian spectrum is invariant under Lorentz transformation, and the effective temperature of the moving object depends on its velocity and the direction of observation (Peebles, ; Heer, ; Henry, ). Another different approach shows only the zero-point temperature term is invariant under Lorentz transformation (Bouyer, ). Attempts to calculate the black-body spectrum in an accelerated frame had been done in (Lee, ), but to the best of our knowledge, the calculation of the spectrum in Rindler coordinates is still lacking in the existing literature. In the present work, we follow the (effective) ’directional-temperature’ approach as in (Henry, ), but instead of using the temperature in Planck’s law, we substitute it with variables that maximize the energy in the spectrum (in our case, we use the ’maximal’ frequency). With this substitution, we avoid the problem associated with temperature transformation, as we are certain about the Lorentz transformation for the ’maximal’ variables.
The main result presented in this article is the black-body spectrum in an accelerated frame. The spectrum has an explicit time dependency on its temperature, Planckian at each instantaneous time, but it is scaled by a factor dependent on the spatial coordinate and the acceleration magnitude. The spatial, coordinate-dependent scale factor is proportional to , while the scale factor related to the acceleration is proportional to , where the sign depending on the observer is either deccelerated or accelerated. This scale factor could be physically interpreted as the emissivity factor of the source, hence, in an accelerated frame, a black-body could be perceived as grey or ’hyper-black’, depending on the magnitude of the acceleration. Furthermore, the variables in Planck’s law, specifically the number of modes and frequency, evolve over time. However, since we are considering all the possible value of frequencies from zero to infinity, the time dependence of the frequencies does not explicitly shown in the law. The Planckian spectrum is continuously red or blue-shifted towards lower (or higher) frequencies as time progresses. In the accelerated frame, we demonstrate that the radiation modes (photons) could be positive or negative, depending on the acceleration or deceleration of the observer, and zero for vanishing acceleration. In the end, assuming the validity of Wien’s displacement law in an accelerated frame, the time-dependent, (directional)-temperature of a body in an accelerated frame is given by . It is worth noting that the spatial, coordinate-dependent scale factor resembles the temperature scale factor in Tolman’s theory of general relativistic thermodynamics (Tolman, ), and for the transformation of the temperature returns to the ones obtained in (Peebles, ; Heer, ; Henry, ).
The paper is structured as follows: Section II contains the derivation of black-body radiation in an accelerated frame. First, we introduce the properties of the Rindler space, followed by the transformation from Minkowski to Rindler coordinates for each variable contained in Planck’s law. With these transformed variables, we derive Planck’s distribution law in the accelerated frame. Section III contains discussions and the physical interpretation of the results, including the aberration of light, relativistic beaming, and red/blueshifts due to the relativistic Doppler effect, creation/annihilation of modes in the accelerated/decelerated frame, the emissivity factor of the black-body source, and the transformation of temperature, assuming Wien’s law holds in the accelerated frame. Finally, in Section IV, we provide a conclusive summary of our work and outline some insights into potential further research. We have included appendices to ensure the self-contained nature of this paper. The appendices contain the derivation of the black-body spectrum in a frame moving with a constant velocity. This supplementary content aids readers in understanding each step of the spectrum derivation in the accelerated frame. One crucial aspect is that we have aimed to keep this paper as classical as possible. The only quantum assumptions we adopt are the explicit use of the de Broglie postulate for photon momentum (with is the photon wave-vector and is the Planck constant) and the assumption of energy discreteness (with is the photon angular frequency and are integers) as the requirement to derive Planck’s law.
II Blackbody Radiation in Accelerated Frame
II.1 Rindler Coordinates
Let be a Minkowski space equipped with a metric . The (global) coordinate =, with is the time coordinate and is the spatial part, is used to parametrize . In this coordinate, the metric is a diagonal metric with signature . Naturally, one could attach an inertial reference of frame to the coordinate , and let be an observer at rest with respect to this inertial frame. Let us call this frame/observer as
Suppose we have another observer that moves with a constant acceleration of magnitude in the direction of the -axis, with respect to . The trajectory of in is , with is a parameter (a proper time with respect to ), satisfying:
(1) | ||||
(2) |
and being constant. The 4-acceleration is where the non-zero components are:
such that the magnitude of is . Note here that is positive definite: . From (1) and (2), the trajectory of satisfies , describing a collection of hyperboloids with asymptotes at the lines and . The asymptotes divide into 4 regions, and let us focus on one of the region where and , usually labeled as the (right) Rindler wedge, see FIG. 1.

One could choose a new coordinate patch =, to parametrize the Rindler wedge as follows:
(3) |
and their inverse:
(4) |
with , , and are constant parameters. The (right) Rindler wedge is covered by the coordinate patch with . For , the coordinate is usually known as the Rindler coordinates in the Lass/radar representation (Carrol, ). In this coordinate, (1)-(2) becomes:
Notice that along the trajectory , is constant and is proportional to the proper time , although it is shifted by a constant amount . An observer moving along constant will experience an acceleration of magnitude:
(5) |
Since is positive definite, then so does : . Naturally, the coordinate is adopted by the observer/frame .
One can obtain the infinitesimal transformation of (3), which is:
(6) | ||||
where we write , and their inverses:
(7) |
with and . is the time-dependent rapidity that is linear to the “proper time” . For , the acceleration vanishes, and (6)-(7) will return to the standard Lorentz transformation (81)-(82).
The line element in Minkowksi space could be written in Rindler coordinate using (7) as follows:
(8) |
The coefficient of the line element gives the Rindler metric in Lass representation. One could also consider the left Rindler wedge (the region IV in FIG. 1), where it could be covered by a coordinate patch similar to the RRW, but with the flip in the sign of (4) as follows:
(9) |
In this region, the future-directed time-like Killing vector is , instead of in region I.
II.2 The Aberration of Light in Accelerated Frame
Suppose with respect to an inertial frame we have an object moving with a constant velocity as follows:
According to the accelerated frame , the velocity of the object is:
Notice that the components of are functions of the Rindler coordinate . Using the transformation (6)-(7), one can obtain the relation between the instantaneous and as follows:
(10) |
(11) |
Relation (10)-(11) could be derived from the standard velocity addition formula (or, similarly, from the relativistic acceleration formula). Their inverses could also be obtained as follows:
In the following paragraphs, we will derive the light aberration formula for an accelerated observer . For a brief review on the light aberration formula for an inertial, moving observer, one could refer to Appendix B, or consult (Johnson, ). Let us consider a black-body source (photons inside a cavity) at rest with respect to an inertial frame Electromagnetic radiations (photons) are emitted by the black-body with propagation velocity in the direction of , with is the (polar) observation (or inclination) angle between the axis and on plane . The velocity of the photon with respect to frame is:
(12) |
The 4-velocity of the photon satisfies the null-vector condition, see Appendix X. To derive the 3-velocity in , let us first write the 4-velocity in Rindler coordinate:
where is a real parameter (usually the proper time with respect to the moving object) and is the 3-velocity in :
Using the null-vector condition for light, we have:
In , the light only propagates in the direction on plane , so even in , the component of the velocity of light is zero. The null condition then becomes:
Writing and the 3-velocity of the photon in is:
(13) |
with and are the propagation direction and the inclination angle of the photon at , respectively. Notice that since the observer is accelerated in the direction , it will receive only photons that are moving towards the observer, namely, the ones that have the component in the velocity. The relation between and can be obtained using the velocity addition formula in the accelerated frame (10)-(11). Inserting (12)-(13) to (10)-(11) will result in the light aberration for an accelerated frame :
(14) | ||||
(15) | ||||
(16) |
where the last equation is obtained from the trigonometry identity . One could also obtained their inverses:
(17) | ||||
(18) | ||||
(19) |
The aberration formula in has exactly the same form with the ones in a constant moving frame (see relation (89)-(91)), however, it should be kept in mind that is time-dependent, namely , while is constant.
II.3 The Doppler Effect in Accelerated Frame
In this paper we will only consider the longitudinal relativistic Doppler effect (LongitudDoppler, ), where the observer/source velocity has component parallel to the wave propagation. To derive this, let us first consider the 4-momentum constraint (107) which is valid for any reference frame. The photon is massless, so using the de Broglie postulate and , (107) becomes the dispersion relation:
(20) |
for the wave vector of the photon satisfies In the dispersion relation becomes:
Since the coordinate axis is perpendicular to , one could write and , with is the inclination angle at Therefore, the 4-momentum of a photon with frequency according to the rest frame (with respect to the black-body source) is , satisfying:
(21) |
Now, let us obtain the photon’s 4-momentum according to the accelerated frame , that is The dipersion relation (20) is also satisfied in , however, written in different coordinate, it becomes:
(22) |
(notice that the dispersion relation comes from hence the metric components needs to be taken accounted). Since the coordinate axis and are also perpendicular to one another, one could defined that and . Notice that is the inclination angle according to . With this, then the 4-momentum at the accelerated frame is:
(23) |
is the frequency of the photon, observed by . Here, we assume that de Broglie postulate is still valid in namely the relation is satisfied for every position in space and each instantaneous time. One could refer to Appendix B for a more detailed explanation on the photon’s momentum.
It needs to be kept in mind that there exists another solution to (20), namely: giving another possible value for and in and and in . However, since we are considering only the right Rindler wedge (RRW), the reasonable case for an observer in RRW to receive the signal is by the waveform that moves to the right (in direction). For an observer in the left Rindler wedge (LRW, which we will consider in Section IV), we should consider the case where
The (generalized) momentum is naturally a covector, however, in this paper, we use its contravariant counterpart to be consistent with the 4-velocity (12)-(13). The entire results do not depends on the choice of vector/covector for the derivation. Since the 4-momentum is a 4-vector, the transformation between and follows (6)-(7), where we write the cofficient of the transformation in terms of hyperbolic functions:
(24) |
with their inverses:
(25) |
and Moreover, by inserting (21)-(23) to the transformation (24), we can obtain the 3 equivalent forms of the relativistic Doppler effect in the accelerated frame as follows:
(26) | ||||
(27) | ||||
(28) |
together with their inverses:
(29) | ||||
Notice that is a function of time and position in the Rindler coordinate. From (26) and (29), one could retrieve the aberration formula (14)-(18).
Let us consider one of the light aberation formula (17). Differentiating (17) with respect to any parameter give:
while using (18), (28), and the chain rule , gives:
(30) |
with . is the rate of change of the inclination angle as observed by the accerelated observer . In , the inclination angle is time-dependent, satisfying equation (14); differentiating (14) with respect to , we obtain:
(31) |
and hence (30) can be simplified into:
(32) |
Using the transformation of the inclination (polar) angle (30), one could obtain the transformation for the solid angle as follows:
(33) |
where we use the Doppler effect (28) and the fact that the transformation of the azimuth angle satisfies .
As explained in the preceeding subsection, observer is accelerated along the direction; causing it to receive only photons approaching from that direction, namely, those with the component in their velocity. Moreover, is moving away from the black-body source at rest with respect to As a consequence, the frequency of the photons received by the accelerated frame is redshifted by equation (26). However, in constrast with the inertial case presented in Appendix B where the redshift remains constant, the redshift in the accelerated frame increases in time, with the wavelength of the photons shift progressively towards the infrared range. To obtain the blueshift case, one could invert the situation by flipping the sign of (or ), effectively moving backward in time.
II.4 Density of State in Accelerated Frame
Another important parameter we must determined in the accelerated frame is the modes distribution of the photons in the cavity. To determine how this quantity transform from the inertial to accelerated frame, first we need to determine transformation of the phase-space volume. In the inertial frame and the accelerated frame , the infinitesimal 3-volume elements are defined as, respectively:
(34) | ||||
To obtain the density of states of our black-body system, we need to calculate how much modes are inside a volume element. Let the finite volume element in be . The spatial length is defined by 2 simultaneous events and along the -axis where , and . According to the length between these 2 event is For an observer at whose accelerated in the -direction with respect to the 2 events and are not simultaneous, but are separated by a time interval . Let us labeled the spatial length of these 2 events in as where could be obtained from (3). Taking the infinitesimal limit , the transformation of the infinitesimal spatial length satisfies (6). Notice that the measurement of length in is obtained from 2 simultaneous event, hence in , this is not the case in Therefore, the (infinitesimal) spatial length of in is:
(35) |
Furthermore, the infinitesimal volume element is perceived by an observer in as the volume swept by the plane from to along an infinitesimal time interval This is the physical interpretation of see also discussion on Appendix B. Inserting (35) to (34) gives the volume transformation:
(36) |
For the next step, we need to determine the infinitesimal volume element in the momentum space, this could be obtained from (21)-(23); in the inertial frame and the accelerated frame , they are, respectively:
(37) | ||||
The 4-momentum is subjected to the energy-momentum constraint, see Appendix B:
(38) |
Taking the differential of (38) gives:
(39) |
Furthermore, using (4), one could show that:
(40) |
so that the transformation of the 4-momentum (24) can be written as follows:
(41) | ||||
with and Differentiating (41) with any parameter will give the infinitesimal version of 4-momentum transformation as follows:
(42) | ||||
Note that in equation (42), there exist infinitesimal coordinates components, namely and . This is due to the fact that the 4-momentum transformation from inertial to accelerated frame (41) varies with coordinates , in constrast to the transformation between inertial frames (98) in Appendix B, which is independent from the temporal and spatial coordinates.
Similar to the inertial case in Appendix B, using the hypersurface constraint and inserting the infinitesimal momentum constraint (39), (42) becomes:
(43) | ||||
and . Now, one can construct altogether the infinitesimal phase-space volume element in frame as follows:
Inserting (36) and (43) into the equation above, we obtain the transformation of the phase-space volume element from the inertial frame to the accelerated frame :
(44) |
Note that by the definition of the wedge product the terms containing equivalent components of differential forms, i.e. , vanishes. Using (40) and then (21), one could simplify (44) into:
(45) |
One the other hand, we have, from (24):
and therefore, using (21)-(23):
(46) |
where we write the transformation (45) in terms of the Doppler factor . (46) is the relation between the phase-space volume element in and
Let us define the relativistic distribution function, or density of state as the number of world-lines that cross the phase-space, i.e, the states , per phase-space volume element (Liboff, ):
(47) |
We assume that the density of state is invariant under coordinate transformation, namely:
with is the number of states per phase-space volume of the accelerated frame . This is a reasonable assumption, and for a detailed explanation, one could refer to Appendix B. Some previous works in the existing literature have assume the invariance of the number of states instead of , however, this assumption is not suitable for our case because our treatment in this works relies only on a coordinate transformations at some part of the Minkowski space, i.e., the Rindler wedge. While these transformations leave the world-lines invariant, they alter the unit volume due to the coordinate transformation, leading to a different count of world-lines crossing the unit volume. Similar arguments are implicitly employed in (Peebles, ; Heer, ; Henry, ) as well.
By the invariance of the density of state under (infinitesimal) coordinate transformation, one could obtain the relation between the number of states/modes in inertial frame and the accelerated frame as follows:
(48) |
II.5 Black-body Radiation in Accelerated Frame
We already had all the Rindler-transformed variables to derive the Planck’s law in a uniformly-accelerated frame. Similar to the inertial frame case that we derived in Appendix B, in the rest frame , we use the Planck’s law version of equation (77) as follows:
(49) |
where the term containing the temperature of the black-body, i.e., , is replaced by , the angular frequency that maximize the radiation energy of the black-body. This is possible by the Wien’s displacement law, see a detailed explanation on the Appendix A. The reason for this substitution is to avoid the problem of the temperature in moving bodies. As explained in the Introduction, currently, there is no consensus regarding the transformation of temperature in moving bodies (Farias, ). However, we convincingly agree on the transformation of the black-body’s frequency, hence, if satisfies the relativistic Doppler effect, then so does the . The dimensionless constant in (49) is the Wien coefficient, obtained from solving the non-linear equation that arise from the maximization of energy with respect to . One could refer to Appendix B for a detailed derivation.
In this section, our objective is to ascertain the validity of the distribution (49) in the accelerated frame . Inserting the relativistic Doppler effect (29), the transformation of solid angle (33), the volume contraction (36), and the transformation of the number of modes distribution (48) to (49), we obtain:
(50) |
One can observe that in the right hand side of (50) there is a scale factor:
(51) |
that prevents the distribution (50) to be the original Planckian (49). The scale factor depends on the acceleration and the spatial coordinate . For the case with where the acceleration vanishes, the distribution becomes Planckian as in (49). The fundamental distinction between the distribution (50) and the distribution (52) in the moving inertial frame lies in the fact that all the variables in the (50) are coordinate-dependent, namely, they vary with and , which represent the proper time and position in . This is due to the fact that the transformations of these variables, namely (29), (33), (36), and (48) are coordinate-dependent. However, since we are considering all the possible value of frequencies from , the time dependence of does not explicitly shown in the law. Consequently, the ’black-body’ spectrum (50) in the accelerated frame experiences a redshift that increase over time and varying with position, where the wavelength of the photon increases toward the infrared direction. This behavior distinguishes (50) from the black-body spectrum in the inertial frames, where the spectrum shifts similarly occur, but the magnitudes of these shifts remain constant, as indicated by equation (100).
Similar with the inertial case in Appendix B, is the “maximal” frequency of the distribution (49) of a black-body in the rest frame, viewed by the accelerated observer . Let us check if also maximizes the energy in distribution (50). By multiplying the left and right hand side of (50) with one obtains:
(52) |
is the energy of modes with frequency observed in the accelerated frame . The value of that maximize can be obtained by solving the following equation:
(53) |
Note that the scale factor does not affect the equation we need to solve. Let us define , the equation from (53) that we need to solve becomes:
(54) |
which is the same as in the inertial cases, namely (74) and (115). The solution is then similar:
with is the Lambert -function (Euler, ). Hence, the Wien’s coefficient is invariant, even under the transformation from the inertial to accelerated frame: . Suppose is the solution to (53), therefore:
or simply : the transformation (26) sends “maximal” frequency in an inertial frame to a “maximal” frequency in an accelerated frame at an instantaneous time. (50) becomes:
(55) |
As previously mentioned, the presence of the scale factor introduces deviations from the Planckian distribution of the black-body radiation in the accelerated frame . For an observer moving along , it will observe a Planckian distribution but with a redshift increasing over time. For another accelerated observer moving along a constant , it will experience a similar increasing redshift, but now the distribution will be scaled by a factor of . Such an observer will interpret this as a change in the emissivity of the black-body source; this facts will be analyzed in detail in the next section.
It needs to be kept in mind that the Planck formula (55) applies in the right Rindler wedge (RRW). With a similar derivation, one could show that the same formula is also valid for the left Rindler wedge, but with the Doppler formula (26)-(27) changed into:
(56) | ||||
(57) |
This is due to the fact that there exists another solution to the dispersion relation (20), namely
III Physics in Rindler Coordinate
III.1 Acceleration and Deceleration
The "Planck’s law" (55) describes the frequency spectrum measured by an observer whose accelerating away from the radiation source along the direction, as described in Section II. This observer moves with an initial velocity and undergoes an acceleration with magnitude . Let us add three possible scenarios besides the ones we had considered in the previous section. First, note that the origin of the rest frame does not coincide to the origin of frame namely, the point at is equivalent to point of at while the point of at is the point of at see FIG. 2 and FIG. 3.


To simplify the analysis, let us set the black-body source to be located at any point in the range of or , depending on the scenarios. With respect to , let frame have an initial point at , that means the clock in is slightly delayed by the amount of . Then moves with the time-dependent rapidity , which could satisfy one of the four following case:
(58) |
noting that the initial rapidity . The sign in front of the initial rapidity depends on the rate of change of the position of observer with respect to the source; it is positive if their distance increases in time, and negative if it decreases. Meanwhile, the sign in front of depends on the rate of change of . If it increases in time (accelerating), then it has the same sign with , and vice versa. Let us list all the possible cases. The first case, where , corresponds to the case described in Section II, used to derive the Planck’s law (55), see FIG. 4.

The second case, where , describes an observer decelerating away from the source; it moves in the direction, but the speed is decreasing in time; we will only consider this case until the observer reach zero velocity as a result of deceleration. See FIG. 5.

The third case with describes an observer that is decelerating towards the source, with a decreasing speed. Similar to the second case, we will only consider the case until the velocity is zero, see FIG. 6.

The last case with describes an observer accelerating towards the source, see FIG. 7.

The concept of acceleration and deceleration is formally defined as the rate of change of the speed, specifically, an acceleration if and deceleration if . It is also important to highlight that the notion of acceleration/deceleration do not necessarily correlate to redshift/blueshift, as elaborated in Subsection III C.
III.2 The Relativistic Beaming
The alteration of the inclination angle in the moving frame leads to a peculiar phenomenon known as relativistic beaming where, in moving a frame, the rays of light shows a tendency to either converge or diverge in the direction of motion, depending on whether the observer is approaching or receding away from the source (Cohen, ). For a frame moving away with constant velocity of , the inclination angle will transform by (90). For the case where the observer is moving towards the source, one could flip the sign in front to become positive. The photons that reach the observer will have the inclination angle within the range of . It can be demonstrated that for the “moving towards” case (), the condition holds, implying . It implies that for an observer approaching the source, the light rays tend to converge to the direction of motion, see FIG. 8.

For the “moving away” case (), there are 2 conditions that affect the light rays bending. First, if holds, then , implying that in this interval, the light rays converge. Second, if the condition holds, the light rays diverge in the direction of motion, namely . This case occurs when , see FIG. 9.

It is worth noting that in a moving inertial frame, is time-independent, resulting in a constant angle deviation.
For the case of an accelerated frame , the inclination angle will be perceived as time-dependent, satisfying (15). Therefore, a similar beaming phenomenon as in the moving inertial cases occurs. In the scenario where the accelerated observer approaching the source, decreases in time, resulting in a convergence at , see FIG. 10.

In the scenario where the accelerated observer receding from the source, will diverge until it reaches the maximal values, namely for and for . From this point, it will stop diverging and start to converge as time increases, as illustrated in FIG. 11.

It is important to emphasize that our analysis of the relativistic beaming phenomenon is restricted to specific cases, namely Cases 1 (accelerated away) and 4 (accelerated toward). The decelerated cases (namely, Case 2 and 3) can be derived by inserting the condition (58) into (15). In the context of relativistic beaming, Case 3 (decelerating toward) will experience a similar phenomenon to Case 1 (accelerating away), while Case 2 (decelerating away) is similar to Case 4 (accelerating towards). Observing only the beaming of lightrays, these two pairs of cases are indistinguishable.
To be more precise, one could obtain the rate of change of the inclination angle by inserting (31) to (15)- as follows:
(59) |
Note that the inclination angle satisfies ; in these ranges, , are positive definite, and the quantity could be positive or negative, depending on whether or vice versa. For the case where the inclination angle at the rest frame satisfies , will be positive definite, so if the quantity is positive, then ; the angle at will diverge. Otherwise, , and the angle will converge. Meanwhile, for the case of , will be negative definite, so the angle at will converge if is positive -resulting in -, otherwise it will diverge.
III.3 The Redshift and Blueshift
Considering the quantity with is a 4-vector and is a 4-momentum for a single photon, one could show that is invariant under Lorentz transformation. A simple way to prove this is to calculate at frame using (96), and at frame using (97). With the help of (78), (89)-(90), and (101), one could show that .
Furthermore, with the invariance of in inertial frames, one can conclude that a plane wave moving in -direction as follows:
(60) |
is invariant under Lorentz transformation. Hence, a plane wave in is also a plane wave in but with the change in frequency satisfying the Doppler effect (101). The Doppler effect causes the frequency in the inertial moving frame to undergo a redshift if it is moving away from, -and blueshift if it is approaching, the source. In the inertial frames, the redshift and blueshift are constant in time.
However, this is not the case in an accelerated frame. Let us take only the temporal part of (60) at for simplicity, namely with is the phase-angle. The accelerated observer will measure the photon of constant frequency at as having frequency satisfying the relativistic Doppler effect (29). Without loosing of generality, let us set so that satisfies:
(61) |
considering only the case of from (58). The phase-angle of the wave of the single photon with respect to is:
(62) |
is time-dependent, since is a function of time . This is different with the phase-angle in the inertial case, where could be written simply as , since in the inertial case is independent of time. Therefore, a single mode of (the temporal part of) the radiation wave with respect to is:
(63) |
(63) is not necessarily a plane-wave, in contrast to its inertial counterpart (60), since the frequency measured by experiences a shift increasing in time. The direction of the frequency’s shift depends on the rate of change of the frequency with respect to the proper time, namely:
(64) |
gives a redshift while gives a blueshift.
Now, having the definition of redshift and blueshift, let us consider the 4 cases of accelerated observers in Section III A. Note that by the relativistic Doppler effect, all these 4 cases will experienced time-dependent shifts on their frequencies: the first and the third cases will have their frequency shifting in time towards the infrared direction (redshift), while the second and the last cases will have the shift toward the ultraviolet direction (blueshift). These pairs of cases could be distinguished by their instantaneous frequency with respect to the frequency of the source at rest: the first case has , while the third has , although they experience increasing time-dependent redshift on their frequencies. A similar condition occurs for the second case, with , and the last case, with ; where both experience increasing blueshift on their frequencies. Comparing the definition on frequency’s shift with the definition of acceleration and deceleration in Section IIIA, it is clear that that they are not correlated, i.e., acceleration does not always correspond to redshift, while deceleration does not always correspond to blueshift.
The "maximal" frequency similarly transformed as (61), therefore the rate of change of satisfies (64). The peak of the spectrum will shift towards higher or lower frequencies as time progresses, depending on which one from the 4 cases is satisfied. See FIG. 12.

One may wonder the relationship between the frequency shift discussed in this work and the gravitational red/blueshift phenomena (Eddington, ). Gravitational frequency shift is an alteration in the photon’s frequency as it travels a gravitational field. The physical clock at different location along the field has different rate of time, therefore, photon appears to be red/blue-shifted relative to the frequency of the clock (Okun, ). On the other hand, the frequency shift considered in this work is due to its velocities; it has the same origin with the relativistic Doppler effect, with the main difference lies in the time-dependent nature of this shift. Now, by the equivalence principle, that asserts the indistinguishability of gravity with acceleration, one might expect an equivalency between the frequency shift due to acceleration and gravitational red/blueshift. However, given a stationary source of gravity, the gravitational redshift remains constant over time, while the frequency shift due to acceleration is time-dependent. It is worth noting that the equivalence principle holds only locally in space and in time, therefore if one considers a gravitational field in a small region in space, it is approximately indistinguishable with a uniform acceleration. This is also valid in reverse: if one consider an acceleration in a (nearly) instantaneous time, it is indistinguishable from a gravitational field. In this limit, the frequency shift due to acceleration is practically indistinguishable to the gravitational red/blueshift.
III.4 Annihilation and Creation of Modes
The number of modes/quanta of inertial moving frame, in general, will differ with the number in a rest frame. This is due to the fact that different inertial coordinates will measure different size of phase-space volume element, hence, they will calculate different number of worldlines that crosses the phase-space volume element. However, there will be no creation or annihilation of quanta/modes, if their number is conserved in time. In contrast, the number of states in the accelerated frame evolves in time. One could obtain the rate of the quanta/modes production as follows. Let us consider only the accelerated and decelerated away case where . Inserting (26) to (48), and setting for simplicity gives:
From here we can obtain the rate of quanta/modes production in , with respect to :
(65) |
For the case where is accelerated away from the black-body source at rest, the number of modes/quanta observed by is decreasing exponentially in time, as described by the minus sign in (65). This corresponds to the increasing redshift experienced by . For the case where is decelerated away, the number of modes increases exponentially: Note that for , the modes will be annihilated, while for , the modes will be created in time.
III.5 The Emissivity Factor
Let us return to the Planck’s law at an accelerated frame (50). The scale factor could be interpreted physically as the emissivity resulted from the acceleration of the observer. The emissivity is the factor that describes the imperfectness of a black-body, i.e., a physical property that describes how efficiently an object emits thermal radiation. The classical range of the emissivity is , where 1 is the emissivity of a black-body and 0 is the emissivity of a perfect thermal mirror, where no radiation is emitted. In this interval, are the region of the grey-body. Classically, and are not defined, except for special cases for particles smaller than the dominant radiation wavelength (Golyk, ).
There are two factors that affect the emissivity due to acceleration as in (51), namely, the spatial coordinate , and the magnitude of the acceleration . Let us consider the case where , namely the case where the observer are respectively, accelerated and decelerated away from the source. One could derive that for these cases, the scale factor is:
(66) |
where the minus sign is for the accelerated away case, and the plus for the decelerated. For the accelerated away case, the region with negative emissivity is obtained when ; at this condition, no radiation will be detected by the accelerated observer. Meanwhile, positive emissivity factor is obtained when . For this case, if it is possible to have the emissivity , hence the black-body will be perceived as “hyperblack” in the accelerated frame; otherwise, for , it will be perceived as grey. See FIG. 13.

For the decelerated away case, the emissivity factor , so the accelerated frame will observe a range of hyperblack, black, and greybody, depending on the value of and . See FIG. 14.

While in the decelerated away case one can perceive the source with different ’blackness’ depending on the value of and , an interesting phenomenon occurs in the case of an accelerated away scenario. For this case, if is satisfied, the emissivity is ; the black-body is perceived as a greybody in , where the signal becomes dimmer as . At , : in the frame of , the black-body stops emitting radiation as if it is a perfect thermal mirror. From this point, the emissivity can only be negative definite as increases. Remarkably, for a constant , the emissivity is independent of time and the position of the observer from the source, which means that no matter how far or close the object is from the observer, theoretically, they will measure the same emissivity. There could be several possible explanations for this phenomenon, which might include the relativistic beaming and the existence of the (Killing) horizon at , however, further research is needed to understand this phenomenon completely.
III.6 Wien’s Displacement Law and Temperature
At the moment, we are not able to derive the transformation of Wien’s displacement law in a moving frame without knowing how the temperature in relation (76) transform under Lorentz transformation. However, if we assume that Wien’s law is valid also in a moving frame, we can define the directional temperature using the Wien’s law (76), simply as a quantity proportional to the frequency that maximies the energy of the source. This, at least, will be useful for the calculation of a temperature for a black-body in moving frames. As we had mentioned in the Introduction, any measurement of temperature in moving bodies are measured via it’s radiation, since we are not able to do a direct measurement of the temperature in moving bodies. If a body is moving, it needs to be at rest with respect to an observer so that it could be thermalize/ in a thermal equilibrium with the measurement apparatus. Here, satisfied or not, the temperature for moving bodies, is treated as a derived variables. Moreover, if we make an assumption that (76) is valid for inertial and moreover, uniformly accelerated frame, then the directional temperature must transform in the same way as the frequency, namely:
(67) |
(67) is a generalization of the ’directional’-(effective) temperature in (Henry, ) which is valid in any inertial frame; it is an extension of the result in (Henry, ) to a uniformly accelerated frame. One could observe that in (67), there is a coordinate-dependent scale factor proportional to similar to the temperature scale factor in Tolman’s theory of general relativistic thermodynamics (Tolman, ).
IV Discussion and Conclusion
IV.1 Discussion
The Planck’s formula in the accelerated frame (50) are based on several assumptions in the derivation, hence, if one or more of these assumptions are not valid, this will also affect the validity of (50). Let us list the assumptions we use for the derivation:
-
1.
. This is implicitly used because we started from the Planck’s law in rest frame (the relation is use to derive Planck’s law).
-
2.
The validity of the de Broglie postulate in the accelerated frame, namely: , for every Rindler time and Rindler position .
-
3.
The invariance of the relativistic distribution function (number of world-lines that crosses a hypersurface) under coordinate transformation.
Starting from these assumptions, together with some standard definitions in the theory of relativity, all the equations we derived, most importantly, the relativistic Doppler effect (29), the transformation of solid angle (33), the volume contraction (36), and the transformation of the number of modes distribution (48), are consequences of the assumptions and definitions we used. Let us check if all these assumptions are reasonable. The first assumption is used to derive Planck’s law in the original form, and since it is not use explicitly to derive the results, this should not concern our work. The second assumption is used explicitly to derive the dispersion relation (20) from the momentum constraint (20) is crucial for deriving the momentum in namely (23). However, one can also obtained the dispersion relation (20) without the de Broglie postulate, if we consider the electromagnetic wave equation , with is the d’Alembertian operator. The form of the wave equation and its solution are invariant under Rindler coordinate transformation, and so does the dispersion relation. Therefore, without the de Broglie postulate, (20) is still valid in as long as the Maxwell equation is satisfied by the black-body radiation.
Finally, opinions on the last assumption are divided among the researchers in the field. The relativistic distribution function is defined as (47). Since our work follows closely the work of (Peebles, ; Heer, ; Henry, ), we use the same assumption that the relativistic distribution function is coordinate-invariant, instead of the invariance of the number of states , which is also employed in many existing works in the literature. The reason for this difference is because there are 2 distinct measurements of volumes in moving frames: the ones that measure all the points in the volume simultaneously at an instant time, and the one that measure all the points in the volume at a time interval . Our work use the second type of measurement, and hence the number of states counted by the ’moving volume’ at time interval is not the same as (the number of states in rest frame); hence we use assumption 3.
Some readers may wonder if a simple coordinate transformation from inertial to Rindler coordinate is sufficient to obtain the physics in an accelerated frame. As far as we understand, in the theory of relativity (special or general), the use of different coordinates on spacetime, in general, will result in different observational perspectives. This is not only valid for coordinates related by Lorentz transformation, but also for more general coordinate systems, including the Rindler coordinate. Therefore, to obtain observational perspective from an accelerated frame, it is sufficient to do a transformation from inertial/Cartesian coordinate to Rindler coordinate. Furthermore, one might also wonder why the Unruh effect is not included in our work. This is out of the scope of the subject in our work, since the Unruh effect is derived from the quantum field theoretic derivation, and we have tried to write our paper as classical as possible. The reason to write the paper as classical as possible is because the subject is full of controversy, so in our opinion, it will be an advantage if we could, at least, clearly understand the subject in the (semi)-classical level. Although it might be possible to include the Unruh effect in the discussion, in the classical level, it can not be obtained just by coordinate transformation from Minkowski/Cartesian coordinate to Rindler coordinate. There is another assumption needed to obtain the Unruh effect in the classical setting, and this is related to the definition of the classical vacuum. This will be an insteresting subject to pursue.
IV.2 Conclusion
Let us conclude our work in this article as follows. We have derived Planck’s law and calculated the black-body spectrum in a uniformly accelerated frame using Rindler coordinates. The spectrum is time-dependent, Planckian at each instantaneous time, but it is scaled by an emissivity factor that depends on the Rindler spatial coordinate and the acceleration magnitude . The spatial, coordinate-dependent scale factor is proportional to while the scale factor related to the acceleration is proportional to , depending on the observer is either accelerated or decelerated. An observer decelerating away from the source will perceive the black-body as hyperblack, black, or grey, while for the accelerating-away observer, there is a limit in the acceleration magnitude in receiving the radiation, namely, if , the accelerated observer will stop receiving radiation from the black-body. Outside of this limit, the black-body is perceived either as grey or hyperblack, depending on the spatial Rindler coordinate .
The variables in the Planck’s law, specifically the number of modes and frequency, evolve over time. The Planckian spectrum is continuously red-shifted towards lower frequencies as time progresses for the case where the observer is accelerating-away or decelerating towards, and blue-shifted towards lower (or higher) frequencies for the case where the observer is accelerating-towards or decelerating-away. In the accelerated frame, the production of radiation modes (photons) can be positive or negative, depending on the acceleration or deceleration of the observer, and zero for vanishing acceleration. Besides this, there exists a peculiar phenomenon perceived by the accelerating observer, the relativistic beaming, where the rays of light tends to converge, or diverge in the direction of motion, depending on whether the observer is moving towards, or away from the source. In the end, assuming Wien’s displacement law also holds in the accelerated frame, the time-dependent, (directional)-temperature of a body within an accelerated frame is given by
Our last comment is on the temperatures of the systems. For the moment, we do not have a universal definition of temperature. However, the (effective) temperature (67) is derived from their frequency spectra, this is also the case for the Hawking-Unruh temperature. Since for a moving body we do not have other choice than treating (effective) temperature as a derived quantity, we argue that one needs to take seriously the directional approach of temperature. This is supported by the argument in (Aldrovandi, ) that the inertial version of effective temperature (67) is not merely a mathematical parameter, but a real transformation law. We are optimistic that this will be a useful approach to understanding the nature of temperature.
Acknowledgements.
S. A. was supported by an appointment to the Young Scientist Training Program at the Asia Pacific Center for Theoretical Physics (APCTP) through the Science and Technology Promotion Fund and Lottery Fund of the Korean Government. This was also supported by the Korean Local Governments - Gyeongsangbuk-do Province and Pohang City. H. L. P. would like to thank Ganesha Talent Assistanship (GTA) Institut Teknologi Bandung for financial support.Appendix A Black-Body Radiation in Rest Frame
A.1 Planck’s Radiation Law
To make this article self-contained, in this subsection of the Appendix A, we derive the Planck’s law in the form of relation (49) from the original equation (Planck2, ):
(68) |
is the intensity or spectral radiance, defined as:
(69) |
with is the frequency of the electromagnetic radiation and is the temperature of the black-body source emitting the radiation. , and are respectively, the Planck constant, the speed of light, and the Boltzmann constant. The intensity is defined as the infinitesimal energy that passes through an infinitesimal area of the surface of a sphere, per time interval , in the frequency range . is the inclination angle: the angle between the velocity of the photon (that is parallel to the normal to ) with the direction of the observation, but with frame replaced with frame . Inserting (69) to (68), and multiplying the right hand side of the equation with gives:
(70) |
The quantity is the infinitesimal volume swept out by the radiation. Using the angular frequency instead of , the Planck’s law can be written as:
(71) |
Moreover, given the energy of each frequency mode as , with is the number (or distribution) of the radiation modes (quanta of radiation/photon, however, we avoid such terminologies because we want this paper to be as classical as possible) having an angular frequency , we could obtain the Planck’s distribution:
(72) |
with the famous Planckian spectrum. (72) is the form of Planck’s law used in (Henry, ).
A.1.1 Wien’s Displacement Law
To derive Wien’s displacement law from the Planck’s law (71), one needs to find the angular frequency that maximize the radiation energy . This can be obtained by solving the equation as follows:
(73) |
Let us define , the equation from (73) that we need to solve becomes:
(74) |
which can be solved by the Lambert function as follows:
(75) |
With this, one could obtain the Wien’s displacement law in terms of angular frequency:
(76) |
with is the angular frequency that maximize the radiation energy , the ’maximal’ frequency. The statement that the temperature of a black-body source is proportional to the frequency that maximizes the energy of the radiation comes directly from (76).
One can insert the Wien’s law (76) to (72) and obtain:
(77) |
We called the dimensionless coefficient satisfying (74) as the Wien constant. Notice that for a different representation of the Planck’s law, i.e., if the law written as a function of the wavelength instead of , the Wien ’constant’ will differ.
The Planck’s law in the form of equation (72) is equivalent with the two equations: the “Planck’s law” in terms of maximal frequency (77), together with the Wien’s law (76). To avoid the problem of the temperature in moving bodies (Farias, ; Derakshani, ), we will use the “Planck’s law” (77) in our analysis for the black-body spectrum in a uniformly accelerated frame.
Appendix B Black-Body Radiation in Inertial Frames
B.1 Lorentz Transformation
In this subsection, we review the basic properties of Lorentz transformation and define the notations used in our article. Let be a frame at rest, with spatial coordinates and time coordinate In the covariant approach, the time and space are regarded in an equal manner, so let us define the ’time’ such that has the same dimension as length; is the speed of light. Let = be an inertial coordinate that parametrizes the Minkowski space.
Suppose another inertial frame is moving with a constant velocity in the -direction. Let be another inertial coordinate of the Minkowski space related to frame . Frames and are related by the coordinate (Lorentz) transformation as follows:
(78) |
where are not affected by the transformation, namely and Here, we use and Let us define the rapidity as , then we have the following relations:
(79) |
With relations (79), the Lorentz transformation (78) can be written in terms of hyperbolic functions:
(80) |
The infinitesimal (4-vector) transformation related to (80) can be obtained as follows:
(81) |
together with their inverses:
(82) |
with and .
B.2 The Relativistic Aberration of Light
Suppose with respect to the rest frame we have an object moving with velocity as follows:
In frame , the velocity of the object is:
Using the transformation (81)-(82), one could obtain the relation between and as follows
(83) |
(84) |
and their inverses:
(85) |
(86) |
Let a black-body source be at rest with respect to an inertial frame The black-body are emitting electromagnetic radiation (photons) with propagation velocity in the direction of , with is the (polar) inclination angle between the axis + and on plane , but with frame replaced with frame . The electric and magnetic part of the radiation lie, respectively, in plane and , so the propagation velocity will not affect their amplitude. The velocity of the photon with respect to frame is:
(87) |
Meanwhile, at frame , the velocity of the photon is:
(88) |
with and are the propagation direction and the inclination angle of the photon according to , respectively. This could be derived from the null (light-like) vector condition for light, where the norm of its 4-velocity is always zero in any coordinate system. This gives:
Using the velocity addition formulas in (83)-(85), one could obtain the transformation between velocities of the photon seen by and , which results in the polar angle formulas as follows:
(89) | ||||
(90) |
Inserting (89) and (90) to a trigonometric identity, , gives:
(91) |
Equation (89)-(91) are the aberration of light formulas (Johnson, ). Their inverses are:
(92) | ||||
(93) | ||||
(94) |
B.3 The Relativistic Doppler Effect
The next step is to derive the relativistic Doppler effect. First, we need to define the 4-momentum of a moving body with respect to as follows:
(95) |
with is the energy of the moving body and is the relativistic 3-momentum. For the black-body radiation case, the moving bodies are photons, which, by de Broglie postulate, has energy and momentum . Here, is the Planck constant, is the (angular) frequency, and is the wave vector satisfying Inserting these information to (95) together with the components of , one could obtain the 4-momentum of a photon with respect to an inertial observer :
(96) |
According to the moving frame , the 4-momentum of the photon is:
(97) |
with is the frequency of the photon, with respect to . One could also obtain (96) and (97) by inserting the de Broglie postulate to the zero mass condition for the photon:
with this, we only consider the positive solution from the dispersion relation
The 4-momentum is an element of the Minkowski space and therefore transform under Lorentz transformation, so the relation of and is:
(98) |
with and . Naturally, the (generalized) momentum is a co-vector instead of a (contravariant) vector, however, here we use the vector transformation of momentum. This will not affect the result as long as the calculations are consistent. Using the Lorentz transformation (98) on the momenta (96) and (97), one could obtain the relations between the frequency of photon as seen by and as seen by , written in 3 equivalent forms as follows (Johnson, ):
(99) | ||||
(100) | ||||
From these equations, one could retrieve the aberration formulas (89)-(91) and their inverses (92)-(94). Inserting (89) to (100), one could rewrite in terms of only variable , the inclination angle seen by observer :
(101) |
Equation (100) and (101) are the formula describing the relativistic Doppler effect for the frequency of the photon in a moving observer with respect to observer at rest (Johnson, ). For our case, the photon frequency observed by the moving observer is redshifted, since observer is moving away from the black-body source. To obtain the blueshifted Doppler effect, one could flip the sign on the velocity of with respect to the source to become .
Since the Planck’s law (77) contains the solid angle term , we need to know how it transforms under Lorentz transformation. First, from (100) and (101) we obtain:
Second, differentiating equation (90) or (93) with respect to a parameter will give:
(102) |
With (102), the solid angle , will transform as follows (Johnson, ):
(103) |
using (99), (102), and the fact that the azimuth angle are not affected by the transformation, .
B.4 The Phase-Space Volume Transformation
The next variable in the Planck’s law that transformed under the Lorentz transformation is , the number or distribution of modes with frequency . To know how this variable changes under different inertial frames, one needs to consider the phase-space volume transformation. The phase-space volume element is not a Lorentz scalar, see a detailed explanation in (Debbasch, ). First, let us obtain the transformation for the 3-volume element of the spatial part of the phase-space. In and , the infinitesimal 3-volume element are defined as, respectively:
(104) | ||||
To derive the transformation relation between these 2 elements, let us consider a measurement of spatial length in To measure spatial length, 2 events must be simultaneous in time with respect to the observer. Let us consider 2 simultaneous events and along the -axis at time , where and . According to the length between these 2 event is:
Now, let these 2 events be perceived in frame , moving with respect to Using Lorentz transformation (78):
it is clear that in these 2 events are not simultaneous. The spatial length between and in is:
Now for our black-body case, we need to calculate how much modes are inside a volume element. Let the finite volume element in be . This volume will be perceived simultaneously by at an instant time , while in it will be perceived as the volume swept by the plane from to along a time interval this is the physical interpretation of Taking the infinitesimal value , we have:
(105) |
Notice that the standard volume contraction formula is which has different interpretation with (105). The standard volume contraction is the comparison between volumes that are both measured instantaneously with respect to each times on each frames, while in (105), the measurement of is done along a time interval . For an alternative derivation for this formula, one could consult (Peebles, ; Heer, ; Henry, ).
Second, let us obtain the transformation for the 3-volume element in the momentum space, which are defined as follows:
(106) | ||||
respectively for and . The 4-momentum are constrained such that it’s norm is constant:
(107) |
Differentiating this constraint with respect to any parameter gives:
(108) |
Let us consider the infinitesimal version of the transformation of 4-momentum; such transformations have similar forms with (98). Inserting the infinitesimal version of (98) to (106), and then using the constraint (108), gives:
(109) |
However, coefficient in (109) is simply by (98), so one has:
(110) |
by the photon momentum (97) and the Doppler effect (102). Now, using (105) and (110), we can obtain the transformation of the phase-space volume element between 2 inertial observers as follows:
B.5 The Relativistic Distribution Function
The trajectory of the photon in spacetime is described by a curve in the Minkowski space. At an instantaneous (constant) time , one could define a hypersurface and extend the hypersurface to the phase-space by attaching the momentum space (the cotangent bundle of ) to The worldine that cross the phase-space will mark a point on the phase-space , describing the state of the photon at time . Moreover, one can construct the phase-space volume element in the phase-space , namely . The relativistic distribution function is defined as the number of wordlines that cross the phase-space, i.e, the states, per volume element (Liboff, ):
Now, if we have another coordinate patch for the Minkowski space, namely , it will perceive different time-constant hypersurface, different phase-space, and different phase-space volume element, namely . Since the phase-space volume element changes, the number of wordlines that cross the volume element will also change accordingly. However, the number of worldlines (and the worldlines themselves) in do not change by coordinate transformation. Hence, it is reasonable to assume that the worldline density is invariant under coordinate (Lorentz) transformation, namely (Liboff, ):
With this reasonable assumption, one could have the number of states/modes transformation between two inertial frames and , which is:
(111) |
Finally, with (101), (103), (105), and (111), we have all the ingredients to show that the Planck’s law are invariant under Lorentz transformation.
B.6 Black-body Radiation in Moving Frame
The objective in this section is to know if the Planckian distribution is invariant under Lorentz transformation. To avoid the problem of temperature in moving bodies, we use the Planck’s Law in the form of equation (77), that is, the one with the term containing , instead of the original form (72) with the term containing . The derivation in this section is based on the derivation in (Henry, ), with some slight modifications. Inserting the relativistic Doppler effect (101), the transformation of solid angle (103), the volume contraction (105), and the transformation of the number of modes distribution (111) to (77), we obtain:
that can be simplified as:
(112) |
which is exactly the form of Planck’s law (77). All the variables inside the equation transform according to the Lorentz transformation, but the relation between these variables is invariant, hence, the observer in a moving frame will still observe the Planckian spectrum. However, there is a subtlety in the relation (112). , the ’maximal’ frequency, also transformed as the angular frequency via the relativistic Doppler effect (101). However , is the ’maximal’ frequency in the distribution (77) of a black-body in the rest frame, observed by a moving observer. It is not necessarily the ’maximal’ frequency obtained from the distribution (112) of a black-body observed by the moving observer. We will see that they are equivalent as follows. Let us multiply the LHS and RHS of (112) with to obtain:
(113) |
is the energy of modes with frequency observed in the moving frame . Let us find the value of that maximize from the equation :
(114) |
Defining as in the previous subsection, the equation from (114) that we need to solve is then:
(115) |
which is exactly similar to (74). Hence it posses a same solution:
namely, the Wien coefficient is invariant under Lorentz transformation: . Then, if is the solution to (114), we have:
or namely : the Lorentz transformation sends ’maximal’ frequency in one inertial frame, to a ’maximal’ frequency in another inertial frame. (113) becomes:
(116) |
With this, we can state that the Planck’s radiation law in the form of (77) is invariant under Lorentz transformation. This result is similar to (Henry, ).
References
- (1) C. Farias, V. A. Pinto, P. S. Moya. What is the temperature of a moving body?. Sci. Rep. 7: 17657. 2017.
- (2) K. Derakhshani. Black Body Radiation in Moving Frames. 2019. https://hal.science/hal-02299780.
- (3) M. Planck. Zur Dynamik bewegter Systeme. Sitzungsber. Preuss. Akad. Wiss. pp. 867–904. 1907.
- (4) A. Einstein. Zur Elektrodynamik bewegter Krper. Ann. Phys. 17: 891. 1905.
- (5) A. Einstein. Uber das Relativitatsprinzip und die aus demselben gezogene Folgerungen. Jahrbuchder Radioaktivitt 4: 411. 1907.
- (6) H. Ott. Lorentz-Transformation der Wrme und der Temperatur. Zeit. Phys. 175: 70. 1963.
- (7) H. Arzelis. Transformation relativiste de la temperature et de quelque autre grandeurs thermodynamique. Nuov. Cim. 35 (3): 792-804. 1965.
- (8) P. T. Landsberg. Does a Moving Body Appear Cool?. Nature 212: 571. 1966.
- (9) P.T. Landsberg. Special relativistic thermodynamics. Proc. Phys. Soc. 89: 1007. 1966.
- (10) T. K. Nakamura. Thermodynamics of extended bodies in special relativity. Phys. Lett. A 352: 175. 2006.
- (11) T. K. Nakamura. Three Views of a Secret in Relativistic Thermodynamics. Prog. Theor. Phys. 128: 463-475. 2012.
- (12) N. G. van Kampen. Relativistic Thermodynamics of Moving Systems. Phys. Rev. 173 (295). 1968.
- (13) W. Israel. Transient relativistic thermodynamics and kinetic theory. Annal. Phys. 100 (310). 1976.
- (14) R. C. Tolman. On the Extension of Thermodynamics to General Relativity. Proc. Nat. Acad. Sci. 14: 268-272. 1928.
- (15) R. C. Tolman. Effect of Inhomogeneity on Cosmological Models. Proc. Nat. Acad. Sci. 20: 169-176. 1928.
- (16) A. Einstein. Relativity: The Special and General Theory. Henry Holt and Co. New York. 1920.
- (17) S. M. Carroll. Spacetime and Geometry: An Introduction to General Relativity. Cambridge University Press. 2019.
- (18) F. W. Sears, G. L. Salinger. Thermodynamics, Kinetic Theory, and Statistical Thermodynamics. 3rd Edition. Addison Wesley. 1975.
- (19) L. Boltzmann. On the Relationship between the Second Fundamental Theorem of the Mechanical Theory of Heat and Probability Calculations Regarding the Conditions for Thermal Equilibrium. Wiss. Abhandlungen II 42: 64-223. 1909.
- (20) R. L. Liboff. Kinetic Theory Classical, Quantum, and Relativistic Descriptions. Springer. 2003.
- (21) W. G. Unruh. Notes on black-hole evaporation. Phys. Rev. D 14 (4): 870–892. 1976.
- (22) S. A. Fulling. Nonuniqueness of Canonical Field Quantization in Riemannian Space-Time. Phys. Rev. D 7 (10): 2850–2862. 1973.
- (23) P. C. W. Davies. Scalar production in Schwarzschild and Rindler metrics. J. Phys. A 8 (4): 609–616. 1975.
- (24) P. M. Alsing, P. W. Milonni. Simplified derivation of the Hawking-Unruh temperature for an accelerated observer in vacuum. Am. J. Phys. 72: 1524-1529. 2004.
- (25) S. W. Hawking. Black hole explosions?. Nature 248 (5443): 30–31. 1974.
- (26) S. W. Hawking. Particle creation by black holes. Commun. Math. Phys. 43 (3): 199–220. 1975.
- (27) S. Deffner, S. Campbell. Quantum Thermodynamics: An introduction to the thermodynamics of quantum information. Morgan & Claypool Publishers. 2019.
- (28) F. Binder, L. A. Correa, C. Gogolin, J. Anders, G. Adesso. Thermodynamics in the Quantum Regime. Fundamental Theories of Physics. Springer. 2018.
- (29) C. Rovelli. Statistical mechanics of gravity and the thermodynamical origin of time. Class. Quant. Grav. 10: 1549–1566. 1993.
- (30) A. Connes, C. Rovelli. Von Neumann algebra automorphisms and time-thermodynamics relation in generally covariant quantum theories. Class. Quant. Grav. 11 (12): 2899–2917. 1994.
- (31) M. Planck. On the Theory of the Energy Distribution Law of the Normal Spectrum. Verhandl. Dtsch. Phys. Ges. 2 (237). 1900.
- (32) N. G. van Kampen. Relativistic thermodynamics. J. Phys. Soc. Japan 26: 316–321. 1969.
- (33) J. J. Mare, P. Hubik, J. estk, V. pika, J. Kri,tofik, J. Stvek. Relativistic transformation of temperature and Mosengeil–Ott’s antinomy. Physica E 42 (3): 484-487. 2010.
- (34) S. S. Costa, G. E. A. Matsas. Temperature and relativity. Phys. Lett. A 209 (155). 1995.
- (35) P. T. Landsberg, G. E. A. Matsas. Laying the ghost of the relativistic temperature transformation. Phys. Lett. A 223 (6): 401-403. 1996.
- (36) P. T. Landsberg, G. E. A. Matsas. The impossibility of a universal relativistic temperature transformation. Physica A 340 (1–3): 92-94. 2004.
- (37) R. Bracewell, E. Conklin. An Observer moving in the K Radiation Field. Nature 219: 1343–1344. 1968.
- (38) W. Wien. ber die Differentialgleichungen der Elektrodynamik fr bewegte Krper. I. Ann. der Phys. 318 (4): 641–662. 1904.
- (39) W. Wien. ber die Differentialgleichungen der Elektrodynamik fr bewegte Krper. II. Ann. der Phys. 318 (4): 663–668. 1904.
- (40) J. Stefan. ber die Beziehung zwischen der Wrmestrahlung und der Temperatur. Sitzungsber. Preuss. Akad. Wiss. 79: 391–428. 1879.
- (41) L. Boltzmann. Ableitung des Stefan’schen Gesetzes, betreffend die Abhngigkeit der Wrmestrahlung von der Temperatur aus der electromagnetischen Lichttheorie. Ann. der Phys. Chem. 258 (6): 291–294. 1884.
- (42) P. J. E. Peebles, D. T. Wilkinson. Comment on the Anisotropy of the Primeval Fireball. Phys. Rev. 174 (2168). 1968.
- (43) C. V. Heer, R. H. Kohl. Theory for the Measurement of the Earth’s Velocity through the K Cosmic Radiation. Phys. Rev. 174 (1611). 1968.
- (44) G. R. Henry, R. B. Feduniak, J. E. Silver, M. A. Peterson. Distribution of Blackbody Cavity Radiation in a Moving Frame of Reference. Phys. Rev. 176 (1451). 1968.
- (45) T. H. Boyer. Derivation of the Blackbody Radiation Spectrum without Quantum Assumptions. Phys. Rev. 182 (1374). 1969
- (46) J. S. Lee, G. B. Cleaver. The relativistic blackbody spectrum in inertial and non-inertial reference frames. New Astronomy 52: 20-28. 2017.
- (47) M. H. Johnson, E. Teller. Intensity changes in the Doppler effect. Proc. Natl. Acad. Sci. 79 (4) :1340. 1982.
- (48) R. P. Feynman, R. B. Leighton, M. Sands. Relativistic Effects in Radiation. The Feynman Lectures on Physics: Volume 1. Addison-Wesley. pp. 34–7. 1977.
- (49) L. Euler. De serie Lambertina Plurimisque eius insignibus proprietatibus. Acta Acad. Scient. Petropol. 2: 29-51. (1783).
- (50) W. Rindler. Relativity: Special, General, and Cosmological. 2nd Ed. Oxford University Press. 2006.
- (51) K. S. Thorne, R. D. Blandford. Modern Classical Physics: Optics, Fluids, Plasmas, Elasticity, Relativity, and Statistical Physics. Princeton University Press, 2017.
- (52) F. Debbasch, J. P. Rivet, W. A. van Leeuwen. Invariance of the relativistic one-particle distribution function. Physica A 301 (1–4): 181-195. 2001.
- (53) M. H. Cohen, M. L. Lister, D. C. Homan, M. Kadler, K. I. Kellermann, Y. Y. Kovalev, R. C. Vermeulen. Relativistic Beaming and the Intrinsic Properties of Extragalactic Radio Jets. The Astro. Jour. 658 (1). 2007.
- (54) A. S. Eddington. Einstein Shift and Doppler Shift. Nature. 117 (2933): 86. (1926).
- (55) L.B. Okun (ITEP, Moscow), K.G. Selivanov (ITEP, Moscow), V.L. Telegdi (EP Division, CERN). On the Interpretation of the Redshift in a Static Gravitational Field. Am.J.Phys.68:115,2000.
- (56) V. A. Golyk, M. Krger, M. Kardar. Heat radiation from long cylindrical objects. Phys. Rev. E 85 (4): 046603. 2012.
- (57) M. Planck. ber die Begrndung des Gesetzes der schwarzen Strahlung. Ann. Phys. 342: 642. 1912.
- (58) S. Weinberg. The Quantum Theory of Fields, Volume 1: Foundations. Cambridge University Press. 2005 .
- (59) R. Jaffe. Casimir effect and the quantum vacuum. Phys. Rev. D. 72 (2): 021301. 2005.
- (60) R. Aldrovandi, J. Gariel. On the riddle of the moving thermometers. Phys. Lett. A 170 (1): 5-10. 1992.