Sunteți pe pagina 1din 7

BLACKBODY RADIATION AND THE RAYLEIGH-JEANS EQUATION

DANIEL MARSHALL PHYS 420H PROFESSOR WILLIS 03 MAY 2012

Marshall 2 INTRODUCTION
When electromagnetic radiation falls upon an opaque body, part of it is reflected and the other part is absorbed. The absorbed part of the radiation imparts a kinetic energy to the constituent particles of the body and therefore causes a small increase in temperature. This increase in kinetic energy accelerates the atoms electrons which in turn emit Figure 1 See reference [2] electromagnetic radiation (which serves to decrease the temperature). Electromagnetic radiation generated in this manner is called thermal radiation. Due to its simplicity, an important and theoretically useful case of thermal radiation is that which is emitted by a blackbody. An ideal blackbody is a body that completely absorbs all incident electromagnetic radiation * (nothing is reflected) . When considering blackbody radiation, one would naturally want to know the spectral distribution and what variables that distribution is dependent on. Figure 1 shows R(), the blackbody thermal radiation intensity as a function of wavelength . The maximum value of each constant temperature curve is of course determined by Wiens displacement law

which was empirically determined by Wilhelm Wien in 1893 [3]. Although the information portrayed in figure 1 was determined experimentally with ease, the theoretical development eluded scientists of the late nineteenth and early twentieth century. One of the most notable failed theoretical derivations of R() is known as the Rayleigh-Jeans equation. In the report that follows, I will give a thorough review of the Rayleigh-Jeans equation and its failed attempt to apply the classical theories of thermodynamics and electromagnetism to the spectral distribution of blackbody radiation.

CAVITY RADIATION
Compared to blackbody radiation, it is much simpler to analyze the spectral distribution function R() of cavity radiation. Conveniently, it turns out that the radiation emitted by a cavity with a small hole is completely equivalent to the thermal radiation emitted by a blackbody. This is certainly plausible when you consider the fact that electromagnetic radiation that enters the metal cavity through a tiny hole has a negligible chance of being reflected directly back out through the hole. That is to say, the radiation that enters the hole is absorbed by the walls of the cavity before exiting through the small hole. Therefore the stream of radiation exiting the cavity is no different from the thermal radiation of a blackbody (i.e., it does not contain any reflected radiation) [4].
*

The reflectance of regular black paint is around 5% to 10%. The darkest material ever created (and therefore the best approximation to an ideal blackbody) is a low density carbon nanotube array created by researchers from Rensselear and Rice University and has a reflectance of 0.045% [1]. Intensity is defined here by the usual definition of power per unit area.

Marshall 3
The Rayleigh-Jeans equation is derived by applying the classical theories of thermodynamics and electromagnetism to cavity radiation. As an introduction to the derivation, it is nice to keep in mind the following characterization of an isothermal cavity (taken from reference [5] page 118): In an isothermal enclosure the stream of radiation in any given direction must be the same as in any other direction; it must be the same at every point inside the enclosure; and it must be the same in all enclosures at a given temperature, irrespective of the materials composing them. Furthermore, all these statements hold for each spectral component of the radiation taken separately. The proof of these statements follows easily by consideration of a proof by contradiction as follows. Assume that any of these statements were not true you could then find two streams of radiation traveling in different directions in which one of the streams of radiation had a higher intensity. We could then situate two absorbers to intersect these streams of differing intensity. The two absorbers would then be maintained at different temperatures and therefore could be used as a source and sink to operate a Carnot engine. This Carnot engine could operate continuously without causing other changes in the system and would therefore violate the second law of thermodynamics. This proof can be generalized to each spectral component of the radiation by considering wavelength sensitive absorbers [5]. Another important property of an isothermal cavity is that the cavity is filled with electromagnetic standing waves. Applying the statements in the box above any radiation that is absorbed is instantly reradiated. This result is the same as if all the radiation was reflected at the walls. Rayleigh assumed the cavity walls were composed of oscillators which were responsible for the absorption and emission of the thermal radiation within the cavity. If these oscillators are to operate continuously and replicate this reflection process, the resulting electromagnetic waves within the cavity must be standing waves [5].

DERIVATION OF THE RAYLEIGH-JEANS EQUATION


Using the idea of cavity radiation to calculate spectral distribution function R(), the development of the Rayleigh-Jeans equation can be broken down into the following two steps: 1. Relate the thermal radiation intensity R() to the energy density distribution function u()

2.

Calculate u() by consideration of the number of modes of oscillation per unit volume n()

STEP ONE

As a result of the electromagnetic standing waves within the cavity, there exists a corresponding mean energy per unit volume within the cavity associated with each wave, let w represent this quantity. As a result of the boxed statements above, beams of radiation are streaming in all directions, toward and away from the cavity surface, all with the same intensity. Let there be N beams of radiation in all. It follows that the total energy density U just in front of the surface is given by Next, in order to determine the relationship between the thermal radiation intensity R() (of the radiation leaving

This step follows the argument of reference [5] - Appendix 5A on pp. 135-137.

Marshall 4
the cavity hole) and the total energy density U (at the surface of the cavity), we need to consider the power I each wave delivers to the surface. By considering figure 2 (on the next page), it is not difficult to see that Therefore, in order to calculate the total intensity of the radiation on the surface of the cavity walls (i.e., the total radiation intensity emitted by the hole in the cavity R), we would need to integrate this result with respect to N. But since half of the waves would be moving away from the surface (by the definition of N), we will also need to multiply this result by . Hence, The next challenge is to find a relationship between dN and d . There is a nice geometric argument that can be developed by considering the hemisphere in figure 3. Let the bit of area on the surface of our cavity in which we have been considering be situated at the origin of figure 3. Imagine the N waves of radiation drawn as lines emanating outward from the origin and piercing the surface. The points of intersection with the hemisphere would be evenly distributed across the hemispherical surface. Now consider the band of differential area as shown in figure 3. Let dN denote the number of points of intersection (of the lines representing the N beams of radiation) that lie within the band of area . Hence, the ratio of dN to N should be the same as the ratio of the differential band area to the area of the entire hemisphere. Since the area of the hemisphere is , we have

Figure 2

By substituting this result back into the expression for R, we are able to obtain the relationship between the blackbody thermal radiation intensity R and the total energy density U within the cavity as follows

Figure 3

By considering the boxed statements on page 3 statements, it should be clear that this result holds for each ** individual wavelength of radiation. Hence ,

STEP TWO

This step involves relating the energy density distribution function u() to n(), the number of modes of oscillation per unit volume with respect to the electromagnetic standing waves corresponding to a specific wavelength of thermal radiation. As a result of classical thermodynamics, an average energy of kbT is assigned to

The areas are given in units of steradians. This relationship holds not only for the radiation exiting the cavity, but also when considering the energy passing through one side of any imaginary unit of area drawn within the cavity. This step follows the argument of reference [6].
**

Marshall 5
each mode of oscillation within the cavity. This is a result of applying the principle of equipartition of energy to the oscillators in the cavity walls (which are responsible for creating the standing waves) [5]. Therefore we are left with the relation In order to explicitly calculate n(), we need to consider mode counting in three dimensions. By first considering the one and two dimensional cases, this process turns out to be relatively painless. The one dimensional case is quite simple. Given a string of length L, we know the only acceptable standing waves are those with nodes at each end of the string, therefore L must equal an integral number of half wavelengths:

where nx is the integer referring to the number of modes across the length L. We want to know how many modes have wavelength between and + d. To do this, let i = and f = + d and subtract their corresponding equations.

This equation resulted from the assumption that there was only one degree of freedom for the standing waves, but of course the standing waves on a string have two degrees of freedom and so the dnx would be twice the value of that given above. Therefore, the linear density of the modes dnL would be given by dnx multiplied by two and divided by L:

Mode counting in the two dimensional case is not much more difficult than in one dimension you just need to consider the x and y components of the wavelength. Figure 4 shows a possible standing wave at arbitrary angles and with the walls of the container. The components of the wave traveling from to (or from to ) would be:

Using these results, we can obtain nx and ny by substitution into equation .

Figure 4 See reference [6]


From here, we obtain the allowed wavelengths by squaring these results and adding them together.

The two dimensional case is easily generalized to three dimensions. The restrictions on nx and ny remain unchanged. The only new addition is an nz component (with a corresponding angle of ) that would be of the same form. Thus we are left with

Marshall 6
The various combinations of nx, ny and nz can be represented as ordered triplets lying within the positive octant of a three dimensional graph (as shown in figure 5). The number of these points contained in a spherical volume of radius possible wavelengths corresponds to the number of modes n for . Hence, the result is

Differentiating this result with respect to

gives

Figure 5 See reference [6]


where we have dropped the minus sign. This is possible since (considering equation ) an increase of dnx corresponds to a decrease in d, so we can absorb the minus sign into the d and assume both differentials are 3 positive. Finally, dividing this result by L gives us the number of modes of oscillation per unit volume n():

Now that we have obtained n(), we are finally able to write the Rayleigh-Jeans equation:

And of course

CONCLUSION
This unfortunate result predicts that (and also ) approaches infinity as approaches zero. This physically absurd and obviously untrue result baffled scientists. The failure was a good indicator of the shortcomings of classical thermodynamics and electromagnetism. It would take until the year 1900 for Max Planck to announce that by making strange assumptions he was able to obtain an expression for that agreed with the experimental result. These strange assumptions were of course related to the discrete energy levels available to the oscillators.

Marshall 7
REFERENCES [1] "Researchers Develop Darkest Man-Made Material." Inside Rensselaer Vol 2. No 2. January 31, 2008 <http://www.rpi.edu/about/inside/issue/v2n2/darkest.html>. Van Der Pol, Edwin. "Method." Infrared spectroscopy. N.p., n.d. Web. 22 Apr 2012. <http://physics.schooltool.nl/irspectroscopy/method.php>. Tipler, P. A., and R. A. Llewellyn. Modern physics. 5. New York: W. H. Freeman, 2008. 119-127. Print. Krane, Kenneth S. Modern Physics. John Wiley & Sons, Inc., 1983. 63. Print. Richtmyer, R. K., E. H. Kennard, and John N. Cooper. Introduction to Modern Physics. Sixth Edition. 1969: McGraw-Hill, Inc., 117-146. Print. Tipler, Paul A., and Ralph A. LLewellyn. "Classical Concept Review, #16." MODERN PHYSICS. W.H.Freeman & Co. Web. 3 May 2012. <http://bcs.whfreeman.com/tiplermodernphysics5e/pages/bcsmain.asp?s=00050&n=03000&i=03050.05&v=category&o=&ns=0&uid=0&rau=0>.

[2]

[3] [4] [5]

[6]

Figures 2 and 3 were made by me in Microsoft Paint.

S-ar putea să vă placă și