Duane's hypothesis

Last updated

In 1923, American physicist William Duane presented [1] a discrete momentum-exchange model of the reflection of X-ray photons by a crystal lattice. Duane showed that such a model gives the same scattering angles as the ones calculated via a wave diffraction model, see Bragg's Law.

Contents

The key feature of Duane's hypothesis is that a simple quantum rule based on the lattice structure alone determines the quanta of momentum that can be exchanged between the crystal lattice and an incident particle.

In effect, the observed scattering patterns are reproduced by a model where the possible reactions of the crystal are quantized, and the incident photons behave as free particles, as opposed to models where the incident particle acts as a wave, and the wave then 'collapses' to one of many possible outcomes.

Duane argued that the way that crystal scattering can be explained by quantization of momentum is not explicable by models based on diffraction by classical waves, as in Bragg's Law.

Duane applied his hypothesis to derive the scattering angles of X-rays by a crystal. Subsequently, the principles that Duane advanced were also seen to provide the correct relationships for optical scattering at gratings, and the diffraction of electrons. [2]

In the early days of diffraction fine details were not observable because the detectors were inefficient, and the sources were also of low intensities. Hence Bragg's law was the only type of diffraction observable, and Duane's approach could model it. Modern electron microscopes and x-ray crystallography instruments are many orders of magnitude brighter, so many find details of electron and x-ray diffraction are now known which cannot be explained by his approach. [3] [4] [5] [6] Hence his approach is no longer used.

Early Developments in Quantum Theory

In 1905, Albert Einstein presented the hypothesis that the photoelectric effect could be explained if a beam of light was composed of a stream of discrete particles (photons), each with an energy (E = hf) the energy (E) of each photon being equal to the frequency (f) multiplied by Planck's constant (h). [7] Later, in 1916 Albert Einstein also showed that the recoil of molecules during the emission and absorption of photons was consistent with, and necessary for, a quantum description of thermal radiation processes. Each photon acts as if it imparts a momentum impulse p equal to its energy divided by the speed of light, (p = E/c). [8]

In 1925, shortly before the development of the full mathematical description of quantum mechanics, Born drew Einstein's attention to the then-new idea of "de Broglie's waves". He wrote "It seems to me that a connection of a completely formal kind exists between these and that other mystical explanation of reflection, diffraction and interference using 'spatial' quantisation which Compton and Duane proposed and which has been more closely studied by Epstein and Ehrenfest." [9] [10] [11] Examining the hypothesis of Duane on quantized translational momentum transfer, as it accounted for X-ray diffraction by crystals, [1] and its follow-up by Compton, [12] Epstein and Ehrenfest had written "The phenomena of Fraunhofer diffraction can be treated as well on the basis of the wave theory of light as by a combination of concept of light quanta with Bohr's principle of correspondence." Later, Born and Biem wrote: "Every physicist must accept Duane's rule." [13]

Using Duane's 1923 hypothesis, the old quantum theory and the de Broglie relation, linking wavelengths and frequencies to energy and momenta, gives an account of diffraction of material particles. [14] [15] [16] [17]

Young's two-slit diffraction experiment, with Fourier analysis

Gregory Breit in 1923 pointed out that such quantum translational momentum transfer, examined by Fourier analysis in the old quantum theory, accounts for diffraction even by only two slits. [18] More recently, two slit particle diffraction has been experimentally demonstrated with single-particle buildup of electron diffraction patterns, as may be seen in the photo in this reference [19] [20] and with helium atoms and molecules. [21]

Bragg diffraction

A wave of wavelength λ is incident at angle θ upon an array of crystal atomic planes, lying in a characteristic orientation, separated by a characteristic distance d. Two rays of the beam are reflected from planes separated by distance nd, where n denotes the number of planes of the separation, and is called the order of diffraction. If θ is such that

then there is constructive interference between the reflected rays, which may be observed in the interference pattern. This is Bragg's law.

The same phenomenon, considered from a different viewpoint, is described by a beam of particles of momentum p incident at angle θ upon the same array of crystal atomic planes. It is supposed that a collective of n such atomic planes reflects the particle, transferring to it a momentum nP, where P is a momentum characteristic of the reflecting planes, in the direction perpendicular to them. The reflection is elastic, with negligible transfer of kinetic energy, because the crystal is massive. The initial momentum of the particle in the direction perpendicular to the reflecting planes was p sin θ. For reflection, the change of momentum of the particle in that direction must be 2p sin θ. Consequently,

This agrees with the observed Bragg condition for the diffraction pattern if θ is such that

or

It is evident that p provides information for a particle viewpoint, while λ provides information for a wave viewpoint. Before the discovery of quantum mechanics, de Broglie in 1923 discovered how to inter-translate the particle viewpoint information and the wave viewpoint information for material particles: [22] [23] use Planck's constant and recall Einstein's formula for photons:

It follows that the characteristic quantum of translational momentum P for the crystal planes of interest is given by

[24] [25]

Quantum mechanics

According to Ballentine, Duane's proposal of quantum translational momentum transfer is no longer needed as a special hypothesis; rather, it is predicted as a theorem of quantum mechanics. [26] It is presented in terms of quantum mechanics by other present day writers also. [27] [28] [29] [30] [31] [32]

Diffraction

One may consider a particle with translational momentum , a vectorial quantity.

In the simplest example of scattering of two colliding particles with initial momenta , resulting in final momenta . The momentum transfer is given by

where the last identity expresses momentum conservation. [33]

In diffraction, the difference of the momenta of the scattered particle and the incident particle is called momentum transfer.

Such phenomena can also be considered from a wave viewpoint, by use of the reduced Planck constant . The wave number is the absolute value of the wave vector , which is related to the wavelength . Often, momentum transfer is given in wavenumber units in reciprocal length

Momentum transfer is an important quantity because is a better measure for the typical distance resolution of the reaction than the momenta themselves.

Bragg diffraction occurs on the atomic crystal lattice. It conserves the particle energy and thus is called elastic scattering. The wave numbers of the final and incident particles, and , respectively, are equal. Just the direction changes by a reciprocal lattice vector with the relation to the lattice spacing . As momentum is conserved, the transfer of momentum occurs to crystal momentum.

For the investigation of condensed matter, neutron, X-ray and electron diffraction are nowadays commonly studied as momentum transfer processes. [34] [35]

Physical accounts of wave and of particle diffraction

The phenomena may be analysed in several appropriate ways. The incoming and outgoing diffracted objects may be treated severally as particles or as waves. The diffracting object may be treated as a macroscopic classical object free of quantum features, or it may be treated as a physical object with essentially quantum character. Several cases of these forms of analysis, of which there are eight, have been considered. For example, Schrödinger proposed a purely wave account of the Compton effect. [36] [37]

Classical diffractor

A classical diffractor is devoid of quantum character. For diffraction, classical physics usually considers the case of an incoming and an outgoing wave, not of particle beams. When diffraction of particle beams was discovered by experiment, it seemed fitting to many writers to continue to think in terms of classical diffractors, formally belonging to the macroscopic laboratory apparatus, and of wave character belonging to the quantum object that suffers diffraction.

It seems that Heisenberg in 1927 was thinking in terms of a classical diffractor. According to Bacciagaluppi & Crull (2009), Heisenberg in 1927 recognized that "the electron is deflected only in the discrete directions that depend on the global properties of the grating." Nevertheless, it seems that this did not lead him to think that the collective global properties of the grating should make it a diffractor with corresponding quantal properties, such as would supply the diffracted electron with a definite trajectory. It seems, rather, that he thought of the diffraction as necessarily a manifestation of wave character belonging to the electron. It seems that he felt this was necessary to explain interference when the electron was detected far from the diffractor. [38] Thus it seems possible that in 1927, Heisenberg was not thinking in terms of Duane's hypothesis of quantal transfer of translative momentum. By 1930, however, Heisenberg thought enough of Duane's hypothesis to expound it in his textbook. [24]

Quantum diffractor

A quantum diffractor has an essentially quantum character. It was first conceived of in 1923 by William Duane, in the days of the old quantum theory, to account for diffraction of X-rays as particles according to Einstein's new conception of them, as carriers of quanta of momentum. The diffractor was imagined as exhibiting quantum transfer of translational momentum, in close analogy with transfer of angular momentum in integer multiples of Planck's constant. The quantum of translational momentum was proposed to be explained by global quantum physical properties of the diffractor arising from its spatial periodicity. This is consonant with present-day quantum mechanical thinking, in which macroscopic physical bodies are conceived as supporting collective modes, [39] manifest for example in quantized quasi-particles, such as phonons. Formally, the diffractor belongs to the quantum system, not to the classical laboratory apparatus.

Related Research Articles

<span class="mw-page-title-main">Double-slit experiment</span> Physics experiment, showing light and matter can be modelled by both waves and particles

In modern physics, the double-slit experiment demonstrates that light and matter can satisfy the seemingly-incongruous classical definitions for both waves and particles, which is considered evidence for the fundamentally probabilistic nature of quantum mechanics. This type of experiment was first performed by Thomas Young in 1801, as a demonstration of the wave behavior of visible light. At that time it was thought that light consisted of either waves or particles. With the beginning of modern physics, about a hundred years later, it was realized that light could in fact show both wave and particle characteristics. In 1927, Davisson and Germer and, independently George Paget Thomson and Alexander Reid demonstrated that electrons show the same behavior, which was later extended to atoms and molecules. Thomas Young's experiment with light was part of classical physics long before the development of quantum mechanics and the concept of wave–particle duality. He believed it demonstrated that Christiaan Huygens' wave theory of light was correct, and his experiment is sometimes referred to as Young's experiment or Young's slits.

<span class="mw-page-title-main">Photon</span> Elementary particle or quantum of light

A photon is an elementary particle that is a quantum of the electromagnetic field, including electromagnetic radiation such as light and radio waves, and the force carrier for the electromagnetic force. Photons are massless, so they always move at the speed of light in vacuum, 299792458 m/s. The photon belongs to the class of boson particles.

<span class="mw-page-title-main">Quantum mechanics</span> Description of physical properties at the atomic and subatomic scale

Quantum mechanics is a fundamental theory in physics that provides a description of the physical properties of nature at the scale of atoms and subatomic particles. It is the foundation of all quantum physics including quantum chemistry, quantum field theory, quantum technology, and quantum information science.

<span class="mw-page-title-main">Quantum electrodynamics</span> Quantum field theory of electromagnetism

In particle physics, quantum electrodynamics (QED) is the relativistic quantum field theory of electrodynamics. In essence, it describes how light and matter interact and is the first theory where full agreement between quantum mechanics and special relativity is achieved. QED mathematically describes all phenomena involving electrically charged particles interacting by means of exchange of photons and represents the quantum counterpart of classical electromagnetism giving a complete account of matter and light interaction.

Wave–particle duality is the concept in quantum mechanics that quantum entities exhibit particle or wave properties according to the experimental circumstances. It expresses the inability of the classical concepts such as particle or wave to fully describe the behavior of quantum objects. During the 19th and early 20th centuries, light was found to behave as a wave, and then later discovered to have a particulate character, whereas electrons were found to act as particles, and then later discovered to have wavelike aspects. The concept of duality arose to name these contradictions.

<span class="mw-page-title-main">Compton scattering</span> Scattering of photons off charged particles

Compton scattering discovered by Arthur Holly Compton, is the scattering of a high frequency photon after an interaction with a charged particle, usually an electron. It results in a decrease in energy of the photon, called the Compton effect. Part of the energy of the photon is transferred to the recoiling particle. Inverse Compton scattering has the opposite effect, occurring when a high-energy charged particle transfers part of its energy to a photon, resulting in an increase in energy of the photon.

<span class="mw-page-title-main">Energy level</span> Different states of quantum systems

A quantum mechanical system or particle that is bound—that is, confined spatially—can only take on certain discrete values of energy, called energy levels. This contrasts with classical particles, which can have any amount of energy. The term is commonly used for the energy levels of the electrons in atoms, ions, or molecules, which are bound by the electric field of the nucleus, but can also refer to energy levels of nuclei or vibrational or rotational energy levels in molecules. The energy spectrum of a system with such discrete energy levels is said to be quantized.

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. All matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

In physics and chemistry, Bragg's law, Wulff–Bragg's condition or Laue–Bragg interference, a special case of Laue diffraction, gives the angles for coherent scattering of waves from a large crystal lattice. It encompasses the superposition of wave fronts scattered by lattice planes, leading to a strict relation between wavelength and scattering angle, or else to the wavevector transfer with respect to the crystal lattice. Such law had initially been formulated for X-rays upon crystals. However, it applies to all sorts of quantum beams, including neutron and electron waves at atomic distances if there are a large number of atoms, as well as visible light with artificial periodic microscale lattices.

In particle, atomic and condensed matter physics, a Yukawa potential is a potential named after the Japanese physicist Hideki Yukawa. The potential is of the form:

The old quantum theory is a collection of results from the years 1900–1925 which predate modern quantum mechanics. The theory was never complete or self-consistent, but was rather a set of heuristic corrections to classical mechanics. The theory is now understood as the semi-classical approximation to modern quantum mechanics. The main and final accomplishments of the old quantum theory were the determination of the modern form of the periodic table by Edmund Stoner and the Pauli Exclusion Principle which were both premised on the Arnold Sommerfeld enhancements to the Bohr model of the atom.

The Davisson–Germer experiment was a 1923-27 experiment by Clinton Davisson and Lester Germer at Western Electric, in which electrons, scattered by the surface of a crystal of nickel metal, displayed a diffraction pattern. This confirmed the hypothesis, advanced by Louis de Broglie in 1924, of wave-particle duality, and also the wave mechanics approach of the Schrödinger equation. It was an experimental milestone in the creation of quantum mechanics.

Quantum mechanics is the study of matter and its interactions with energy on the scale of atomic and subatomic particles. By contrast, classical physics explains matter and energy only on a scale familiar to human experience, including the behavior of astronomical bodies such as the moon. Classical physics is still used in much of modern science and technology. However, towards the end of the 19th century, scientists discovered phenomena in both the large (macro) and the small (micro) worlds that classical physics could not explain. The desire to resolve inconsistencies between observed phenomena and classical theory led to a revolution in physics, a shift in the original scientific paradigm: the development of quantum mechanics.

In particle physics, wave mechanics and optics, momentum transfer is the amount of momentum that one particle gives to another particle. It is also called the scattering vector as it describes the transfer of wavevector in wave mechanics.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

The ensemble interpretation of quantum mechanics considers the quantum state description to apply only to an ensemble of similarly prepared systems, rather than supposing that it exhaustively represents an individual physical system.

The history of quantum mechanics is a fundamental part of the history of modern physics. The major chapters of this history begin with the emergence of quantum ideas to explain individual phenomena -- blackbody radiation, the photoelectric effect, solar emission spectra -- an era called the Old or Older quantum theories. The invention of wave mechanics by Schrodinger and expanded by many others triggers the "modern" era beginning around 1925. Dirac's relativistic quantum theory work lead him to explore quantum theories of radiation, culminating in quantum electrodynamics, the first quantum field theory. The history of quantum mechanics continues in the history of quantum field theory. The history of quantum chemistry, theoretical basis of chemical structure, reactivity, and bonding, interlaces with the events discussed in this article.

The Bohr–Kramers–Slater theory was perhaps the final attempt at understanding the interaction of matter and electromagnetic radiation on the basis of the so-called old quantum theory, in which quantum phenomena are treated by imposing quantum restrictions on classically describable behaviour. It was advanced in 1924, and sticks to a classical wave description of the electromagnetic field. It was perhaps more a research program than a full physical theory, the ideas that are developed not being worked out in a quantitative way. The purpose of BKS Theory was to disprove Einstein's hypothesis of the light quantum.

The Kapitza–Dirac effect is a quantum mechanical effect consisting of the diffraction of matter by a standing wave of light. The effect was first predicted as the diffraction of electrons from a standing wave of light by Paul Dirac and Pyotr Kapitsa in 1933. The effect relies on the wave–particle duality of matter as stated by the de Broglie hypothesis in 1924.

The Planck constant, or Planck's constant, denoted by , is a fundamental physical constant of foundational importance in quantum mechanics: a photon's energy is equal to its frequency multiplied by the Planck constant, and the wavelength of a matter wave equals the Planck constant divided by the associated particle momentum.

References

  1. 1 2 Duane, W. (1923). The transfer in quanta of radiation momentum to matter, Proc. Natl. Acad. Sci.9(5): 158–164.
  2. Bitsakis, E.(1997). The wave-particle duality, pp. 333–348 in The Present Status of the Quantum Theory of Light: Proceedings of a Symposium in Honour of Jean-Pierre Vigier, edited by Whitney, C.K., Jeffers, S., Roy, S., Vigier, J.-P., Hunter, G., Springer, ISBN   978-94-010-6396-8, p. 338.
  3. COWLEY, JOHN M. (1995), "Diffraction from crystals", Diffraction Physics, Elsevier, pp. 123–144, doi:10.1016/b978-044482218-5/50008-0, ISBN   9780444822185 , retrieved 2023-08-13
  4. Cullity, Bernard D.; Stock, Stuart R. (2001). Elements of X-ray diffraction (3rd ed.). Upper Saddle River, NJ: Prentice Hall. ISBN   978-0-201-61091-8.
  5. Warren, Bertram Eugene (1990). X-ray diffraction. Dover books on physics and chemistry. New York: Dover. ISBN   978-0-486-66317-3.
  6. Peng, L.-M.; Dudarev, S. L.; Whelan, M. J. (2011). High energy electron diffraction and microscopy. Monographs on the physics and chemistry of materials (1. publ. in paperback ed.). Oxford: Oxford Univ. Press. ISBN   978-0-19-960224-7.
  7. Einstein, A. (1905). "Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt". Annalen der Physik . 17 (6): 132–148. Bibcode:1905AnP...322..132E. doi: 10.1002/andp.19053220607 . Translated in Arons, A. B.; Peppard, M. B. (1965). "Einstein's proposal of the photon concept: A translation of the Annalen der Physik paper of 1905" (PDF). American Journal of Physics . 33 (5): 367. Bibcode:1965AmJPh..33..367A. doi:10.1119/1.1971542. Archived from the original (PDF) on 2016-03-04. Retrieved 2014-09-14.
  8. Einstein, A. (1916). "Zur Quantentheorie der Strahlung". Mitteilungen der Physikalischen Gesellschaft Zürich. 18: 47–62. and a nearly identical version Einstein, A. (1917). "Zur Quantentheorie der Strahlung". Physikalische Zeitschrift . 18: 121–128. Bibcode:1917PhyZ...18..121E. Translated here and in ter Haar, D. (1967). The Old Quantum Theory . Pergamon Press. pp.  167–183. LCCN   66029628.
  9. Born, M. (1925/1971). Letter of 15 July 1925, pp. 84–85 in The Born-Einstein Letters, translated by I. Born, Macmillan, London.
  10. Epstein, P.S., Ehrenfest, P., (1924). The quantum theory of the Fraunhofer diffraction, Proc. Natl. Acad. Sci.10: 133–139.
  11. Ehrenfest, P., Epstein, P.S. (1924/1927). Remarks on the quantum theory of diffraction, Proc. Natl. Acad. Sci.13: 400–408.
  12. Compton, A.H. (1923). The quantum integral and diffraction by a crystal, Proc. Natl. Acad. Sci.9(11): 360–362.
  13. Landé, A., Born, M., Biem, W. (1968). 'Dialog on dualism', Physics Today, 21(8): 55–56; doi : 10.1063/1.3035103.
  14. Heisenberg, W. (1930). The Physical Principles of the Quantum Theory, translated by C. Eckart and F.C. Hoyt, University of Chicago Press, Chicago, pp. 77–78.
  15. Pauling, L.C., Wilson, E.B. (1935). Introduction to Quantum Mechanics: with Applications to Chemistry, McGraw-Hill, New York, pp. 34–36.
  16. Landé, A. (1951). Quantum Mechanics, Sir Isaac Pitman and Sons, London, pp. 19–22.
  17. Bohm, D. (1951). Quantum Theory, Prentice Hall, New York, pp. 71–73.
  18. Breit, G. (1923). The interference of light and the quantum theory, Proc. Natl. Acad. Sci.9: 238–243.
  19. Tonomura, A., Endo, J., Matsuda, T., Kawasaki, T., Ezawa, H. (1989). Demonstration of single‐electron buildup of an interference pattern, Am. J. Phys. 57(2): 117–120.
  20. Dragoman, D. Dragoman, M. (2004). Quantum–Classical Analogies, Springer, Berlin, ISBN   3-540-20147-5, pp. 170–175.
  21. Schmidt, L.P.H., Lower, J., Jahnke, T., Schößler, S., Schöffler, M.S., Menssen, A., Lévêque, C., Sisourat, N., Taïeb, R., Schmidt-Böcking, H., Dörner, R. (2013). Momentum transfer to a free floating double slit: realization of a thought experiment from the Einstein-Bohr debates, Physical Review Letters111: 103201, 1–5.
  22. Bohr, N. (1948). On the notions of causality and complementarity, Dialectica2: 312–319; p. 313: "It is further important to realize that any determination of Planck's constant rests upon the comparison between aspects of the phenomena which can be described only by means of pictures not combinable on the basis of classical theories."
  23. Messiah, A. (1961). Quantum Mechanics, volume 1, translated by G.M. Temmer from the French Mécanique Quantique, North-Holland, Amsterdam, p. 52, "relations between dynamical variables of the particle and characteristic quantities of the associated wave".
  24. 1 2 Heisenberg, W. (1930). The Physical Principles of the Quantum Theory, translated by C. Eckart and F.C. Hoyt, University of Chicago Press, Chicago, p. 77.
  25. Landé, A. (1951). Quantum Mechanics, Sir Isaac Pitman and Sons, London, p. 20.
  26. Ballentine, L.E. (1998). Quantum Mechanics: a Modern Development, World Scientific, Singapore, ISBN   981-02-2707-8, p. 136.
  27. Van Vliet, K. (1967). Linear momentum quantization in periodic structures, Physica, 35: 97–106, doi:10.1016/0031-8914(67)90138-3.
  28. Van Vliet, K. (2010). Linear momentum quantization in periodic structures ii, Physica A, 389: 1585–1593, doi:10.1016/j.physa.2009.12.026.
  29. Thankappan, V.K. (1985/2012). Quantum Mechanics, third edition, New Age International, New Delhi, ISBN   978-81-224-3357-9, pp. 6–7.
  30. Wennerstrom, H. (2014). Scattering and diffraction described using the momentum representation, Advances in Colloid and Interface Science, 205: 105–112.
  31. Mehra, J., Rechenberg, H. (2001). The Historical Development of Quantum Theory, volume 1, part 2, Springer, pp. 555–556 here.
  32. Hickey, T.J. (2014). Twentieth-Century Philosophy of Science:a History, self-published by the author, here.
  33. Prigogine, I. (1962). Non-equilibrium Statistical Mechanics, Wiley, New York, pp. 258–262.
  34. Squires, G.L. (1978/2012). Introduction to the Theory of Thermal Neutron Scattering, third edition, Cambridge University Press, Cambridge UK, ISBN   978-110-764406-9.
  35. Böni, P., Furrer, A. (1999). Introduction to neutron scattering, Chapter 1, pp. 1–27 of Frontiers of Neutron Scattering, edited by A. Furrer, World Scientific, Singapore, ISBN   981-02-4069-4.
  36. Schrödinger, E. (1927). Über den Comptoneffekt, Annalen der Physik series 4, 82<387(2)>: 257–264. Translated from the second German edition by J.F. Shearer, W.M. Deans at pp. 124–129 in Collected papers on Wave Mechanics, Blackie & Son, London (1928).
  37. Landé, A. (1951). Quantum Mechanics, Sir Isaac Pitman and Sons, London pp. 16–18.
  38. Bacciagaluppi, G., Crull, E. (2009). Heisenberg (and Schrödinger, and Pauli) on hidden variables, Studies in History and Philosophy of Modern Physics, 40: 374–382.
  39. Heisenberg, W. (1969/1985) The concept of "understanding" in theoretical physics, pp. 7–10 in Properties of Matter Under Unusual Conditions (In Honor of Edward Teller's 60th Birthday), edited by H. Mark, S. Fernbach, Interscience Publishers, New York, reprinted at pp. 335–339 in Heisenberg, W., Collected Works, series C, volume 3, ed. W. Blum, H.-P. Dürr, H. Rechenberg, Piper, Munich, ISBN   3-492-02927-2, p. 336.