You are on page 1of 4

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/251154378

Atomic Models, J.J. Thomson's “Plum Pudding” Model

Article · July 2009


DOI: 10.1007/978-3-540-70626-7_9

CITATIONS READS

5 27,271

1 author:

Klaus Hentschel
Universität Stuttgart
618 PUBLICATIONS   1,288 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

History of Spectroscopy View project

history of "history of science & technology" View project

All content following this page was uploaded by Klaus Hentschel on 16 June 2017.

The user has requested enhancement of the downloaded file.


Appeared in D. Greenberger, K. Hentschel & F. Weinert (eds.) Compendium of Quantum
Physics – Concerpts, Experiments, History and Philosophy, Springer 2009, 18-21.

Atomic models, J.J. Thomson’s “plum pudding” model

In 1897, Joseph John Thomson (1856–1940) announced the discovery of a corpuscle. Others
soon called it →electron, despite Thomson’s stubborn preference for his original term,
borrowed from Robert Boyle (1627–91) to denote any particle-like structure. Very soon
afterwards, Thomson began to think about how to explain the periodicity of the chemical
elements in terms of these negatively charged corpuscles. Chemical properties would thus
have to depend on the number and constellations of these atomic constituents. The corpuscles
would have to have stable positions inside the atom, bound by electrostatic and possibly
kinetic forces. Because under normal conditions chemical atoms are electrically neutral, the
total electric charge of these negatively charged electrons had to be compensated for by an
equally strong amount of positive charge. For Thomson it was natural to assume that this
positive charge was continuously distributed over the atom, whose radius was estimated at the
time to be around 1012. He visualized the very small negatively charged electrons
(contemporary estimates indicated an order of magnitude of 10-15m) as dotting the atom like
raisins inside a cake or like plums in a pudding, whence the popular nickname for Thomson’s
atomic model as the “plum pudding model”.

In order to get a better idea of the stable configurations of these corpuscles inside the atom,
Thomson drew an analogy with experiments by Alfred Marshall Mayer (1836–1897).
Piercing small magnetic needles into corks and letting them float in water below a strong
magnet (see figure 1, left), Mayer had observed in 1878/79 that the magnetized floating
needles quasi-automatically positioned themselves in characteristic configurations depending
on their number. With more than six magnetic needles present, a seventh and eighth would
inevitably position itself inside the outer ring of six (see the third row of fig. 1, middle). As
the number of floating magnets increased, more and more rings would form. Thomson hoped
that a similar ring-structure composed of corpuscules could be found inside chemical atoms,
and suspected that each of these rings would be analogous to the chemical periods in the
periodic table of the elements. Specific configurations of the innermost ring would determine
the chemical properties of the chemical element at hand. Two chemical elements with
differing numbers of outer rings of corpuscles but similar innermost configurations would
thus share similar chemical properties, like elements situated beneath each other in a column
of the periodic table. To stabilize these configurations, Thomson also assumed that the
concentric rings would all rotate around their common center.

Figure 1 Left: From A.M. Mayer (1878), p. 248; right: from Mayer (1879), pp. 100-101.
Around 1904 Thomson believed each chemical atom would contain a very large number of
electrons, something in the order of magnitude of 1000 or more. With such high numbers he
hoped to explain the puzzle of the exceedingly many spectral lines in each atom’s spectrum and
the fact that the masses of atoms proved to be several thousand times the mass of an electron.
Radioactive decay, very often correlated with the emission of negatively charged β-rays which
turned out to be nothing but highly accelerated electrons, Thomson thus interpreted as a
mechanical instability of these electron configurations. A slight disturbance of the carefully
balanced equilibrium position would result in electrostatic repulsion taking over and the
expulsion of individual electrons or whole groups of electrons from the atom, where they
would be experimentally observable as β-rays. Thomson also tried to explore the atomic
structure by using corpuscles/electrons as projectiles in β-ray scattering experiments onto a thin
foil. The scattering angles observed by him and his students were predominantly very small,
with a Gaussian distribution peaking sharply around zero-degree refraction and a width
proportional to the thickness of the target layer. This experimental finding was interpreted as
evidence for small-angle scattering, with successive layers of matter in thicker foils inducing an
increasing, but still relatively small probability of multiple scattering, with occasional larger
scattering angles resulting.

When Ernest Rutherford (1871–1937) started to conduct → scattering experiments, he varied


Thomson’s set-up by also using the positively charged and much heavier α-rays as projectiles.
As discussed in detail in the entries on → large-angle scattering and the → Rutherford
nucleus model, Rutherford’s experiments showed that large-angle scattering was far more
frequent than would be expected on the basis of J.J. Thomson’s plum pudding model. Ruther-
ford decided to modify J.J. Thomson’s atomic model: instead of assuming a continuous
smeared-out positive charge, Rutherford postulated a concentrated → atomic nucleus model
with positive charge surrounded by a diffuse sphere of negative electricity (cf. fig. 2). Quanti-
tative analysis of his α-ray scattering experiments showed this atomic nucleus model was
consistent with his data if the positive charge of the core was of the order of A/2 · e, with A
being the atomic number of the chemical element and e equal to the charge of J.J. Thomson’s
corpuscles, the elementary charge quantum. Thus Rutherford’s estimate (which proved to be
correct) drastically reduced the number of electrons inside atoms compared to J.J. Thomson’s.

Fig. 2: Rutherford’s first calculations on the passage of α-particles through atoms: “Theory of
structure of atoms / Suppose atoms consist of + charge ne at centre & – charge as electrons
distributed throughout sphere of radius ρ.” From the Rutherford papers, Cambridge University
Library, reproduced from Heilbron (1977), p. 24.
When the young Niels Bohr (1885–1962) finished writing his Ph.D. thesis at the University of
Copenhagen and obtained a fellowship for postgraduate study abroad, he chose to go to
Cambridge, hoping to work more closely with J.J. Thomson, who was director of the
Cavendish laboratory since 1884. The two personalities did not match, however, and Bohr
soon decided to move on to Manchester where Ernest Rutherford introduced him to the
intricacies of scattering experiments with α-rays and discussed his brand new nuclear core
model of the atom. In the atomic model Bohr introduced in 1913, later refined by Arnold
Sommerfeld (1868–1951) and others (→Bohr-Sommerfeld atom model; → Sommerfeld
school), Bohr masterfully merged ideas from J.J. Thomson, Rutherford and Nagaoka (→
Saturnian atomic model). He also superimposed quantum conditions introduced by Max
Planck (1858–1947) in 1900 and first employed in atomic models from 1910 on, by Arthur
Erich Haas (1884–1941) and John William Nicholson (1881–1955) [cf., e.g. McCormmach
1975, and Hermann (ed.) 1965]. While Bohr and Rutherford soon looked back on the older
atomic models by J.J. Thomson and others as “a museum of scientific curiosities”, J.J.
Thomson for his part rejected Bohr’s advances as “meretricious superficialities obtained
without, or at the price of, an understanding of the mechanism of atoms” (Heilbron 1977, p.
23). Today we know that J.J. Thomson’s hope to arrive at an intuitive, quasi-mechanical
understanding of the atom was in vain—but at the time no one could be sure.
Klaus Hentschel, Univ. Stuttgart

Literature

Primary:

Alfred M. Mayer: Floating magnets, American Journal of Science 116 (1878), 248f., also in
Nature 17, 487-488.
– On the morphological laws of the configurations formed by magnets floating vertically and
subjected to the attraction of a superposed magnet, Philosophical Magazine (5th series) 17
(1879), 98-108.
Joseph John Thomson: On the structure of the atom, Philosophical Magazine (6th series) 7
(1904), 237-265.
– The structure of the atom (an evening lecture at the Royal Institution of London on March
10, 1905), published in the Proceedings of the Royal Institution of London 1905, 1-15.
– The Corpuscular Theory of Matter, London: Constable & Co., 1907.

Secondary:

Edward Arthur Davis & Elisabeth Falconer: J.J. Thomson and the Discovery of the Electron,
London: Taylor & Francis 1997, esp. chap. 6, with a reprint of Thomson 1905 on pp. 215-229.
John Heilbron: J.J. Thomson and the Bohr atom, Physics Today 30, April 1977, 23-30.
Armin Hermann (ed.) Der erste Quantenansatz für das Atom: Arthur Erich Haas, Stuttgart:
Battenberg, 1965 (Dokumente der Naturwissenschaft, Abt. Physik, issue 10)
Helge Kragh: The first subatomic explanations of the periodic system, Foundations of
Chemistry 3, issue 2 (2001), 129-143.
Russell McCormmach: The atomic theory of John William Nicholson, Archive for History of
Exact Sciences 3, issue 2 (1975), 160-184.

View publication stats

You might also like