Karl Schwarzschild’s Radius: How Fame Eclipsed a Physicist’s own Legacy

In an ironic twist of the history of physics, Karl Schwarzschild’s fame has eclipsed his own legacy.  When asked who was Karl Schwarzschild (1873 – 1916), you would probably say he’s the guy who solved Einstein’s Field Equations of General Relativity and discovered the radius of black holes.  You may also know that he accomplished this Herculean feat while dying slowly behind the German lines on the Eastern Front in WWI.  But asked what else he did, and you would probably come up blank.  Yet Schwarzschild was one of the most wide-ranging physicists at the turn of the 20th century, which is saying something, because it places him into the same pantheon as Planck, Lorentz, Poincaré and Einstein.  Let’s take a look at the part of his career that hides in the shadow of his own radius.

A Radius of Interest

Karl Schwarzschild was born in Frankfurt, Germany, shortly after the Franco-Prussian war thrust Prussia onto the world stage as a major political force in Europe.  His family were Jewish merchants of longstanding reputation in the city, and Schwarzschild’s childhood was spent in the vibrant Jewish community.  One of his father’s friends was a professor at a university in Frankfurt, whose son, Paul Epstein (1871 – 1939), became a close friend of Karl’s at the Gymnasium.  Schwarzshild and Epstein would partially shadow each other’s careers despite the fact that Schwarzschild became an astronomer while Epstein became a famous mathematician and number theorist.  This was in part because Schwarzschild had large radius of interests that spanned the breadth of current mathematics and science, practicing both experiments and theory. 

Schwarzschild’s application of the Hamiltonian formalism for quantum systems set the stage for the later adoption of Hamiltonian methods in quantum mechanics. He came dangerously close to stating the uncertainty principle that catapulted Heisenberg to fame.

By the time Schwarzschild was sixteen, he had taught himself the mathematics of celestial mechanics to such depth that he published two papers on the orbits of binary stars.  He also became fascinated in astronomy and purchased lenses and other materials to construct his own telescope.  His interests were helped along by Epstein, two years older and whose father had his own private observatory.  When Epstein went to study at the University of Strasbourg (then part of the German Federation) Schwarzschild followed him.  But Schwarzschild’s main interest in astronomy diverged from Epstein’s main interest in mathematics, and Schwarzschild transferred to the University of Munich where he studied under Hugo von Seeliger (1849 – 1924), the premier German astronomer of his day.  Epstein remained at Strasbourg where he studied under Bruno Christoffel (1829 – 1900) and eventually became a professor, but he was forced to relinquish the post when Strasbourg was ceded to France after WWI. 

The Birth of Stellar Interferometry

Until the Hubble space telescope was launched in 1990 no star had ever been resolved as a direct image.  Within a year of its launch, using its spectacular resolving power, the Hubble optics resolved—just barely—the red supergiant Betelgeuse.  No other star (other than the Sun) is close enough or big enough to image the stellar disk, even for the Hubble far above our atmosphere.  The reason is that the diameter of the optical lenses and mirrors of the Hubble—as big as they are at 2.4 meter diameter—still produce a diffraction pattern that smears the image so that stars cannot be resolved.  Yet information on the size of a distant object is encoded as phase in the light waves that are emitted from the object, and this phase information is accessible to interferometry.

The first physicist who truly grasped the power of optical interferometry and who understood how to design the first interferometric metrology systems was the French physicist Armand Hippolyte Louis Fizeau (1819 – 1896).  Fizeau became interested in the properties of light when he collaborated with his friend Léon Foucault (1819–1868) on early uses of photography.  The two then embarked on a measurement of the speed of light but had a falling out before the experiment could be finished, and both continued the pursuit independently.  Fizeau achieved the first measurement using a toothed wheel rotating rapidly [1], while Foucault came in second using a more versatile system with a spinning mirror [2].  Yet Fizeau surpassed Foucault in optical design and became an expert in interference effects.  Interference apparatus had been developed earlier by Augustin Fresnel (the Fresnel bi-prism 1819), Humphrey Lloyd (Lloyd’s mirror 1834) and Jules Jamin (Jamin’s interferential refractor 1856).  They had found ways of redirecting light using refraction and reflection to cause interference fringes.  But Fizeau was one of the first to recognize that each emitting region of a light source was coherent with itself, and he used this insight and the use of lenses to design the first interferometer.

Fizeau’s interferometer used a lens with a with a tight focal spot masked off by an opaque screen with two open slits.  When the masked lens device was focused on an intense light source it produced two parallel pencils of light that were mutually coherent but spatially separated.  Fizeau used this apparatus to measure the speed of light in moving water in 1859 [3]

Fig. 1  Optical configuration of the source element of the Fizeau refractometer.

The working principle of the Fizeau refractometer is shown in Fig. 1.  The light source is at the bottom, and it is reflected by the partially-silvered beam splitter to pass through the lens and the mask containing two slits.  (Only the light paths that pass through the double-slit mask on the lens are shown in the figure.)  The slits produce two pencils of mutually coherent light that pass through a system (in the famous Fizeau ether drag experiment it was along two tubes of moving water) and are returned through the same slits, and they intersect at the view port where they produce interference fringes.  The fringe spacing is set by the separation of the two slits in the mask.  The Rayleigh region of the lens defines a region of spatial coherence even for a so-called “incoherent” source.  Therefore, this apparatus, by use of the lens, could convert an incoherent light source into a coherent probe to test the refractive index of test materials, which is why it was called a refractometer. 

Fizeau became adept at thinking of alternative optical designs of his refractometer and alternative applications.  In an address to the French Physical Society in 1868 he suggested that the double-slit mask could be used on a telescope to determine sizes of distant astronomical objects [4].  There were several subsequent attempts to use Fizeau’s configuration in astronomical observations, but none were conclusive and hence were not widely known.

An optical configuration and astronomical application that was very similar to Fizeau’s idea was proposed by Albert Michelson in 1890 [5].  He built the apparatus and used it to successfully measure the size of several moons of Jupiter [6].  The configuration of the Michelson stellar interferometer is shown in Fig. 2.  Light from a distant star passes through two slits in the mask in front of the collecting optics of a telescope.  When the two pencils of light intersect at the view port, they produce interference fringes.  Because of the finite size of the stellar source, the fringes are partially washed out.  By adjusting the slit separation, a certain separation can be found where the fringes completely wash out.  The size of the star is then related to the separation of the slits for which the fringe visibility vanishes.  This simple principle allows this type of stellar interferometry to measure the size of stars that are large and relatively close to Earth.  However, if stars are too far away even this approach cannot be used to measure their sizes because telescopes aren’t big enough.  This limitation is currently being bypassed by the use of long-baseline optical interferometers.

Fig. 2  Optical configuration of the Michelson stellar interferometer.  Fringes at the view port are partially washed out by the finite size of the star.  By adjusting the slit separation, the fringes can be made to vanish entirely, yielding an equation that can be solved for the size of the star.

One of the open questions in the history of interferometry is whether Michelson was aware of Fizeau’s proposal for the stellar interferometer made in 1868.  Michelson was well aware of Fizeau’s published research and acknowledged him as a direct inspiration of his own work in interference effects.  But Michelson also was unaware of the undercurrents in the French school of optical interference.  When he visited Paris in 1881, he met with many of the leading figures in this school (including Lippmann and Cornu), but there is no mention or any evidence that he met with Fizeau.  By this time Fizeau’s wife had passed away, and Fizeau spent most of his time in seclusion at his home outside Paris.  Therefore, it is unlikely that he would have been present during Michelson’s visit.  Because Michelson viewed Fizeau with such awe and respect, if he had met him, he most certainly would have mentioned it.  Therefore, Michelson’s invention of the stellar interferometer can be considered with some confidence to be a case of independent discovery.  It is perhaps not surprising that he hit on the same idea that Fizeau had in 1868, because Michelson was one of the few physicists who understood coherence and interference at the same depth as Fizeau.

Schwarzschild’s Stellar Interferometer

The physics of the Michelson stellar interferometer is very similar to the physics of Young’s double slit experiment.  The two slits in the aperture mask of the telescope objective act to produce a simple sinusoidal interference pattern at the image plane of the optical system.  The size of the stellar diameter is determined by using the wash-out effect of the fringes caused by the finite stellar size.  However, it is well known to physicists who work with diffraction gratings that a multiple-slit interference pattern has a much greater resolving power than a simple double slit. 

This realization must have hit von Seeliger and Schwarzschild, working together at Munich, when they saw the publication of Michelson’s theoretical analysis of his stellar interferometer in 1890, followed by his use of the apparatus to measure the size of Jupiter’s moons.  Schwarzschild and von Seeliger realized that by replacing the double-slit mask with a multiple-slit mask, the widths of the interference maxima would be much narrower.  Such a diffraction mask on a telescope would cause a star to produce a multiple set of images on the image plane of the telescope associated with the multiple diffraction orders.  More interestingly, if the target were a binary star, the diffraction would produce two sets of diffraction maxima—a double image!  If the “finesse” of the grating is high enough, the binary star separation could be resolved as a doublet in the diffraction pattern at the image, and the separation could be measured, giving the angular separation of the two stars of the binary system.  Such an approach to the binary separation would be a direct measurement, which was a distinct and clever improvement over the indirect Michelson configuration that required finding the extinction of the fringe visibility. 

Schwarzschild enlisted the help of a fine German instrument maker to create a multiple slit system that had an adjustable slit separation.  The device is shown in Fig. 3 from Schwarzschild’s 1896 publication on the use of the stellar interferometer to measure the separation of binary stars [7].  The device is ingenious.  By rotating the chain around the gear on the right-hand side of the apparatus, the two metal plates with four slits could be raised or lowered, cause the projection onto the objective plane to have variable slit spacings.  In the operation of the telescope, the changing height of the slits does not matter, because they are near a conjugate optical plane (the entrance pupil) of the optical system.  Using this adjustable multiple slit system, Schwarzschild (and two colleagues he enlisted) made multiple observations of well-known binary star systems, and they calculated the star separations.  Several of their published results are shown in Fig. 4.

Fig. 3  Illustration from Schwarzschild’s 1896 paper describing an improvement of the Michelson interferometer for measuring the separation of binary star systems Ref. [7].
Fig. 4  Data page from Schwarzschild’s 1896 paper measuring the angular separation of two well-known binary star systems: gamma Leonis and chsi Ursa Major. Ref. [7]

Schwarzschild’s publication demonstrated one of the very first uses of stellar interferometry—well before Michelson himself used his own configuration to measure the diameter of Betelgeuse in 1920.  Schwarzschild’s major achievement was performed before he had received his doctorate, on a topic orthogonal to his dissertation topic.  Yet this fact is virtually unknown to the broader physics community outside of astronomy.  If he had not become so famous later for his solution of Einstein’s field equations, Schwarzschild nonetheless might have been famous for his early contributions to stellar interferometry.  But even this was not the end of his unique contributions to physics.

Adiabatic Physics

As Schwarzschild worked for his doctorate under von Seeliger, his dissertation topic was on new theories by Henri Poincaré (1854 – 1912) on celestial mechanics.  Poincaré had made a big splash on the international stage with the publication of his prize-winning memoire in 1890 on the three-body problem.  This is the publication where Poincaré first described what would later become known as chaos theory.  The memoire was followed by his volumes on “New Methods in Celestial Mechanics” published between 1892 and 1899.  Poincaré’s work on celestial mechanics was based on his earlier work on the theory of dynamical systems where he discovered important invariant theorems, such as Liouville’s theorem on the conservation of phase space volume.  Schwarzshild applied Poincaré’s theorems to problems in celestial orbits.  He took his doctorate in 1896 and received a post at an astronomical observatory outside Vienna. 

While at Vienna, Schwarzschild performed his most important sustained contributions to the science of astronomy.  Astronomical observations had been dominated for centuries by the human eye, but photographic techniques had been making steady inroads since the time of Hermann Carl Vogel (1841 – 1907) in the 1880’s at the Potsdam observatory.  Photographic plates were used primarily to record star positions but were known to be unreliable for recording stellar intensities.  Schwarzschild developed a “out-of-focus” technique that blurred the star’s image, while making it larger and easier to measure the density of the exposed and developed photographic emulsions.  In this way, Schwarzschild measured the magnitudes of 367 stars.  Two of these stars had variable magnitudes that he was able to record and track.  Schwarzschild correctly explained the intensity variation caused by steady oscillations in heating and cooling of the stellar atmosphere.  This work established the properties of these Cepheid variables which would become some of the most important “standard candles” for the measurement of cosmological distances.  Based on the importance of this work, Schwarzschild returned to Munich as a teacher in 1899 and subsequently was appointed in 1901 as the director of the observatory at Göttingen established by Gauss eighty years earlier.

Schwarzschild’s years at Göttingen brought him into contact with some of the greatest mathematicians and physicists of that era.  The mathematicians included Felix Klein, David Hilbert and Hermann Minkowski.  The physicists included von Laue, a student of Woldemar Voigt.  This period was one of several “golden ages” of Göttingen.  The first golden age was the time of Gauss and Riemann in the mid-1800’s.  The second golden age, when Schwarzschild was present, began when Felix Klein arrived at Göttingen and attracted the top mathematicians of the time.  The third golden age of Göttingen was the time of Born and Jordan and Heisenberg at the birth of quantum mechanics in the mid 1920’s.

In 1906, the Austrian Physicist Paul Ehrenfest, freshly out of his PhD under the supervision of Boltzmann, arrived at Göttingen only weeks before Boltzmann took his own life.  Felix Klein at Göttingen had been relying on Boltzmann to provide a comprehensive review of statistical mechanics for the Mathematical Encyclopedia, so he now entrusted this project to the young Ehrenfest.  It was a monumental task, which was to take him and his physicist wife Tatyanya nearly five years to complete.  Part of the delay was the desire by the Ehrenfests to close some open problems that remained in Boltzmann’s work.  One of these was a mechanical theorem of Boltzmann’s that identified properties of statistical mechanical systems that remained unaltered through a very slow change in system parameters.  These properties would later be called adiabatic invariants by Einstein. 

Ehrenfest recognized that Wien’s displacement law, which had been a guiding light for Planck and his theory of black body radiation, had originally been derived by Wien using classical principles related to slow changes in the volume of a cavity.  Ehrenfest was struck by the fact that such slow changes would not induce changes in the quantum numbers of the quantized states, and hence that the quantum numbers must be adiabatic invariants of the black body system.  This not only explained why Wien’s displacement law continued to hold under quantum as well as classical considerations, but it also explained why Planck’s quantization of the energy of his simple oscillators was the only possible choice.  For a classical harmonic oscillator, the ratio of the energy of oscillation to the frequency of oscillation is an adiabatic invariant, which is immediately recognized as Planck’s quantum condition .  

Ehrenfest published his observations in 1913 [8], the same year that Bohr published his theory of the hydrogen atom, so Ehrenfest immediately applied the theory of adiabatic invariants to Bohr’s model and discovered that the quantum condition for the quantized energy levels was again the adiabatic invariants of the electron orbits, and not merely a consequence of integer multiples of angular momentum, which had seemed somewhat ad hoc

After eight exciting years at Göttingen, Schwarzschild was offered the position at the Potsdam Observatory in 1909 upon the retirement from that post of the famous German astronomer Carl Vogel who had made the first confirmed measurements of the optical Doppler effect.  Schwarzschild accepted and moved to Potsdam with a new family.  His son Martin Schwarzschild would follow him into his profession, becoming a famous astronomer at Princeton University and a theorist on stellar structure.  At the outbreak of WWI, Schwarzschild joined the German army out of a sense of patriotism.  Because of his advanced education he was made an officer of artillery with the job to calculate artillery trajectories, and after a short time on the Western Front in Belgium was transferred to the Eastern Front in Russia.  Though he was not in the trenches, he was in the midst of the chaos to the rear of the front.  Despite this situation, he found time to pursue his science through the year 1915. 

Schwarzschild was intrigued by Ehrenfest’s paper on adiabatic invariants and their similarity to several of the invariant theorems of Poincaré that he had studied for his doctorate.  Up until this time, mechanics had been mostly pursued through the Lagrangian formalism which could easily handle generalized forces associated with dissipation.  But celestial mechanics are conservative systems for which the Hamiltonian formalism is a more natural approach.  In particular, the Hamilton-Jacobi canonical transformations made it particularly easy to find pairs of generalized coordinates that had simple periodic behavior.  In his published paper [9], Schwarzschild called these “Action-Angle” coordinates because one was the action integral that was well-known in the principle of “Least Action”, and the other was like an angle variable that changed steadily in time (see Fig. 5). Action-angle coordinates have come to form the foundation of many of the properties of Hamiltonian chaos, Hamiltonian maps, and Hamiltonian tapestries.

Fig. 5  Description of the canonical transformation to action-angle coordinates (Ref. [9] pg. 549). Schwarzschild names the new coordinates “Wirkungsvariable” and “Winkelvariable”.

During lulls in bombardments, Schwarzschild translated the Hamilton-Jacobi methods of celestial mechanics to apply them to the new quantum mechanics of the Bohr orbits.  The phrase “quantum mechanics” had not yet been coined (that would come ten years later in a paper by Max Born), but it was clear that the Bohr quantization conditions were a new type of mechanics.  The periodicities that were inherent in the quantum systems were natural properties that could be mapped onto the periodicities of the angle variables, while Ehrenfest’s adiabatic invariants could be mapped onto the slowly varying action integrals.  Schwarzschild showed that action-angle coordinates were the only allowed choice of coordinates, because they enabled the separation of the Hamilton-Jacobi equations and hence provided the correct quantization conditions for the Bohr electron orbits.  Later, when Sommerfeld published his quantized elliptical orbits in 1916, the multiplicity of quantum conditions and orbits had caused concern, but Ehrenfest came to the rescue, showing that each of Sommerfeld’s quantum conditions were precisely Schwarzschild’s action-integral invariants of the classical electron dynamics [10].

The works by Schwarzschild, and a closely-related paper that amplified his ideas published by his friend Paul Epstein several months later [11], were the first to show the power of the Hamiltonian formulation of dynamics for quantum systems, foreshadowing the future importance of Hamiltonians for quantum theory.  An essential part of the Hamiltonian formalism is the concept of phase space.  In his paper, Schwarzschild showed that the phase space of quantum systems was divided into small but finite elementary regions whose areas were equal to Planck’s constant h-bar (see Fig. 6).  The areas were products of a small change in momentum coordinate Delta-p and a corresponding small change in position coordinate Delta-x.  Therefore, the product DxDp = h-bar.  This observation, made in 1915 by Schwarzschild, was only one step away from Heisenberg’s uncertainty relation, twelve years before Heisenberg discovered it.  However, in 1915 Born’s probabilistic interpretation of quantum mechanics had not yet been made, nor the idea of measurement uncertainty, so Schwarzschild did not have the appropriate context in which to have made the leap to the uncertainty principle.  However, by introducing the action-angle coordinates as well as the Hamiltonian formalism applied to quantum systems, with the natural structure of phase space, Schwarzschild laid the foundation for the future developments in quantum theory made by the next generation.

Fig. 6  Expression of the division of phase space into elemental areas of action equal to h-bar (Ref. [9] pg. 550).

All Quiet on the Eastern Front

Towards the end of his second stay in Munich in 1900, prior to joining the Göttingen faculty, Schwarzschild had presented a paper at a meeting of the German Astronomical Society held in Heidelberg in August.  The topic was unlike anything he had tackled before.  It considered the highly theoretical question of whether the universe was non-Euclidean, and more specifically if it had curvature.  He concluded from observation that if the universe were curved, the radius of curvature must be larger than between 50 light years and 2000 light years, depending on whether the geometry was hyperbolic or elliptical.  Schwarzschild was working out ideas of differential geometry and applying them to the universe at large at a time when Einstein was just graduating from the ETH where he skipped his math classes and had his friend Marcel Grossmann take notes for him.

The topic of Schwarzschild’s talk tells an important story about the warping of historical perspective by the “great man” syndrome.  In this case the great man is Einstein who is today given all the credit for discovering the warping of space.  His development of General Relativity is often portrayed as by a lone genius in the wilderness performing a blazing act of creation out of the void.  In fact, non-Euclidean geometry had been around for some time by 1900—five years before Einstein’s Special Theory and ten years before his first publications on the General Theory.  Gauss had developed the idea of intrinsic curvature of a manifold fifty years earlier, amplified by Riemann.  By the turn of the century alternative geometries were all the rage, and Schwarzschild considered whether there were sufficient astronomical observations to set limits on the size of curvature of the universe.  But revisionist history is just as prevalent in physics as in any field, and when someone like Einstein becomes so big in the mind’s eye, his shadow makes it difficult to see all the people standing behind him.

This is not meant to take away from the feat that Einstein accomplished.  The General Theory of Relativity, published by Einstein in its full form in 1915 was spectacular [12].  Einstein had taken vague notions about curved spaces and had made them specific, mathematically rigorous and intimately connected with physics through the mass-energy source term in his field equations.  His mathematics had gone beyond even what his mathematician friend and former collaborator Grossmann could achieve.  Yet Einstein’s field equations were nonlinear tensor differential equations in which the warping of space depended on the strength of energy fields, but the configuration of those energy fields depended on the warping of space.  This type of nonlinear equation is difficult to solve in general terms, and Einstein was not immediately aware of how to find the solutions to his own equations.

Therefore, it was no small surprise to him when he received a letter from the Eastern Front from an astronomer he barely knew who had found a solution—a simple solution (see Fig. 7) —to his field equations.  Einstein probably wondered how he could have missed it, but he was generous and forwarded the letter to the Reports of the Prussian Physical Society where it was published in 1916 [13].

Fig. 7  Schwarzschild’s solution of the Einstein Field Equations (Ref. [13] pg. 194).

In the same paper, Schwarzschild used his exact solution to find the exact equation that described the precession of the perihelion of Mercury that Einstein had only calculated approximately. The dynamical equations for Mercury are shown in Fig. 8.

Fig. 8  Explanation for the precession of the perihelion of Mercury ( Ref. [13]  pg. 195)

Schwarzschild’s solution to Einstein’s Field Equation of General Relativity was not a general solution, even for a point mass. He had constants of integration that could have arbitrary values, such as the characteristic length scale that Schwarzschild called “alpha”. It was David Hilbert who later expanded upon Schwarzschild’s work, giving the general solution and naming the characteristic length scale (where the metric diverges) after Schwarzschild. This is where the phrase “Schwarzschild Radius” got its name, and it stuck. In fact it stuck so well that Schwarzschild’s radius has now eclipsed much of the rest of Schwarzschild’s considerable accomplishments.

Unfortunately, Schwarzschild’s accomplishments were cut short when he contracted an autoimmune disease that may have been hereditary. It is ironic that in the carnage of the Eastern Front, it was a genetic disease that caused his death at the age of 42. He was already suffering from the effects of the disease as he worked on his last publications. He was sent home from the front to his family in Potsdam where he passed away several months later having shepherded his final two papers through the publication process. His last paper, on the action-angle variables in quantum systems , was published on the day that he died.

Schwarzschild’s Legacy

Schwarzschild’s legacy was assured when he solved Einstein’s field equations and Einstein communicated it to the world. But his hidden legacy is no less important.

Schwarzschild’s application of the Hamiltonian formalism of canonical transformations and phase space for quantum systems set the stage for the later adoption of Hamiltonian methods in quantum mechanics. He came dangerously close to stating the uncertainty principle that catapulted Heisenberg to later fame, although he could not express it in probabilistic terms because he came too early.

Schwarzschild is considered to be the greatest German astronomer of the last hundred years. This is in part based on his work at the birth of stellar interferometry and in part on his development of stellar photometry and the calibration of the Cepheid variable stars that went on to revolutionize our view of our place in the universe. Solving Einsteins field equations was just a sideline for him, a hobby to occupy his active and curious mind.


Read all the stories behind the history of general relativity, in Galileo Unbound from Oxford Press:


References


[1] Fizeau, H. L. (1849). “Sur une expérience relative à la vitesse de propagation de la lumière.” Comptes rendus de l’Académie des sciences 29: 90–92, 132.

[2] Foucault, J. L. (1862). “Détermination expérimentale de la vitesse de la lumière: parallaxe du Soleil.” Comptes rendus de l’Académie des sciences 55: 501–503, 792–596.

[3] Fizeau, H. (1859). “Sur les hypothèses relatives à l’éther lumineux.” Ann. Chim. Phys.  Ser. 4 57: 385–404.

[4] Fizeau, H. (1868). “Prix Bordin: Rapport sur le concours de l’annee 1867.” C. R. Acad. Sci. 66: 932.

[5] Michelson, A. A. (1890). “I. On the application of interference methods to astronomical measurements.” The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 30(182): 1-21.

[6] Michelson, A. A. (1891). “Measurement of Jupiter’s Satellites by Interference.” Nature 45(1155): 160-161.

[7] Schwarzschild, K. (1896). “Über messung von doppelsternen durch interferenzen.” Astron. Nachr. 3335: 139.

[8] P. Ehrenfest, “Een mechanische theorema van Boltzmann en zijne betrekking tot de quanta theorie (A mechanical theorem of Boltzmann and its relation to the theory of energy quanta),” Verslag van de Gewoge Vergaderingen der Wis-en Natuurkungige Afdeeling, vol. 22, pp. 586-593, 1913.

[9] Schwarzschild, K. (1916). “Quantum hypothesis.” Sitzungsberichte Der Koniglich Preussischen Akademie Der Wissenschaften: 548-568.

[10] P. Ehrenfest, “Adiabatic invariables and quantum theory,” Annalen Der Physik, vol. 51, pp. 327-352, Oct 1916.

[11] Epstein, P. S. (1916). “The quantum theory.” Annalen Der Physik 51(18): 168-188.

[12] Einstein, A. (1915). “On the general theory of relativity.” Sitzungsberichte Der Koniglich Preussischen Akademie Der Wissenschaften: 778-786.

[13] Schwarzschild, K. (1916). “Über das Gravitationsfeld eines Massenpunktes nach der Einstein’schen Theorie.” Sitzungsberichte der Königlich-Preussischen Akademie der Wissenschaften: 189.

Freeman Dyson’s Quantum Odyssey

In the fall semester of 1947, a brilliant young British mathematician arrived at Cornell University to begin a yearlong fellowship paid by the British Commonwealth.  Freeman Dyson (1923 –) had received an undergraduate degree in mathematics from Cambridge University and was considered to be one of their brightest graduates.  With strong recommendations, he arrived to work with Hans Bethe on quantum electrodynamics.  He made rapid progress on a relativistic model of the Lamb shift, inadvertently intimidating many of his fellow graduate students with his mathematical prowess.  On the other hand, someone who intimidated him, was Richard Feynman.

Initially, Dyson considered Feynman to be a bit of a buffoon and slacker, but he started to notice that Feynman could calculate QED problems in a few lines that took him pages.

Freeman Dyson at Princeton in 1972.

I think like most science/geek types, my first introduction to the unfettered mind of Freeman Dyson was through the science fiction novel Ringworld by Larry Niven. The Dyson ring, or Dyson sphere, was conceived by Dyson when he was thinking about the ultimate fate of civilizations and their increasing need for energy. The greatest source of energy on a stellar scale is of course a star, and Dyson envisioned an advanced civilization capturing all that emitted stellar energy by building a solar collector with a radius the size of a planetary orbit. He published the paper “Search for Artificial Stellar Sources of Infra-Red Radiation” in the prestigious magazine Science in 1960. The practicality of such a scheme has to be seriously questioned, but it is a classic example of how easily he thinks outside the box, taking simple principles and extrapolating them to extreme consequences until the box looks like a speck of dust. I got a first-hand chance to see his way of thinking when he gave a physics colloquium at Cornell University in 1980 when I was an undergraduate there. Hans Bethe still had his office at that time in the Newman laboratory. I remember walking by and looking into his office getting a glance of him editing a paper at his desk. The topic of Dyson’s talk was the fate of life in the long-term evolution of the universe. His arguments were so simple they could not be refuted, yet the consequences for the way life would need to evolve in extreme time was unimaginable … it was a bazaar and mind blowing experience for me as an undergrad … and and example of the strange worlds that can be imagined through simple physics principles.

Initially, as Dyson settled into his life at Cornell under Bethe, he considered Feynman to be a bit of a buffoon and slacker, but he started to notice that Feynman could calculate QED problems in a few lines that took him pages.  Dyson paid closer attention to Feynman, eventually spending more of his time with him than Bethe, and realized that Feynman had invented an entirely new way of calculating quantum effects that used cartoons as a form of book keeping to reduce the complexity of many calculations.  Dyson still did not fully understand how Feynman was doing it, but knew that Feynman’s approach was giving all the right answers.  Around that time, he also began to read about Schwinger’s field-theory approach to QED, following Schwinger’s approach as far as he could, but always coming away with the feeling that it was too complicated and required too much math—even for him! 

Road Trip Across America

That summer, Dyson had time to explore America for the first time because Bethe had gone on an extended trip to Europe.  It turned out that Feynman was driving his car to New Mexico to patch things up with an old flame from his Los Alamos days, so Dyson was happy to tag along.  For days, as they drove across the US, they talked about life and physics and QED.  Dyson had Feynman all to himself and began to see daylight in Feynman’s approach, and to understand that it might be consistent with Schwinger’s and Tomonaga’s field theory approach.  After leaving Feynman in New Mexico, he travelled to the University of Michigan where Schwinger gave a short course on QED, and he was able to dig deeper, talking with him frequently between lectures. 

At the end of the summer, it had been arranged that he would spend the second year of his fellowship at the Institute for Advanced Study in Princeton where Oppenheimer was the new head.  As a final lark before beginning that new phase of his studies he spent a week at Berkeley.  The visit there was uneventful, and he did not find the same kind of open camaraderie that he had found with Bethe in the Newman Laboratory at Cornell, but it left him time to think.  And the more he thought about Schwinger and Feynman, the more convinced he became that the two were equivalent.  On the long bus ride back east from Berkeley, as he half dozed and half looked out the window, he had an epiphany.  He saw all at once how to draw the map from one to the other.  What was more, he realized that many of Feynman’s techniques were much simpler than Schwinger’s, which would significantly simplify lengthy calculations.  By the time he arrived in Chicago, he was ready to write it all down, and by the time he arrived in Princeton, he was ready to publish.  It took him only a few weeks to do it, working with an intensity that he had never experienced before.  When he was done, he sent the paper off to the Physical Review[1].

Dyson knew that he had achieved something significant even though he was essentially just a second-year graduate student, at least from the point of view of the American post-graduate system.  Cambridge was a little different, and Dyson’s degree there was more than the standard bachelor’s degree here.  Nonetheless, he was now under the auspices of the Institute for Advanced Study, where Einstein had his office, and he had sent off an unsupervised manuscript for publication without any imprimatur from the powers at be.  The specific power that mattered most was Oppenheimer, who arrived a few days after Dyson had submitted his manuscript.  When he greeted Oppenheimer, he was excited and pleased to hand him a copy.  Oppenheimer, on the other hand, was neither excited nor pleased to receive it.  Oppenheimer had formed a particularly bad opinion of Feynman’s form of QED at the conference held in the Poconos (to read about Feynman’s disaster at the Poconos conference, see my blog) half-a-year earlier and did not think that this brash young grad student could save it.  Dyson, on his part, was taken aback.  No one who has ever met Dyson would ever call him brash, but in this case he fought for a higher cause, writing a bold memo to Oppenheimer—that terrifying giant of a personality—outlining the importance of the Feynman theory.

Battle for the Heart of Quantum Field Theory 

Oppenheimer decided to give Dyson a chance, and arranged for a series of seminars where Dyson could present the story to the assembled theory group at the Institute, but Dyson could make little headway.  Every time he began to make progress, Oppenheimer would bring it crashing to a halt with scathing questions and criticisms.  This went on for weeks, until Bethe visited from Cornell.  Bethe by then was working with the Feynman formalism himself.  As Bethe lectured in front of Oppenheimer, he seeded his talk with statements such as “surely they had all seen this from Dyson”, and Dyson took the opportunity to pipe up that he had not been allowed to get that far.  After Bethe left, Oppenheimer relented, arranging for Dyson to give three seminars in one week.  The seminars each went on for hours, but finally Dyson got to the end of it.  The audience shuffled out of the seminar room with no energy left for discussions or arguments.  Later that day, Dyson found a note in his box from Oppenheimer saying “Nolo Contendre”—Dyson had won!

With that victory under his belt, Dyson was in a position to communicate the new methods to a small army of postdocs at the Institute, supervising their progress on many outstanding problems in quantum electrodynamics that had resisted calculations using the complicated Schwinger-Tomonaga theory.  Feynman, by this time, had finally published two substantial papers on his approach[2], which added to the foundation that Dyson was building at Princeton.  Although Feynman continued to work for a year or two on QED problems, the center of gravity for these problems shifted solidly to the Institute for Advanced Study and to Dyson.  The army of postdocs that Dyson supervised helped establish the use of Feynman diagrams in QED, calculating ever higher-order corrections to electromagnetic interactions.  These same postdocs were among the first batch of wartime-trained theorists to move into faculty positions across the US, bringing the method of Feynman diagrams with them, adding to the rapid dissemination of Feynman diagrams into many aspects of theoretical physics that extend far beyond QED [3].

As a graduate student at Berkeley in the 1980’s I ran across a very simple-looking equation called “the Dyson equation” in our graduate textbook on relativistic quantum mechanics by Sakurai. The Dyson equation is the extraordinarily simple expression of an infinite series of Feynman diagrams that describes how an electron interacts with itself through the emission of virtual photons that link to virtual electron-positron pairs. This process leads to the propagator Green’s function for the electron and is the starting point for including the simple electron in more complex particle interactions.

The Dyson equation for the single-electron Green’s function represented as an infinite series of Feynman diagrams.

I had no feel for the use of the Dyson equation, barely limping through relativistic quantum mechanics, until a few years later when I was working at Lawrence Berkeley Lab with Mirek Hamera, a visiting scientist from Warwaw Poland who introduced me to the Haldane-Anderson model that applied to a project I was working on for my PhD. Using the theory, with Dyson’s equation at its heart, we were able to show that tightly bound electrons on transition-metal impurities in semiconductors acted as internal reference levels that allowed us to measure internal properties of semiconductors that had never been accessible before. A few years later, I used Dyson’s equation again when I was working on small precipitates of arsenic in the semiconductor GaAs, using the theory to describe an accordion-like ladder of electron states that can occur within the semiconductor bandgap when a nano-sphere takes on multiple charges [4].

The Coulomb ladder of deep energy states of a nano-sphere in GaAs calculated using self-energy principles first studied by Dyson.

I last saw Dyson when he gave the Hubert James Memorial Lecture at Purdue University in 1996. The title of his talk was “How the Dinosaurs Might Have Been Saved: Detection and Deflection of Earth-Impacting Bodies”. As always, his talk was wild and wide ranging, using the simplest possible physics to derive the most dire consequences of our continued existence on this planet.


[1] Dyson, F. J. (1949). “THE RADIATION THEORIES OF TOMONAGA, SCHWINGER, AND FEYNMAN.” Physical Review 75(3): 486-502.

[2] Feynman, R. P. (1949). “THE THEORY OF POSITRONS.” Physical Review 76(6): 749-759.  Feynman, R. P. (1949). “SPACE-TIME APPROACH TO QUANTUM ELECTRODYNAMICS.” Physical Review 76(6): 769-789.

[3] Kaiser, D., K. Ito and K. Hall (2004). “Spreading the tools of theory: Feynman diagrams in the USA, Japan, and the Soviet Union.” Social Studies of Science 34(6): 879-922.

[4] Nolte, D. D. (1998). “Mesoscopic Point-like Defects in Semiconductors.” Phys. Rev. B58(12): pg. 7994

Physicists in Revolution: Arago, Riemann, Jacobi and Doppler

The opening episode of Victoria on Masterpiece Theatre (PBS) this season finds the queen confronting widespread unrest among her subjects who are pressing for more freedoms and more say in government. Louis-Phillipe, former King of France, has been deposed in the February Revolution of 1848 in Paris and his presence at the Royal Palace does not help the situation.

In 1848 a wave of spontaneous revolution swept across Europe.  It was not a single revolution of many parts, but many separate revolutions with similar goals.  Two essential disruptions of life occurred in the early 1800’s.  The first was the partitioning of Europe at the Congress of Vienna from 1814 to 1815, presided over by Prince Metternich of Austria, that had carved up Napoleon’s conquests and sought to establish a stable order based on the old ideal of absolute monarchy.  In the process, nationalities were separated or suppressed.  The second was the industrialization of Europe in the early 1800’s that created economic upheaval, with masses of working poor fleeing effective serfdom in the fields and flocking to the cities.  Wages fell, food became scarce, legions of the poor and starving bloomed.  Because of these influences, European society had become unstable, supercooled beyond a phase transition and waiting for a seed or catalyst to crystalize the continent into a new state of matter. 

When the wave came, physicists across Europe were caught in the upheaval.  Some were caught up in the fervor and turned their attention to national service, some lost their standing and their positions during the inevitable reactionary backlash, others got the opportunities of their careers.  It was difficult for anyone to be untouched by the 1848 revolutions, and physicist were no exception.

The Spontaneous Fire of Revolution

The extraodinary wave of revolution was sparked by a small rebellion in Sicily in January 1848 that sought to overturn the ruling Bourbons.  It was a small rebellion of little direct consequence to Europe, but it succeeded in establishing a liberal democracy in an independent state that stood as a symbol of what could be achieved by a determined populace.  The people of Paris took notice, and in the sudden and unanticipated February Revolution, the French constitutional monarchy under Louis-Phillipe was overthrown in a few days and replaced by the French Second Republic.  The shock of Louis-Phillipe’s fall reverberated across Europe, feared by those in power and welcomed by those who sought a new world order.  Nationalism, liberalism, socialism and communism were on the rise, and the opportunity to change the world seemed to have arrived.  The Five Days of Milan in Italy, the March Revolution of the German states, the Polish rebellion against Prussia, and the Young Irelander Rebellion in Ireland were all consequences of the unstable conditions and the unprecidented opportunities for the people to enact change.  None of these uprisings were coordinated by any central group.  It was a spontaneous consequence of similar preconditions that existed across nearly all the states of Europe.

Arago and the February Revolution in Paris

The French were no newcomers to street rebellions.  Paris had a history of armed conflict between citizens manning barricades and the superior forces of the powers at be.  The unforgettable scene in Les Misérables of Marius at the barricade and Jean Valjean’s rescue through the sewers of Paris was based on the 1832 June Rebellion in Paris.  Yet this event was merely an echo of the much larger rebellion of 1830 that had toppled the unpopular monarchy of Charles X, followed by the ascension of the Bourgeois Monarch Louis Phillipe at the start of the July Monarchy.  Eighteen years later, Louis Phillipe was still on the throne and the masses were ready again for a change.  Alexis de Tocqueville saw the change coming and remarked, “We are sleeping together in a volcano. … A wind of revolution blows, the storm is on the horizon.”  The storm would sweep up a generation of participants, including the French physicist Francois Arago (1786 – 1853).

Lamartine in front of the Town Hall of Paris on 25 February 1848 (Image by Henri Félix Emmanuel Philippoteaux in public domain).

Arago is one of the under-appreciated French physicists of the 1800’s.  This may be because so many of his peers have become icons in the history of physics: Fourier, Fresnel, Poisson, Laplace, Malus, Biot and Foucault.  The one place where his name appears—the Spot of Arago—was not exclusively his discovery, but rather was an experimental demonstration of an effect derived by Poisson using Fresnel’s new theory of diffraction.  Poisson derived the phenomenon as a means to show the absurdity of Fresnel’s undulatory theory of light, but Arago’s experimental demonstration turned the tables on Poisson and the emissionists (followers of Newton’s particulate theory of light).  Yet Arago played a role behind the scenes as a supporter and motivator of some of the most important discoveries in optics.  In particular, it was Arago’s encouragement and support of the (at that time) unknown Fresnel, that helped establish the Fresnel theory of diffraction and the wave nature of light.  Together, Arago and Fresnel established the transverse nature of the light wave, and Arago is also the little-known discoverer of optical rotation.  As a young scientist, he attempted to measure the drift of the ether, which was a null experiment that foreshadowed the epochal experiments of Michelson and Morley 80 years later.  In his later years, Arago proposed the methodology for measuring the speed of light in both stationary and moving materials, which became the basis for the important measurements of the speed of light by Fizeau and Foucault (who also attempted to measure ether drift).

In addition to his duties as the director of the National Observatory and as the perpetual secretary of the Academie des Sciences (replacing Fourier), he entered politics in 1830 when he was elected as a member of the chamber of deputies.  At the fall of Louis-Phillipe in the February Revolution of 1848, he was appointed as a member of the steering committee of the newly formed government of the French Second Republic, and he was named head of the Marine and Colonies as well as the head of the Department of War.  Although he was a staunch republican and supporter of the people, his position put him in direct conflict with the later stages of the revolutions of 1848. 

The population of Paris became disenchanted with the conservative trends in the Second Republic.  In June of 1848 barricades were again erected in the streets of Paris, this time in opposition to the Republic.  Forces were drawn up on both sides, although many of the Republican forces defected to the insurgents, and attempts were made to mediate the conflict.  At the barricade on the rue Soufflot near the Pantheon, Arago himself approached the barricades to implore defenders to disperse.  It is a measure of the respect Arago held with the people when they replied, “Monsieur Arago, we are full of respect for you, but you have no right to reproach us.  You have never been hungry.  You don’t know what poverty is.” [1] When Arago finally withdrew, he feared that death and carnage were inevitable.  They came at noon on June 23 when the barricade at Porte Saint-Denis was attacked by the National Guards.  This started a general onslaught of all the barricades by Republican forces that left 1,500 workers dead in the streets and more than 11,000 arrested.  Arago resigned from the steering committee on June 24, although he continued to work in the government until the coup d’Etat by Louis Napolean, the nephew of Napoleon Bonaparte, in 1852 when he became Napoleon III, Emperor of the Second French Empire. Louis Napoleon demanded that all government workers take an oath of allegiance to him, but Arago refused.  Yet such was the respect that Arago commanded that Louis Napoleon let him continue unmolested as the astronomer of the Bureau des Longitudes.

Riemann and Jacobi and the March Revolution in Berlin

The February Revolution of Paris was followed a month later by the March Revolutions of the German States.  The center of the German-speaking world at that time was Vienna, and a demonstration by students broke out in Vienna on March 13. Emperor Ferdinand, following the advice of Metternich, called out the army who fired on the crowd, killing several protestors.  Throngs rallied to the protest and arms were distributed, readying for a fight.  Rather than risk unreserved bloodshed, the emperor dismissed Metternich who went into exile to London (following closely the footsteps of the French Louis-Phillipe).  Within the week, the revolutionary fervor had spread to Berlin where a student uprising marched on the royal palace of King Frederick Wilhelm IV on March 18.  They were met by 20,000 troops. 

The March 1848 revolution in Berlin (Image in the public domain).

Not all university students were liberals and revolutionaries, and there were numerous student groups that formed to support the King.  One of the students in one of these loyalist groups was a shy mathematician who joined a loyalist student militia to protect the King.  Bernhard Riemann (1826 – 1866) had come to the University of Berlin after spending a short time in the Mathematics department at the University in Göttingen.  Despite the presence of Gauss there, the mathematics department was not considered strong (this would change dramatically in about 50 years when Göttingen became the center of German mathematics with the arrival of Felix Klein, Karl Schwarzschild and Hermann Minkowski).  At Berlin, Riemann attended lectures by Steiner, Jacobi, Dirichlet and Eisenstein. 

On the night of the uprising, a nervous Riemann found himself among a group of students, few more than 20 years old, guarding the quarters of the King, not knowing what would unfold.  They spent a sleepless night that dawned on the chaos and carnage at the barricades at Alexander Platz with hundreds of citizens dead.  King Wilhelm was caught off guard by the events, and he assured the citizens that he would reorganize the government and yield to the demonstrator’s demands for parliamentary elections, a constitution, and freedom of the press.  Two days later the king attended a mass funeral for the fallen, attended by his generals and ministers who wore the german revolutionary tricolor of black, red and gold.  This ploy worked, and the unrest in Berlin died away before the king was forced to abdicate.  This must have relieved Riemann immensely, because this entire episode was entirely outside his usual meek and mild character.  Yet the character of all the unrelated 1848 revolutions had one thing in common: a sharp division among the populace between the liberals and the conservatives.  As Riemann had elected to join with the loyalists, one of his professors picked the other side.

Carl Gustav Jacob Jacobi (1804 – 1851) had been born in Potsdam and had obtained his first faculty position at the University of Königsberg where he was soon ranked among the top mathematicians in Europe.  However, in his early thirties he was stricken with diabetes, and the harsh winters of Königsberg became to difficult to bear.  He returned to the milder climate of Berlin to a faculty position at the university when the wave of revolution swept over the city.  Jacobi was a liberal thinker and was caught up in the movement, attending meetings at the Constitution Club.  Once the danger to Wilhelm IV had passed, the reactionary forces took their revenge, and Jacobi’s teaching stipend was suspended.  When he threatened to move to the University of Vienna, the royalists relented, so Jacobi too was able to weather the storm. 

The surprising footnote to this story is that Jacobi delivered lectures on a course on the application of differential equations to mechanics in the winter semester of 1847 – 1848 right in the midst of the political turmoil.  His participation in the extraordinary political events of that time apparently did not hamper him from giving one of the most extraordinary sets of lectures in mathematical physics.  Jacobi’s lectures of 1848 were the greatest advance in mathematical physics since Euler had reinterpreted Newton a hundred years earlier.  This is where Jacobi expanded on the work of Hamilton, establishing what is today called the Hamilton-Jacobi theory of dynamics.  He also derived and proved, using Liouville’s theorem of 1838, that the volume of phase space was an invariant in a conservative dynamical system [2].  It is tempting to imagine Jacobi returning home late at night, after rousing discussions of revolution at the Constitution Club, to set to work on his own revolutionary theories in physics.

Doppler and the Hungarian Revolution

Among all the states of Europe, the revolutions of 1848 posed the greatest threat to the Austrian Empire, which was a beaurocratic state entangling scores of diverse nationalities sprawled across the largest state of Europe.  The Austrian Empire was the remnant of the Holy Roman Empire that had succumbed to the Napoleonic invasion.  The lands that were controlled by Austria, after Metternich engineered the Congress of Vienna, included Poles, Ukranians, Romanians, Germans, Czechs, Slovaks, Hungarians, Slovenes, Serbs, Albanians and more.  Holding this diverse array of peoples together was already a challenge, and the revolutions of 1848 carried with them strong feelings of nationalism.  The revolutions spreading across Europe were the perfect catalyst to set off the Hungarian Revolution that grew into a war for independence, and the fierce fighting across Hungary could not be avoided even by cloistered physicists.

Christian Doppler (1803 – 1853) had moved in 1847 from Prague (where he had proposed what came to be called the Doppler effect in 1842 to the Royal Bohemian Society of Sciences) to the Academy of Mines and Forests in Schemnitz (modern Banská Štiavnica in Slovakia, but then part of the Kingdom of Hungary) with more pay and less work.  His health had been failing, and the strenuous duties at Prague had taken their toll.  If the goal of this move to an obscure school far from the center of Austrian power had been to lead a peaceful life, Doppler’s plans were sorely upset.

The news of the protests in Vienna arrived in Schemnitz on the 17th of March, and student demonstrations commenced immediately.  Amidst the uncertainty, Doppler requested a leave of absence from the summer semester and returned to Vienna.  It is not clear why he went there, whether to be near the center of excitement, or to take advantage of the free time to pursue his own researches.  While in Vienna he read a treatise before the Academy on galvano-electric effects.  He returned to Schemnitz in the Fall to relative peace, until the 12th of December, when the Hungarians rejected to acknowledge the new Emperor Franz Josef in Vienna, replacing his Uncle Ferdinand who was forced to abdicate, and the Hungarian war for independence began.

Görgey’s troops crossing the Sturec pass. Their ability to evade the Austrian pursuit was legendary (Image by Keiss Károly in the public domain).

One of Doppler’s former students from his days in Prague was appointed to command the newly formed Hungarian army.  General Arthur Görgey (1818 – 1916) moved to take possession of the northern mining towns (present day Slovakia) and occupied Schemnitz.  When Görgey learned that his old teacher was in the town he sent word to Doppler to meet him at his headquarters.  Meeting with a revolutionary and rebel could have marked Doppler as a traitor in Vienna, but he decided to meet him anyway, taking along one of his colleagues as a “witness” that the discussion were purely academic.  This meeting opens an interesting unsolved question in the history of physics. 

Around this time Doppler was interested in the dynamical properties of the pendulum for cases when the suspension wire was exceptionally long.  Experiments on such extreme pendula could provide insight into changes in gravity with height as well as the effects of the motion of the Earth.  For instance, Coriolis had published his paper on forces in rotating frames many years earlier in 1835.  Because Schemnitz was a mining town, there was ample access to deep mine shafts in which to set up a pendulum with a very long wire.  This is where the story becomes murky.  Within the family of Doppler’s descendants there are stories of Doppler setting up such an experiment, and even a night time visit to the Doppler house by Görgey.  The pendulum was thought to be one of the topics discussed by Doppler and Görgey at their first meeting, and Görgey (from his life as a scientist prior to becoming a revolution general) had arrived to help with the experiment [3]

This story is significant for two reasons.  First, it would be astounding to think of General Görgey taking a break from the revolution to do some physics for fun.  Görgey has not been graced by history with a benevolent reputation.  He was known as a hard and sometimes vicious leader, and towards the end of the short-lived Hungarian Revolution he displaced the President Kossuth to become the dictator of Hungary.  The second reason, which is important for the history of physics, is that if Doppler had performed this experiment in 1848, it would have preceded the famous experiment by Foucault by more than two years.  However, the paper published by Doppler around this time on the dynamics of the pendulum did not mention the experiment, and it remains an open question in the history of physics whether Doppler may have had priority over Foucault.

The Austrian Imperial Army laid siege to Schemnitz and commenced a short bombardment that displaced Görgey and his troops from the town.  Even as Schemnitz was being liberated, a letter arrived informing Doppler that his old mentor Stampfer at the University of Vienna was retiring and that he had been chosen to be his replacement.  The March Revolution had led to the abdication of the previous Austrian emperor and his replacement by the more liberal-minded Franz Josef who was interested in restructuring the educational system in the Austrian empire.  On the advice of Doppler’s supporters who were in the new government, the Institute of Physics was formed and Doppler was named as its first director.  He arrived in the spring of 1850 to take up his new post.

The Legacy of 1848

Despite the early successes and optimism of the revolutions of 1848, reactionary forces were quick to reverse many of the advances made for universal suffrage, constitutional government, freedom of the press, and freedom of expression.  In most cases, monarchs either retained power or soon returned.  Even the reviled Metternich returned to Vienna from exile in London in 1851.  Yet as is so often the case, once a door has been opened it is difficult to shut it again.  The pressure for reforms continued long after the revolutions faded away, and by 1870 many of the specific demands of the people had been instituted by most of the European states.  Russia was an exception, which may explain why the inevitable Russian Revolution half a century later was so severe.            

The revolutions of 1848 cannot be said to have had a long-lasting impact on the progress of physics, although they certainly had a direct impact on the lives of selected physicists.  The most lasting effect of the revolutions on science was the restructuring of educational systems, not only in Austria, but in many of the European states.  This was perhaps one of the first times when the social and economic benefits of science education to the national welfare was understood and implemented across Europe, although a similar recognition had occurred earlier during the French Revolution, for instance leading to the founding of the Ecole Polytechnique.  The most important, though subtle, effect of the revolutions of 1848 on society was the shift away from autocratic rule to democracy, and the freeing of expression and thought from rigid bounds.  The coming revolution in physics at the turn of the next century may have been helped a little by the revolutionary spirit that still echoed from 1848.


[1] pg. 201, Mike Rapport, “1848: Year of Revolution” (Basic Books, 2008)

[2] D. D. Nolte, The Tangled Tale of Phase Space, Chap. 6 in Galileo Unbound (Oxford University Press, 2018)

[3] Schuster, P. Moving the stars : Christian Doppler, his life, his works and principle, and the world after. Pöllauberg, Austria, Living Edition. (2005)



George Green’s Theorem

For a thirty-year old miller’s son with only one year of formal education, George Green had a strange hobby—he read papers in mathematics journals, mostly from France.  This was his escape from a dreary life running a flour mill on the outskirts of Nottingham, England, in 1823.  The tall wind mill owned by his father required 24-hour attention, with farmers depositing their grain at all hours and the mechanisms and sails needing constant upkeep.  During his one year in school when he was eight years old he had become fascinated by maths, and he nurtured this interest after leaving school one year later, stealing away to the top floor of the mill to pore over books he scavenged, devouring and exhausting all that English mathematics had to offer.  By the time he was thirty, his father’s business had become highly successful, providing George with enough wages to become a paying member of the private Nottingham Subscription Library with access to the Transactions of the Royal Society as well to foreign journals.  This simple event changed his life and changed the larger world of mathematics.

Green’s windmill in Sneinton, England.

French Analysis in England

George Green was born in Nottinghamshire, England.  No record of his birth exists, but he was baptized in 1793, which may be assumed to be the year of his birth.  His father was a baker in Nottingham, but the food riots of 1800 forced him to move outside of the city to the town of Sneinton, where he bought a house and built an industrial-scale windmill to grind flour for his business.  He prospered enough to send his eight-year old son to Robert Goodacre’s Academy located on Upper Parliament Street in Nottingham.  Green was exceptionally bright, and after one year in school he had absorbed most of what the Academy could teach him, including a smattering of Latin and Greek as well as French along with what simple math that was offered.  Once he was nine, his schooling was over, and he took up the responsibility of helping his father run the mill, which he did faithfully, though unenthusiastically, for the next 20 years.  As the milling business expanded, his father hired a mill manager that took part of the burden off George.  The manager had a daughter Jane Smith, and in 1824 she had her first child with Green.  Six more children were born to the couple over the following fifteen years, though they never married.

Without adopting any microscopic picture of how electric or magnetic fields are produced or how they are transmitted through space, Green could still derive rigorous properties that are independent of any details of the microscopic model.

            During the 20 years after leaving Goodacre’s Academy, Green never gave up learning what he could, teaching himself to read French readily as well as mastering English mathematics.  The 1700’s and early 1800’s had been a relatively stagnant period for English mathematics.  After the priority dispute between Newton and Leibniz over the invention of the calculus, English mathematics had become isolated from continental advances.  This was part snobbery, but also part handicap as the English school struggled with Newton’s awkward fluxions while the continental mathematicians worked with Leibniz’ more fruitful differential notation.  One notable exception was Brook Taylor who developed the Taylor’s Series (and who grew up on the opposite end of the economic spectrum from Green, see my Blog on Taylor). However, the French mathematicians in the early 1800’s were especially productive, including such works as those by Lagrange, Laplace and Poisson.

            One block away from where Green lived stood the Free Grammar School overseen by headmaster John Topolis.  Topolis was a Cambridge graduate on a minor mission to update the teaching of mathematics in England, well aware that the advances on the continent were passing England by.  For instance, Topolis translated Laplace’s mathematically advanced Méchaniqe Celéste from French into English.  Topolis was also well aware of the work by the other French mathematicians and maintained an active scholarly output that eventually brought him back to Cambridge as Dean of Queen’s College in 1819 when Green was 26 years old.  There is no record whether Topolis and Green knew each other, but their close proximity and common interests point to a natural acquaintance.  One can speculate that Green may even have sought Topolis out, given his insatiable desire to learn more mathematics, and it is likely that Topolis would have introduced Green to the vibrant French school of mathematics.             

By the time Green joined the Nottingham Subscription Library, he must already have been well trained in basic mathematics, and membership in the library allowed him to request loans of foreign journals (sort of like Interlibrary Loan today).  With his library membership beginning in 1823, Green absorbed the latest advances in differential equations and must have begun forming a new viewpoint of the uses of mathematics in the physical sciences.  This was around the same time that he was beginning his family with Jane as well as continuing to run his fathers mill, so his mathematical hobby was relegated to the dark hours of the night.  Nonetheless, he made steady progress over the next five years as his ideas took rough shape and were refined until finally he took pen to paper, and this uneducated miller’s son began a masterpiece that would change the history of mathematics.

Essay on Mathematical Analysis of Electricity and Magnetism

By 1827 Green’s free-time hobby was about to bear fruit, and he took out a modest advertisement to announce its forthcoming publication.  Because he was an unknown, and unknown to any of the local academics (Topolis had already gone back to Cambridge), he chose vanity publishing and published out of pocket.   An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism was printed in March of 1828, and there were 51 subscribers, mostly from among the members of the Nottingham Subscription Library who bought it at 7 shillings and 6 pence per copy, probably out of curiosity or sympathy rather than interest.  Few, if any, could have recognized that Green’s little essay contained several revolutionary elements.

Fig. 1 Cover page of George Green’s Essay

            The topic of the essay was not remarkable, treating mathematical problems of electricity and magnetism, which was in vogue at that time.  As background, he had read works by Cavendish, Poisson, Arago, Laplace, Fourier, Cauchy and Thomas Young (probably Young’s Course of Lectures on Natural Philosopy and the Mechanical Arts (1807)).  He paid close attention to Laplace’s treatment of celestial mechanics and gravitation which had obvious strong analogs to electrostatics and the Coulomb force because of the common inverse square dependence. 

            One radical contribution in Green’s essay was his introduction of the potential function—one of the first uses of the concept of a potential function in mathematical physics—and he gave it its modern name.  Others had used similar constructions, such as Euler [1], D’Alembert [2], Laplace[3] and Poisson [4], but the use had been implicit rather than explicit.  Green shifted the potential function to the forefront, as a central concept from which one could derive other phenomena.  Another radical contribution from Green was his use of the divergence theorem.  This has tremendous utility, because it relates a volume integral to a surface integral.  It was one of the first examples of how measuring something over a closed surface could determine a property contained within the enclosed volume.  Gauss’ law is the most common example of this, where measuring the electric flux through a closed surface determines the amount of enclosed charge.  Lagrange in 1762 [5] and Gauss in 1813 [6] had used forms of the divergence theorem in the context of gravitation, but Green applied it to electrostatics where it has become known as Gauss’ law and is one of the four Maxwell equations.  Yet another contribution was Green’s use of linear superposition to determine the potential of a continuous charge distribution, integrating the potential of a point charge over a continuous charge distribution.  This was equivalent to defining what is today called a Green’s function, which is a common method to solve partial differential equations.

            A subtle contribution of Green’s Essay, but no less influential, was his adoption of a mathematical approach to a physics problem based on the fundamental properties of the mathematical structure rather than on any underlying physical model.  Without adopting any microscopic picture of how electric or magnetic fields are produced or how they are transmitted through space, he could still derive rigorous properties that are independent of any details of the microscopic model.  For instance, the inverse square law of both electrostatics and gravitation is a fundamental property of the divergence theorem (a mathematical theorem) in three-dimensional space.  There is no need to consider what space is composed of, such as the many differing models of the ether that were being proposed around that time.  He would apply this same fundamental mathematical approach in his later career as a Cambridge mathematician to explain the laws of reflection and refraction of light.

George Green: Cambridge Mathematician

A year after the publication of the Essay, Green’s father died a wealthy man, his milling business having become very successful.  Green inherited the family fortune, and he was finally able to leave the mill and begin devoting his energy to mathematics.  Around the same time he began working on mathematical problems with the support of Sir Edward Bromhead.  Bromhead was a Nottingham peer who had been one of the 51 subscribers to Green’s published Essay.  As a graduate of Cambridge he was friends with Herschel, Babbage and Peacock, and he recognized the mathematical genius in this self-educated miller’s son.  The two men spent two years working together on a pair of publications, after which Bromhead used his influence to open doors at Cambridge.

            In 1832, at the age of 40, George Green enrolled as an undergraduate student in Gonville and Caius College at Cambridge.  Despite his concerns over his lack of preparation, he won the first-year mathematics prize.  In 1838 he graduated as fourth wrangler only two positions behind the future famous mathematician James Joseph Sylvester (1814 – 1897).  Based on his work he was elected as a fellow of the Cambridge Philosophical Society in 1840.  Green had finally become what he had dreamed of being for his entire life—a professional mathematician.

            Green’s later papers continued the analytical dynamics trend he had established in his Essay by applying mathematical principles to the reflection and refraction of light. Cauchy had built microscopic models of the vibrating ether to explain and derive the Fresnel reflection and transmission coefficients, attempting to understand the structure of ether.  But Green developed a mathematical theory that was independent of microscopic models of the ether.  He believed that microscopic models could shift and change as newer models refined the details of older ones.  If a theory depended on the microscopic interactions among the model constituents, then it too would need to change with the times.  By developing a theory based on analytical dynamics, founded on fundamental principles such as minimization principles and geometry, then one could construct a theory that could stand the test of time, even as the microscopic understanding changed.  This approach to mathematical physics was prescient, foreshadowing the geometrization of physics in the late 1800’s that would lead ultimately to Einsteins theory of General Relativity.

Green’s Theorem and Greens Function

Green died in 1841 at the age of 49, and his Essay was mostly forgotten.  Ten years later a young William Thomson (later Lord Kelvin) was graduating from Cambridge and about to travel to Paris to meet with the leading mathematicians of the age.  As he was preparing for the trip, he stumbled across a mention of Green’s Essay but could find no copy in the Cambridge archives.  Fortunately, one of the professors had a copy that he lent Thomson.  When Thomson showed the work to Liouville and Sturm it caused a sensation, and Thomson later had the Essay republished in Crelle’s journal, finally bringing the work and Green’s name into the mainstream.

            In physics and mathematics it is common to name theorems or laws in honor of a leading figure, even if the they had little to do with the exact form of the theorem.  This sometimes has the effect of obscuring the historical origins of the theorem.  A classic example of this is the naming of Liouville’s theorem on the conservation of phase space volume after Liouville, who never knew of phase space, but who had published a small theorem in pure mathematics in 1838, unrelated to mechanics, that inspired Jacobi and later Boltzmann to derive the form of Liouville’s theorem that we use today.  The same is true of Green’s Theorem and Green’s Function.  The form of the theorem known as Green’s theorem was first presented by Cauchy [7] in 1846 and later proved by Riemann [8] in 1851.  The equation is named in honor of Green who was one of the early mathematicians to show how to relate an integral of a function over one manifold to an integral of the same function over a manifold whose dimension differed by one.  This property is a consequence of the Generalized Stokes Theorem (named after George Stokes), of which the Kelvin-Stokes Theorem, the Divergence Theorem and Green’s Theorem are special cases.

Fig. 2 Green’s theorem and its relationship with the Kelvin-Stokes theorem, the Divergence theorem and the Generalized Stokes theorem (expressed in differential forms)

            Similarly, the use of Green’s function for the solution of partial differential equations was inspired by Green’s use of the superposition of point potentials integrated over a continuous charge distribution.  The Green’s function came into more general use in the late 1800’s and entered the mainstream of physics in the mid 1900’s [9].

Fig. 3 The application of Green’s function so solve a linear operator problem, and an example applied to Poisson’s equation.

By David D. Nolte, Dec. 26, 2018


[1] L. Euler, Novi Commentarii Acad. Sci. Petropolitanae , 6 (1761)

[2] J. d’Alembert, “Opuscules mathématiques” , 1 , Paris (1761)

[3] P.S. Laplace, Hist. Acad. Sci. Paris (1782)

[4] S.D. Poisson, “Remarques sur une équation qui se présente dans la théorie des attractions des sphéroïdes” Nouveau Bull. Soc. Philomathique de Paris , 3 (1813) pp. 388–392

[5] Lagrange (1762) “Nouvelles recherches sur la nature et la propagation du son” (New researches on the nature and propagation of sound), Miscellanea Taurinensia (also known as: Mélanges de Turin ), 2: 11 – 172

[6] C. F. Gauss (1813) “Theoria attractionis corporum sphaeroidicorum ellipticorum homogeneorum methodo nova tractata,” Commentationes societatis regiae scientiarium Gottingensis recentiores, 2: 355–378

[7] Augustin Cauchy: A. Cauchy (1846) “Sur les intégrales qui s’étendent à tous les points d’une courbe fermée” (On integrals that extend over all of the points of a closed curve), Comptes rendus, 23: 251–255.

[8] Bernhard Riemann (1851) Grundlagen für eine allgemeine Theorie der Functionen einer veränderlichen complexen Grösse (Basis for a general theory of functions of a variable complex quantity), (Göttingen, (Germany): Adalbert Rente, 1867

[9] Schwinger, Julian (1993). “The Greening of quantum Field Theory: George and I”: 10283. arXiv:hep-ph/9310283

Wave-Particle Duality and Hamilton’s Physics

Wave-particle duality was one of the greatest early challenges to quantum physics, partially clarified by Bohr’s Principle of Complementarity, but never easily grasped even today.  Yet long before Einstein proposed the indivisible quantum  of light (later to be called the photon by the chemist Gilbert Lewis), wave-particle duality was firmly embedded in the foundations of the classical physics of mechanics.

Light led the way to mechanics more than once in the history of physics.

 

Willebrord Snel van Royen

The Dutch physicist Willebrord Snel van Royen in 1621 derived an accurate mathematical description of the refraction of beams of light at a material interface in terms of sine functions, but he did not publish.  Fifteen years later, as Descartes was looking for an example to illustrate his new method of analytic geometry, he discovered the same law, unaware of Snel’s prior work.  In France the law is known as the Law of Descartes.  In the Netherlands (and much of the rest of the world) it is known as Snell’s Law.  Both Snell and Descartes based their work on Newton’s corpuscles of light.  The brilliant Fermat adopted corpuscles when he developed his principle of least time to explain the law of Descartes in 1662.  Yet Fermat was forced to assume that the corpuscles traveled slower in the denser material even though it was generally accepted that light should travel faster in denser media, just as sound did.  Seventy-five years later, Maupertuis continued the tradition when he developed his principle of least action and applied it to light corpuscles traveling faster through denser media, just as Descartes had prescribed.

HuygensParticle-02

The wave view of Snell’s Law (on the left). The source resides in the medium with higher speed. As the wave fronts impinge on the interface to a medium with lower speed, the wave fronts in the slower medium flatten out, causing the ray perpendicular to the wave fronts to tilt downwards. The particle view of Snell’s Law (on the right). The momentum of the particle in the second medium is larger than in the first, but the transverse components of the momentum (the x-components) are conserved, causing a tilt downwards of the particle’s direction as it crosses the interface. [i]

Maupertuis’ paper applying the principle of least action to the law of Descartes was a critical juncture in the development of dynamics.  His assumption of faster speeds in denser material was wrong, but he got the right answer because of the way he defined action for light.  Encouraged by the success of his (incorrect) theory, Maupertuis extended the principle of least action to mechanical systems, and this time used the right theory to get the right answers.  Despite Maupertuis’ misguided aspirations to become a physicist of equal stature to Newton, he was no mathematician, and he welcomed (and  somewhat appropriated) the contributions of Leonid Euler on the topic, who established the mathematical foundations for the principle of least action.  This work, in turn, attracted the attention of the Italian mathematician Lagrange, who developed a general new approach (Lagrangian mechanics) to mechanical systems that included the principle of least action as a direct consequence of his equations of motion.  This was the first time that light led the way to classical mechanics.  A hundred years after Maupertuis, it was time again for light to lead to the way to a deeper mechanics known as Hamiltonian mechanics.

Young Hamilton

William Rowland Hamilton (1805—1865) was a prodigy as a boy who knew parts of thirteen languages by the time he was thirteen years old. These were Greek, Latin, Hebrew, Syriac, Persian, Arabic, Sanskrit, Hindoostanee, Malay, French, Italian, Spanish, and German. In 1823 he entered Trinity College of Dublin University to study science. In his second and third years, he won the University’s top prizes for Greek and for mathematical physics, a run which may have extended to his fourth year—but he was offered the position of Andrew’s Professor of Astronomy at Dublin and Royal Astronomer of Ireland—not to be turned down at the early age of 21.

Hamilton1

Title of Hamilton’s first paper on his characteristic function as a new method that applied his theory from optics to the theory of mechanics, including Lagrangian mechanics as a special case.

His research into mathematical physics  concentrated on the theory of rays of light. Augustin-Jean Fresnel (1788—1827) had recently passed away, leaving behind a wave theory of light that provided a starting point for many effects in optical science, but which lacked broader generality. Hamilton developed a rigorous mathematical framework that could be applied to optical phenomena of the most general nature. This led to his theory of the Characteristic Function, based on principles of the variational calculus of Euler and Lagrange, that predicted the refraction of rays of light, like trajectories, as they passed through different media or across boundaries. In 1832 Hamilton predicted a phenomenon called conical refraction, which would cause a single ray of light entering a biaxial crystal to refract into a luminous cone.

Mathematical physics of that day typically followed experimental science. There were so many observed phenomena in so many fields that demanded explanation, that the general task of the mathematical physicist was to explain phenomena using basic principles followed by mathematical analysis. It was rare for the process to work the other way, for a theorist to predict a phenomenon never before observed. Today we take this as very normal. Einstein’s fame was primed by his prediction of the bending of light by gravity—but only after the observation of the effect by Eddington four years later was Einstein thrust onto the world stage. The same thing happened to Hamilton when his friend Humphrey Lloyd observed conical refraction, just as Hamilton had predicted. After that, Hamilton was revered as one of the most ingenious scientists of his day.

Following the success of conical refraction, Hamilton turned from optics to pursue a striking correspondence he had noted in his Characteristic Function that applied to mechanical trajectories as well as it did to rays of light. In 1834 and 1835 he published two papers On a General Method in Mechanics( I and II)[ii], in which he reworked the theory of Lagrange by beginning with the principle of varying action, which is now known as Hamilton’s Principle. Hamilton’s principle is related to Maupertuis’ principle of least action, but it was more rigorous and a more general approach to derive the Euler-Lagrange equations.  Hamilton’s Principal Function allowed the trajectories of particles to be calculated in complicated situations that were challenging for a direct solution by Lagrange’s equations.

The importance that these two papers had on the future development of physics would not be clear until 1842 when Carl Gustav Jacob Jacobi helped to interpret them and augment them, turning them into a methodology for solving dynamical problems. Today, the Hamiltonian approach to dynamics is central to all of physics, and thousands of physicists around the world mention his name every day, possibly more often than they mention Einstein’s.

[i] Reprinted from D. D. Nolte, Galileo Unbound: A Path Across Life, the Universe and Everything (Oxford, 2018)

[ii] W. R. Hamilton, “On a general method in dynamics I,” Phil. Trans. Roy. Soc., pp. 247-308, 1834; W. R. Hamilton, “On a general method in dynamics II,” Phil. Trans. Roy. Soc., pp. 95-144, 1835.

Galileo Unbound

Book Outline Topics

  • Chapter 1: Flight of the Swallows
    • Introduction to motion and trajectories
  • Chapter 2: A New Scientist
    • Galileo’s Biography
  • Chapter 3: Galileo’s Trajectory
    • His study of the science of motion
    • Publication of Two New Sciences
  • Chapter 4: On the Shoulders of Giants
    • Newton’s Principia
    • The Principle of Least Action: Maupertuis, Euler, and Voltaire
    • Lagrange and his new dynamics
  • Chapter 5: Geometry on my Mind
    • Differential geometry of Gauss and Riemann
    • Vector spaces rom Grassmann to Hilbert
    • Fractals: Cantor, Weierstrass, Hausdorff
  • Chapter 6: The Tangled Tale of Phase Space
    • Liouville and Jacobi
    • Entropy and Chaos: Clausius, Boltzmann and Poincare
    • Phase Space: Gibbs and Ehrenfest
  • Chapter 7: The Lens of Gravity
    • Einstein and the warping of light
    • Black Holes: Schwarzschild’s radius
    • Oppenheimer versus Wheeler
    • The Golden Age of General Relativity
  • Chapter 8: On the Quantum Footpath
    • Heisenberg’s matrix mechanics
    • Schrödinger’s wave mechanics
    • Bohr’s complementarity
    • Einstein and entanglement
    • Feynman and the path-integral formulation of quantum
  • Chapter 9: From Butterflies to Hurricanes
    • KAM theory of stability of the solar system
    • Steven Smale’s horseshoe
    • Lorenz’ butterfly: strange attractor
    • Feigenbaum and chaos
  • Chapter 10: Darwin in the Clockworks
    • Charles Darwin and the origin of species
    • Fibonnacci’s bees
    • Economic dynamics
    • Mendel and the landscapes of life
    • Evolutionary dynamics
    • Linus Pauling’s molecular clock and Dawkins meme
  • Chapter 11: The Measure of Life
    • Huygens, von Helmholtz and Rayleigh oscillators
    • Neurodynamics
    • Euler and the Seven Bridges of Königsberg
    • Network theory: Strogatz and Barabasi

In June of 1633 Galileo was found guilty of heresy and sentenced to house arrest for what remained of his life. He was a renaissance Prometheus, bound for giving knowledge to humanity. With little to do, and allowed few visitors, he at last had the uninterrupted time to finish his life’s labor. When Two New Sciences was published in 1638, it contained the seeds of the science of motion that would mature into a grand and abstract vision that permeates all science today. In this way, Galileo was unbound, not by Hercules, but by his own hand as he penned the introduction to his work:

. . . what I consider more important, there have been opened up to this vast and most excellent science, of which my work is merely the beginning, ways and means by which other minds more acute than mine will explore its remote corners.

            Galileo Galilei (1638) Two New Sciences

Galileo Unbound (Oxford University Press, 2018) explores the continuous thread from Galileo’s discovery of the parabolic trajectory to modern dynamics and complex systems. It is a history of expanding dimension and increasing abstraction, until today we speak of entangled quantum particles moving among many worlds, and we envision our lives as trajectories through spaces of thousands of dimensions. Remarkably, common themes persist that predict the evolution of species as readily as the orbits of planets. Galileo laid the foundation upon which Newton built a theory of dynamics that could capture the trajectory of the moon through space using the same physics that controlled the flight of a cannon ball. Late in the nineteenth-century, concepts of motion expanded into multiple dimensions, and in the 20th century geometry became the cause of motion rather than the result when Einstein envisioned the fabric of space-time warped by mass and energy, causing light rays to bend past the Sun. Possibly more radical was Feynman’s dilemma of quantum particles taking all paths at once—setting the stage for the modern fields of quantum field theory and quantum computing. Yet as concepts of motion have evolved, one thing has remained constant—the need to track ever more complex changes and to capture their essence—to find patterns in the chaos as we try to predict and control our world. Today’s ideas of motion go far beyond the parabolic trajectory, but even Galileo might recognize the common thread that winds through all these motions, drawing them together into a unified view that gives us the power to see, at least a little, through the mists shrouding the future.

Huygens’ Tautochrone

In February of 1662, Pierre de Fermat wrote a paper Synthesis ad refractiones that explained Descartes-Snell’s Law of light refraction by finding the least time it took for light to travel between two points. This famous approach is now known as Fermat’s principle, and it motivated other searches for minimum principles. A few years earlier, in 1656, Christiaan Huygens had invented the pendulum clock [1], and he began a ten-year study of the physics of the pendulum. He was well aware that the pendulum clock does not keep exact time—as the pendulum swings wider, the period of oscillation slows down. He began to search for a path of the pendular mass that would keep the period the same (and make pendulum clocks more accurate), and he discovered a trajectory along which a mass would arrive at the same position in the same time no matter where it was released on the curve. That such a curve could exist was truly remarkable, and it promised to make highly accurate time pieces.

It made minimization problems a familiar part of physics—they became part of the mindset, leading ultimately to the principle of least action.

This curve is known as a tautochrone (literally: same or equal time in Greek) and Huygens provided a geometric proof in his Horologium Oscillatorium sive de motu pendulorum (1673) that the curve was a cycloid [2]. A cycloid is the curve traced by a point on the rim of a circular wheel as the wheel rolls without slipping along a straight line. Huygens invented such a pendulum in which the mass executed a cycloid curve. It was a mass on a flexible yet inelastic string that partially wrapped itself around a solid bumper on each half swing. In principle, whether the pendulum swung gently, or through large displacements, the time would be the same. Unfortunately, friction along the contact of the string with the bumper prevented the pendulum from achieving this goal, and the tautochronic pendulum did not catch on.

HuygensIsochron

Fig. 1 Huygens’ isochronous pendulum.  The time it takes the pendulum bob to follow the cycloid arc is independent of the pendulum’s amplitude, unlike for the circular arc, as the pendulum slows down for larger excursions.

The solution of the tautochrone curve of equal time led naturally to a search for the curve of least time, known as the brachistochrone curve for a particle subject to gravity, like a bead sliding on a frictionless wire between two points. Johann Bernoulli published a challenge to find the brachistochrone in 1696 in the scientific journal Acta Eruditorum that had been founded in 1682 by Leibniz in Germany in collaboration with Otto Mencke. Leibniz envisioned the journal to be a place where new ideas in the natural sciences and mathematics could be published and disseminated rapidly, and it included letters and commentaries, acting as a communication hub to help establish a community of scholars across Europe. In reality, it was the continental response to the Proceedings of the Royal Society in England.  Naturally, the Acta and the Proceedings would later take partisan sides in the priority dispute between Leibniz and Newton for the development of the calculus.

When Bernoulli published his brachistochrone challenge in the June issue of 1696, it was read immediately by the leading mathematicians of the day, many of whom took up the challenge and replied. The problem was solved and published in the May 1697 issue of the Acta by no less than five correspondents, including Johann Bernoulli, Jakob Bernoulli (Johann’s brother), Isaac Newton, Gottfried Leibniz and Ehrenfried Walther von Tschirnhaus. Each of them varied in their approaches, but all found the same solution. Johann and Jakob each considered the problem as the path of a light beam in a medium whose speed varied with depth. Just as in the tautochrone, the solved curve was a cycloid. The path of fastest time always started with a vertical path that allowed the fastest acceleration, and the point of greatest depth always was at the point of greatest horizontal speed.

The brachistrochrone problem led to the invention of the variational calculus, with first steps by Jakob Bernoulli and later more rigorous approaches by Euler.  However, its real importance is that it made minimization problems a familiar part of physics—they became part of the mindset, leading ultimately to the principle of least action.

Mathematical Description of the Tautochrone Curve

If the length of the pendulum is L, the position coordinates (relative to the cusp of the cycloid) are

where the angle θ is between the straight section of the string and the vertical and depends on time as

for a maximum angle θ0. The angular frequency of the isochronous pendulum is given by the standard expression for a small-amplitude pendulum as

For small-angle oscillations, the tautochrone and the conventional pendulum have the same periods and amplitudes. The difference is that the tautochrone has that same period regardless of the amplitude, so no approximation is needed.

Huygens’ proof of the tautochrone curve was made geometrically without the use of calculus, requiring almost 18 pages and 16 figures.

The modern proof (due to Euler) takes only a few lines of calculus. All that is needed is to recognize that the equation of motion of the pendulum bob along its path of length s should be that of a simple-harmonic oscillator

because a simple harmonic oscillator always oscillates at the same frequency regardless of amplitude. Of course, a simple pendulum is not a simple harmonic oscillator because it becomes nonlinear and slows down for large amplitudes. The right-hand-side is the standard force on the bob given by the gradient of the potential energy along the curve

yielding the simple differential equation

which is integrated to give a quadratic dependence of height on path length

To get an equation for the curve itself, plug this back in to the differential equation and square it to give

Euler recognized this as the differential equation of a cycloid

Notes

[1] Galileo conceived the pendulum clock in 1641, and his son Vincenzo started construction, but it was never finished.  Huygens submitted and received a patent in 1657 for a practical escape mechanism on pendulum clocks that is still used today.

[2] Huygens, Christiaan; Blackwell,, Richard J., trans. (1986). Horologium Oscillatorium (The Pendulum Clock, or Geometrical demonstrations concerning the motion of pendula as applied to clocks). Ames, Iowa: Iowa State University Press

Geometry as Motion

Nothing seems as static and as solid as geometry—there is even a subfield of geometry known as “solid geometry”. Geometric objects seem fixed in time and in space. Yet the very first algebraic description of geometry was born out of kinematic constructions of curves as René Descartes undertook the solution of an ancient Greek problem posed by Pappus of Alexandria (c. 290 – c. 350) that had remained unsolved for over a millennium. In the process, Descartes’ invented coordinate geometry.

Descartes used kinematic language in the process of drawing  curves, and he even talked about the speed of the moving point. In this sense, Descartes’ curves are trajectories.

The problem of Pappus relates to the construction of what were known as loci, or what today we call curves or functions. Loci are a smooth collection of points. For instance, the intersection of two fixed lines in a plane is a point. But if you allow one of the lines to move continuously in the plane, the intersection between the moving line and the fixed line sweeps out a continuous succession of points that describe a curve—in this case a new line. The problem posed by Pappus was to find the appropriate curve, or loci, when multiple lines are allowed to move continuously in the plane in such a way that their movements are related by given ratios. It can be shown easily in the case of two lines that the curves that are generated are other lines. As the number of lines increases to three or four lines, the loci become the conic sections: circle, ellipse, parabola and hyperbola. Pappus then asked what one would get if there were five such lines—what type of curves were these? This was the problem that attracted Descartes.

What Descartes did—the step that was so radical that it reinvented geometry—was to fix lines in position rather than merely in length. To us, in the 21st century, such an act appears so obvious as to remove any sense of awe. But by fixing a line in position, and by choosing a fixed origin on that line to which other points on the line were referenced by their distance from that origin, and other lines were referenced by their positions relative to the first line, then these distances could be viewed as unknown quantities whose solution could be sought through algebraic means. This was Descartes’ breakthrough that today is called “analytic geometry”— algebra could be used to find geometric properties.

Newton too viewed mathematical curves as living things that changed in time, which was one of the central ideas behind his fluxions—literally curves in flux.

Today, we would call the “locations” of the points their “coordinates”, and Descartes is almost universally credited with the discovery of the Cartesian coordinate system. Cartesian coordinates are the well-known grids of points, defined by the x-axis and the y-axis placed at right angles to each other, at whose intersection is the origin. Each point on the plane is defined by a pair of numbers, usually represented as (x, y). However, there are no grids or orthogonal axes in Descartes’ Géométrie, and there are no pairs of numbers defining locations of points. About the most Cartesian-like element that can be recognized in Descartes’ La Géométrie is the line of reference AB, as in Fig. 1.

Descartesgeo5

Fig. 1 The first figure in Descartes’ Géométrie that defines 3 lines that are placed in position relative to the point marked A, which is the origin. The point C is one point on the loci that is to be found such that it satisfies given relationships to the 3 lines.

 

In his radical new approach to loci, Descartes used kinematic language in the process of drawing the curves, and he even talked about the speed of the moving point. In this sense, Descartes’ curves are trajectories, time-dependent things. Important editions of Descartes’ Discourse were published in two volumes in 1659 and 1661 which were read by Newton as a student at Cambridge. Newton also viewed mathematical curves as living things that changed in time, which was one of the central ideas behind his fluxions—literally curves in flux.

 

Descartes’ Odd Geometry

Rene Descartes was an unlikely candidate to revolutionize geometry. He began his career as a mercenary soldier, his mind wrapped around things like war and women, which are far from problems of existence and geometry. Descartes’ strange conversion from a life of action to a life of mind occurred on the night of November 10-11 in 1619 while he was bivouacked in an army encampment in Bavaria as a mercenary early in the Thirty Years’ War (1618—1648). On that night, Descartes dreamed that exact rational thought, even mathematical method, could be applied to problems of philosophy. This became his life’s work, and because he was a man of exceptional talent, he succeeded in exceptional ways.

Even Descartes’ footnotes were capable of launching new fields of thought.

Descartes left his mercenary employment and established himself in the free-thinking republic of the Netherlands which was ending the long process of casting off the yolk of Spanish rule towards the end of the Eighty Years War (1568—1648). In 1623, he settled in The Hague, a part of the republic that had been free of Spanish troops for many years, and after a brief absence (during which he witnessed the Siege of Rochelle by Cardinal Richelieu), he returned to the Netherlands in 1628, at the age of 32. He remained in the Netherlands, moving often, taking classes or teaching classes at the Universities of Leiden and Utrecht until 1649, when he was enticed away by Queen Christina of Sweden to colder climes and ultimately to his death.

Descartes3Char

Descartes’ original curve (AC), constructed on non-orthogonal (oblique) x and y coordinates (La Géométrie, 1637)

Descartes’ years in the Netherlands were the most productive epoch of his life as he created his philosophy and pursued his dream’s promise. He embarked on an ambitious project to codify his rational philosophy to gain a full understanding of natural philosophy. He called this work Treatise on the World, known in short as Le Monde, and it quite naturally adopted Copernicus’ heliocentric view of the solar system, which by that time had become widely accepted in learned circles even before Galileo’s publication in 1632 of his Dialogue Concerning the Two Chief World Systems. However, when Galileo was convicted in 1633 of suspicion of heresy (See Galileo Unbound, Oxford University Press, 2018), Descartes abruptly abandoned his plans to publish Le Monde, despite being in the Netherlands where he was well beyond the reach of the Church. It was, after all, the Dutch publisher Elzevir who published Galileo’s last work on the Two Sciences in 1638 when no Italian publishers would touch it. However, Descartes remained a devout Catholic throughout his life and had no desire to oppose its wishes. Despite this setback, Descartes continued to work on less controversial parts of his project, and in 1637 he published three essays preceded by a short introduction.

The introduction was called the Discourse on the Method (which contained his famous cogito ergo sum), and the three essays were La Dioptrique on optics, Les Météores on atmosphere and weather and finally La Géométrie on geometry in which he solved a problem posed by Pappus of Alexandria in the fourth century AD. Descartes sought to find a fundamental set of proven truths that would serve as the elements one could use in a deductive method to derive higher-level constructs. It was partially as an exercise in deductive reasoning that he sought to solve the classical mathematics problem posed by Pappus. La Géométrie was published as an essay following the much loftier Discourse, so even Descartes’ footnotes were capable of launching new fields of thought. The new field is called analytical geometry, also known as Cartesian or coordinate geometry, in which algebra is applied to geometric problems. Today, coordinates and functions are such natural elements of mathematics and physics, that it is odd to think that they emerged as demonstrations of abstract philosophy.

Bibliography:  R. Descartes, D. E. Smith, and M. L. Latham, The geometry of René Descartes. Chicago: Open Court Pub. Co., 1925.

 

The Oxford Scholars

 

Oxford University, and specifically Merton College, was a site of intense intellectual ferment in the middle of the Medieval Period around the time of Chaucer. A string of natural philosophers, today called the Oxford Scholars or the Oxford Calculators, began to explore early ideas of motion, taking the first bold steps beyond Aristotle. They were the first “physicists” (although that term would not be used until our own time) and laid the foundation upon which Galileo would build the first algebraic law of physics.

It is hard to imagine today what it was like doing mathematical physics in the fourteenth century. Mathematical symbolism did not exist in any form. Nothing stood for anything else, as we routinely use in algebra, and there were no equations, only proportions.

Thomas Bradwardine (1290 – 1349) was the first of the Scholars, arriving at Oxford around 1320. He came from a moderately wealthy family from Sussex on the southern coast of England not far from where the Anglo Saxon king Harold lost his kingdom and his life at the Battle of Hastings. The life of a scholar was not lucrative, so Bradwardine supported himself mainly through the royal patronage of Edward III, for whom he was chaplain and confessor during Edward’s campaigns in France, eventually becoming the Archbishop of Canterbury, although he died of the plague returning from Avignon before he could take up the position. When not campaigning or playing courtier, Bradwardine found time to develop a broad-ranging program of study that spanned from logic and theology to physics.

UK-2014-Oxford-Merton_College_05

Merton College, Oxford (attribution: Andrew Shiva / Wikipedia)

Bradwardine began a reanalysis of an apparent paradox that stemmed from Aristotle’s theory of motion. As anyone with experience pushing a heavy box across a floor knows, the box does not move until sufficient force is applied. Today we say that the applied force must exceed the static force of friction. However, this everyday experience is at odds with Aristotle’s theory that placed motion in inverse proportion to the resistance. In this theory, only an infinite resistance could cause zero motion, yet the box does eventually move if enough force is applied. Bradwardine sought to resolve this paradox. Within the scholastic tradition, Aristotle was always assumed to have understood the truth, even if fourteenth-century scholars could not understand it themselves. Therefore, Bradwardine constructed a mathematical “explanation” that could preserve Aristotle’s theory of proportion while still accounting for the fact that the box does not move.

It is hard to imagine today what it was like doing mathematical physics in the fourteenth century. Mathematical symbolism did not exist in any form. Nothing stood for anything else, as we routinely use in algebra, and there were no equations, only proportions. The introduction of algebra into Europe through Arabic texts was a hundred years away. Not even Euclid or Archimedes had been rediscovered by Bradwardine’s day, so all he had to work with was Pythagorean theory of ratios and positive numbers—even negative numbers did not exist—and the only mathematical tools at his disposal were logic and language. Nonetheless, armed only with these sparse tools, Bradwardine constructed a verbal argument that the proportion of impressed force to resistance must itself be as a proportionality between speeds. As awkward as this sounds in words, it is a first intuitive step towards the concept of an exponential relationship—a power law. In Bradwardine’s rule for motion, the relationships among force, resistance and speed were like compounding interest on a loan. Bradwardine’s rule is not correct physics, because the exponential function is never zero, and because motion does not grow exponentially, but it did introduce the intuitive idea that a physical process could be nonlinear (using our modern terminology), changing from small effects to large effects disproportionate to the change in the cause. Therefore, the importance of Bradwardine was more his approach than his result. He applied mathematical reasoning to a problem of kinetics and set the stage for mathematical science.

A few years after Bradwardine had devised his rule of motion, a young mathematician named William Heytesbury (1313—1373) arrived as a fellow of Merton College. In the years that they overlapped at Oxford, one can only speculate what was transmitted from the senior to the junior fellow, but by 1335 Heytesbury had constructed a theory of continuous magnitudes and their continuous changes that included motion as a subset. The concept of the continuum had been a great challenge for Aristotelian theory, leading to many paradoxes or sophisms, like Zeno’s paradox that supposedly proved the impossibility of motion. Heytesbury shrewdly recognized that the problem was the ill-defined idea of instantaneous rate of change.

Heytesbury was just as handicapped as Bradwardine in his lack of mathematical notation, but worse, he was handicapped by the Aristotelian injunction against taking ratios of unlike qualities. According to Aristotle, proportions must only be taken of like qualities, such as one linear length to another, or one mass to another. To take a proportion of a mass to a length was nonsense. Today we call it “division” (more accurately a distribution), and mass divided by length is a linear mass density. Therefore, because speed is distance divided by time, no such ratio was possible in Heytesbury’s day because distance and time are unlike qualities. Heytesbury ingeniously got around this restriction by considering the linear distances covered by two moving objects in equal times. The resulting linear distances were similar qualities and could thereby be related by proportion. The ratio of the distances become the ratio of speeds, even though speed itself could not be defined directly. This was a first step, a new tool. Using this conceit, Heytesbury was able to go much farther, to grapple with the problem of nonuniform motion and hence the more general concept of instantaneous speed.

In the language of calculus (developed by Newton and Leibniz 300 years later), instantaneous speed is a ratio of an element of length to an element of time in the limit as the elements vanish uniformly. In the language of Heytesbury (Latin), instantaneous speed is simply the average speed between two neighboring speeds (still working with ratios of distances traversed in equal times). And those neighboring speeds are similarly the averages of their neighbors, until one reaches zero speed on one end and final speed on the other. Heytesbury called this kind of motion difform as opposed to uniform motion.

A special case of difform motion was uniformly difform motion—uniform acceleration. Acceleration was completely outside the grasp of Aristotelian philosophers, even Heytesbury, but he could imagine a speed that changed uniformly. This requires that the extra distance travelled during the succeeding time unit relative to the distance travelled during the current time unit has a fixed value. He then showed, without equations, using only his math-like language, that if a form changes uniformly in time (constant rate of change) then the average value of the form over a fixed time is equal to the average of the initial and final values. This work had a tremendous importance, not only for the history of mathematics, but also for the history of physics, because when the form in question is speed, then this represents the discovery of the mean speed theorem for the case of uniform acceleration. The mean speed theorem is often attributed to Galileo, who proved the theorem as part of his law of fall, and he deserves the attribution because there is an important difference in context. Heytesbury was not a scientist nor even a natural philosopher. He was a logician interested in sophisms that arose in discussions of Aristotle. The real purpose of Heytesbury’s analysis was to show that paradoxes like that of Zeno could be resolved within the Aristotelian system. He certainly was not thinking of falling bodies, whereas Galileo was.

Not long after Heytesbury demonstrated the mean speed theorem, he was joined at Merton College by yet another young fellow, Richard Swineshead (fl. 1340-1354). Bradwardine was already gone, but his reputation survived, as well as his memory, in the person of Heytesbury, and Swineshead became another member in the tradition of the Merton Scholars. He was perhaps the most adept at mathematics of the three, and he published several monumental treatises that mathematically expanded upon both Bradwardine and Heytesbury, systematizing their results and disseminating them in published accounts that spread across scholastic Europe—all still without formulas, symbols or equations. For these works, he became known as The Calculator. By consolidating and documenting the work of the Oxford Scholars, his influence on the subsequent history of thought was considerable, as he was widely read by later mathematicians, including Leibniz, who had a copy of Heytesbury in his personal library.

( To read more about the Oxford Scholars, and their connections with members of their contemporaries in Paris, see Chapter 3 of Galileo Unbound (Oxford University Press, 2018).)