Image Anal Stereol 2012;31:1-16 Review Article EARLY HISTORY OF GEOMETRIC PROBABILITY AND STEREOLOGY Magdalena Hyksova^3'1, Anna Kalousova2 and Ivan Saxl^3 institute of Applied Mathematics, Faculty of Transportation Sciences, Czech Technical University in Prague, Na Florenci 25, CZ-110 00 Praha 1, Czech Republic; department of Mathematics, Faculty of Electrical Engineering, Czech Technical University in Prague, Technicka 2, CZ-166 27 Praha 6, Czech Republic; 3In memoriam e-mail: hyksova@fd.cvut.cz, kalous@math.feld.cvut.cz (Received October 27, 2011; revised January 2, 2012; accepted January 2, 2012) ABSTRACT This paper provides an account of the history of geometric probability and stereology from the time of Newton to the early 20th century. It depicts the development of two parallel paths. On the one hand, the theory of geometric probability was formed with minor attention paid to applications other than those concerning spatial chance games. On the other hand, practical rules for the estimation of area or volume fraction and other characteristics, easily deducible from the geometric probability theory, were proposed without knowledge of this branch. Special attention is paid to the paper of J.-E. Barbier, published in 1860, which contained the fundamental stereological formulas, but remained almost unnoticed by both mathematicians and practitioners. Keywords: geometric probability, history, stereology. INTRODUCTION The first known problem related to geometric probability can be found in a private manuscript of Isaac Newton (1643-1727) previously written between the years 1664-1666, but not published until the 20th century (Newton, 1967). It consists of calculating the chance of hitting one of two unequal areas of a circle by a ball (of negligible size): If ye Proportion of chances for any stake bee irrational, the interest may bee found after ye same manner. As if ye Radij oh. or. divide ye horizontal circle bed in two pts abec & abed in such proportion as 2 to \/5. And if a ball falling perpendicularly upon ye center a doth tumble into ye portion abec I win (a): but if ye other portion, I win b, my hopes is worth (2a+'V5b)/(2 + V5). ... if a die bee not a regular body but a Parallelepipedon or otherwise unequal] sided, it may be found how much one cast is more easily gotten than other. (Newton, 1967, p. 60-61) Fig. 1. Newton 's illustration (Newton, 1967). Newton wrote his note after reading the treatise (Huygens, 1657) to point out that probability can be irrational. What is even more remarkable is his claim that chance is proportional to area fraction and the proposal of a frequency experiment for chance estimation. More generally, applying Newton's idea to the ball of negligible size falling N times on an area C = A U B and hitting n times the region A, and setting a = 1, b = 0, Newton's hope would be n/N, which estimates the area fraction of A with respect to C. Thus, the first 2D stereological formula, Aa = Pp. was born. In other words, Newton's ideas immediately lead to replacing the counting of events with their measure and to estimating an area fraction by a point count. We can therefore say that Newton also invented stereology. In a published form, some ideas of geometric probability were applied by Edmond Halley (1656 — 1742) in a paper that is generally appreciated as having laid the foundations of a correct theory of life annuities (Halley, 1693). Halley deduced formulas for various annuities first analytically, and then gave their geometric illustration. One year earlier, John Arbuthnot (1667-1735) published the first English work on probability (Arbuthnot, 1692), which contained the translation of Huygens (1657) and some other problems concerning various chance games. Among them we can find an unsolved problem of a completely different nature, which was solved half a century later by means of geometric probability: In a Parallelopipedon, whose Sides are to one another in the Ratio of a. h. c : to find at how many Throws any one may undertake that any given Plane, viz. oh. may arise.1 The problem was solved by Thomas Simpson (1710-1761) who considered a sphere described round the parallelepiped and imagined a radius passing round the boundary of the given plane. He said that the chance of the given plane being uppermost in a single throw was equal to the ratio of the spherical surface bounded by the moving radius to the whole sphere (Simpson, 1740). In this way, he reduced the problem to finding the area of a certain portion of the surface of a sphere and thus anticipated the measure of the bundle of lines in space passing through a given point. BUFFON'S PROBLEMS Geometric probability is inseparably connected with the name of Georges-Louis Leclerc, later Comte de Buffon (1707-1788), who is generally regarded to be the founder of the discipline. Fontenelle (1735, pp. 43-35) reports about Buffon's presentation of the solutions of franc-carreau and needle problems at the French Royal Academy of Sciences in 1733,2 and some generalizations can be found in Buffon (1777).3 Recall that in the franc-carreau game, popular at the French court, a round coin is tossed at random on a large plane area covered by regular tiles (in the 1733 presentation, Buffon considered squares; the 1777 paper also dealt with triangles, rhombi and hexagons); one of the players bets that the coin hits only one tile while the other bets that it hits more of them. Finding the odds of both players in this game is clearly similar to the above described Newton's problem: each player needs the coin centre to hit a certain subset of a given set. But the needle problem is a completely different and new one - here an interaction of two ID sets in 2D is examined:4 In a room, the floor of which is simply divided by parallel joints, a rod is thrown in the air, and one of the players bets that the rod will not cross any of the lines while the other bets that it will cross several of them. The odds for these two players are required. We can play this game on a board with a sewing needle or a pin without a head. (Buffon, 1777, pp. 100-101) Denote by / the rod (needle) length and by d > 1 the distance of unbounded parallels. Using integral calculus, Buffon obtained the odds (l - : and used them to calculate the ratio l/d ensuring a fair game for players betting on hits or misses; obviously, the odds are equal for l/d = nj\. Buffon (1777) also considered a generalization where the rod was thrown down on a square grid, but he gave an incorrect result. The needle problem deserves several comments concerning its relation to stereology. The determined odds correspond to the hitting probability nd (1) Noting that l/d is the length intensity I,a of the system of parallels and estimating P by the relative frequency N/n, where n is the number of randomly thrown needles J? of total length L = nl, and N the number of successful hits .7 *| - 7, Buffon's result can be rewritten as _ which is the equation for estimating the length of a curve (Steinhaus, 1930b), but in the present case it relates to a compact random segment (fibre) set. Besides sections, projections are also important tools of geometric sampling and subsequent inference (Davy and Miles, 1977). Moreover, by the Hadwiger characterization theorem (Hadwiger, 1957), the set of all possible mean projection measures together with the Lebesgue measure and Euler-Poincare characteristics form the basis for the (in + 1)-dimensional vector space of all convex-continuous motion invariant valuations defined on the convex ring in ¿%m (Klain and Rota, 1997). By the Cauchy-Crofton formula (Eq. 7) for convex bodies in 2D the relation L= kw holds between the length of the perimeter L and the mean projection (also called the mean width) w. Consequently, the hitting probability (Eq. 13) can be written as P = w/d, which is just the way the formula was derived for the needle, namely by calculating its mean projection into the normal of parallels J? (note that the perimeter of a ID segment embedded in equals twice its length). Hence if such a relation was valid in general for convex figures, it would be possible to estimate the mean width w or the perimeter L of arbitrary figures by their repeated throwing on the test Quotation from (Simpson. 1740. p. 67); the original formulation in Arbuthnot (1692) was in Latin. Arbuthnot wrote that he left the Solution to those who think it merits their pains (Todhunter, 1865. p. 53). 2The franc-carreau with square tiles was already described in the letter written by Gabriel Cramer to James Stirling on February 22. 1732. published in (Tweedie. 1922. pp. 122-128). Cramer remarked that its solution for a round coin was not difficult, but for a square one. he was not able to solve it. On the other hand. Buffon wrote the first letter to Cramer in 1727 and in 1731 he visited him in Geneva. 3The first pages of many of works discussed in this paper, including Buffon (1777). are reproduced in Miles and Serra (1978). 4Similarly to Cramer. Buffon mentioned already in the 1733 presentation that the franc-carreau problem would be more complicated if the round coin was replaced e.g.. by a square one (or a Spanish pistol). Since he did not know how to solve it. he considered only a wand; to simplify the problem further, he started with parallels and only in the later paper he tried to generalize the solution to the square grid. Let us further remark that Buffon (1777) also considered more players betting on hitting different numbers of commissures. lines (or equivalently by laying on them repeatedly an isotropic and uniform random test system J?) and counting the fraction of successful hits. Buffon's problem remained unnoticed by his contemporaries. It received considerable attention only after Pierre-Simon de Laplace (1749 -1827) presented - without any reference to its author - both the needle problem and its generalization with two systems of parallels forming a rectangular (not only squared) grid with correct solutions. Laplace (1812) introduced the discussion of these problems with the remark that probability theory could also be used for curve rectification and surface quadrature. Nevertheless, he gave only one example, namely the estimation of the perimeter 2n of the unit circle, based on repeatedly throwing a narrow cylinder on a system of parallels. Unfortunately, since Todhunter (1865), Laplace's result was often referred to as the application of Buffon's needle problem to the estimation of k (in school mathematics, it is usually the only "application" of the needle problem until today). Since then, other mathematicians introduced some generalizations. For example, Isaac Todhunter (18201884), a lecturer at St John's College, Cambridge, who popularized geometric probability by several exercises in his textbook on integral calculus, considered a cube, a rod of the length equal to the multiple of d, and an ellipse to be thrown on the plane ruled with equidistant parallels at a distance d (Todhunter, 1857), later also a closed curve without singular points with the greatest diameter5 less than d (1862 edition of the same book). With the reference to the earlier presentation in Fontenelle (1735), Buffon's problems were also analysed in a historical overview (Todhunter, 1865), where a simple derivation of a correct hitting probability for a rectangular grid was given as well. While Laplace (1812) divided rectangles into particular parts and for each of them he "measured" the favourite positions of a needle, Todhunter (1865) used a unique integral over an angle formed by the needle and one of the parallel systems. Gabriel Lamé (1795-1870) included Buffon's needle problem and its generalizations to a circle, an ellipse and regular polygons in his lectures held at the Sorbonne. Inspired by these lectures, Joseph-Émile Barbier (1839-1889) published a general theorem concerning the mean number of intersections of an arbitrary curve with a system of parallels and, what is remarkable, he replaced equidistant parallels by an arbitrary system of lines or even a unique curve of constant length per unit area, and came to the estimator Eq. 16. Moreover, he extended his results to 3D and formulated three more theorems that express other contemporary fundamental stereological formulas (Eqs. 17-19) for surface area and curve length estimation. Unfortunately, his contribution remained unknown for some time to Todhunter as well as to Crofton, and its immediate stereological applications were not appreciated until the beginning of the 21st century - see Kalousova (2009). The section after next will therefore describe Barbier's contribution in more detail. AUGUSTIN-LOUIS CAUCHY A.-L. Cauchy (1789-1857) did not speak about "geometric probability", but he derived several theorems with interesting practical implications for length and surface area estimation which were later rediscovered or reformulated in the terms of geometric probability. The lithograph (Cauchy, 1832) starts with the theorem stating that the equality6 (3) holds for the length L of a system J? of arbitrary curves, where W(8) denotes the length of the total [orthogonal linear] projection of J? onto a straight line forming an angle 6 with a fixed axis (each subinterval of the projection is counted as many times as there are points on the curve which project onto it). Cauchy gives a brief direct demonstration and then shows that Eq. 3 can also be derived from another theorem stating that L can be approximated by K - LPÜ — -Wn (4) where Wn denotes the mean value of the total projection of J? onto n radial lines separated by equal angles, and the error of this approximation is smaller than jnWn/n1. In a detailed demonstration, Cauchy decomposes J? into infinitesimal line segments and for any of them he investigates the projections. Increasing n to infinity, Cauchy obtains (in our notation) K — (5) where W is the mean value _ £xw(0)de w /-* de — Jw(d)dd, (6) which implies Eq. 3. In a corollary to the first theorem Cauchy remarks that for a closed convex curve, W(6) 5 Todhunter used the term diameter in the sense of our width and greatest diameter in the sense of our diameter. 6The notation is slightly changed with respect to the following parts of this paper. is reduced to twice the ["usual" orthogonal linear] projection w{6). From today's point of view we can notice that Eq. 3 and Eq. 5 immediately imply the so-called Cauchy-Crofton formula, which enables the rectification of a closed convex curve by finding its mean width w, i.e., the mean length of its orthogonal projections onto an isotropic bundle of directions (see Fig. 2):7 L= f w{d)dd = Kw. (7) Jo Fig. 2. Illustration of the notation used in Eq. 7. Cauchy continues into 3D and proposes a practical procedure for surface area estimation: for any surface y with surface area S, he considers its total projections into n planes containing faces of a convex polyhedron that lies between concentric spheres of radii r, r( 1 + e), and shows that S can be approximated by twice the mean area A,', of these projections, S^2A[V (8) with an error less than 2Aln • [(1 + e)2 — l] . Increasing n to infinity, Cauchy replaces a polyhedral face by a differential element of a unit sphere and gets 5 = 2-A7, (9) where A' is the mean value & = t- r I*A'{,d)sm, 6) denotes the area of the total projection of the surface S? into a plane, the normal of which is given by the co-latitude and the longitude 6 in a fixed coordinate system (see Fig. 3).8 Eqs. 9,10 obviously imply the formula for the surface area of -/ : 5 = J- I*A'{,d)sm,0) is reduced to twice the ["usual" orthogonal planar] projection. Thus the well-known formula for convex bodies, S = 4-A, (12) where A denotes the mean projected area of S? (taken over all possible orientations in space), again immediately follows from the theorems derived by Cauchy already in 1832. Fig. 3. Illustration of the notation used in Eq. 10. Cauchy was aware of the importance of his results for the rectification of curves and the quadrature of surfaces, although he gave only several examples concerning circle, ellipse, sphere and ellipsoid. Unfortunately, the lithograph (Cauchy, 1832) was not easily accessible to a wide audience and even though it was reprinted in Mémoires de VAcadémie des Sciences in 1850, many authors later referred only to the communication (Cauchy, 1841) published in Comptes Rendus, where the theorems providing Eqs. 3,11 are merely stated and remarks on the approximations (Eqs. 4, 8) are missing (although there are some other theorems concerning, e.g., the upper bound of the length or surface area estimates). One of the earliest exceptions was J.-É. Barbier, to whom the next section is devoted. JOSEPH-ÉMILE BARBIER J.-É. Barbier (1839-1889) studied at the École Normale Supérieure, where he attended (among others) mathematical lectures by Joseph Bertrand (1822-1900). Moreover, he attended the mentioned Lamé's lectures at the Sorbonne. He started his career in 1860 as a professor at a Lycée in Nice and 7Note that for a circle we obtain L = nd, where d is its diameter; for a square with the side a. the mean width is Aa/n. sThe figure is adapted from Czuber (1884a); Cauchy (1832) gave no illustrations. later he worked as an assistant astronomer at the Observatoire de Paris. In 1865, due to developed mental problems, he left Paris, broke all contacts with colleagues and friends and was not heard from again for the next fifteen years. In 1880, his former lecturer Joseph Bertrand found him at an asylum in Charenton-St-Maurice and encouraged him to return to mathematical research. Two years later Barbier won (thanks to Bertrand's intervention) the Francoeur' prize for an article published in Comptes Rendus (Barbier, 1882). This enabled him to leave hospital and to spend the rest of his life more pleasantly. During the 1880s, he published (with one exception in Comptes Rendus) 14 communications upon various branches of mathematics; three of them (Barbier, 1882; 1887a;b) concerned the probability theory (estimation of n based on binomial distribution, a ballot problem etc.) and are referred to and discussed in the recent literature in the context of Barbier's relation to J. Bertrand. From the stereological point of view, the most important one is the paper that Barbier (1860) wrote during his studies. It starts with the exposition of the needle problem, attributed to Laplace, and its generalizations due to Lamé. Then Barbier turns to a convex disk of arbitrary shape with perimeter L that cannot in any position in the plane intersect more than one dividing parallel, and proves that the probability of hitting some of the lines is L nd (13) To prove it, Barbier follows the ideas typical for the analysis of hazards, acknowledging Bertrand's (Bayesian) preference for expectation over probability: consider a convex polygon having m sides of the length c and a diameter less than the parallels distance d\ evidently, each of the sides has the same chance to cross one of the parallels. Now imagine a game of m players, in which each side belongs to one of them and the fact that a side hits some parallel is connected with certain prize. Before each throw, the "mathematical expectations" of all the players are equal, say E. If a player buys n sides, his expected value is therefore nE, i.e., proportional to the number of sides, and it is not changed by any deformation maintaining the sides length and convexity of the polygon (so that its boundary has two intersections with any line it crosses) and keeping the diameter less than d. Thus the hitting probability (multiplied by the award to give nE) is also proportional to the number of sides. Approximating any convex disk with the perimeter L and a diameter less than d by such a deformed polygon, the hitting probability Eq. 13 follows immediately from Lame's result for regular polygons. It is worth noting that Barbier showed that Eq. 13 can also be derived directly from the aforementioned considerations: since the hitting probability is equal for all convex figures with the same perimeter L and diameter less than d, it is sufficient to consider the simplest case, namely a circle with radius r where L = 2nr, r ) by flats (points, lines, planes etc.). Its immediate consequence is another relation given by Crofton (1885): _ = ff^Cdpdd = ttA = A L w ' where C is the mean chord length of a convex region of area A bounded by a curve J? of length L, and w is the mean width of «5? - recall Eq. 7. Also important is the 2D Crofton-Hostinsky formula for the third power of chord lengths C3dpde = 3A2 (23) from which the estimator LC3/ 3 of the squared area A2 immediately follows. Formulae of this type are extremely useful in the stereological analysis of particle aggregates (Miles, 1983). Crofton (1868) did not investigate the chord lengths yet, but in a remark to Eq. 21 we can find another important stereological result, namely the Cauchy-Crofton formula Eq. 7.18 From other problems considered by Crofton (1868), let us at least mention the generalization of Buffon's problem to a "needle" consisting of two rigidly connected figures of diameters not exceeding the distance between parallel lines, later extended by Sylvester (1890) to an arbitrarily long chain of such figures. Only the 2D problems are solved by Crofton, but once formulated they allow for a generalization to higher dimensions. Crofton can therefore be credited for laying the foundations of geometric probability and for undertaking the first systematic attempt to relate measures of intersections of bodies to their properties. As for the generalization to 3D, Crofton himself outlined it at the end of his 1868 paper; in full detail, it was done by Czuber (1884a;b), whose book Crofton (1885) appreciated in the concluding remark: We have now done enough to give the reader some idea of the subject of local probability. We refer him for further information to the very interesting work just published by Emanuel Czuber of Prague, Geometrische Wahrscheinlichkeiten und Mittelwerte, Leipsic 1884... (Crofton, 1885, p. 788) EMANUELCZUBER The above-mentioned "very interesting work" (Czuber, 1884a) published by Emanuel Czuber (18511925), at that time a secondary school teacher in Prague, later a professor at the German Technical University in Brno and at the Technical University in Vienna, represented the first monograph summarizing the state of the art of geometric probability of that time and it played an important role in this field for several decades. In 1902 it was translated into French and it remained a classic until Robert Deltheil (1890-1972), a professor at the Toulouse University, published his book (Deltheil, 1926). The first and more extensive part of Czuber's monograph is devoted to geometric probability itself and it provides not only a detailed exposition of results achieved by French and British predecessors, supplemented with historical remarks and many exercises, but also new results and generalizations. For example, Crofton (1868) 16One of them was the formula ff(a — sin a) dvdy = \L2 — kA, where a denotes the angle between two tangents from an exterior point (x,y) to the curve of the length L that forms the boundary of a convex region of the area A, and the integral is taken over the whole plane outside the boundary. This result and several other relating theorems had already been briefly announced in (Crofton. 1867b). Crofton (1868) appreciated Buffon and Laplace; without mentioning (Barbier. 1860). he remarked that no real attention was devoted to geometric probability till English mathematicians as. e.g., Sylvester and Woolhouse entered this field of research. 17Crofton wrote limits only to single integrals; the notation J? j has been added by the authors of this paper for the sake of clarity. 18Crofton (1868) gave no reference to Cauchy; the fact that Eq. 7 immediately followed from the theorem stated by Cauchy was pointed out by Czuber (1884a). although the reference concerned only the Comptes rendus report (Cauchy. 1841). derived key theorems concerning sets of points and straight lines in a plane and briefly outlined a possible generalization to 3D; this generalization was given in full detail by Czuber. Moreover, although the particular results are derived in 2D and 3D, the introductory chapter contains a definition of geometric probability as a content ratio in ¿Ü'1: = JJ...JK,dxldx2...dxn (24) JJ... JKdxidx2 ■ ■ .dxn ' where K' C K c - . Inspired by the discussion in the Educational Times (referred to elsewhere in the book), Czuber recalls that it is often possible to find different solutions of problems concerning geometric probability, and points out that this diversity arises from different conceptions of the random sampling. Thus he also foreshadowed Bertrand's paradoxes. The second part of the monograph contains the original exposition of the determination of mean values of geometric variables based on geometric probability. As for stereological applications, the book contains only one explicit remark that concerns an experimental rectification of a closed convex curve in 2D. After the proof of the Crofton formula Eq. 21, Czuber recalls another result contained in (Crofton, 1868), namely that the probability that a line hitting a closed convex curve J? of length L hits also a closed convex curve £ of length / that lies inside J? is p = l/L. Then he remarks that this result provides an experimental rectification of a closed convex curve: The curve that has to be rectified is surrounded by another closed convex curve (circle, polygon) of the known length L, a great number s of arbitrary straight lines intersecting L are drawn in the plane of both curves, and those intersecting also the curve of the unknown length I are counted; let their number be in. The higher s, the more accurate the equality m/s = l/L holds, which implies I = Lm/s. (Czuber, 1884a, p. 116) We may regret that Czuber does not explicitly mention an analogous surface area estimation. Nevertheless, it immediately follows from his theorem stating that the measure of all lines hitting a closed convex surface S? is proportional by n/2 to its surface area S. More specifically, in accordance with Crofton's outline, Czuber (1884a) proved: i* I* A(,6) sin dd0 = (25) Jo Jo 2 where A(,d) denotes the area of the projection of S? into the plane whose normal has co-latitude é and longitude 6 in a fixed coordinate system. In a remark to this theorem Czuber refers to (Cauchy, 1841) where the formula Eq. 11 was stated without any demonstration; probably unaware of the older treatise, Czuber provides the direct proof corresponding to the brief outline mentioned by (Cauchy, 1832) in a note to this formula (see page 4 in this paper) and shows that Eq. 25 can simply be deduced from Eq. 11. Some additional original results concerning geometric mean values in 3D are contained in the paper (Czuber, 1884b). As an example, let us mention the formula for the mean chord length C of a convex region of volume V, bounded by a closed convex surface S? of surface area S : V = --SC. (26) 4 Czuber returned to geometric probability also in his later publications, e.g., in a comprehensive treatise (Czuber, 1899) devoted to the history of probability theory and its applications, or in the textbook on probability theory and its applications for life insurance (Czuber, 1903); in the second edition of this book published in 1908, the exposition of geometric probability was enriched by the discussion of set theory and its use in probability theory, as it was systematically done by Rudolph Lammel (1879-1962) in his Ph.D. thesis (Lammel, 1904). FURTHER DEVELOPMENT Considerable attention to geometric probability was also paid by Joseph Bertrand. In the second volume of (Bertrand, 1864; 1870), he devoted a separate section to Crofton's theorems. Among other results, he returned to the problem of a generalized Buffon's needle consisting of two rigidly connected convex figures of diameters less than the distance of parallels solved by Crofton (1868), and he gave its solution based on Barbier's expectation approach. However, more famous are the so-called Bertrand's paradoxes concerning a random selection from infinite populations (Bertrand, 1889), formulated to warn against a careless use of infinity,19 especially his three possibilities of sampling chords in a circle leading to three different answers to the question about the probability that a chord chosen at random is longer than the side a of the inscribed equilateral triangle:20 First, one of the chord endpoints is known (by symmetry, this knowledge should not change the outcome) and its direction determined by the angle a 19Bertrand (1889. p. 4) wrote in the introduction: infinity is not a number, it shall not be, without explanation, introduced in the reasoning. 20Bertrand (1889. p. 4) formulated the problem as follows: A chord is drawn at random in a circle. What is the probability that it is smaller than the side of the inscribed equilateral triangle? But the solution was formulated for chords longer than that side. from the tangent at the endpoint is chosen at random; the chord is longer than a for a € (k/3,2k/3), the required probability is therefore Pi = 1/3 (note that the same result arises from the independent choice of chord endpoints with a uniform distribution over the circumference). Second, the chord direction is known and its distance from the circle centre is chosen at random; the chord is longer than a if the distance is less than half of the circle radius r, which leads to P2 = 1/2. Third, the midpoint is chosen at random; now the chord is longer than a if its midpoint lies inside the concentric circle of the radius r/2, which gives the probability P3 = 1/4 equal to the ratio of circle areas. Bertrand (1889, p. 5) then concludes: Among these three answers, which one is right? None of the three is incorrect, none is correct, the question is ill-posed. We have already mentioned a similar problem formulated by Godfray (1866) and the answer of Woolhouse and Crofton (see p. 7). Bertrand's questions provoked another discussion on the foundations of geometric probability. Henri Poincare (1854-1912), a professor at the Sorbonne, treated this topic in (Poincare, 1896); he introduced the concept of probability density, derived its form for the first two cases considered by Bertrand and stated that the problem followed from the fact that they were different. In accord with his conventionalism, Poincare claims that in general we do not know the nature of the density function, which can be arbitrary, and we must set it at the beginning of our considerations by a meaningful convention. Then he investigates a generalized needle problem. He says that the probability that a given plane figure satisfies certain conditions concerning its position is proportional to the integral J J J dvdvdoj. where (x. y) are the Cartesian coordinates of a fixed point M of ^ and a> is the angle between the .v-axis and a straight line passing through M and rigidly attached to . since - as he shows - this integral is invariant under rotations and translations. If this convention is adopted and a random chord is regarded as a segment of one of the parallels with the distance d < 2r on which the given circle with the radius r is thrown, then the chord is longer than the side of the inscribed equilateral triangle when the line hits also a concentric circle with the radius r/2; since the probability of this event is equal to the ratio of perimeters of the two circles, we obtain P2 = 1/2. The paradox was later discussed by Czuber (1903; from the second edition published in 1908) who presented three more possibilities,21 calculated the corresponding probabilities and pointed out that only 21 One endpoint of the chord is given on the circumference, then ai chosen with uniform distribution over the circle circumference, wl two points are independently chosen inside the circle. the second Bertrand's alternative with P2 = 1/2 corresponded to the concept of randomly chosen straight line as it had been introduced by Crofton. Similarly Borel (1909) asserted that the majority of conceivable natural procedures led to P2. We have already seen that this solution also corresponds to the motion invariant sampling scheme, which is now generally accepted in geometric probability; nevertheless, this does not answer completely the philosophical background of the problem. Marinoff (1994) argues that Bertrand's answers can be construed as replies to three different questions: the random chord is either generated by a procedure on the circumference of the circle, by a procedure outside the circle, or by a procedure inside the circle (compare again Woolhouse, 1866). He criticises the former literature for the little recognition of the distinction between them and demonstrates that clearly stated variations lead to different, but theoretically and empirically self-consistent solutions. Further, see Plato (1994); Sheynin (1994; 2003), and a study paying attention to Bertrand's teaching and probabilistic thinking (Bru, 2006). Let us note that independently of Poincare, the motion invariance requirement was formulated by Elie Cartan (1869-1951), at that time a lecturer at the University at Montpellier, later a professor at the University of Nancy and at the Sorbonne. Cartan (1896) studied multiple integrals over systems of lines in the plane and systems of lines and planes in space and introduced measures of such systems independent (as he proved) of translation and rotation, corresponding to those proposed by Crofton and Czuber, including Eq. 21 and other relations. Nevertheless, there is not a word about geometric probability and its proponents as Cartan was solely dealing with the theory of integrals. In the framework of geometric probability, this topic was investigated by Georg Pölya (1887-1985), a lecturer and later a professor at ETH Zürich. The aim of his paper (Pölya, 1917) was to show (without a reference to Cartan or Poincare) that the measures of sets of lines and planes on which Crofton and Czuber based the geometric probability theory, were the only legitimate ones, and the reason was again the motion invariance. At the beginning of the 20th century, interesting contributions to geometric probability were presented by Bohuslav Hostinsky (1884-1951), a private associate professor at Charles University in Prague, later a full professor at Masaryk University in Brno. Hostinsky (1917; 1920) criticized the traditional ier point is chosen inside the circle; two endpoints are independently is shown to be equivalent with the first possibility with Pi = 1 /3; solution of Buff on's needle problem for being based on an unrealistic assumption that parallels were drawn on an unbounded board and the probability that the needle midpoint hit a region of a given area was proportional to this area and independent of the position of the region. Hostinsky argued that no real experiment could satisfy such an assumption and replaced it by a more realistic one: parallels are drawn on a square table board and the experiment requires the needle to fall on it; now the probability that the needle midpoint hits a square of a given area near the edge of the table is lower than the probability that it hits a square of the same area near the middle. To solve this problem, Hostinsky generalizes the method of arbitrary functions introduced by Poincare (1896)22 and supposes that the probability that the needle midpoint falls into a region M inside a square C (the table) is proportional to the integral over M of the form //