Some results about equichordal convex bodies
Abstract
Let and be two convex bodies in , , with . We say that is an equichordal body for if every chord of tangent to has length equal to a given fixed value . In [1], J. Barker and D. Larman proved that if is a ball, then is a ball concentric with . In this paper we prove, derived from the proof of Theorem 1, that there exist an infinite number of closed curves, different from circles, which possess an equichordal convex body. If the dimension of the space is more than or equal to 3, then only Euclidean balls possess an equichordal convex body. We also prove some results about isoptic curves and give relations between isoptic curves and convex rotors in the plane.
1 Introduction
Let be a convex body in the plane, i.e., a compact and convex set with non-empty interior, and let be a convex polygon. It is said that is a rotor in if for every rotation , there is a translate of that contains to and all sides of are tangent to . There are many results about rotors in regular polygons, see for instance [6], and for the particular case of rotors in equilateral triangles see [21]. The case of rotors in squares is well-known, indeed, bodies of constant width are a very important topic in Convex Geometry and have many interesting properties and applications in mechanisms (see the quite nice book [13]).
Another topic, apparently not related to rotors, is the Equichordal Problem. Let be a point in the interior of a convex body , we say that is an equichordal point if every chord of through have the same length. The famous Equichordal Problem, due to W. Blaschke, W. Rothe, and R. Weitzenböck [2], asks about the existence of a convex body with two equichordal points. There are many false proofs about the non existence of such a body, however, M. Rychlik finally gave a complete proof about the non existence of a body with two equichordal points in [16]. It is worth mentioning that there are many convex bodies, different from the disc, which have exactly one equichordal point. Here we are interested in a generalization of the notion of equichordal point in the following way: Let and be two convex bodies in , , with . We say that is an equichordal body for if every chord of tangent to have length equal to a given fixed value . In [1], J. Barker and D. Larman proved that if is a convex body that possesses an equichordal ball then it is also a ball. However, we wonder if there exist convex bodies different from balls which possess an equichordal convex body in its interior. It seems that bodies which float in equilibrium in every position provide examples of such bodies in the plane (see for instance [20]), however, it is not clear if the considered bodies are convex or not.
In sections 3 and 4 of this paper we study convex bodies for which the chords of one of its isoptic curves (defined in the following section), that are tangent to have length equal to a constant number . Moreover, these bodies are examples of rotors in regular polygons and if we fix the convex body and the circumscribed polygon is rotated, while maintained circumscribed to , its vertices describe the isoptic curve of . In section 6 we also prove that in dimension or higher, only Euclidean balls (or simply balls) possess and equichordal convex body in its interior.
2 Preliminary concepts
We give first some definitions and notation. Let be a given planar convex body; for every real number we denote by the support line of with outward normal vector . The function , defined as , is known by the name of support function of . When the origin is contained in , is nothing else than the distance from to the support line The distance between the support lines and is called the width of in direction and it is denoted by , in other words, . If is constant, independently of , we say that is a body of constant width. For any , the -isoptic of is defined as the locus of points at which two tangent lines to intersect at an angle . Using the support function, is parameterized (see for instance [19]) by
(1) |
By Cauchy’s formula, the perimeter of can be obtained by (see [19])
(3) |
For any we define (see Fig. 1)
(4) | ||||
(5) | ||||
(6) | ||||
(7) | ||||
(8) |
We also define to be the distance between the points obtained by projecting the origin onto the support lines of at and .

By some tedious but simple calculations we can express the lengths of all these chords in terms of the support function of :
(9) | ||||
(10) | ||||
(11) | ||||
(12) |
Finally, the support function of a convex body is a periodic function, with period , and it is also absolutely continuous, so we can consider its expansion in terms of the Fourier series (see [6]), i.e.,
(13) |
The first and second derivatives of are expressed as
(14) | ||||
(15) |
3 Some results about isoptic curves in the plane
Our first result about isoptic chords is the following.
Theorem 1.
Let be a strictly convex body in the plane with differentiable boundary and let be a fixed angle such that is an irrational number. Suppose , for every , for a positive number Then is a disc.
Proof. Since we have by (11) that
it follows that
(16) |
If we substitute the Fourier coefficients of in the differential equation (16), by (14) we have
Since this holds for every real number , we must have that for every
The coefficients of and must be both equal to 0, hence we have that
The determinant of the matrix above is given by
This determinant is zero only if
(17) |
and
(18) |
Since , for every we have that none of (17) and (18) are satisfied if is an irrational number. Hence, we have that the determinant is non zero and then for every . It follows that , i.e., is the support function of a disc (see for instance [19]).
Remark 1.
If is a rational number, then there exist convex bodies different from discs for which is constant. For instance: for the angles and let be the convex body whose support function is given by (see Fig. 2). In this case the isoptic curves and have the property that its chords tangent to have constant values and , respectively, with , where . Moreover, is homothetic to with ratio of homothety equal to .

However, if we impose the additional condition that is also constant, then must be a disc.
Theorem 2.
Let be a strictly convex body in the plane with differentiable boundary , and let be a fixed angle. Suppose , and , for every , and for two positive numbers and . Then is a disc.
Proof. By some simple calculations we have that (see for instance [4]). Since is also constant, we have
hence
It follows that
which implies that , where is the angle between the vectors and , and is the angle between the vectors and . It follows that the angle between the chord and the tangent lines at and , are equal (see Fig. 3). Similarly, we obtain that the angles between the chord and the tangent lines at and , are equal. Since the length of the tangent vector is constant for every , and all the triangles are congruent, we also have that the angles between the chord and the tangent lines at and , are equal. By elementary geometry we have that the circle circumscribed to and the body , share the tangent lines at the points , , and ; under this condition it was proved in Lemma 3.3 in [8] that must be a disc. Now we use Theorem 2 (b) in [9] and conclude that is a disc.

Let denote the length of the segment from to the projection of into the support line of at . It is easy to show that
(19) |

We have the following result.
Theorem 3.
Let be a strictly convex body in the plane with differentiable boundary and let be a fixed angle. Suppose , for every , for a positive number Then is a disc.
Proof. By (19) we have that
(20) |
Using the expansion in Fourier series for (see equation (14)) we have that
Using trigonometric identities and simplifying we conclude that for all and for all we must have
This yields to the system of equations
Notice that the determinant of the matrix above is given by
This determinant is never zero, for if , then , for some integer . Nonetheless, in this case and since , it is impossible to have . It follows that the only solutions for the system of equations is for all . Thus, the solutions of the differential equation (20) are constant functions and must be a disc centred at .
Theorem 4.
Let be a strictly convex body in the plane and let be a fixed angle. Suppose , for every , for a positive number and has rotational symmetry of angle or . Then is a circle.
Proof. The vector can be expressed using the parametrization given in (2) by
If has rotational symmetry of angle or we have that
or equivalently
In this case can be easily calculated as
hence
i.e., the value of

Let and be the projections of into the support lines of at and , respectively (see Fig. 5). The quadrilateral is cyclic, i.e., there exist a circle which passes through its four vertices, hence It follows that is a circle centred at .
4 Some comments about rotors in the plane
In this section we give some words about how the results obtained in this work are related to rotors in polygons. Moreover, in all the examples shown below, if we fix the convex body and the circumscribed polygon is rotated, while maintained circumscribed to , the vertices describe a isoptic of .
When has constant value, using the Fourier series for the support function of in the proof of theorem 1 we arrived to the equations
and
If is even, then
and
Both of them are zero if or after some trigonometric transformations
It follows that or , where is any integer number. In this case the determinant of the associated matrix is zero and we can choose the coefficients , , arbitrarily. For instance, for we select , , , , and for any other natural number we have that , i.e., the support function of is (see Fig. 6). The body in this example is centrally symmetric.

If is odd, then
and
Both of them are zero if or after some trigonometric transformations
It follows that or , where is any integer number. Notice that we can obtain an example for any angle of the form , where and are integers such that . We just use this case, since , we select and or . For instance, for , and then we select . In Fig. 7 we show the body with its isoptic with the property that has a constant value. The support function for this example is . The body in this example has constant width.

The example shown in Fig. 8 has coefficients different from zero for and , consequently, the body obtained is neither of constant width or centrally symmetric. The support function of is

Finally, we give an example of a rotor in the square. The support function is .

5 An inequality about the length of some special chords
The following inequality and characterization of the disc, in terms of the length , was proved in [7].
Theorem JY. Let be a strictly convex body in the plane with minimal width . For any fixed there exists at least one value of the parameter , , such that . Moreover, if there is not such a chord with length exceeding , then is a disc.
Here we prove the following similar result.
Theorem 5.
Let be a convex body in the plane with minimal width . For any fixed there exists at least one value of the parameter , , such that . Moreover, if there is not such a chord with length exceeding , then is a convex body of constant width .
Proof. The mean value of the chords of that are tangent to is
Indeed, by the same argument, the mean value of is
Let be any number and consider the triangle with sides of length , , , and angle , as shown in Fig. 10. By Problem C4 in the book by I. Niven [15], we have that the minimum of , among all triangles with a fixed angle and the sum , for a constant number , is obtained if and only if . It follows that for every it holds that
with equality if and only if Let be such that we have that
(21) |

Using Cauchy’s formula for the perimeter of we can also prove that , with equality if and only is a body of constant width Hence
Now, if there is no chord with , then , which by (21) implies that i.e., must be a body of constant width (see Fig. 11).

6 Some results in higher dimensions
The following lemma is needed for some of the subsequent results.
Lemma 1.
Let , , be convex bodies with , such that every -dimensional section of tangent to is an -dimensional ball, then is a ball. If additionally, the centre of every tangent ball is at the point of contact with , then and are concentric balls.
Proof. There are several ways to prove this lemma. Here we give the following proof. First note that if all the -dimensional sections of a convex body through a given point in its interior are -dimensional balls, then it is an -dimensional ball. We consider any point in the interior of for such a point and since the hypothesis of the theorem is inheritated to every -dimensional section, we have that it is sufficient to prove the theorem in the -dimensional case.
Let be any point in and let be any line supporting at . Consider the two support planes of which share the line , to say and . Let be any other support plane of through the point . By hypothesis is a disc which intersects each one of the circles and at two points. The circle passes through three points of the set , and since there is a unique sphere which contains the two circles and , it holds that this sphere contains . Since is any support plane of through , we have that is a closed subset of a sphere. Let be any point such that , then is contained in the same sphere. Continuing in this way, since is a compact set, we can prove that is a sphere.
Now, let be any hyperplane supporting at a point and suppose the centre of the -dimensional ball is . The line orthogonal to through passes through the centre of the ball , this implies that indeed is the tangent plane of through , i.e., is a differentiable surface. We have that all the normal lines of passes through the centre of , this implies (see for instance [18]) that is a sphere with centre at the centre of . We conclude that and are concentric balls.
Let . In [5], J. W. Green proved that is a Euclidean disc if is a circle and any of the following conditions hold: is an irrational multiple of , or , with and relatively prime integer numbers. On the other side, M.S. Klamkin conjectured [11] and J. Nitsche proved [14] that two different values , , such that and are circles, are enough to prove that is a Euclidean disc. A similar result was proved by Á. Kurusa and T. Ódor in [12] in , , where the isoptic surface is defined using the solid angle under which is seen a convex body . They proved that if two isoptic surfaces of a given convex body are concentric spheres, then it is a ball. The following is a variant of this kind of results.
Theorem 6.
Let be a strictly convex body contained in the interior of a sphere . Suppose that for every -dimensional plane , with , it holds that is a isoptic curve of Then is a ball concentric with .
Proof. We first observe the following: If for some angle the -isoptic curve of a planar body is a circle , then its centre belongs to . Suppose this is not the case. Let be the point which is closest to . Let be the point where the ray intersects to and let be the point in such that is a diameter. Let be the line through which is orthogonal to . The central projection of into the line is a segment that contains the centre . The angle formed by the lines and contains , however, at least one of the lines or does not intersect . Since the angles and are equal, we obtain that the angles under which is seen from the points and are not equal. This contradicts that is the isoptic curve of angle of . Hence, we have that belongs to .

Now let be any arbitrary point and let be a plane supporting at . By hypothesis we have that is a circle; we are going to prove that its centre is . Consider a sequence of planes parallel to in such a way that the sequence converges to . Let be the centre of the circle , for every natural number . By the comment above, we have that , for every , and since we have that when . In other words, the centre of is . Now we apply Lemma 1 and conclude that is a ball concentric with .
The following result shows that in dimension 3 or higher, only Euclidean balls have an equichordal convex body.
Theorem 7.
Let , , be a strictly convex body which possesses an equichordal convex body in its interior. Then and are concentric balls.
Proof. Suppose the length of the chords of tangent to is . First we prove that is strictly convex, i.e., has not segments in its boundary. Suppose to the contrary that there is a segment and consider any -dimensional plane supporting at . By hypothesis, every chord of through has length , and the same happens for every chord through . This implies that has two equichordal points, but as were proved by M. R. Rychlik [16] this is not possible.
Now, let be any -dimensional plane supporting at a point . We will prove that the diameter of is . Suppose this is not the case and there is a chord of with Clearly, Consider a -dimensional plane that contains to and such that Let such that the lines through and , respectively, that are parallel to are support lines of . Since , we have that there exist a chord of that is separated from by the chord and . We have three chords of parallel to and with length . This contradicts that is strictly convex, then the diameter of must be .
Every chord of through has length equal to , then every chord of through is a binormal of ; it follows that is a disc centred at (see for instance [3]). Since this is true for every -dimensional plane through , we have that any -dimensional section of tangent to at is an -dimensional sphere centred at . This is also true for every , so we apply Lemma 1 and conclude that and are concentric balls.
Corollary 1.
Let , , be two convex bodies. Suppose that for every hyperplane that intersects to it holds that is an equichordal body of . Then and are concentric balls.
Proof. It is quite simple to prove that all the chords of tangent to have the same length. The conclusion follows applying Theorem 7.
Suppose now that is a strictly convex body in . For , we will say that the segment is an -chord of , if there exist tangent planes at and meeting at an angle . We present the following analogue to Theorem JY.
Theorem 8.
Let be a strictly convex body in with boundary of class and with minimal width . For any fixed there exists an -chord of with length at least . Moreover, if there is not an -chord with length exceeding , then is a ball.
Proof. Let be a direction such that the width of in this direction has minimal value . Let be a vector orthogonal to and let’s denote the orthogonal projection on the plane by . Clearly, the minimum width of is also and every -chord of is obtained as the projection of some -chord of . By Theorem JY there is an -chord of with length at least and so there is an -chord of with length larger than or equal to .
Suppose now that all the -chords of have length less than or equal to . Then all the -chords of , for every , have length less than or equal to . It follows from Theorem JY that is a disc. Since is an arbitrary direction, we have that every -dimensional projection of is a disc and therefore is a ball.
Theorem 9.
Let be a strictly convex body in with boundary of class and with -chords of constant length , then is a ball.
Proof. For any point , let denote the tangent plane of at and let . For each point in , we have that intersects with angle . Let be two points such that is normal to and . For every , the triangle is isosceles with sides and of equal length. Then is a circle centred at . Moreover, given any two antipodal points and of , the segment must be orthogonal to the planes and . By repeating the previous argument to each pair of antipodal points of , we have that any section of containing is a disc. This implies that is a ball.
References
- [1] J. A. Barker, and D. G. Larman, Determination of convex bodies by certain sets of sectional volumes. Discrete Math. 241 (2001), .
- [2] W. Blaschke, W. Rothe, and R. Weitzenböck, Aufgabe 552, Arch. Math. Phys. 27 (1917), .
- [3] G. D. Chakerian, and H. Groemer. Convex bodies of constant width, in Convexity and its Applications, Eds. P.M. Gruber and J.M. Wills, Birkhäuser, Basel, (1983).
- [4] W. Cieślak, A. Miernowski, and W. Mozgawa. Isoptics of a closed strictly convex curve, Lecture Notes in Mathematics 1481 (1991), .
- [5] J. W. Green. Sets subtending a constant angle on a circle, Duke Math. J. 17 (1950), .
- [6] H. Groemer. Geometric applications of Fourier series and spherical harmonics, Cambridge Univ. Press, Cambridge (1996).
- [7] J. Jerónimo-Castro, and C. Yee-Romero, An inequality for the length of isoptic chords of convex bodies. Aequat. Math. 93 (3) (2019), .
- [8] J. Jerónimo-Castro, and F. G. Jimenez-Lopez. Symmetries of convex sets in the hyperbolic plane, J. Conv. Analysis 26 (2019), .
- [9] J. Jerónimo-Castro, F. G. Jimenez-Lopez, and R. Jiménez-Sánchez. On convex bodies with isoptic chords of constant length, Aequationes Math. 94 (2020), no. 6, .
- [10] J. Jerónimo-Castro, M. A. Rojas-Tapia, U. Velasco-García, and C. Yee-Romero. An isoperimetric type inequality for isoptic curves of convex bodies, Results Math. 75 (2020).
- [11] M. S. Klamkin. Conjectured isoptic characterization of a circle, Amer. Math. Monthly 95 (1988), .
- [12] Á. Kurusa, and T. Ódor. Isoptic characterization of spheres, J. Geom. 106 (2015), .
- [13] H. Martini, L. Montejano, D. Oliveros. Bodies of Constant Width: An Introduction to Convex Geometry with Applications, Birkhäuser, (2019).
- [14] J. Nitsche. Isoptic Characterization of a Circle (Proof of a conjecture of M. S. Klamkin), Amer. Math. Monthly 97 (1990), .
- [15] I. Niven. Maxima and Minima without Calculus, Dolciani Mathematical Expositions No. 6, (1981).
- [16] M. R. Rychlik. A complete solution to the equichordal point problem of Fujiwara, Blaschke, Rothe and Weitzenböck, Invent. Math. 129 (1997), no. 1, .
- [17] C. Schütt, and E. Werner. The convex floating body, Math. Scand. 66 (1990),
- [18] V. A. Toponogov. Differential geometry of curves and surfaces, a concise guide, Birkhäuser, Boston-Basel-Berlin (2006).
- [19] F. A. Valentine. Convex sets. McGraw-Hill Series in Higher Mathematics, McGraw-Hill, New York, (1964).
- [20] F. Wegner. Floating bodies of equilibrium, Stud. Appl. Math. 111 (2003), no. 2, .
- [21] I. M. Yaglom, and V. G. Boltyanski. Convex Figures, Holt, Rinehart and Winston (1961)
This manuscript has not any associated data.