This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Loci of 3-periodics in an Elliptic Billiard:
Why so many ellipses?

Ronaldo Garcia Ronaldo Garcia
Inst. de Matemática e Estatística
Univ. Federal de Goiás
Goiânia, GO, Brazil
ragarcia@ufg.br
Jair Koiller Jair Koiller
Dept. de Matemática
Univ. Federal de Juiz de Fora
Juiz de Fora, MG, Brazil
jairkoiller@gmail.com
 and  Dan Reznik Dan Reznik
Data Science Consulting
Rio de Janeiro, RJ, Brazil
dan@dat-sci.com
Abstract.

A triangle center such as the incenter, barycenter, etc., is specified by a function thrice- and cyclically applied on sidelengths and/or angles. Consider the 1d family of 3-periodics in the elliptic billiard, and the loci of its triangle centers. Some will sweep ellipses, and others higher-degree algebraic curves. We propose two rigorous methods to prove if the locus of a given center is an ellipse: one based on computer algebra, and another based on an algebro-geometric method. We also prove that if the triangle center function is rational on sidelengths, the locus is algebraic.

Keywords: Poncelet, porism, locus, resultants, algebraic

MSC2010 37-40 and 51N20 and 51M04 and 51-04

1. Introduction

Classic notable points of a triangle include the incenter, barycenter, orthocenter, and circumcenter, see Appendix A.1 for a refresher. While these are traditionally obtained via geometric constructions, Kimberling has generalized this to triangle centers [15], specified by a triangle function, thrice-applied (in cyclical fashion) to the sidelengths and/or angles of a triangle, thus forming a triple of trilinear coordinates. These are reviewed in Appendix A.2.

Thousands of such centers are catalogued in Kimberling’s Encyclopedia of Triangle Centers (ETC) [16]. Centers are labeled XiX_{i}, e.g., the incenter is X1X_{1}, the barycenter X2X_{2}, etc. For each center the corresponding triangle function is provided. A quick perusal reveals these to be either rational, irrational, or more rarely, transcendental, on the sidelengths and/or angles of a triangle.

Consider the 1d family of 3-periodic orbits in the elliptic billiard (EB), Figure 1. The vertices are bisected by ellipse normals and the sides are tangent to a virtual confocal caustic; see Appendix B for a review. We have been drawn to this family because unexpectedly, the locus of the incenter is an ellipse and that of the Mittenpunkt111This point was discovered by Nagel in 1836 as the point of concurrence of lines from the excenters through side midpoints [33], see Figure 10 (right). X9X_{9} is the billiard center [23]. Additionally, many other curious invariants have been detected and/or proved [1, 2, 24].

Refer to caption
Figure 1. Three 3-periodic orbits (one solid with vertex PP and two dashed with vertices PP^{\prime}, P′′P^{\prime\prime}). The vertices are bisected by ellipse normals and the sides are dynamically tangent to a virtual confocal caustic (brown). Remarkably, the family conserves perimeter [31]. app

In general, triangle centers sweep such curves as ellipses, quartics, sextics, etc., with or without self-intersections, etc.; see Figure 2. The central question here is: given a triangle center, is it possible to predict its locus curve type over billiard 3-periodics based on the triangle function?

Refer to caption
Figure 2. Left: A sample 3-periodic orbit (blue), the elliptic locus of the Incenter X1X_{1} (green) and of that of the Intouchpoints I1,I2,I3I_{1},I_{2},I_{3} (brown), the points of contact of the Incircle (dashed green) with the orbit sides. These produce a curve with two internal lobes whose degree is at least 6. Y1(ρ)Y_{1}(\rho) is a convex combination of X1X_{1} and I1I_{1} referred to in Section 2.5. app, Video 1, Video 2. Right: The locus of X59X_{59} is a curve with four self-intersections. A vertical line epsilon away from the origin intersects the locus at 6 points. app

Main Results

We present a hybrid numerical-CAS (Computer Algebra System) method to rigorously verify if the locus of given triangle center is an ellipse or not. We apply it to the first 100 centers listed in [16], finding that 29 are elliptic. We have derived explicit expressions for their axes [9] (Theorem 1).

We consider the curious case of the Symmedian Point X6X_{6}, whose locus closely approximates an ellipse but using CAS-based manipulation, we show it is actually a quartic (Theorem 2), which we derive explicitly. Interestingly, its triangle function is rational on the sidelengths and is one of the simplest on the whole of [16]; see Table 6 in Appendix A.1.

Using algebro-geometric techniques, we prove (Theorem 3) that when the trilinear coordinates (defined in Appendix A.2) of a triangle center are rational on the orbit’s sidelengths, the locus is an algebraic curve (elliptic or otherwise). We also describe an algorithm based on the method of resultants which computes the zero set of a polynomial in two variables corresponding to the locus. After extensive experimentation, we haven’t yet found a non-rational triangle function which results in an elliptic locus, so it is likely that the latter requires a rational triangle function.

Related Work

Odehnal extensively studied loci of triangle centers over the poristic triangle family [19]. The early experimental result that the locus of the Incenter X1X_{1} of billiard 3-periodics is an ellipse was subsequently proven [25, 7]. Proofs soon followed for the ellipticity of both X2X_{2} [28] and X3X_{3} [5, 7]; see Figures 16 and 15 in Appendix E. More recently, a theory for the ellipticity of a locus based on certain properties of a triangle is being developed, see [14]. For a textbook on Poncelet-based phenomena see [8].

Outline

Our main methods appear in Sections 2 and 3. We conclude in Section 4 with a list of questions and interesting links and videos. The Appendices contain supporting material. Most figures contain clickable links to relevant videos and/or the experiment displayed on our browser-based app [3].

2. Detecting Elliptic Loci

Let the boundary of the EB be given by (a>b>0a>b>0):

(1) f(x,y)=(xa)2+(yb)2=1.f(x,y)=\left(\frac{x}{a}\right)^{2}+\left(\frac{y}{b}\right)^{2}=1.

We describe a numerically-assisted method222We started with visual inspection, but this is both laborious and unreliable, some loci (take X30X_{30} and X6X_{6} as examples) are indistinguishable from ellipses to the naked eye. which proves that the locus of a given triangle center is elliptic or otherwise. We then apply it to the first 100 triangle centers. listed in [16]. Indeed, a much larger list could be tested.

2.1. Proof Method

Our proof method consists of two phases, one numeric, and one symbolic, Figure 4. It makes use of the following Lemmas, whose proofs appear in Appendix D:

Lemma 1.

The locus of a triangle center XiX_{i} is symmetric about both EB axes and centered on the latter’s origin.

Proof.

Given a 3-periodic T=P1P2P3T=P_{1}P_{2}P_{3}. Since the EB is symmetric about its axes, the family will contain the reflection of TT about said axes, call these TT^{\prime} and T′′T^{\prime\prime}. Since triangle centers are invariant with respect to reflections, Appendix A, XiX_{i}^{\prime} and Xi′′X_{i}^{\prime\prime} will be found at similary reflected locations. ∎

Note that if the locus is elliptic, the above implies it will be concentric and axis-aligned with the EB.

Lemma 2.

Any triangle center XiX_{i} of an isosceles triangle is on the axis of symmetry of said triangle.

Lemma 3.

A parametric traversal of P1P_{1} around the EB boundary triple covers the locus of triangle center XiX_{i}, elliptic or not.

See Section 2.5 for more details.

A first phase fits a concentric, axis-aligned ellipse to a fine sampling of the locus of some triangle center XiX_{i}. A good fit occurs when the error is several orders of magnitude333Robust fitting of ellipses to a cloud of points is not new [6]. In our case, the only source of error in triangle center coordinates is numerical precision, whose propagation can be bounded by Interval Analysis [18, 29]. less than the sum of axes regressed by the process. False negatives are eliminated by setting the error threshold to numeric precision. False positives can be produced by adding arbitrarily small noise to samples of a perfect ellipse, though this type of misclassification does not survive the next, symbolic phase.

A second phase attempts to symbolically verify via a Computer Algebra System (CAS) if the parametric locus of the XiX_{i} satisfies the equation of a concentric, axis-aligned ellipse. Expressions for its semi-axes are obtained by evaluating XiX_{i} at isosceles orbit configurations, Figure 3. The method is explained in detail in Figure 5.

Refer to caption
Figure 3. With P1P_{1} at the right (resp. top) vertex of the EB, the orbit is a sideways (resp. upright) isosceles triangle, solid red (resp. solid blue). Not shown are their two symmetric reflections. Also shown (green) is the locus of a sample triangle center, X1X_{1} in this case. At the isosceles positions, vertices will lie on the axis of symmetry of the triangle, Lemma 2. When the locus is elliptic, the x,yx,y coordinates of Xi(0),Xi(π/2)X_{i}(0),X_{i}(\pi/2) are the semi-axes ai,bia_{i},b_{i}, respectively.
Refer to caption
Figure 4. Our method as a flow-chart, modules labeled from I to V: (I) Once the ith center is specified, its trilinears p:q:rp:q:r (see Appendix A) are obtained from the Encyclopedia of triangle centers (ETC) [16]. (II) Given a symbolic parameter tt or a numeric sample tkt_{k}, obtain cartesian coordinates for orbit vertices and the sidelengths, Appendix C.2. (III) Combine trilinears and orbit data to obtain, via (3), numeric cartesian coordinates for Xi(tk)X_{i}(t_{k}) or symbolically as Xi(t)X_{i}(t). (IV) Least-squares fit an axis-aligned, concentric ellipse to the (xk,yk)(x_{k},y_{k}) samples and accept XiX_{i} as candidate if the fit error is sufficiently small. (V) Verify, via a Computer Algebra System (CAS), if the parametric locus Xi(t)X_{i}(t) satisfies the equation of an ellipse whose axes ai,bia_{i},b_{i} are obtained symbolically in II-III by setting t=0,π/2t=0,\pi/2. If the CAS is successful, Xi(t)X_{i}(t) is deemed elliptic, otherwise the result is “undecided”. The CAS was successful for all 29 candidates selected by IV out of the first 100 Kimberling centers.

1. Select Candidates: Let the EB have axes a,ba,b such that a>b>0a>b>0. Calculate P1(tk)=(acos(tk),bsin(tk))P_{1}(t_{k})=\left(a\cos(t_{k}),b\sin(t_{k})\right), for MM equally-spaced samples tk[0,2π)t_{k}\in[0,2\pi), k=1,2,,Mk=1,2,...,M. Obtain the cartesian coordinates for the orbit vertices P2(tk)P_{2}(t_{k}) and P3(tk)P_{3}(t_{k}), k\forall k (Appendix C.2). Obtain the cartesian coordinates for triangle center XiX_{i} from its trilinears (3), for tk\forall{t_{k}}. If analyzing the vertex of a derived triangle, convert a row of its trilinear matrix to cartesian coordinates, Appendix A.3. Least-squares fit an origin-centered, axis-aligned ellipse (2 parameters) to the Xi(tk)X_{i}(t_{k}) samples, Lemma 1. Accept the locus as potentially elliptic if the numeric fit error is negligible, rejecting it otherwise. 2. Verify with CAS Taking a,ba,b as symbolic variables, calculate Xi(0)X_{i}(0) (resp. X(π/2)X(\pi/2)), placing P1P_{1} at the right (resp. top) EB vertex. The orbit will be a sideways (resp. upright) isosceles triangle, Figure 3. By Lemma 2, XiX_{i} will fall along the axis of symmetry of either isosceles. The xx coordinate of Xi(0)X_{i}(0) (resp. the yy of Xi(π/2)X_{i}(\pi/2)) will be symbolic expressions in a,ba,b. Use them as candidate locus semiaxes’ lengths ai,bia_{i},b_{i}. Taking tt as a symbolic variable, use a computer algebra system (CAS) to verify if Xi(t)=(xi(t),yi(t))X_{i}(t)=\left(x_{i}(t),y_{i}(t)\right), as parametrics on tt, satisfy (xi(t)/ai)2+(yi(t)/bi)2=1(x_{i}(t)/a_{i})^{2}+(y_{i}(t)/b_{i})^{2}=1, t{\forall}t. If the CAS is successful, Lemma 3 guarantees Xi(t)X_{i}(t) is will cover the entire ellipse, so assert that the locus of XiX_{i} is an ellipse. Else, locus ellipticity is indeterminate.

Figure 5. Method for detecting ellipticity of a triangle center locus.

2.2. Phase 1: Least-Squares-Based Candidate Selection

Let the position of Xi(t)X_{i}(t) be sampled (randomly or uniformly) at tk[0,2π]t_{k}\in[0,2\pi], k=1Mk=1{\ldots}M. If the locus is an ellipse, than the latter is concentric and axis-aligned with the EB, Lemma 1. Express the squared error as the sum of squared sample deviations from an implicit ellipse:

err2(ai,bi)=k=1M[(xkai)2+(ykbi)21]2\text{err}^{2}(a_{i},b_{i})=\sum_{k=1}^{M}{\left[\left(\frac{x_{k}}{a_{i}}\right)^{2}+\left(\frac{y_{k}}{b_{i}}\right)^{2}-1\right]^{2}}

Least-squares can be used to estimate the semi-axes:

(ai^,bi^)=argmina,b{err2(ai,bi)}(\hat{a_{i}},\hat{b_{i}})=\operatorname*{arg\,min}_{a,b}\left\{\text{err}^{2}(a_{i},b_{i})\right\}

The first 100 Kimberling centers separate into two distinct clusters: 29 with negligible least-squares error, and 71 with finite ones. These are shown in ascending order of error in our companion website [9, Part II].

A gallery of loci generated by X1X_{1} to X100X_{100} (as well as vertices of several derived triangles) is provided in [20].

2.3. Phase 2: Symbolic Verification with a CAS

A CAS was successful in symbolically verifying that all 29 candidates selected in Phase 1 satisfy the equation of an ellipse (none were undecided). As an intermediate step, explicit expressions for their elliptic semi-axes were computed and appear in [9, Part I].

Theorem 1.

Out of the first 100 centers in [16], exactly 29 produce elliptic loci, all of which are concentric and axis-aligned with the EB. These are XiX_{i},i=1, 2, 3, 4, 5, 7, 8, 10, 11, 12, 20, 21, 35, 36, 40, 46, 55, 57, 63, 65, 72, 78, 79, 80, 84, 88, 90. Specifically:

  • The loci of Xi,i=2,7,57,63X_{i},i=2,7,57,63 are ellipses homothetic to the EB.

  • The loci of Xi,i=4,10,40X_{i},i=4,10,40 are ellipses homothetic to a 9090^{\circ}-rotated copy of the EB.

  • The loci of Xi,i=88,100X_{i},i=88,100 are ellipses identical to the EB.

  • The loci of X55X_{55} is an ellipse homothetic to the N=3N=3 caustic.

  • The loci of Xi,i=3,84X_{i},i=3,84 are ellipses homothetic to a 9090^{\circ}-rotated copy of the N=3N=3 caustic.

  • The locus of X11X_{11} is an ellipse identical to the N=3N=3 caustic.

The above above are summarized on Table 1. The least-square fit errors for the first 100 Kimberling are shown in Figure 6.

Remark 1.

The loci of XiX_{i}, i=1,2,3,4i=1,2,3,4 are the ellipses x2/ai2+y2/bi2=1x^{2}/a_{i}^{2}+y^{2}/b_{i}^{2}=1. Via CAS, their semi-axes (ai,bi)(a_{i},b_{i}) are given by:

a1\displaystyle a_{1} =δb2a\displaystyle=\frac{\delta-b^{2}}{a} b1\displaystyle b_{1} =a2δb\displaystyle=\frac{a^{2}-\delta}{b}
a2\displaystyle a_{2} =k2a\displaystyle=k_{2}a b2\displaystyle b_{2} =k2b\displaystyle=k_{2}b
a3\displaystyle a_{3} =a2δ2b\displaystyle=\frac{a^{2}-\delta}{2b} b3\displaystyle b_{3} =δb22a\displaystyle=\frac{\delta-b^{2}}{2a}
a4\displaystyle a_{4} =k4a\displaystyle=\frac{k_{4}}{a} b4\displaystyle b_{4} =k4b\displaystyle=\frac{k_{4}}{b}

where δ2=a4+b4a2b2\delta^{2}=a^{4}+b^{4}-a^{2}b^{2}, k2=(2δa2b2)/c2k_{2}=(2\delta-a^{2}-b^{2})/c^{2}, k4=[(a2+b2)δ2a2b2)]/c2k_{4}=[(a^{2}+b^{2})\delta-2a^{2}b^{2})]/c^{2}, and c2=a2b2c^{2}=a^{2}-b^{2}.

Explicit expressions for the locus semi-axes for abovementioned centers appear in [9, Part I].

Refer to caption
Figure 6. Log of Least-squares error for first 100 Kimberling centers in ascending order of error, for three values of a/ba/b, M=1500M=1500. Elliptic vs. non-elliptic centers are clearly separated in two groups whose errors differ by several orders of magnitude. A table in [9, Part II] shows that X37X_{37} and X6X_{6} are have ranks 3030 and 3131 respectively, i.e., they are the centers whose non-elliptic loci are closest to a perfect ellipse.
rowXidefinitionsim11IncenterJt22BarycenterB33CircumcenterCt44OrthocenterBt559-Point Center67Gergonne PointB78Nagel Point810Spieker CenterBt911Feuerbach PointC+1012{X1,5}-Harm.Conj. of X111120de Longchamps Point1221Schiffler Point1335{X1,3}-Harm.Conj. of X361436Inverse-in-Circumc. of X11540Bevan PointBt\scriptsize\begin{array}[]{|c|c|l|c|}\hline\cr\text{row}&X_{i}&\text{definition}&\text{sim}\\ \hline\cr 1&{1}&\text{Incenter}&\text{J}^{t}\\ 2&{2}&\text{Barycenter}&\text{B}\\ 3&{3}&\text{Circumcenter}&\text{C}^{t}\\ 4&{4}&\text{Orthocenter}&\text{B}^{t}\\ 5&{5}&\text{9-Point Center}&\\ 6&{7}&\text{Gergonne Point}&\text{B}\\ 7&{8}&\text{Nagel Point}&\\ 8&{10}&\text{Spieker Center}&\text{B}^{t}\\ 9&{11}&\text{Feuerbach Point}&\text{C}^{+}\\ 10&{12}&\text{$\{X_{1,5}\}$-Harm.Conj. of $X_{11}$}&\\ 11&{20}&\text{de Longchamps Point}&\\ 12&{21}&\text{Schiffler Point}&\\ 13&{35}&\text{$\{X_{1,3}\}$-Harm.Conj. of $X_{36}$}&\\ 14&{36}&\text{Inverse-in-Circumc. of $X_{1}$}&\\ 15&{40}&\text{Bevan Point}&\text{B}^{t}\\ \hline\cr\end{array}
rowXidefinitionsim1646X4-Ceva Conj. of X11755Insimilictr(Circumc.,Incir.)C1856Exsimilictr(Circumc.,Incir.)1957Isogonal Conj. of X9B2063Isogonal Conj. of X19B2165Intouch Triangle’s X42272Isogonal Conj. of X28J2378Isogonal Conj. of X342479Isogonal Conj. of X352580Refl. of X1 about X11Jt2684Isogonal Conj. of X40Ct2788Isogonal Conj. of X44B+2890X3-Cross Conj. of X129100Anticomplement of X11B+\scriptsize\begin{array}[]{|c|c|l|c|}\hline\cr\text{row}&X_{i}&\text{definition}&\text{sim}\\ \hline\cr 16&{46}&\text{$X_{4}$-Ceva Conj. of $X_{1}$}&\\ 17&{55}&\text{Insimilictr(Circumc.,Incir.)}&\text{C}\\ 18&{56}&\text{Exsimilictr(Circumc.,Incir.)}&\\ 19&{57}&\text{Isogonal Conj. of $X_{9}$}&\text{B}\\ 20&{63}&\text{Isogonal Conj. of $X_{19}$}&\text{B}\\ 21&{65}&\text{Intouch Triangle's $X_{4}$}&\\ 22&{72}&\text{Isogonal Conj. of $X_{28}$}&\text{J}\\ 23&{78}&\text{Isogonal Conj. of $X_{34}$}&\\ 24&{79}&\text{Isogonal Conj. of $X_{35}$}&\\ 25&{80}&\text{Refl. of $X_{1}$ about $X_{11}$}&\text{J}^{t}\\ 26&{84}&\text{Isogonal Conj. of $X_{40}$}&C^{t}\\ 27&{88}&\text{Isogonal Conj. of $X_{44}$}&\text{B}^{+}\\ 28&{90}&\text{$X_{3}$-Cross Conj. of $X_{1}$}&\\ 29&{100}&\text{Anticomplement of $X_{11}$}&\text{B}^{+}\\ \hline\cr\end{array}
Table 1. The 29 Kimberling centers within X1X_{1} to X100X_{100} with elliptic loci. Under column “sim.”, letters B,C,J indicate the locus is similar to EB, caustic, or Excentral locus, respectively. An additional + (resp. t) exponent indicates the locus is identical (resp. similar to a perpendicular copy) to the indicated ellipse. Note: the ellipticity of XiX_{i},i=1,2,3,4i=1,2,3,4 was previously proven [25, 32, 5, 7].

2.4. The quartic locus of the Symmedian Point

The construction of the Symmedian point X6X_{6} of a triangle is shown in Figure 7.

Refer to caption
Figure 7. Given a triangle ABCABC, the Symmedian point X6X_{6} is the point of concurrence of the three symmedians (pink), where the latter is a reflection of a median (brown) about the corresponding angular bisector (green). Medians (resp. angular bisectors) concur at the barycenter X2X_{2} (resp. the incenter X1X_{1}).

Over 3-periodic orbits in an EB with a/b=1.5a/b=1.5, the locus of X6X_{6} is visually indistinguishable from an ellipse, Figure 8. Fortunately, its fit error is 10 orders of magnitude higher than the ones produced by true elliptic loci, see [9, Part II]. So it is easily rejected by the least-squares phase. Indeed, symbolic manipulation with a CAS yields:

Theorem 2.

The locus of X6X_{6} is a convex quartic given by:

𝒳6(x,y)=c1x4+c2y4+c3x2y2+c4x2+c5y2=0\mathcal{X}_{6}(x,y)=c_{1}x^{4}+c_{2}y^{4}+c_{3}x^{2}y^{2}+c_{4}x^{2}+c_{5}y^{2}=0

where:

c1=b4(5δ24(a2b2)δa2b2)c2=a4(5δ2+4(a2b2)δa2b2)c3=2a2b2(a2b2+3δ2)c4=a2b4(3b4+2(2a2b2)δ5δ2)c5=a4b2(3a4+2(2b2a2)δ5δ2)δ=a4a2b2+b4\begin{array}[]{rlrl}c_{1}=&b^{4}(5\delta^{2}-4(a^{2}-b^{2})\delta-a^{2}b^{2})&c_{2}=&a^{4}(5\delta^{2}+4(a^{2}-b^{2})\delta-a^{2}b^{2})\\ c_{3}=&2a^{2}b^{2}(a^{2}b^{2}+3\delta^{2})&c_{4}=&a^{2}b^{4}(3b^{4}+2(2a^{2}-b^{2})\delta-5\delta^{2})\\ c_{5}=&a^{4}b^{2}(3a^{4}+2(2b^{2}-a^{2})\delta-5\delta^{2})&\delta=&\sqrt{a^{4}-a^{2}b^{2}+b^{4}}\end{array}
Proof.

Using a CAS, obtain symbolic expressions for the coefficients of a quartic symmetric about both axes (no odd-degree terms), passing through 5 known-points. Still using a CAS, verify the symbolic parametric for the locus satisfies the quartic. ∎

Note the above is also satisfied by a degenerate level curve (x,y)=(0,0)(x,y)=(0,0), which we ignore.

Remark 2.

The axis-aligned ellipse 6\mathcal{E}_{6} with semi-axes a6,b6a_{6},b_{6} is internally tangent to 𝒳6(x,y)=0\mathcal{X}_{6}(x,y)=0 at the four vertices where:

(2) a6=[(3a2b2)δ(a2+b2)b2]aa2b2+3δ2,b6=[(a23b2)δ+(a2+b2)a2]ba2b2+3δ2\displaystyle a_{6}=\frac{\left[(3\,a^{2}-b^{2})\delta-(a^{2}+b^{2})b^{2}\right]a}{a^{2}b^{2}+3\delta^{2}},\;\;\;b_{6}=\frac{\left[(a^{2}-3\,b^{2})\delta+(a^{2}+b^{2})a^{2}\right]b}{a^{2}b^{2}+3\delta^{2}}
Refer to caption
Figure 8. An a/b=1.5a/b=1.5 EB is shown (black) as well as a sample 3-periodic (blue). At this aspect ratio, the locus of X6X_{6} (green) is indistinguishable to the naked eye from a perfect ellipse. To see it is non-elliptic, consider the locus of a point X6(t)=X6(t)+k|X6(t)Y6(t)|{\sim}X_{6}(t)=X_{6}(t)+k|X_{6}(t)-Y_{6}^{\prime}(t)|, with k=2×106k=2{\times}10^{6} and Y6(t)Y_{6}^{\prime}(t) the intersection of OX6(t)OX_{6}(t) with a best-fit ellipse (visually indistinguishable from green), (2). app

Table 2 shows the above coefficients numerically for a few values of a/ba/b.

a/ba6b6c1/c3c2/c3c4/c3c5/c3A(6)/A(𝒳6)1.250.4330.2820.2111.1850.0400.0950.99991.500.8740.4270.1142.1840.0870.3990.99982.001.6120.5490.0524.8500.1341.4610.99833.002.7910.6200.02012.4230.1574.7690.9949\begin{array}[]{|c|c|c|c|c|c|c|c|}\hline\cr\text{a/b}&a_{6}&b_{6}&c_{1}/c_{3}&c_{2}/c_{3}&c_{4}/c_{3}&c_{5}/c_{3}&A(\mathcal{E}_{6})/A(\mathcal{X}_{6})\\ \hline\cr 1.25&0.433&0.282&0.211&1.185&-0.040&-0.095&0.9999\\ 1.50&0.874&0.427&0.114&2.184&-0.087&-0.399&0.9998\\ 2.00&1.612&0.549&0.052&4.850&-0.134&-1.461&0.9983\\ 3.00&2.791&0.620&0.020&12.423&-0.157&-4.769&0.9949\\ \hline\cr\end{array}
Table 2. Coefficients ci/c3c_{i}/c_{3}, i=1,2,4,5i=1,2,4,5 for the quartic locus of X6X_{6} as well as the axes a6,b6a_{6},b_{6} for the best-fit ellipse, for various values of a/ba/b. The last-column reports the area ratio of the internal ellipse 6\mathcal{E}_{6} (with axes a6,b6a_{6},b_{6}) to that of the quartic locus 𝒳6\mathcal{X}_{6}, showing an almost exact match.

2.5. Locus Triple Winding

As an illustration of Lemma 3, consider the elliptic locus of X1X_{1}, the Incenter444The same argument is valid for the non-elliptic locus of, e.g., X59X_{59}, Figure 2 (right).. Consider the locus of a point Y1Y_{1} located on between X1X_{1} and an Intouchpoint I1I_{1}, Figure 2 (left):

Y1(t;ρ)=(1ρ)X1(t)+ρI1(t),ρ[0,1]Y_{1}(t;\rho)=(1-{\rho})X_{1}(t)+{\rho}I_{1}(t),\;\;\;\rho\in[0,1]

When ρ=1\rho=1 (resp. 0), Y1(t)Y_{1}(t) is the two-lobe locus of the Intouchpoints (resp. the elliptic locus of X1X_{1}). With ρ\rho just above zero, Y1Y_{1} winds thrice around the EB center. At ρ=0\rho=0, the two lobes and the remainder of the locus become one and the same: Y1Y_{1} winds thrice over the locus of the Incenter, i.e., the latter is the limit of such a convex combination.

It can be shown that at ρ=ρ\rho=\rho^{*}, with ρ=1(b/a)2\rho^{*}=1-(b/a)^{2}, the two Y1(t)Y_{1}(t) lobes touch at the the EB center. When ρ>ρ\rho>\rho^{*} (resp. ρ<ρ\rho<\rho^{*}), the locus of Y1(t)Y_{1}(t) has winding number 1 (resp. 3) with respect to the EB center, see Figure 9.

A similar phenomenon occurs for loci of convex combinations of the following pairs: (i) Barycenter X2X_{2} and a side midpoint, (ii) Circumcenter X3X_{3} and a side midpoint, (iii) Orthocenter X4X_{4} and altitude foot, etc., see [21, pl#11,12].

Refer to caption
Figure 9. An a/b=1.5a/b=1.5 EB is shown (black) as well as the elliptic locus of X1X_{1} (green) and the locus of Y1(t)Y_{1}(t) (pink), the convex combination of X1(t)X_{1}(t) and an Intouchpoint given by a parameter ρ[0,1]\rho\in[0,1], see Figure 2(left). At ρ=1\rho=1 (top-left), Y1(t)Y_{1}(t) is the two-lobe locus of the Intouchpoint. For every tour of an orbit vertex P1(t)P_{1}(t) around the EB, Y1(t)Y_{1}(t) winds once over its locus. At ρ=0.8\rho=0.8 (top-right) the lobes approach each other but still lie in different half planes. At ρ=ρ=1(b/a)2\rho=\rho^{*}=1-(b/a)^{2}, the lobes touch at the EB center (bottom-left). If ρ(0,ρ)\rho\in(0,\rho^{*}), the two lobes self-intersect twice. As ρ0\rho{\rightarrow}0, the two lobes become nearly coincidental (bottom-right). At ρ=0\rho=0, the Y1Y_{1} locus with its two lobes all collapse to the Incenter locus ellipse (green), in such a way that for every tour of P1(t)P_{1}(t) around the EB, X1X_{1} winds thrice over its locus. Video 1, Video 2

3. Toward a Typification of Loci

Our use of a few dozen centers listed in [16] was a means to validate our approach. In general we would like to predict locus type based on any triangle function, hand-curated or not. Below we take a few steps toward building a practical Computational Algebraic Geometry context useful for practitioners.

3.1. Trilinears: No Apparent Pattern

When one looks at a few examples of triangle centers whose loci are elliptic vs non, one finds no apparent algebraic pattern in said trilinears, Table 6 in Appendix A.1.

A few observations include:

  • The locus of a triangle center is symmetric about both EB axes, Lemma 1, Section 3.

  • Some trilinear centers rational on the sidelengths which produce (i) elliptic loci (e.g., X1,X2X_{1},X_{2}, etc.) as well as (ii) non-elliptic (e.g., X6X_{6}, X19X_{19}, etc.).

  • No locus has been found with more than 6 intersections with a straight line, suggesting the degree is at most 6.

  • No center has been found with non-rational trilinears whose locus is an ellipse555Not shown, but also tested were non-rational triangle centers XjX_{j}, j=14, 16, 17, 18, 359, 360, 364, 365, 367j=14,\,16,\,17,\,18,\,359,\,360,\,364,\,365,\,367., suggesting that the locus of non-rational centers is always non-elliptic.

3.2. An Algebro-Geometric Ambient

Given EB semi-axes a,ba,b, our problem can be described by the following 14 variables:

  • 6 triangle vertex coordinates, Pi=(xi,yi),i=1,2,3P_{i}=(x_{i},y_{i}),\,i=1,2,3;

  • 3 sidelengths s1,s2,s3s_{1},s_{2},s_{3};

  • 3 trilinears p,q,rp,q,r;

  • 2 locus coordinates x,yx,y.

These are related by the system of 14 polynomial equations defined in Table 3.

eqns.descriptionzero set of3vertices on the EB(xi/a)2+(yi/b)21,i=1,2,33reflection law at Pjj,k,cyclic,𝒜=diag(1/a2;1/b2)(𝒜Pj.P𝒜Pj.Pj)|PkPj|(𝒜Pj.Pk𝒜Pj.Pj)|PPj|3sidelengths(xixj)2+(yiyj)2sk22locus cartesians, (3)(ps1+qs2+rs3)(x,y)ps1P1+qs2P2+rs3P33trilinears (must rationalize)ph(s1,s2,s3);qh(s2,s3,s1);rh(s3,s1,s2)\begin{array}[]{|c|l|l|}\hline\cr\textbf{eqns.}&\textbf{description}&\textbf{zero set of}\\ \hline\cr 3&\text{vertices on the EB}&(x_{i}/a)^{2}+(y_{i}/b)^{2}-1\;,i=1,2,3\\ \hline\cr 3&\begin{array}[]{l}\text{reflection law at $P_{j}$}\\ j,k,\ell\;\text{cyclic},\mathcal{A}=\text{diag}(1/a^{2};1/b^{2})\end{array}&\begin{array}[]{l}(\mathcal{A}P_{j}.P_{\ell}-\mathcal{A}P_{j}.P_{j})|P_{k}-P_{j}|\\ \;\;\;-(\mathcal{A}P_{j}.P_{k}-\mathcal{A}P_{j}.P_{j})|P_{\ell}-P_{j}|\end{array}\\ \hline\cr 3&\text{sidelengths}&(x_{i}-x_{j})^{2}+(y_{i}-y_{j})^{2}-s_{k}^{2}\\ \hline\cr 2&\text{locus cartesians, \eqref{eqn:trilin-cartesian}}&(ps_{1}+qs_{2}+rs_{3})(x,y)-ps_{1}P_{1}+qs_{2}P_{2}+rs_{3}P_{3}\\ \hline\cr 3&\text{trilinears (must rationalize)}&p-h(s_{1},s_{2},s_{3});\;q-h(s_{2},s_{3},s_{1});\;r-h(s_{3},s_{1},s_{2})\\ \hline\cr\end{array}
Table 3. System of 14 equations which the locus must satisfy. The three equations in the third line are obtained from the reflection law – angle of incidence equals the angle of reflection – imposed on the vertices of a triangular orbit.

Since the 3-periodic family of orbits is one dimensional, out of the first 9 equations, one is functionally dependent on the rest. Therefore, we have 13 independent equations in 14 variables, yielding a 1d algebraic variety, which can be complexified if desired, as in [5, 10, 11, 25].

Can tools from computational Algebraic Geometry [27, 30] be used to eliminate 12 variables automatically, thus obtaining a single polynomial equation (x,y)=0\mathcal{L}(x,y)=0 whose Zariski closure contains the locus? Below we provide a method based on the theory of resultants [17, 30] to compute \mathcal{L} for a subset of triangle centers.

3.3. When Trilinears are Rational

Consider a triangle center XX whose trilinears p:q:rp:q:r are rational on the sidelengths s1,s2,s3s_{1},s_{2},s_{3}, i.e., the triangle center function hh is rational, equation (4).

Theorem 3.

The locus of a rational triangle center is an algebraic curve.

We thank one of the referees for kindly contributing the outline for this proof.

Proof.

Consider the complexified variables of the first 4 rows of the table 3: the xi,yix_{i},y_{i} and sis_{i}. We have a 9 dimensional complex space. Consider the Zariski closure VV in 9{\mathbb{P}}_{9}\mathbb{C} of the complex algebraic subset defined by the 9 first equations given in the Table 3. Explicitly we have the equations:

fi\displaystyle f_{i} =xi2a2+yi2b21=0,(i=1,2,3)\displaystyle=\frac{x_{i}^{2}}{a^{2}}+\frac{y_{i}^{2}}{b^{2}}-1=0,\;(i=1,2,3)
f4\displaystyle f_{4} =(x1x2a2+y1y2b21)s2(x1x3a2+y1y3b21)s3=0\displaystyle=(\frac{x_{1}x_{2}}{a^{2}}+\frac{y_{1}y_{2}}{b^{2}}-1)s_{2}-(\frac{x_{1}x_{3}}{a^{2}}+\frac{y_{1}y_{3}}{b^{2}}-1)s_{3}=0
f5\displaystyle f_{5} =(x1x2a2+y1y2b21)s1(x2x3a2+y2y3b21)s3=0\displaystyle=(\frac{x_{1}x_{2}}{a^{2}}+\frac{y_{1}y_{2}}{b^{2}}-1)s_{1}-(\frac{x_{2}x_{3}}{a^{2}}+\frac{y_{2}y_{3}}{b^{2}}-1)s_{3}=0
f6\displaystyle f_{6} =(x1x3a2+y1y3b21)s1(x2x3a2+y2y3b21)s2=0\displaystyle=(\frac{x_{1}x_{3}}{a^{2}}+\frac{y_{1}y_{3}}{b^{2}}-1)s_{1}-(\frac{x_{2}x_{3}}{a^{2}}+\frac{y_{2}y_{3}}{b^{2}}-1)s_{2}=0
f7\displaystyle f_{7} =(x2x1)2+(y2y1)2s32=0\displaystyle=(x_{2}-x_{1})^{2}+(y_{2}-y_{1})^{2}-s_{3}^{2}=0
f8\displaystyle f_{8} =(x2x3)2+(y2y3)2s12=0\displaystyle=(x_{2}-x_{3})^{2}+(y_{2}-y_{3})^{2}-s_{1}^{2}=0
f9\displaystyle f_{9} =(x3x1)2+(y3y1)2s22=0\displaystyle=(x_{3}-x_{1})^{2}+(y_{3}-y_{1})^{2}-s_{2}^{2}=0

By construction, it is a complex projective algebraic subset of 9{\mathbb{P}}_{9}\mathbb{C}. We have that s2f5s3f6s1f4=0s_{2}f_{5}-s_{3}f_{6}-s_{1}f_{4}=0. Computing its dimension it follows that:

  1. (i)

    the rank of the Jacobian matrix given by the 5 equations f1=f2=f3=f4=f5=0f_{1}=f_{2}=f_{3}=f_{4}=f_{5}=0 in relation to the variables (x1,y1,x2,y2,x3)(x_{1},y_{1},x_{2},y_{2},x_{3}) or (x1,y1,x2,y2,y3)(x_{1},y_{1},x_{2},y_{2},y_{3}) is generically 5.

  2. (ii)

    the rank of the Jacobian matrix given by the 9 equations is generically 8.

  3. (iii)

    the sks_{k} are determined by the xi,yix_{i},y_{i} (last 3 equations of Table 3) and that the set of complex triangular orbits is of dimension 1, see [5, 10, 11, 25], and [7] for an explicit parametrization of VV. Longer (explicit) expressions appear in Appendix C.

This implies that VV is a projective algebraic curve of 9{\mathbb{P}}_{9}\mathbb{C}, the complex projective space of dimension 9. The map which associates any point of VV with X=(x,y)2X=(x,y)\in{\mathbb{P}}_{2}\mathbb{C} as in Equation 3 (where p,q,rp,q,r are given rational maps of the sis_{i}) is well-defined and holomorphic on an algebraic open subset of VV, and hence, on the complement of a discrete subset of VV. Therefore, it is defined and holomorphic everywhere on VV. Now its image is an analytic subset of 2{\mathbb{P}}_{2}\mathbb{C} (Remmert proper mapping theorem) hence an algebraic subset (Chow theorem), see [13, Ch.V]. Its dimension is less or equal than 1, otherwise it would be the whole 2{\mathbb{P}}_{2}\mathbb{C} (an impossibility, since it is bounded for real points). When the dimension of the image is zero the locus degenerates to a point. ∎

A typical example of a point-locus is that of the Mittenpunkt X9X_{9}, which remains stationary at the center of the EB. In this case p=h(s1,s2,s3)=s2+s3s1p=h(s_{1},s_{2},s_{3})=s_{2}+s_{3}-s_{1}, q=h(s2,s3,s1)=s1+s3s2q=h(s_{2},s_{3},s_{1})=s_{1}+s_{3}-s_{2} and r=h(s3,s1,s2)=s1+s2s3r=h(s_{3},s_{1},s_{2})=s_{1}+s_{2}-s_{3}. In the EB we have that X9=(0,0)X_{9}=(0,0).

Proposition 1.

Over 3-periodics in the EB, the Mittenpunkt X9X_{9} is the only triangle center whose locus is a point.

Proof.

Given the four-fold symmetry in the one-dimensional family of 3-periodics in the EB, the locus of any triangle center must be symmetric about both the vertical and horizontal semi-axes of the EB. So if some center’s locus is a point, said point must be at the center of the EB. Over a continuous set of generic triangles (such as those in the 3-periodic family), two distinct triangle centers cannot always coincide (other than in a discrete set of configurations), so the claim follows. ∎

3.4. Algorithm to compute the locus polynomial

Our algorithm is based on the following 3-steps which yield an algebraic curve (x,y)=0\mathcal{L}(x,y)=0 which contains the locus. We refer to Lemmas 4 and 5 appearing below. Appendix C contains supporting expressions.

Step 1.

Introduce the symbolic variables u,u1,u2u,u_{1},u_{2}:

u2+u12=1,ρ1u2+u22=1.u^{2}+u_{1}^{2}=1,\;\;\;\rho_{1}\,u^{2}+u_{2}^{2}=1.

The vertices will be given by rational functions of u,u1,u2u,u_{1},u_{2}

P1=(au,bu1),P2=(p2x,p2y)/q2,P3=(p3x,p3y)/q3P_{1}=(a\,u,b\,u_{1}),\;\;P_{2}=(p_{2x},p_{2y})/q_{2},\;\;\;P_{3}=(p_{3x},p_{3y})/q_{3}

Expressions for P1,P2,P3P_{1},P_{2},P_{3} appear in Appendix C as do equations gi=0g_{i}=0, i=1,2,3i=1,2,3, polynomial in si,u,u1,u2s_{i},u,u_{1},u_{2}.

Step 2.

Express the locus XX as a rational function on u,u1,u2,s1,s2,s3u,u_{1},u_{2},s_{1},s_{2},s_{3}.

Convert p:q:rp:q:r to cartesian coordinates X=(x,y)X=(x,y) via Equation (3). From Lemma 4, it follows that (x,y)\left(x,y\right) is rational on u,u1,u2,s1,s2,s3u,u_{1},u_{2},s_{1},s_{2},s_{3}.

x=𝒬/,y=𝒮/𝒯x=\mathcal{Q}/\mathcal{R},\;\;\;y=\mathcal{S}/\mathcal{T}

To obtain the polynomials 𝒬,,𝒮,𝒯\mathcal{Q,R,S,T} on said variables u,u1,u2,s1,s2,s3u,u_{1},u_{2},s_{1},s_{2},s_{3}, one substitutes the p,q,rp,q,r by the corresponding rational functions of s1,s2,s3s_{1},s_{2},s_{3} that define a specific triangle center XX. Other than that, the method proceeds identically.

Step 3.

Computing resultants. Our problem is now cast in terms of the polynomial equations:

E0=𝒬x=0,F0=𝒮y𝒯=0E_{0}=\mathcal{Q}-x\,\mathcal{R}=0,\;\;\;F_{0}=\mathcal{S}-y\,\mathcal{T}=0

Firstly, compute the resultants, in chain fashion:

E1=\displaystyle E_{1}= Res(g1,E0,s1)=0,F1=Res(g1,F0,s1)=0\displaystyle\textrm{Res}(g_{1},E_{0},s_{1})=0,\;\;\;F_{1}=\textrm{Res}(g_{1},F_{0},s_{1})=0
E2=\displaystyle E_{2}= Res(g2,E1,s2)=0,F2=Res(g2,F1,s2)=0\displaystyle\textrm{Res}(g_{2},E_{1},s_{2})=0,\;\;\;F_{2}=\textrm{Res}(g_{2},F_{1},s_{2})=0
E3=\displaystyle E_{3}= Res(g3,E2,s3)=0,F3=Res(g3,F2,s3)=0\displaystyle\textrm{Res}(g_{3},E_{2},s_{3})=0,\;\;\;F_{3}=\;\textrm{Res}(g_{3},F_{2},s_{3})=0

It follows that E3(x,u,u1,u2)=0E_{3}(x,u,u_{1},u_{2})=0 and F3(y,u,u1,u2)=0F_{3}(y,u,u_{1},u_{2})=0 are polynomial equations. In other words, s1,s2,s3s_{1},s_{2},s_{3} have been eliminated.

Now eliminate the variables u1u_{1} and u2u_{2} by taking the following resultants:

E4(x,u,u2)=\displaystyle E_{4}(x,u,u_{2})= Res(E3,u12+u21,u1)=0\displaystyle\textrm{Res}(E_{3},u_{1}^{2}+u^{2}-1,u_{1})=0
F4(y,u,u2)=\displaystyle F_{4}(y,u,u_{2})= Res(F3,u12+u21,u1)=0\displaystyle\textrm{Res}(F_{3},u_{1}^{2}+u^{2}-1,u_{1})=0
E5(x,u)=\displaystyle E_{5}(x,u)= Res(E4,u22+ρ1u21,u2)=0\displaystyle\textrm{Res}(E_{4},u_{2}^{2}+\rho_{1}u^{2}-1,u_{2})=0
F5(y,u)=\displaystyle F_{5}(y,u)= Res(F4,u22+ρ1u21,u2)=0\displaystyle\textrm{Res}(F_{4},u_{2}^{2}+\rho_{1}u^{2}-1,u_{2})=0

This yields two polynomial equations E5(x,u)=0E_{5}(x,u)=0 and F5(y,u)=0F_{5}(y,u)=0.

Finally compute the resultant

=Res(E5,F5,u)=0{\mathcal{L}}=\textrm{Res}(E_{5},F_{5},u)=0

that eliminates uu and gives the implicit algebraic equation for the locus XX.

Remark 3.

In practice, after obtaining a resultant, a human assists the CAS by factoring out spurious branches (when recognized), in order to get the final answer in more reduced form.

When non-rational in the sidelengths, except a few cases (e.g., X359,X360X_{359},X_{360} which are transcendental), triangle centers in Kimberling’s list have explicit trilinears involving fractional powers and/or terms containing the triangle area. Those can be made implicit, i.e, given by zero sets of polynomials involving p,q,r,s1,s2,s3p,q,r,s_{1},s_{2},s_{3}. The chain of resultants to be computed will be increased by three, in order to eliminate the variables p,q,rp,q,r before (or after) s1,s2,s3s_{1},s_{2},s_{3}.

Lemma 4.

Let P1=(au,b1u2).P_{1}=({a}{u},b\sqrt{1-u^{2}}). The coordinates of P2P_{2} and P3P_{3} of the 3-periodic billiard orbit are rational functions in the variables u,u1,u2u,u_{1},u_{2}, where u1=1u2u_{1}=\sqrt{1-u^{2}}, u2=1ρ1u2u_{2}=\sqrt{1-\rho_{1}u^{2}} and ρ1=c4(b2+δ)2/a6\rho_{1}=c^{4}(b^{2}+\delta)^{2}/a^{6}.

Proof.

Follows directly from the parametrization of the billiard orbit, Appendix C.2. In fact, P2=(x2(u),y2(u))=(p2x/q2,p2y/q2)P_{2}=(x_{2}(u),y_{2}(u))=(p_{2x}/q_{2},p_{2y}/q_{2}) and P3=(x3(u),y3(u))P_{3}=(x_{3}(u),y_{3}(u)) =(p3x/q3,p3y/q3)=(p_{3x}/q_{3},p_{3y}/q_{3}), where p2xp_{2x}, p2yp_{2y}, p3xp_{3x} and p3yp_{3y} have degree 44 in (u,u1,u2)(u,u_{1},u_{2}) and q2q_{2}, q3q_{3} are algebraic of degree 44 in uu. Expressions for u1,u2u_{1},u_{2} appear in Appendix C.1. ∎

Lemma 5.

Let P1=(au,b1u2).P_{1}=(au,b\sqrt{1-u^{2}}). Let s1s_{1}, s2s_{2} and s3s_{3} the sides of the triangular orbit P1P2P3{P_{1}}{P_{2}}{P_{3}}. Then g1(u,s1)=0g_{1}(u,s_{1})=0, g2(s2,u2,u)=0g_{2}(s_{2},u_{2},u)=0 and g3(s3,u2,u)=0g_{3}(s_{3},u_{2},u)=0 for polynomial functions gig_{i}, defined in Appendix C.3.

Proof.

Using the parametrization of the 3-periodic billiard orbit it follows that s12|P2P3|2=0s_{1}^{2}-|P_{2}-P_{3}|^{2}=0 is a rational equation in the variables u,s1u,s_{1}. Simplifying, leads to g1(s1,u)=0.g_{1}(s_{1},u)=0.

Analogously for s2s_{2} and s3s_{3}. In this case, the equations s22|P1P3|2=0s_{2}^{2}-|P_{1}-P_{3}|^{2}=0 and s32|P1P2|2=0s_{3}^{2}-|P_{1}-P_{2}|^{2}=0 have square roots u2=1ρ1u2u_{2}=\sqrt{1-\rho_{1}u^{2}} and u1=1u2u_{1}=\sqrt{1-u^{2}} and are rational in the variables s2,u2,u1,us_{2},u_{2},u_{1},u and s3,u2,u1,us_{3},u_{2},u_{1},u respectively. It follows that the degrees of g1g_{1}, g2g_{2}, and g3g_{3} are 1010. Simplifying, leads to g2(s2,u2,u1,u)=0g_{2}(s_{2},u_{2},u_{1},u)=0 and g3(s3,u2,u1,u)=0g_{3}(s_{3},u_{2},u_{1},u)=0. ∎

3.5. Examples

Table 4 shows the Zariski closure obtained contained the elliptic locus of a few triangle centers. Notice one factor is always of the form [(x/ai)2+(y/bi)2)1]3[(x/a_{i})^{2}+(y/b_{i})^{2})-1]^{3}, related to the triple cover described in Section 2.5. The expressions shown required some manual simplification during the symbolic calculations.

XiNameSpurious FactorsElliptic Factoraibi1Incentervery long expression[36(91+6161)x2+324(139+1961)y2+22761396961]60.635040.297442Barycenter(91500x2+4992261370993)2(100x2+225y2+5261413)30.262050.17473Circumcenterx(3600x46380x21539)(40x2+56143)2(1104500y2+591136614633685)2(65880x2+85276164649)6(5832x2+(275232061)y2575172961)30.0991460.4763\scriptsize\begin{array}[]{|c|l|l|l|l|l|}\hline\cr X_{i}&\text{Name}&\text{Spurious Factors}&\text{Elliptic Factor}&a_{i}&b_{i}\\ \hline\cr 1&\text{Incenter}&\text{very long expression}&\begin{array}[]{l}[36(91+61\sqrt{61})x^{2}\\ +324(139+19\sqrt{61})y^{2}\\ +22761-3969\sqrt{61}]^{6}\end{array}&0.63504&0.29744\\ \hline\cr 2&\text{Barycenter}&\begin{array}[]{l}(91500x^{2}\\ +49922\sqrt{61}-370993)^{2}\end{array}&\begin{array}[]{l}(100x^{2}\\ +225y^{2}\\ +52\sqrt{61}-413)^{3}\end{array}&0.26205&0.1747\\ \hline\cr 3&\text{Circumcenter}&\begin{array}[]{l}x\\ (3600x^{4}-6380x^{2}-1539)\\ (40x^{2}+5\sqrt{61}-43)^{2}\\ \begin{array}[]{l}(-1104500y^{2}\\ +591136\sqrt{61}-4633685)^{2}\end{array}\\ (65880x^{2}+8527\sqrt{61}-64649)^{6}\end{array}&\begin{array}[]{l}(5832x^{2}\\ +(2752-320\sqrt{61})y^{2}\\ 5751-729\sqrt{61})^{3}\end{array}&0.099146&0.4763\\ \hline\cr\end{array}
Table 4. Method of Resultants applied to obtain the zero set for a few sample triangle centers with elliptic loci, for the specific case of a/b=1.5a/b=1.5. Both spurious and an elliptic factor are present. The latter are raised to powers multiple of three suggesting a phenomenon related to the triple cover, Section 2.5. Also shown are semi-axes ai,bia_{i},b_{i} implied by the elliptic factor. These have been checked to be in perfect agreement with the values predicted for those semi-axes in [7].

4. Conclusion

A few interesting questions are posed to the reader.

  • Can the degree of the locus of a triangle center or derived triangle vertex be predicted based on its trilinears?

  • Is there a triangle center such that its locus intersects a straight line more than 6 times?

  • Certain triangle centers have non-convex loci (e.g., X67X_{67} at a/b=1.5a/b=1.5 [20]). What determines non-convexity?

  • What determines the number of self-intersections of a given locus?

  • In the spirit of [5, 25], how would one determine via complex analytic geometry, that X6X_{6} is a quartic?

  • What is the non-elliptic locus described by the summits of equilaterals erected over each orbit side (used in the construction of the Outer Napoleon Triangle [33]). [21, pl#13]. What kind of curve is it?

  • Within X1X_{1} and X100X_{100} only XiX_{i}, i=1318i=13{\ldots}18 have irrational trilinears. XjX_{j}, j=359, 360j=359,\,360 are transcendental and XkX_{k}, k=364, 365, 367k=364,\,365,\,367 are irrational. In the latter two cases, the loci are numerically non-elliptic. Can any non-rational triangle center produce an ellipse?

  • X6X_{6} is the isogonal conjugate666Given a point PP, reflect the three PP-cevians (lines through vertices and point PP) about the angular bisectors. These meet at the isogonal conjugate of PP. of X2X_{2}. Though the latter’s locus is an ellipse, the former’s is a quartic. In the case of the isogonal pair X3X_{3} and X4X_{4} both are ellipses. What is the connection with isogonal (and/or isotomic777Given vertices PiP_{i}, i=1,2,3i=1,2,3, of a triangle and a point XX, let QiQ_{i} denote the intersections of XX-cevians with the opposite side. Let QiQ_{i}^{\prime} be the reflection of QiQ_{i} about the midpoint of the corresponding side. Lines PiQiP_{i}Q_{i}^{\prime} meet at XX^{\prime}, the isotomic conjugate of XX. transformations) and ellipticity?

4.1. Videos and Media

The reader is encouraged to browse our companion paper [22] where intriguing locus phenomena are investigated. Additionally, loci can be explored interactively with our browser-based app [3].

Videos mentioned herein are on a playlist [21], with links provided on Table 5.

id Title Section youtu.be/<.>
01 Locus of X1X_{1} is an Ellipse 1 BsyM7RnswA
02 Locus of Intouchpoints is non-elliptic 1, E 9xU6T7hQMzs
03 X9X_{9} stationary at EB center 1 tMrBqfRBYik
04 Stationary Excentral Cosine Circle 1 ACinCf-D_Ok
05 Loci for X1X5X_{1}\ldots{X_{5}} are ellipses E sMcNzcYaqtg
06 Elliptic locus of Excenters similar to rotated X1X_{1} E Xxr1DUo19_w
07 Loci of X11X_{11}, X100X_{100} and Extouchpoints are the EB E TXdg7tUl8lc
08 Family of Derived Triangles 2 xyroRTEVNDc
09 Loci of Vertices of Derived Triangles E,2 OGvCQbYqJyI
10 Peter Moses’ 29 Billiard Points 2 JdcJt5PExsw
11 Convex Comb.: X1X_{1}-Intouch and X2X_{2}-Midpoint 2.5 3Gr3Nh5-jHs
12 Convex Comb.: X3X_{3}-Midpoint and X4X_{4}-Altfoot 2.5 HZFjkWD_CnE
13 Oval Locus of the Outer Napoleon Summits 4 70-E-NZrNCQ
Table 5. Videos mentioned in the paper available in a Youtube playlist [21]. The last column contains a clickable YouTube code.

Acknowledgments

We warmly thank Clark Kimberling, Peter Moses, Sergei Tabachnikov, Richard Schwartz, Arseniy Akopyan, Olga Romaskevich, Ethan Cotterill, for their invaluable input during this research. We would like to thank the referees for all very useful suggestions, including an alternative proof for Theorem 3.

The first author is fellow of CNPq and coordinator of Project PRONEX/ CNPq/ FAPEG 2017 10 26 7000 508.

Appendix A Triangle Centers

A.1. Triangle Centers

Constructions for a few basic triangle centers are shown in Figure 10, using Kimberling’s XiX_{i} notation [16].

Refer to caption
Figure 10. Constructions for triangle centers XiX_{i}, i=1,2,3,4,5,9,11i=1,2,3,4,5,9,11, taken from [23].
  • The Incenter X1X_{1} is the intersection of angular bisectors, and center of the Incircle (green), a circle tangent to the sides at three Intouchpoints (green dots), its radius is the Inradius rr.

  • The Barycenter X2X_{2} is where lines drawn from the vertices to opposite sides’ midpoints meet. Side midpoints define the Medial Triangle (red).

  • The Circumcenter X3X_{3} is the intersection of perpendicular bisectors, the center of the Circumcircle (purple) whose radius is the Circumradius RR.

  • The Orthocenter X4X_{4} is where altitudes concur. Their feet define the Orthic Triangle (orange).

  • X5X_{5} is the center of the 9-Point (or Euler) Circle (pink): it passes through each side’s midpoint, altitude feet, and Euler points [33].

  • The Feuerbach Point X11X_{11} is the single point of contact between the Incircle and the 9-Point Circle.

  • Given a reference triangle P1P2P3P_{1}P_{2}P_{3} (blue), the Excenters P1P2P3P_{1}^{\prime}P_{2}^{\prime}P_{3}^{\prime} are pairwise intersections of lines through the PiP_{i} and perpendicular to the bisectors. This triad defines the Excentral Triangle (green).

  • The Excircles (dashed green) are centered on the Excenters and are touch each side at an Extouch Point ei,i=1,2,3e_{i},i=1,2,3.

  • Lines drawn from each Excenter through sides’ midpoints (dashed red) concur at the Mittenpunkt X9X_{9}.

  • Also shown (brown) is the triangle’s Mandart Inellipse, internally tangent to each side at the eie_{i}, and centered on X9X_{9}. This is identical to the N=3N=3 caustic.

A.2. Trilinear Coordinates

Trilinear coordinates are signed distances to the sides of a triangle T=P1P2P3T=P_{1}P_{2}P_{3}, i.e. they are invariant with respect to similarity/reflection transformations, see Figure 11.

Refer to caption
Figure 11. Trilinear coordinates p:q:rp:q:r for a point XX on the plane of a generic triangle T=P1P2P3T=P_{1}P_{2}P_{3} are homogeneous signed distances to each side whose lengths are s1,s2,s3s_{1},s_{2},s_{3}. The red dots are known as the pedal points of XX [33]. The left (resp. right) figure shows an interior (resp. exterior) point. In both cases p,rp,r are positive however qq is positive (resp. negative) in the former (resp. latter) case.

Trilinears can be easily converted to cartesian coordinates using [33]:

(3) X=ps1P1+qs2P2+rs3P3ps1+qs2+rs3X=\frac{ps_{1}P_{1}+qs_{2}P_{2}+rs_{3}P_{3}}{p{s_{1}}+q{s_{2}}+r{s_{3}}}

where s1=|P3P2|s_{1}=|P_{3}-P_{2}|, s2=|P1P3|s_{2}=|P_{1}-P_{3}| and s3=|P2P1|s_{3}=|P_{2}-P_{1}| are the sidelengths, Figure 11. For instance, the trilinears for the barycenter are p=s11,q=s21,r=s31p=s_{1}^{-1},\,q=s_{2}^{-1},\,r=s_{3}^{-1}, yielding the familiar X2=(P1+P2+P3)/3X_{2}=(P_{1}+P_{2}+P_{3})/3.

A triangle center is such a triple obtained by applying a triangle center function hh thrice to the sidelengths s1,s2,s3s_{1},s_{2},s_{3} cyclically [15]:

(4) p:q:rh(s1,s2,s3):h(s2,s3,s1):h(s3,s1,s2)p:q:r{\iff}h(s_{1},s_{2},s_{3}):h(s_{2},s_{3},s_{1}):h(s_{3},s_{1},s_{2})

hh must (i) be bi-symmetric, i.e., h(s1,s2,s3)=h(s1,s3,s2)h(s_{1},s_{2},s_{3})=h(s_{1},s_{3},s_{2}), and (ii) homogeneous, h(ts1,ts2,ts3)=tnh(s1,s2,s3)h(ts_{1},ts_{2},ts_{3})=t^{n}h(s_{1},s_{2},s_{3}) for some nn [15].

See Table 6 for triangle center functions for selected centers. Trilinears can be converted to cartesian coordinates using (3).

Trilinear coordinates for selected triangle centers appear in Table 6.

center name h(s1,s2,s3)h(s_{1},s_{2},s_{3})
X1X_{1} Incenter 11
X2X_{2} Barycenter 1/s11/s_{1}
X3X_{3} Circumcenter s1(s22+s32s12)s_{1}(s_{2}^{2}+s_{3}^{2}-s_{1}^{2})
X4X_{4} Orthocenter 1/[s1(s22+s32s12)]1/[s_{1}(s_{2}^{2}+s_{3}^{2}-s_{1}^{2})]
X5X_{5} 9-Point Center s2s3[s12(s22+s32)(s22s32)2]{s_{2}}{s_{3}}[s_{1}^{2}(s_{2}^{2}+s_{3}^{2})-(s_{2}^{2}-s_{3}^{2})^{2}]
X11X_{11} Feuerbach Point s2s3(s2+s3s1)(s2s3)2{s_{2}}{s_{3}}(s_{2}+s_{3}-s_{1})(s_{2}-s_{3})^{2}
X88X_{88} Isog. Conjug. of X44X_{44} 1/(s2+s32s1)1/(s_{2}+s_{3}-2{s_{1}})
X100X_{100} Anticomplement of X11X_{11} 1/(s2s3)1/(s_{2}-s_{3})
𝐗𝟔\mathbf{X_{6}} Symmedian Point 𝐬𝟏\mathbf{s_{1}}
𝐗𝟏𝟑\mathbf{X_{13}^{*}} Fermat Point 𝐬𝟏𝟒𝟐(𝐬𝟐𝟐𝐬𝟑𝟐)𝟐+𝐬𝟏𝟐(𝐬𝟐𝟐+𝐬𝟑𝟐+𝟒𝟑𝐀)\mathbf{s_{1}^{4}-2(s_{2}^{2}-s_{3}^{2})^{2}+s_{1}^{2}(s_{2}^{2}+s_{3}^{2}+4\sqrt{3}A)}
𝐗𝟏𝟓\mathbf{X_{15}^{*}} 2nd Isodynamic Point 𝐬𝟏[𝟑(𝐬𝟏𝟐𝐬𝟐𝟐𝐬𝟑𝟐)𝟒𝐀]\mathbf{s_{1}[\sqrt{3}(s_{1}^{2}-s_{2}^{2}-s_{3}^{2})-4A]}
𝐗𝟏𝟗\mathbf{X_{19}} Clawson Point 𝟏/(𝐬𝟐𝟐+𝐬𝟑𝟐𝐬𝟏𝟐)\mathbf{1/(s_{2}^{2}+s_{3}^{2}-s_{1}^{2})}
𝐗𝟑𝟕\mathbf{X_{37}} Crosspoint of 𝐗𝟏,𝐗𝟐\mathbf{X_{1},X_{2}} 𝐬𝟐+𝐬𝟑\mathbf{s_{2}+s_{3}}
𝐗𝟓𝟗\mathbf{X_{59}} Isog. Conj. of 𝐗𝟏𝟏\mathbf{X_{11}} 𝟏/[𝐬𝟐𝐬𝟑(𝐬𝟐+𝐬𝟑𝐬𝟏)(𝐬𝟐𝐬𝟑)𝟐]\mathbf{1/[{s_{2}}{s_{3}}(s_{2}+s_{3}-s_{1})(s_{2}-s_{3})^{2}]}
X9X_{9} Mittenpunkt s2+s3s1s_{2}+s_{3}-s_{1}
Table 6. Triangle center function hh for a few selected XiX_{i}’s, taken from [16]. The first 8 centers produce elliptic loci, whereas the remainder (boldfaced) do not. X13X_{13} and X15X_{15} are starred to indicate their trilinears are irrational: these contain AA, the area the triangle, known (e.g., from Heron’s formula) to be irrational on the sidelengths. We haven’t yet detected an algebraic pattern which differentiates both groups, nor have we detected an irrational center whose locus is elliptic. Regarding the last row, the Mittenpunkt, we don’t consider its locus to be elliptic since it degenerates to a point at the EB center.

A.3. Derived Triangles

A derived triangle TT^{\prime} is constructed from vertices of a reference triangle TT. These can be represented as a 3x3 matrix, where each row, taken as trilinears, is a vertex of TT^{\prime}. For example, the Excentral, Medial, and Intouch Triangles TexcT^{\prime}_{exc}, TmedT^{\prime}_{med}, and TintT^{\prime}_{int} are given by [33]:

[111111111],[0s21s31s110s31s11s210],[0s1s3s1s2+s3s1s2s1+s2s3s2s3s1+s2+s30s1s2s1+s2s3s2s3s1+s2+s3s1s3s1s2+s30]\left[\begin{matrix}-1&1&1\\ 1&-1&1\\ 1&1&-1\end{matrix}\right],\;\left[\begin{matrix}0&s_{2}^{-1}&s_{3}^{-1}\\ s_{1}^{-1}&0&s_{3}^{-1}\\ s_{1}^{-1}&s_{2}^{-1}&0\end{matrix}\right],\;\left[\begin{matrix}0&\frac{s_{1}s_{3}}{s_{1}-s_{2}+s_{3}}&\frac{s_{1}s_{2}}{s_{1}+s_{2}-s_{3}}\\ \frac{s_{2}s_{3}}{-s_{1}+s_{2}+s_{3}}&0&\frac{s_{1}s_{2}}{s_{1}+s_{2}-s_{3}}\\ \frac{s_{2}s_{3}}{-s_{1}+s_{2}+s_{3}}&\frac{s_{1}s_{3}}{s_{1}-s_{2}+s_{3}}&0\end{matrix}\right]

A few derived triangles are shown in Figure 12, showing a property similar to Lemma 2, Appendix D, namely, when the 3-periodic is an isosceles, one vertex of the derived triangle lies on the orbit’s axis of symmetry.

Refer to caption
Figure 12. When the orbit is an isosceles triangle (solid blue), any derived triangle will contain one vertex on the axis of symmetry of the orbit. Video

Appendix B Review: Elliptic Billiards

An Elliptic Billiard (EB) is a particle moving with constant velocity in the interior of an ellipse, undergoing elastic collisions against its boundary [26, 31], Figure 13. For any boundary location, a given exit angle (e.g., measured from the normal) may give rise to either a quasi-periodic (never closes) or NN-periodic trajectory [31], where NN is the number of bounces before the particle returns to its starting location. All trajectory segments are tangent to a confocal caustic [31]. The EB is a special case of Poncelet’s Porism [4]: if one NN-periodic trajectory can be found departing from some boundary point, any other such point will initiate an NN-periodic, i.e., a 1d family of such orbits will exist. A classic result is that NN-periodics conserve perimeter [31].

Refer to caption
Figure 13. Trajectory regimes in an Elliptic Billiard, taken from [23]. Top left: first four segments of a trajectory departing at P1P_{1} and toward P2P_{2}, bouncing at Pi,i=2,3,4P_{i},i=2,3,4. At each bounce the normal n^i\hat{n}_{i} bisects incoming and outgoing segments. Joachimsthal’s integral [31] means all segments are tangent to a confocal caustic (brown). Top right: a 3-periodic trajectory. All 3-periodics in this Billiard will be tangent to a special confocal caustic (brown). Bottom: first 50 segments of a non-periodic trajectory starting at P1P_{1} and directed toward P2P_{2}. Segments are tangent to a confocal ellipse (left) or hyperbola (right). The former (resp. latter) occurs if P1P2P_{1}P_{2} passes outside (resp. between) the EB’s foci (black dots).

Appendix C Expressions used in Section 3.3

Let the boundary of the Billiard satisfy Equation (1). Assume, without loss of generality, that aba{\geq}b. Here we provide expressions used in Section 3. Let P1,P2,P3P_{1},P_{2},P_{3} be an orbit’s vertices.

C.1. Exit Angle Required for 3-Periodicity

Consider a starting point P1=(x1,y1)P_{1}=(x_{1},y_{1}) on a Billiard with semi-axes a,ba,b. The cosine of the exit angle α\alpha (measured with respect to the normal at P1P_{1}, i.e., α=θ1/2\alpha=\theta_{1}/2) required for the trajectory to close after 3 bounces is given by [7]:

cos2α=d12δ12d2=k1,sinαcosα=δ1d12d2d2d14δ12=k2\cos^{2}{\alpha}=\frac{d_{1}^{2}\delta_{1}^{2}}{\,d_{2}}=k_{1},\;\;\;\sin{\alpha}\cos\alpha=\frac{\delta_{1}d_{1}^{2}}{d_{2}}\sqrt{d_{2}-d_{1}^{4}\delta_{1}^{2}}=k_{2}

with c2=a2b2c^{2}=a^{2}-b^{2}, d1=(ab/c)2d_{1}=(a\,b/c)^{2}, d2=b4x12+a4y12d_{2}={b}^{4}x_{1}^{2}+{a}^{4}y_{1}^{2}, δ=a4+b4a2b2\delta=\sqrt{a^{4}+b^{4}-a^{2}b^{2}}, and δ1=2δa2b2\delta_{1}=\sqrt{2\delta-a^{2}-b^{2}}.

C.2. Orbit Vertices

Given a starting vertex P1=(x1,y1)P_{1}=(x_{1},y_{1}) on the EB, P2=(p2x,p2y)/q2P_{2}=(p_{2x},p_{2y})/q_{2}, and P3=(p3x,p3y)/q3P_{3}=(p_{3x},p_{3y})/q_{3} where [7]:

p2x=\displaystyle\small p_{2x}= b4((a2+b2)k1a2)x132a4b2k2x12y1\displaystyle-{b}^{4}\left(\left(a^{2}+{b}^{2}\right)k_{1}-{a}^{2}\right)x_{1}^{3}-2\,{a}^{4}{b}^{2}k_{2}x_{1}^{2}{y_{1}}
+a4((a23b2)k1+b2)x1y122a6k2y13\displaystyle+{a}^{4}\left(({a}^{2}-3\,{b}^{2})k_{1}+{b}^{2}\right){x_{1}}\,y_{1}^{2}-2{a}^{6}k_{2}y_{1}^{3}
p2y=\displaystyle p_{2y}= 2b6k2x13+b4((b23a2)k1+a2)x12y1\displaystyle 2{b}^{6}k_{2}x_{1}^{3}+{b}^{4}\left(({b}^{2}-3\,{a}^{2})k_{1}+{a}^{2}\right)x_{1}^{2}{y_{1}}
+2a2b4k2x1y12a4((a2+b2)k1b2)y13\displaystyle+2\,{a}^{2}{b}^{4}k_{2}{x_{1}}y_{1}^{2}-{a}^{4}\left(\left(a^{2}+{b}^{2}\right)k_{1}-{b}^{2}\right)y_{1}^{3}
q2=\displaystyle q_{2}= b4(a2c2k1)x12+a4(b2+c2k1)y122a2b2c2k2x1y1\displaystyle{b}^{4}\left(a^{2}-c^{2}k_{1}\right)x_{1}^{2}+{a}^{4}\left({b}^{2}+c^{2}k_{1}\right)y_{1}^{2}-2\,{a}^{2}{b}^{2}{c^{2}}k_{2}{x_{1}}\,{y_{1}}
p3x=\displaystyle p_{3x}= b4(a2(b2+a2))k1x13+2a4b2k2x12y1\displaystyle{b}^{4}\left({a}^{2}-\left({b}^{2}+{a}^{2}\right)\right)k_{1}x_{1}^{3}+2\,{a}^{4}{b}^{2}k_{2}x_{1}^{2}{y_{1}}
+a4(k1(a23b2)+b2)x1y12+2a6k2y13\displaystyle+{a}^{4}\left(k_{1}\left({a}^{2}-3\,{b}^{2}\right)+{b}^{2}\right){x_{1}}\,y_{1}^{2}+2\,{a}^{6}k_{2}y_{1}^{3}
p3y=\displaystyle p_{3y}= 2b6k2x13+b4(a2+(b23a2)k1)x12y1\displaystyle-2\,{b}^{6}k_{2}x_{1}^{3}+{b}^{4}\left({a}^{2}+\left({b}^{2}-3\,{a}^{2}\right)k_{1}\right){{x_{1}}}^{2}{y_{1}}
2a2b4k2x1y12+a4(b2(b2+a2)k1)y13,\displaystyle-2\,{a}^{2}{b}^{4}k_{2}x_{1}y_{1}^{2}+{a}^{4}\left({b}^{2}-\left({b}^{2}+{a}^{2}\right)k_{1}\right)\,y_{1}^{3},
q3=\displaystyle q_{3}= b4(a2c2k1)x12+a4(b2+c2k1)y12+2a2b2c2k2x1y1.\displaystyle{b}^{4}\left({a}^{2}-{c^{2}}k_{1}\right)x_{1}^{2}+{a}^{4}\left({b}^{2}+c^{2}k_{1}\right)y_{1}^{2}+2\,{a}^{2}{b}^{2}c^{2}k_{2}\,{x_{1}}\,{y_{1}}.

C.3. Polynomials Satisfied by the Sidelengths

Let sidelengths s1=|P3P2|,s2=|P1P3|,s3=|P2P1|s_{1}=|P_{3}-P_{2}|,s_{2}=|P_{1}-P_{3}|,s_{3}=|P_{2}-P_{1}|. The following are polynomial expressions on sis_{i} and u,u1,u2u,u_{1},u_{2}:

g1=\displaystyle g_{1}= h1s12+h0,g2=h1s22h2u1u2+h3,g3=h1s32+h2u1u2+h3\displaystyle-h_{1}\,s_{1}^{2}+h_{0},\;\;g_{2}=-h_{1}\,s_{2}^{2}-h_{2}\,u_{1}\,u_{2}+h_{3},\;\;\;g_{3}=-h_{1}\,s_{3}^{2}+h_{2}\,u_{1}\,u_{2}+h_{3}
h0=\displaystyle h_{0}= 12c12(a2+b2+2δ)u8+24c10(a2b2+2b42δc2u6\displaystyle 12c^{12}(a^{2}+b^{2}+2\delta)u^{8}+24c^{10}(a^{2}b^{2}+2b^{4}-2\delta c^{2}u^{6}
\displaystyle- 4c4[10a1012a8b2+11a6b47a4b6+24a2b88b10\displaystyle 4c^{4}[10a^{10}-12a^{8}b^{2}+11a^{6}b^{4}-7a^{4}b^{6}+24a^{2}b^{8}-8b^{10}
\displaystyle- (2(4a42a2b2+b4))(a42a2b2+4b4)δ]u4\displaystyle(2(4a^{4}-2a^{2}b^{2}+b^{4}))(a^{4}-2a^{2}b^{2}+4b^{4})\delta]u^{4}
+\displaystyle+ 8a6c2[4a67a4b2+11a2b42b6(2(a2+ab+b2))(a2ab+b2)δ]u2\displaystyle 8a^{6}c^{2}[4a^{6}-7a^{4}b^{2}+11a^{2}b^{4}-2b^{6}-(2(a^{2}+ab+b^{2}))(a^{2}-ab+b^{2})\delta]u^{2}
+\displaystyle+ 4a12δ12\displaystyle 4a^{12}\delta_{1}^{2}
h1=\displaystyle h_{1}= c2(3c4u42c2(a22b2u2a4))2\displaystyle c^{2}\left(3c^{4}u^{4}-2c^{2}(a^{2}-2b^{2}u^{2}-a^{4})\right)^{2}
h2=\displaystyle h_{2}= 2a(b2δ)δ1u((3c6δ+(6(a2+b2))c6)u6+((3(a2+4b2))c4δ\displaystyle-2{a(b^{2}-\delta)}\delta_{1}u((3c^{6}\delta+(6(a^{2}+b^{2}))c^{6})u^{6}+((3(a^{2}+4b^{2}))c^{4}\delta
\displaystyle- (3(2a4a2b24b4))c4)u4+((b2a2)(7a48b4)δ\displaystyle(3(2a^{4}-a^{2}b^{2}-4b^{4}))c^{4})u^{4}+((b_{2}-a^{2})(7a^{4}-8b^{4})\delta
+\displaystyle+ 2c2(a22b2)(a42b4))u2+a4(a24b2)δa4(2a43a2b2+4b4))\displaystyle 2c^{2}(a^{2}-2b^{2})(a^{4}-2b^{4}))u^{2}+a^{4}(a^{2}-4b^{2})\delta-a^{4}(2a^{4}-3a^{2}b^{2}+4b^{4}))
h3=\displaystyle h_{3}= c6(9c6u10+1)((18(a4+b4))δ(3(7a616a4b2+29a2b48b6))u8\displaystyle c^{6}(9c^{6}u^{10}+1)((18(a^{4}+b^{4}))\delta-(3(7a^{6}-16a^{4}b^{2}+29a^{2}b^{4}-8b^{6}))u^{8}
+\displaystyle+ 2c2[(13a812a6b2+49a4b454a2b6+16b8)\displaystyle 2c^{2}\left[(13a^{8}-12a^{6}b^{2}+49a^{4}b^{4}-54a^{2}b^{6}+16b^{8})\right.
\displaystyle- 2(10a67a4b2+11a2b48b6)δ]u6\displaystyle\left.2(10a^{6}-7a^{4}b^{2}+11a^{2}b^{4}-8b^{6})\delta\right]u^{6}
+\displaystyle+ ((28a4δ240a2b6+16b8)δ26a10+12a8b2+42a4b648a2b8+16b10)u4\displaystyle((28a^{4}\delta^{2}-40a^{2}b^{6}+16b^{8})\delta-26a^{10}+12a^{8}b^{2}+42a^{4}b^{6}-48a^{2}b^{8}+16b^{10})u^{4}
+\displaystyle+ a(13a69a4b28b4c2(4(2a4+a2b22b4))δ)4u2+2a8(δa8c2)\displaystyle a(13a^{6}-9a^{4}b^{2}-8b^{4}c^{2}-(4(2a^{4}+a^{2}b^{2}-2b^{4}))\delta)^{4}u^{2}+2a^{8}(\delta-a^{8}c^{2})

Appendix D Proof of Lemmas used in Section 2

Lemma (3).

A parametric traversal of P1P_{1} around the EB boundary is a triple cover of the locus of a triangle center XiX_{i}.

Proof.

Let P1(t)=(acost,bsint)P_{1}(t)=\left(a\cos{t},b\sin{t}\right). Let tt^{*} (resp. t)t^{**}) be the value of tt for which the 3-periodic orbit is an isosceles with a horizontal (resp. vertical) axis of symmetry, Figure 14, t>tt^{*}>t^{**}. For such cases one can easily derive888Indeed, sint=Jb\sin{t^{*}}=Jb and cost=Ja\cos{t^{**}}=Ja, where JJ is Joachmisthal’s constant (N=3N=3) [31].:

tant=b2δa2+2b2a2,tant=2δ2a2+b23a\tan{t^{*}}=\frac{b\sqrt{2\delta-{a}^{2}+2\,{b}^{2}}}{a^{2}},\;\;\;\tan{t^{**}}=\frac{\sqrt{2\,\delta-2\,{a}^{2}+{b}^{2}}}{\sqrt{3}\,a}

with δ=a4a2b2+b4\delta=\sqrt{a^{4}-a^{2}b^{2}+b^{4}} as above. Referring to Figure 14:

Affirmation 1.

A continuous counterclockwise motion of P1(t)P_{1}(t) along the intervals [t,t)[-t^{*},-t^{**}), [t,0)[-t^{**},0), [0,t)[0,t^{**}), and [t,t)[t^{**},t^{*}) will each cause XiX_{i} to execute a quarter turn along its locus, i.e., with tt varying from t-t^{*} to tt^{*}, XiX_{i} will execute one complete revolution on its locus999The direction of this revolution depends on XiX_{i} and its not always monotonic. The observation is valid for elliptic or non-elliptic loci alike. See [14] for details..

Affirmation 2.

A continuous counterclockwise motion of P1(t)P_{1}(t) in the [t,πt)[t^{*},\pi-t^{*}) (resp. [πt,π+t)[\pi-t^{*},\pi+t^{*}), and [π+t,2πt)[\pi+t^{*},2\pi-t^{*})), visits the same 3-periodics as when tt sweeps [t,t)[-t^{*},-t^{**}) (resp. [t,t)[-t^{*},t^{*}), and [t,πt)[t^{*},\pi-t^{*})).

Therefore, a complete turn of P1(t)P_{1}(t) around the EB visits the 3-periodic family thrice, i.e., XiX_{i} will wind thrice over its locus. ∎

Refer to caption
Figure 14. Counterclockwise motion of P1(t)P_{1}(t), for t=[t,t)t=[-t^{*},t^{*}) can be divided in four segments delimited by t=(t,t,0,t,t)t=(-t^{*},-t^{**},0,t^{**},t^{**}). Orbit positions for the first four are shown (red polygons) in the top-left, top-right, bot-left, and bot-right pictures. At P1(±t)P_{1}({\pm}t^{*}) (resp. P1(±t)P_{1}({\pm}t^{**})) the orbit is an isosceles triangle with a horizontal (resp. vertical) axis of symmetry. Observations 1,2 assert that a counterclockwise motion of P1(t)P_{1}(t) along each of the four intervals causes a triangle center XiX_{i} to execute a quarter turn along its locus (elliptic or not), and that a complete revolution of P1(t)P_{1}(t) around the EB causes XiX_{i} to wind thrice on its locus. For illustration, the locus of X1(t)X_{1}(t) is shown (green) at each of the four interval endpoints.

Below we provide proofs of Lemmas 2, 1 and 3.

Lemma (2).

Any triangle center XiX_{i} of an isosceles triangle is on the axis of symmetry of said triangle.

Proof.

Consider a sideways isosceles triangle with vertices P1=(x1,0)P_{1}=(x_{1},0), P2=(x2,y2)P_{2}=(-x_{2},y_{2}) and P3=(x2,y2)P_{3}=(-x_{2},-y_{2}), i.e., its axis of symmetry is the xx axis. Let XiX_{i} have trilinears p:q:r=h(s1,s2,s3):h(s2,s3,s1):h(s3,s1,s2)p:q:r=h(s_{1},s_{2},s_{3}):h(s_{2},s_{3},s_{1}):h(s_{3},s_{1},s_{2}). Its cartesian coordinates are given by Equation (3). As s2=s3s_{2}=s_{3} and hh is symmetric on its last two variables h(s1,s2,s3)=h(s1,s3,s2)h(s_{1},s_{2},s_{3})=h(s_{1},s_{3},s_{2}) it follows from equation (3) that yi=0y_{i}=0. ∎

Lemma (1).

If the locus of triangle center XiX_{i} is elliptic, said ellipse must be concentric and axis-aligned with the EB.

Proof.

This follows from Lemma 1. ∎

Lemma (3).

If the locus of XiX_{i} is an ellipse, when P1(t)P_{1}(t) is at either EB vertex, its non-zero coordinate is equal to the corresponding locus semi-axis length.

Proof.

The family of 3-periodic orbits contains four isosceles triangles, Figure 3. Parametrize P1(t)=(acost,bsint)P_{1}(t)=(a\cos t,b\sin t). It follows from Lemma 1 that Xi(0)=(±ai,0)X_{i}(0)=(\pm a_{i},0), Xi(π/2)=(0,±bi),X_{i}(\pi/2)=(0,\pm b_{i}), Xi(π)=(ai,0)X_{i}(\pi)=(\mp a_{i},0) and Xi(3π/2)=(0,bi)X_{i}(3\pi/2)=(0,\mp b_{i}), for some ai,bia_{i},b_{i}. This ends the proof. ∎

Appendix E Review: Early Observations

Figure 15 (left) shows the elliptic loci of XkX_{k}, k=1,,5k=1,...,5. The latter two are new results.

Refer to caption
Figure 15. Left: The loci of Incenter X1X_{1}, Barycenter X2X_{2}, Circumcenter X3X_{3}, Orthocenter X4X_{4}, and center of the 9-Point Circle X5X_{5} are all ellipses, Video, app. Also shown is the Euler Line (dashed black) which for any triangle, passes through all of XiX_{i}, i=15i=1...5 [33]. Right: A 3-periodic orbit starting at P1(t)P_{1}(t) is shown (blue). The locus of X11X_{11}, where the Incircle (green) and 9-Point Circle (pink) meet, is the caustic (brown), also swept by the Extouchpoints eie_{i}. X100X_{100} (double-length reflection of X11X_{11} about X2X_{2}) is the EB Video, app.

We have also observed that the locus of the Feuerbach Point X11X_{11} coincides with the N=3N=3 caustic, a confocal ellipse to which the 3-periodic family is internally tangent, Appendix B. Additionally, the locus of X100X_{100}, the anticomplement101010Double-length reflection about the Barycenter X2X_{2}. of X11X_{11} is the EB boundary. These phenomena appear in Figure 15 (right).

Indeed, some these facts will be known to triangle specialists if one regards the EB as the circumellipse centered on the Mittenpunkt X9X_{9} [16, X(9)] and [12]. For example, a full 57 triangle centers lie on said circumellipse. An animation of some of these points is viewable in [21, pl#10].

Refer to caption
Figure 16. Loci generated by the vertices of selected orbit-derived triangles, namely: the Intouch (green), Feuerbach (blue, not to be confused with the Feuerbach Point X11X_{11}), and Medial (red) Triangles are non-elliptic. However, those of the Extouch Triangle (brown), are identical to the N=3N=3 caustic (a curve also swept by X11X_{11}). Not shown is the locus of the Excentral Triangle, an ellipse similar to a rotated copy of the Incenter locus. app, Video 1, Video 2, Video 3, Video 4.

Additionally, a few observations have been made [23] about the loci of vertices of orbit-derived triangles (see Appendix A.3), some of which are elliptic and others non, illustrated in Figure 16.

References

  • [1] Akopyan, A., Schwartz, R., Tabachnikov, S. (2020). Billiards in ellipses revisited. Eur. J. Math. doi.org/10.1007/s40879-020-00426-9.
  • [2] Bialy, M., Tabachnikov, S. (2020). Dan Reznik’s identities and more. Eur. J. Math. doi.org/10.1007/s40879-020-00428-7.
  • [3] Darlan, I., Reznik, D. (2020). Ellipse mounted triangles. GitHub Pages. dan-reznik.github.io/ellipse-mounted-loci-p5js/.
  • [4] Dragović, V., Radnović, M. (2011). Poncelet Porisms and Beyond: Integrable Billiards, Hyperelliptic Jacobians and Pencils of Quadrics. Frontiers in Mathematics. Basel: Springer.
  • [5] Fierobe, C. (2021). On the circumcenters of triangular orbits in elliptic billiard. J. Dyn. Control Syst., 27(4): 693–705.
  • [6] Fitzgibbon, A., Pilu, M., Fisher, R. (1999). Direct least-square fitting of ellipses. Pattern Analysis and Machine Intelligence, 21(5).
  • [7] Garcia, R. (2019). Elliptic billiards and ellipses associated to the 3-periodic orbits. Amer. Math. Monthly, 126(06): 491–504.
  • [8] Garcia, R., Reznik, D. (2021). Discovering Poncelet Invariants in the Plane. Rio de Janeiro: IMPA.
  • [9] Garcia, R., Reznik, D., Koiller, J. (2021). Companion website to “loci of 3-periodics in an elliptic billiard: Why so many ellipses?”. GitHub Pages. dan-reznik.github.io/why-so-many-ellipses/.
  • [10] Glutsyuk, A. (2014). On quadrilateral orbits in complex algebraic planar billiards. Moscow Math. J., 14(2): 239–289, 427.
  • [11] Griffiths, P., Harris, J. (1978). On Cayley’s explicit solution to Poncelet’s porism. Enseign. Math. (2), 24(1-2): 31–40.
  • [12] Grozdev, S., Dekov, D. (2014). The computer program “Discoverer” as a tool of mathematical investigation. Intl. J. of Comp. Disc. Math. (IJCDM). www.ddekov.eu/j/2014/JCGM201405.pdf.
  • [13] Gunning, R. C., Rossi, H. (1965). Analytic functions of several complex variables. Englewood Cliffs, NJ: Prentice-Hall.
  • [14] Helman, M., Laurain, D., Reznik, D., Garcia, R. (2022). Poncelet triangles: a theory for locus ellipticity. Beitr. Algebra Geom. doi:10.1007/s13366-021-00620-0.
  • [15] Kimberling, C. (1993). Triangle centers as functions. Rocky Mountain J. Math., 23(4): 1269–1286.
  • [16] Kimberling, C. (2019). Encyclopedia of triangle centers. faculty.evansville.edu/ck6/encyclopedia/ETC.html.
  • [17] Lang, S. (2002). Algebra. Graduate Texts in Mathematics. Springer-Verlag, New York, 3rd ed.
  • [18] Moore, R., Kearfott, R. B., Cloud, M. J. (2009). Introduction to Interval Analysis. Philadelphia: SIAM.
  • [19] Odehnal, B. (2011). Poristic loci of triangle centers. J. Geom. Graph., 15(1): 45–67.
  • [20] Reznik, D. (2019). Triangular orbits in elliptic billards: Loci of points X(1) to X(100). GitHub Pages. dan-reznik.github.io/billiard-loci/.
  • [21] Reznik, D. (2020). Playlist for “Loci of Triangular Orbits in an Elliptic Billiard: Elliptic? Algebraic?”. bit.ly/2REOigc.
  • [22] Reznik, D., Garcia, R., Koiller, J. (2020). The ballet of triangle centers on the elliptic billiard. J. Geom. and Graphics, 24(1): 079–101.
  • [23] Reznik, D., Garcia, R., Koiller, J. (2020). Can the elliptic billiard still surprise us? Math Intelligencer, 42: 6–17.
  • [24] Reznik, D., Garcia, R., Koiller, J. (2021). Fifty new invariants of NN-periodics in the elliptic billiard. Arnold Math. J., 7(3): 341–355.
  • [25] Romaskevich, O. (2014). On the incenters of triangular orbits on elliptic billiards. Enseign. Math., 60(3-4): 247–255.
  • [26] Rozikov, U. A. (2018). An Introduction To Mathematical Billiards. Singapore: World Scientific Publishing Company.
  • [27] Schenck, H. (2003). Computational Algebraic Geometry, London Mathematical Society Student Text, vol. 58. Cambridge: Cambridge University Press.
  • [28] Schwartz, R., Tabachnikov, S. (2016). Centers of mass of Poncelet polygons, 200 years after. Math. Intelligencer, 38(2): 29–34.
  • [29] Snyder, J. M. (1992). Interval analysis for computer graphics. Computer Graphics, 26(2): 121–129.
  • [30] Sturmfels, B. (1997). Introduction to resultants, in applications of computational algebraic geometry. In: Proceedings of Symposia in Applied Mathematics, vol. 53. Providence, RI: American Mathematical Society, pp. 25–40.
  • [31] Tabachnikov, S. (2005). Geometry and Billiards, vol. 30 of Student Mathematical Library. Providence, RI: American Mathematical Society.
  • [32] Tabachnikov, S. (2019). Projective configuration theorems: old wine into new wineskins. In: S. Dani, A. Papadopoulos, eds., Geometry in History. Basel: Springer Verlag, pp. 401–434.
  • [33] Weisstein, E. W. (2003). CRC concise encyclopedia of mathematics. Chapman & Hall/CRC, Boca Raton, FL, 2nd ed.