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

A Table Theorem for Surfaces with Odd Euler Characteristic

Ali Naseri Sadr
Abstract

We use the square peg problem for smooth curves to prove a generalized table Theorem for real valued functions on Riemannian surfaces with odd Euler characteristic. We then use this result to prove the table conjecture for even functions on the two sphere.

1 Introduction

In 19511951, Freeman Dyson proved a remarkable result: for every continuous real-valued function on the unit sphere in 3\mathbb{R}^{3}, it is possible to find the vertices of a square on a great circle (diameter d=2d=2) at which the function takes the same value; see [2] for more details. Dyson conjectures in the paper that the same result holds for some circle with diameter dd, for every 0<d20<d\leq 2. This conjecture reappears as The Table Problem in [6, conjecture 1616]. Roger Fenn proved an analogous result for positive functions defined on a convex disk in the plane; see [3] for more details.

The table problem admits the following natural generalization. Define a square with diameter d>0d>0 on a Riemannian surface (Σ,g)(\Sigma,g) to be the image, under the exponential map expg:TΣΣ\exp_{g}\colon T\Sigma\to\Sigma, of the vertices of a square of diameter dd (with respect to gg) centered at the origin in some tangent plane. A table for a continuous function ff is a square such that ff takes the same value at the four vertices of this square. We establish the following results:

Theorem 1.1 (Table Theorem for Surfaces with Odd Euler Characteristic).

Let (Σ,g)(\Sigma,g) be a Riemannian surface with χ(Σ)\chi(\Sigma) odd and ff a continuous real valued function on it. Then for every d>0d>0, ff admits a table with diameter dd.

As an immediate corollary of this Theorem, we will get the result for even functions on S2S^{2} and any Riemannian metric that is invariant under the antipodal map. In particular, this resolves the table problem for even functions.

Corollary 1.2.

Let gg be a Riemannian metric on S2S^{2} so that the antipodal map is an isometry of this metric and ff an even function on S2S^{2}. Then for every positive dd, ff admits a table with diameter dd.

The proof of Theorem 1.1 is based on the following topological idea. By compactness, it suffices to prove the Theorem for positive C2C^{2} functions. Now for a fixed diameter dd and a positive C2C^{2} function ff, we deform all the circles with diameter dd in TΣT\Sigma by mapping each vector (x,v)(x,v) in TΣT\Sigma to (x,f(exp(x,v))v)(x,f(exp(x,v))\cdot v). The result is a subbundle of TΣT\Sigma where each fiber is a star-shaped C2C^{2} Jordan curve. Let CC be the subspace of TΣT\Sigma consisting of the center points of the inscribed squares in these curves. Every table for ff with diameter dd is in a one-to-one correspondence with the intersection points of CC and the zero section of TΣT\Sigma. Within CC is the subspace C0C_{0} of center points of gracefully inscribed squares: these are the squares whose vertices appear in the same cyclic order around the square and the curve; see [10] for more details. We use Sard-Smale Theorem for Fredholm maps to prove for a generic choice of f,f, the subspace C0C_{0} gives us a 2\mathbb{Z}_{2}-cycle representing the non-zero element in H2(TΣ;2)H_{2}(T\Sigma;\mathbb{Z}_{2}) since every generic C2C^{2} star-shaped curve inscribes an odd number of graceful squares; this is proved for generic PL curves in [10]. Hence, C0C_{0} intersects the zero section in at least one point when χ(Σ)\chi(\Sigma) is odd, which yields Theorem 1.1 by a convergence argument.

In section 22, we show how our definition of tables for a function on a surface is a generalization of the corresponding one on the two sphere and propose a table problem for Riemannian surfaces. We define the mappings and spaces we need for our proof in section 33. In section 44, we establish the technical aspects of our proof, while a formal proof for Theorem 1.1 and corollary 1.2 is given in section 55. Related transversality arguments appear in [12, 5, 1]. Finally, we reprove the square-peg problem for C2C^{2} star-shaped Jordan curves by similar transversality arguments to sections 33 and 44 in the appendix.

Acknowledgments

The author is grateful to his advisors, John Baldwin and Josh Greene, for their invaluable guidance, support, and insightful conversations about this work. He would also like to express his gratitude to Peter Feller and Joaquin Lema for inspiring conversations about this work.

2 Tables on Riemannian Surfaces

In the following, we will view S2S^{2} as the set of points with distance 11 from the origin in 3\mathbb{R}^{3} and consider the induced Riemannian metric on it.

Definition.

For a continuous real-valued function ff on the round two sphere, we say p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} on S2S^{2} are the basis of a table for ff if they are four vertices of a square in 3\mathbb{R}^{3} and we have

f(p1)=f(p2)=f(p3)=f(p4).f(p_{1})=f(p_{2})=f(p_{3})=f(p_{4}).

For every four vertices of a square p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} on the two sphere, we can find a point xx and vectors vv and ww in TxS2T_{x}S^{2} such that

expx(v)=p1,expx(w)=p2,\displaystyle\exp_{x}(v)=p_{1},\hskip 2.84526pt\exp_{x}(w)=p_{2},
expx(v)=p3,expx(w)=p4,\displaystyle\exp_{x}(-v)=p_{3},\hskip 2.84526pt\exp_{x}(-w)=p_{4},
vw=0,v=w.\displaystyle v\cdot w=0,\hskip 2.84526pt\|v\|=\|w\|.

Note that xx lies on the line that goes through the origin and the center of square in 3\mathbb{R}^{3}. In particular, there are two points on the two sphere that satisfy the previous equations, but if we also require that w=vπ2\|w\|=\|v\|\leq\frac{\pi}{2}, then xx becomes unique unless the square lies on a great circle. This gives us a way to parameterize all the squares with a fixed side length as pairs of (x,v)(x,v) in TS2TS^{2} where vv has a fixed length. We can also use this observation to define tables for arbitrary closed Riemannian surfaces.

Definition.

Let (Σ,g)(\Sigma,g) be a Riemannian surface and ff a continuous real valued function on Σ\Sigma. We say (x,v)(x,v) in TΣT\Sigma is basis of a table for ff if

f(exp(x,v))=f(exp(x,w))=f(exp(x,v))=f(exp(x,w)),f(\exp(x,v))=f(\exp(x,w))=f(\exp(x,-v))=f(\exp(x,-w)), (1)

where ww is a vector perpendicular to vv and has the same length as vv.

Remark.

There are two choices for ww in the previous definition, but our definition is independent of this choice.

The following is a reformulation of the table problem for S2S^{2}.

Conjecture 2.1 (The Table Problem for S2S^{2}).

Fix a positive real number aa. A continuous function on the two sphere endowed with the round metric admits a table with v=a\|v\|=a.

Using the generalization given in equation (1), one can ask a similar question for other Riemannian surfaces. In particular, we have the following problem.

Problem 2.1.

Let (Σ,g)(\Sigma,g) be a closed Riemannian surface and fix a positive real number aa. Does every continuous function on Σ\Sigma admit a table defined by equation (1) and v=a\|v\|=a?

We note that our main Theorem answers this problem in affirmative for surfaces with odd Euler characteristic.

Remark.

Since we work with compact surfaces and solving problem 2.1 for a function ff is the same as solving it for f+cf+c where cc is a constant real number, we only need to solve the problem for positive functions. The other important point is that in contrast to peg problems for Jordan curves in the plane, this problem can be proved using a convergence argument because we fix the diameter of our table beforehand. Therefore, if one wants to prove problem 2.1 for a surface Σ\Sigma, it suffices to prove it for a dense subset of positive functions on Σ\Sigma.

3 A Submersion

In this section, we will assume that (Σ,g)(\Sigma,g) is a fixed Riemannian surface. Consider a positive real number aa and let

Ua(Σ){(x,v)TΣ:v=a}.U_{a}(\Sigma)\coloneqq\{(x,v)\in T\Sigma:\|v\|=a\}.

We are going to work with the fourth symmetric product of each fiber in Ua(Σ)U_{a}(\Sigma); this space is a manifold itself, but we prefer to work with an open submanifold of it; see [9] for more details.

Let XX be a topological space and consider sym4(X)sym^{4}(X); we define the fat diagonal Δ\Delta to be the subset of points in sym4(X)sym^{4}(X) for which at least two of the coordinates are equal. We define Ka(Σ)K_{a}(\Sigma) to be a fiber bundle over Σ\Sigma where each fiber over a point xx is the fourth symmetric product of the circle with radius aa in TxΣT_{x}\Sigma minus its fat diagonal. Note that the fibers are open non-orientable four manifolds. By abuse of notation, we let sym4(TΣ)sym^{4}(T\Sigma) denote the fiber bundle where each fiber over a point xx is sym4(TxΣ)sym^{4}(T_{x}\Sigma); we cut out the fat diagonal from each fiber and denote the resulting fiber bundle by Q(Σ)Q(\Sigma). The fibers of Q(Σ)Q(\Sigma) are open orientable eight dimensional manifolds and Ka(Σ)K_{a}(\Sigma) is a subbundle of Q(Σ)Q(\Sigma).

Let C2(Ua(Σ))C^{2}(U_{a}(\Sigma)) denote the space of C2C^{2} functions on Ua(Σ)U_{a}(\Sigma). This function space can be endowed with a norm that makes it a Banach space; see [8] for more details. We will work with the open subset of positive functions in C2(Ua(Σ))C^{2}(U_{a}(\Sigma)) and denote it by C+2(Ua(Σ))C^{2}_{+}(U_{a}(\Sigma)).

Definition.

Define a map Ψ:C+2(Ua(Σ))×Ka(Σ)Q(Σ)\Psi\colon C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma)\to Q(\Sigma) by

(h,[θ1,θ2,θ3,θ4])[h(θ1)θ1,h(θ2)θ2,h(θ3)θ3,h(θ4)θ4].(h,[\theta_{1},\theta_{2},\theta_{3},\theta_{4}])\mapsto[h(\theta_{1})\cdot\theta_{1},h(\theta_{2})\cdot\theta_{2},h(\theta_{3})\cdot\theta_{3},h(\theta_{4})\cdot\theta_{4}]. (2)

This map is well defined and the image avoids the diagonal in sym4(TΣ)sym^{4}(T\Sigma) because we are working with positive functions. In particular, this is a smooth map from a Banach manifold to a finite dimensional manifold.

Remark.

Fix a positive function hh in C+2(UaΣ)C^{2}_{+}(U_{a}\Sigma) and let Ψh\Psi_{h} denote the restriction of Ψ\Psi to {h}×Ka(Σ)Ka(Σ)Q(Σ)\{h\}\times K_{a}(\Sigma)\cong K_{a}(\Sigma)\to Q(\Sigma); this map covers the identity on Σ\Sigma and it is an embedding since we are only considering positive functions. Moreover, this map scales each fiber of Ka(Σ)K_{a}(\Sigma) according to the function hh; thus the image of Ψh\Psi_{h} in each fiber of Q(Σ)Q(\Sigma) over a point xx is the fourth symmetric product of a star-shaped curve in TxΣT_{x}\Sigma minus its fat diagonal.

Lemma 3.1.

The map Ψ\Psi is a submersion.

Proof.

Consider a pair (h,θ)(h,\theta) in C+2(Ua(Σ))×Ka(Σ)C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma) and let ξ\xi be its image under Ψ\Psi. Since Ψh\Psi_{h} covers the identity map on Σ\Sigma, if θ\mathcal{H}_{\theta} is a horizontal subspace of TθKa(Σ)T_{\theta}K_{a}(\Sigma), then dΨh(θ)d\Psi_{h}(\mathcal{H}_{\theta}) is also a horizontal subspace of TξQ(Σ)T_{\xi}Q(\Sigma). Therefore, we only need to check that Ψ\Psi is a submersion when we restrict the map to a fiber over an arbitrary point xx. Let θi\theta_{i} denote the components of θ\theta. Fix all the θis\theta_{i}^{\prime}s except θ1\theta_{1} and change θ1\theta_{1} along a curve δ(t)\delta(t) in Ua(Σ)U_{a}(\Sigma) such that the curve γ(t)=[δ(t),θ2,θ3,θ4]\gamma(t)=[\delta(t),\theta_{2},\theta_{3},\theta_{4}] avoids the diagonal in sym4(TΣ)sym^{4}(T\Sigma). We have

Ψ(h,γ(t))=[h(δ(t))δ(t),h(θ2)θ2,h(θ3)θ3,h(θ4)θ4].\Psi(h,\gamma(t))=[h(\delta(t))\delta(t),h(\theta_{2})\theta_{2},h(\theta_{3})\theta_{3},h(\theta_{4})\theta_{4}].

Hence, we get

dΨ(0,γ˙(0))=[h(θ1)δ˙(0)+d(h(δ(t)))dt|t=0θ1,0,0,0]Im(dΨ),d\Psi(0,\dot{\gamma}(0))=[h(\theta_{1})\cdot\dot{\delta}(0)+\frac{d(h(\delta(t)))}{dt}\Big{|}_{t=0}\cdot\theta_{1},0,0,0]\in\text{Im}(d\Psi),

where we used δ(0)=θ1\delta(0)=\theta_{1}. Since δ(t)\delta(t) has constant length, δ˙(0)\dot{\delta}(0) is a non-zero vector orthogonal to θ1\theta_{1}. Repeating this argument for the other coordinates proves we have vectors of the previous form in the image of dΨd\Psi where all the components are zero except one and the non-zero component is of the form riωi+siθir_{i}\cdot\omega_{i}+s_{i}\cdot\theta_{i} with ri=h(θi)r_{i}=h(\theta_{i}) greater than zero, ωi\omega_{i} a vector orthogonal to θi\theta_{i}, and sis_{i} some arbitrary real number. Now fix the four-tuple θ\theta and pick a function gg such that g(θ1)=1g(\theta_{1})=1 and gg vanishes on the other three coordinates. Let gt=h+tgg_{t}=h+tg and note that for tt small enough all the functions gtg_{t} are positive. We get

dΨ(g,0)=[g(θ1)θ1,g(θ2)θ2,g(θ3)θ3,g(θ4)θ4]=[θ1,0,0,0]Im(dΨ).d\Psi(g,0)=[g(\theta_{1})\cdot\theta_{1},g(\theta_{2})\cdot\theta_{2},g(\theta_{3})\cdot\theta_{3},g(\theta_{4})\cdot\theta_{4}]=[\theta_{1},0,0,0]\in\text{Im}(d\Psi).

The same argument shows we have vectors of the form [0,,θi,,0][0,\dots,\theta_{i},\dots,0] in the image of dΨd\Psi. We conclude the lemma because the set of vectors

[0,,θi,,0],\displaystyle[0,\dots,\theta_{i},\dots,0],
[0,,ωi,,0]\displaystyle[0,\dots,\omega_{i},\dots,0]

generate all the vertical vectors over xx in Q(Σ)Q(\Sigma). ∎

Let A(Σ)A(\Sigma) denote the subbundle of Q(Σ)Q(\Sigma) where over each point xx, A(Σ)|xA(\Sigma)\big{|}_{x} is the set of four tuples of vectors that are vertices of a square in TxΣT_{x}\Sigma with respect to the metric on Σ\Sigma. This subbundle has codimension four in Q(Σ)Q(\Sigma).

Corollary 3.2.

Let 𝒮\mathcal{S} denote Ψ1(A(Σ))\Psi^{-1}(A(\Sigma)). Then 𝒮\mathcal{S} is a codimension four smooth submanifold of C+2(Ua(Σ))×Ka(Σ)C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma).

Proof.

This follows from the fact that Ψ\Psi is a submersion and A(Σ)A(\Sigma) is a codimension 44 submanifold. ∎

We will call 𝒮\mathcal{S} the space of star-shaped squares since if we consider ξ=Ψ(h,θ)\xi=\Psi(h,\theta) for a point (h,θ)(h,\theta) in 𝒮\mathcal{S}, then ξi\xi_{i}’s are vertices of a square in the fiber of TΣT\Sigma over a point xx and the four points lie on a star-shaped curve around the origin in TxΣT_{x}\Sigma; this curve is defined by sending each (x,v)(x,v) in Ua(Σ)U_{a}(\Sigma) to (x,h(x,v)v)(x,h(x,v)\cdot v) in TxΣT_{x}\Sigma.

4 Star-Shaped Squares

Definition.

We define a map F:𝒮C+2(Ua(Σ))F\colon\mathcal{S}\to C_{+}^{2}(U_{a}(\Sigma)) by restricting the first projection map on C+2(Ua(Σ))×Ka(Σ)C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma) to 𝒮\mathcal{S}. This is a smooth map on 𝒮\mathcal{S} since 𝒮\mathcal{S} is a smooth submanifold of C+2(Ua(Σ))×Ka(Σ)C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma).

Remark.

For simplicity, we will denote C+2(Ua(Σ))×Ka(Σ)C_{+}^{2}(U_{a}(\Sigma))\times K_{a}(\Sigma) by 𝒩\mathcal{N} and C+2(Ua(Σ))C_{+}^{2}(U_{a}(\Sigma)) by \mathcal{F} in the following sections.

Lemma 4.1.

The map FF is Fredholm.

Proof.

Fix a point (h,θ)(h,\theta) in 𝒮\mathcal{S} and consider dF:T(h,θ)𝒮ThdF\colon T_{(h,\theta)}\mathcal{S}\to T_{h}\mathcal{F}. We need to show ker(dF)\ker(dF) and coker(dF)\text{coker}(dF) are finite dimensional. Let pr1pr_{1} denote the first projection map on 𝒩\mathcal{N} and note that ker(dpr1)\ker(dpr_{1}) at (h,θ)(h,\theta) is TθKa(Σ)T_{\theta}K_{a}(\Sigma) which has dimension 66; we conclude that dim(ker(dF))\dim(\ker(dF)) is finite. Let WW denote T(h,θ)𝒩T_{(h,\theta)}\mathcal{N} and VV denote T(h,θ)𝒮T_{(h,\theta)}\mathcal{S}. We define a map L:W/Vcoker(dF)L\colon W/V\to\text{coker}(dF) by

[w][dpr1(w)].[w]\mapsto[dpr_{1}(w)].

This map is surjective because pr1pr_{1} is a submersion and we know dim(W/V)=codim(𝒮)=4\dim(W/V)=\text{codim}(\mathcal{S})=4. Hence, dim(coker(dF))\dim(\text{coker}(dF)) is finite. ∎

Our next step is to compute the index of FF and since index is constant on each connected component of 𝒮\mathcal{S}, we need to compute the index on each connected component. Fortunately, we only need one of these connected components to prove our Theorem.

Definition.

Let ξ\xi be a square inscribed in a curve γ\gamma. We say ξ\xi is graceful if we orient the curve γ\gamma and consider the induced order on the vertices of ξ\xi, this order agrees with the one induced from the circle inscribing ξ\xi. Consider (h,θ)(h,\theta) in 𝒮\mathcal{S} and let ξ=Ψ(h,θ)\xi=\Psi(h,\theta). We say (h,θ)(h,\theta) is graceful if ξ\xi is a graceful square inscribed inside the corresponding curve to hh in TΣT\Sigma over π(θ)\pi(\theta), where π:Ka(Σ)Σ\pi\colon K_{a}(\Sigma)\to\Sigma is the bundle projection map.

We denote the subset of graceful squares in 𝒮\mathcal{S} by 𝒮0\mathcal{S}_{0}. We will prove 𝒮0\mathcal{S}_{0} is a connected component of 𝒮\mathcal{S}. Then we find the index of FF over this component. We expect the index to be two because 𝒮\mathcal{S} has codimension four and Ka(Σ)K_{a}(\Sigma) is six dimensional and indeed, this is what we will show.

Lemma 4.2.

Assume (h1,θ1)(h_{1},\theta_{1}) is in 𝒮0\mathcal{S}_{0} and there is a path γ\gamma in 𝒮\mathcal{S} from (h1,θ1)(h_{1},\theta_{1}) to (h2,θ2)(h_{2},\theta_{2}). Then (h2,θ2)(h_{2},\theta_{2}) is also graceful.

Proof.

Since the square corresponding to (h1,θ1)(h_{1},\theta_{1}) is graceful, if we consider this square in the fiber of TΣT\Sigma over π(θ1)\pi(\theta_{1}), the origin cannot lie in the regions A,B,C,A,B,C, and DD determined by this square in figure 11.

AABBCCDD
Figure 1: The four regions that cannot contain the origin

Suppose γ(s)=(hs,θs)\gamma(s)=(h_{s},\theta_{s}) is a path in 𝒮\mathcal{S} starting from (h1,θ1)(h_{1},\theta_{1}) and ending at (h2,θ2)(h_{2},\theta_{2}). By contradiction, assume the square corresponding to (h2,θ2)(h_{2},\theta_{2}) is not graceful. Hence, the origin will enter one of the regions A,B,C,A,B,C, or DD. Without loss of generality, let this region be AA. Then there is a time s0s_{0} in between the two ends where the origin lies on the line ll given in figure 22.

oollAAv1v_{1}v2v_{2}
Figure 2: Line ll and origin oo

Now the two vertices v1v_{1} and v2v_{2} of the square corresponding to (hs0,θs0)(h_{s_{0}},\theta_{s_{0}}) lie on the same side of the origin in ll and this contradicts the fact that the curve inscribing this square is star-shaped (this follows from positivity of hs0h_{s_{0}}). ∎

Proposition 4.3.

The space 𝒮0\mathcal{S}_{0} is connected. In particular, it is a connected component of 𝒮\mathcal{S}; therefore, it is a Banach manifold of codimension four in 𝒩\mathcal{N}.

Proof.

Consider a point (g,θ)(g,\theta) in 𝒮0\mathcal{S}_{0}. This point corresponds to a graceful square that is inscribed in a star-shaped curve around the origin in some fiber TxΣT_{x}\Sigma. We can also consider the constant function 11 and four vertices of a square δx\delta_{x} on the circle with radius aa around the origin in TxΣT_{x}\Sigma. This gives us a point (1,δx)(1,\delta_{x}) in 𝒮0\mathcal{S}_{0} and all the points of this form in 𝒮0\mathcal{S}_{0} can be connected to each other by parallel transport and rotation in their corresponding fibers. Hence, it suffices to prove there is a path between (g,θ)(g,\theta) and (1,δx)(1,\delta_{x}) in 𝒮0\mathcal{S}_{0}.

Since (g,θ)(g,\theta) corresponds to a graceful square, the origin in TxΣT_{x}\Sigma cannot lie in the four region defined in terms of this square given in figure 11. For any point outside of these four regions, we can find a sufficiently large ellipse going through the four vertices of the square such that the point is inside the ellipse. Consider such an ellipse for the origin; see the figure below.

oo
Figure 3: The blue curve is our star-shaped curve.

There is a positive function hh defined on the circle with radius aa around the origin so that the map

ηh(η)η\eta\mapsto h(\eta)\cdot\eta

takes the circle into the ellipse; extend this function to a positive function on Ua(Σ)U_{a}(\Sigma) and note that gg restricted to the circle of radius aa in TxΣT_{x}\Sigma will give us the corresponding function for our original curve in TxΣT_{x}\Sigma. We define a path of positive functions by ht=th+(1t)gh_{t}=th+(1-t)g. Note that hth_{t} is constant for each θi\theta_{i} in θ\theta because we have fixed the square and only move the curves. Hence, (ht,θ)(h_{t},\theta) gives a path from (g,θ)(g,\theta) to (h,θ)(h,\theta). Now we can translate the ellipse to an ellipse centered around the origin. This gives us a continuous path of functions gtg_{t} and tuples of points θt\theta_{t} with g0=hg_{0}=h and θ0=θ\theta_{0}=\theta because each ellipse inscribes a unique square and all the ellipses in the translation contain the origin. Thus we get a path from (g,θ)(g,\theta) to (u,σ)(u,\sigma) where Ψ(u,σ)\Psi(u,\sigma) is a square inscribed inside an ellipse centered around the origin in TxΣT_{x}\Sigma. Now we can take (u,σ)(u,\sigma) to (1,δx)(1,\delta_{x}) by first a homotopy fixing the four vertices of (u,σ)(u,\sigma) and then scaling and rotation.

We will need the two following lemmas for computing the index and the transversality argument after that.

Lemma 4.4.

Fix a point (h,θ)(h,\theta) in 𝒮\mathcal{S}. The space Ψh(Ka(Σ))\Psi_{h}(K_{a}(\Sigma)) is a six dimensional submanifold of Q(Σ)Q(\Sigma) and A(Σ)A(\Sigma) is also a six dimensional submanifold of Q(Σ)Q(\Sigma). There is a one to one correspondence between the kernel of dF:T(h,θ)𝒮ThdF\colon T_{(h,\theta)}\mathcal{S}\to T_{h}\mathcal{F} and TξΨh(Ka(Σ))TξA(Σ)T_{\xi}\Psi_{h}(K_{a}(\Sigma))\cap T_{\xi}A(\Sigma) where ξ=Ψ(h,θ)\xi=\Psi(h,\theta).

Proof.

Suppose vv is a vector in the kernel of dFdF over (h,θ)(h,\theta). Then we can write vv as a pair (0,η)(0,\eta) where η\eta is a vector in TθKa(Σ)T_{\theta}K_{a}(\Sigma) since FF is restriction of the first projection map to 𝒮\mathcal{S}. Now since vv lies in the tangent space of 𝒮\mathcal{S} over (h,θ)(h,\theta), we have

dΨ(h,θ)[(0,η)]=dΨh[η]TξA.d\Psi_{(h,\theta)}[(0,\eta)]=d\Psi_{h}[\eta]\in T_{\xi}A.

A similar argument shows if a vector δ\delta is in TξΨh(Ka(Σ))TξAT_{\xi}\Psi_{h}(K_{a}(\Sigma))\cap T_{\xi}A, then δ=dΨh[η]\delta=d\Psi_{h}[\eta] for some vector η\eta in TθKa(Σ)T_{\theta}K_{a}(\Sigma) and (0,η)(0,\eta) lies in the kernel of dFdF. ∎

Lemma 4.5.

Let (h,θ)(h,\theta) be a point in 𝒮\mathcal{S} and assume ξ=Ψ(h,θ)\xi=\Psi(h,\theta) is a transverse intersection point of Ψh(Ka(Σ))\Psi_{h}(K_{a}(\Sigma)) and A(Σ)A(\Sigma). Then dF:T(h,θ)𝒮ThdF\colon T_{(h,\theta)}\mathcal{S}\to T_{h}\mathcal{F} is surjective.

Proof.

We can identify ThT_{h}\mathcal{F} with the space of all C2C^{2} functions on Ua(Σ)U_{a}(\Sigma); let gg be an arbitrary function in ThT_{h}\mathcal{F}. Consider v=(g,0)v=(g,0) in T(h,θ)𝒩T_{(h,\theta)}\mathcal{N} and define w=dΨ(h,θ)[v]w=d\Psi_{(h,\theta)}[v] to be its image in TξQa(Σ)T_{\xi}Q_{a}(\Sigma). We can write ww as wA+wKw_{A}+w_{K} where wAw_{A} is in TξA(Σ)T_{\xi}A(\Sigma) and wKw_{K} is in TξKa(Σ)T_{\xi}K_{a}(\Sigma) because we assumed ξ\xi is a transverse intersection point. There exists a vector η\eta in TθKa(Σ)T_{\theta}K_{a}(\Sigma) such that wK=dΨh[η]w_{K}=d\Psi_{h}[\eta]. Now consider the vector v~=(g,η)\tilde{v}=(g,-\eta) in T(h,θ)𝒩T_{(h,\theta)}\mathcal{N}; we get

dΨ[v~]=dΨ[v]+dΨ[(0,η)]=wwK=wATξA(Σ).d\Psi[\tilde{v}]=d\Psi[v]+d\Psi[(0,-\eta)]=w-w_{K}=w_{A}\in T_{\xi}A(\Sigma).

Hence, v~\tilde{v} lies in T(h,θ)𝒮T_{(h,\theta)}\mathcal{S} and we have dF[v~]=gdF[\tilde{v}]=g. ∎

Proposition 4.6.

The map FF has index 22 over 𝒮0\mathcal{S}_{0}.

Proof.

Let xx be a point on Σ\Sigma. Assume e1,e2e_{1},e_{2} are an orthonormal basis for TΣT\Sigma in a disk DD around xx and let v1v_{1} and v2v_{2} denote the corresponding coordinates for TΣT\Sigma above DD. Define a local fiber bundle of ellipses around xx by

E{(y,v1,v2)|v12+2v22=1}.E\coloneqq\{(y,v_{1},v_{2})|v_{1}^{2}+2\cdot v_{2}^{2}=1\}.

There is a positive function hh defined on Ua(S2)U_{a}(S^{2}) above DD such that the map

ηh(η)η\eta\mapsto h(\eta)\cdot\eta

takes Ua(S2)U_{a}(S^{2}) for each point in DD to the ellipse above that point. Extend hh to a positive function h~\tilde{h} in \mathcal{F} and let θ\theta be the four-tuple of points in Ka(Σ)K_{a}(\Sigma) over xx that corresponds to the unique graceful square inscribed in ExE_{x}. Then ξ=Ψ(h~,θ)\xi=\Psi(\tilde{h},\theta) is a transverse intersection point of Ψh~(Ka(Σ))\Psi_{\tilde{h}}(K_{a}(\Sigma)) and A(Σ)A(\Sigma) because around xx all the corresponding curves are ellipses and these meet the space of squares transversely; in fact, Ψh~(Ka(Σ))A(Σ)\Psi_{\tilde{h}}(K_{a}(\Sigma))\cap A(\Sigma) is locally a two dimensional disk around ξ\xi. Therefore, by Lemma 4.4, we know ker(dF)\ker(dF) has dimension two at (h~,θ)(\tilde{h},\theta) and by Lemma 4.5, we know dFdF is surjective at this point. We conclude that FF has index two since 𝒮0\mathcal{S}_{0} is connected. ∎

Lemma 4.7.

The map FF restricted to 𝒮0\mathcal{S}_{0} is surjective.

Proof.

Consider a function hh in \mathcal{F}. We know Ψh(Ka(Σ))A(Σ)\Psi_{h}(K_{a}(\Sigma))\cap A(\Sigma) is non-empty and at least on of these intersection points corresponds to a graceful square since each of the star-shaped curves corresponding to hh in fibers of TΣT\Sigma inscribes at least one graceful square. We prove this Theorem in the appendix (Theorem A.5); see also [12, 4] for other versions of this result. Hence, there is a θKa(Σ)\theta\in K_{a}(\Sigma) such that (h,θ)(h,\theta) lies in 𝒮0\mathcal{S}_{0}. ∎

Proposition 4.8.

If hh is a regular value of FF restricted to 𝒮0\mathcal{S}_{0}, then Ψh(Ka(Σ))\Psi_{h}(K_{a}(\Sigma)) and A(Σ)A(\Sigma) intersect transversely in Q(Σ)Q(\Sigma) at graceful intersection points; in particular, the subset of graceful squares in Ψh(Ka(Σ))A(Σ)\Psi_{h}(K_{a}(\Sigma))\cap A(\Sigma) is a two dimensional manifold.

Proof.

Consider (h,θ)(h,\theta) in 𝒮0\mathcal{S}_{0} and ξ=Ψ(h,θ)\xi=\Psi(h,\theta) in A(Σ)A(\Sigma). Since hh is a regular value of FF, the kernel of dFdF at (h,θ)(h,\theta) is two dimensional and this proves TξΨh(Ka(Σ))TξA(Σ)T_{\xi}\Psi_{h}(K_{a}(\Sigma))\cap T_{\xi}A(\Sigma) is two dimensional by Lemma 4.4. Both TξΨh(Ka(Σ))T_{\xi}\Psi_{h}(K_{a}(\Sigma)) and TξA(Σ)T_{\xi}A(\Sigma) are six dimensional subspaces of TξQa(Σ)T_{\xi}Q_{a}(\Sigma) which has dimension ten. We conclude the two subspaces meet transversely because their intersection has dimension two so the subset of graceful squares in Ψh(Ka(Σ))A(Σ)\Psi_{h}(K_{a}(\Sigma))\cap A(\Sigma) is a surface. ∎

Corollary 4.9.

There exists a dense subset of functions in \mathcal{F} such that Ψh(Ka(Σ))\Psi_{h}(K_{a}(\Sigma)) intersects A(Σ)A(\Sigma) transversely at graceful squares for every function hh in this subset.

Proof.

Consider the regular values of FF; by Proposition 4.8, we know the intersection is transverse at graceful squares for every element in this subset. The map FF is a surjective CC^{\infty} Fredholm map of index two between connected second countable Banach manifolds. Thus its regular values are dense in \mathcal{F} by Sard-Smale Theorem; see [11] for more details. ∎

5 Proof of the Main Theorem

Lemma 5.1.

Suppose hh is a regular value of FF restricted to 𝒮0\mathcal{S}_{0}. Then the subset of graceful squares in Ψh(Ka(Σ))A(Σ)\Psi_{h}(K_{a}(\Sigma))\cap A(\Sigma) is a compact surface.

Proof.

Let Σh\Sigma_{h} denote the subset of graceful squares in Ψh(Ka(Σ))A(Σ)\Psi_{h}(K_{a}(\Sigma))\cap A(\Sigma) and note that this is a surface by proposition 4.8. Now consider a sequence an=Ψh(θn)a_{n}=\Psi_{h}(\theta_{n}) in Σh\Sigma_{h} and since θn\theta_{n} is in Ka(Σ)K_{a}(\Sigma), we can assume θn\theta_{n} converges to a point θ\theta in sym4(TΣ)sym^{4}(T\Sigma) after passing to a subsequence. The point θ\theta is not in the fat diagonal of sym4(TΣ)sym^{4}(T\Sigma) because the function hh is C2C^{2} and we can uniformly bound the curvature of all the star-shaped curves corresponding to hh; thus there is a positive lower bound for the side length of all the squares ana_{n} and Ψh(θ)\Psi_{h}(\theta) is a non-degenerate square. The square Ψh(θ)\Psi_{h}(\theta) is graceful since it is the limit of a sequence of graceful squares. ∎

Remark.

Note that for a function hh in regular values of F|𝒮0F\big{|}_{\mathcal{S}_{0}}, the surface Σh\Sigma_{h} is not necessarily orientable because Ka(Σ)K_{a}(\Sigma) is non-orientable.

Proposition 5.2.

Assume hh is a regular value of F|𝒮0F\big{|}_{\mathcal{S}_{0}}. Then [Σh][\Sigma_{h}] is non-zero in H2(sym4(TΣ);2)2H_{2}(sym^{4}(T\Sigma);\mathbb{Z}_{2})\cong\mathbb{Z}_{2}.

Proof.

Let π\pi denote the bundle projection map from sym4(TΣ)sym^{4}(T\Sigma) to Σ\Sigma; this is a deformation retract onto Σ\Sigma. Consider the restriction of π\pi to Σh\Sigma_{h}; we will show this map has non-zero mod 22 degree. Suppose xΣx\in\Sigma is a regular value of this map. Then for every θ\theta in π1(x)Σh\pi^{-1}(x)\cap\Sigma_{h}, TθΣhT_{\theta}\Sigma_{h} is a horizontal subspace of Tθsym4(TΣ)T_{\theta}sym^{4}(T\Sigma) by assumption. In particular, TθΨh(Ka(Σ))T_{\theta}\Psi_{h}(K_{a}(\Sigma)) and TθA(Σ)T_{\theta}A(\Sigma) have no vertical intersection and this proves the manifold of squares in TxΣT_{x}\Sigma and fourth symmetric product of the star-shaped curve corresponding to hh above xx meet transversely at every θ\theta in π1(x)Σh\pi^{-1}(x)\cap\Sigma_{h}. Therefore, this curve has finitely many graceful squares and the mod 22 degree of π\pi is equal to the number of graceful squares inscribed inside this curve mod 22. It is proved in the appendix (corollary A.6) that a generic star-shaped curve has an odd number of graceful squares; this was originally proved in [12] for generic smooth curves and it was proved in [10] for generic PL curves. The curve corresponding to hh above xx is a generic one because of the transversal intersection and we conclude the proposition. ∎

We define a map c:sym4(TΣ)TΣc\colon sym^{4}(T\Sigma)\to T\Sigma by

[θ1,θ2,θ3,θ4]θi4.[\theta_{1},\theta_{2},\theta_{3},\theta_{4}]\mapsto\frac{\sum\theta_{i}}{4}.

We call cc the center map; note that cc is a homotopy equivalence.

Proposition 5.3.

Suppose hh is a regular value of F|𝒮0F\big{|}_{\mathcal{S}_{0}} and Σ\Sigma is a surface with odd Euler characteristic. Then cc vanishes at some point on the surface Σh=Ψ(F1(h)𝒮0)\Sigma_{h}=\Psi(F^{-1}(h)\cap\mathcal{S}_{0}).

Proof.

By Proposition 5.2, we know [Σh][\Sigma_{h}] is a non-zero homology class. Hence, c[Σh]c_{*}[\Sigma_{h}] is also a non-zero homology class in TΣT\Sigma. The mod 22 intersection number of such homology classes with the zero section in TΣT\Sigma is equal to the second Stiefel–Whitney number of Σ\Sigma and this is equal to χ(Σ)\chi(\Sigma) mod 22; see [7] for more details. We conclude that this intersection number is non-zero because we assumed Σ\Sigma has odd Euler characteristic. ∎

Remark.

Proposition 5.3 shows for every regular value hh of F|𝒮0F\big{|}_{\mathcal{S}_{0}}, there is a point xx in Σ\Sigma so that the star shaped curve corresponding to hh above xx inscribes a graceful square centered around the origin in TxΣT_{x}\Sigma.

Lemma 5.4.

Consider a positive function hh in regular values of F|𝒮0F\big{|}_{\mathcal{S}_{0}} and let

Ψ(h,θ)=ξ=[ξ1,ξ2,ξ3,ξ4]\Psi(h,\theta)=\xi=[\xi_{1},\xi_{2},\xi_{3},\xi_{4}]

be a four-tuple in A(Σ)Ψh(Ka(Σ))A(\Sigma)\cap\Psi_{h}(K_{a}(\Sigma)) over a point xx in Σ\Sigma. Assume we have ξ1+ξ2+ξ3+ξ4=0\xi_{1}+\xi_{2}+\xi_{3}+\xi_{4}=0. Then we must have

h(θ1)=h(θ2)=h(θ3)=h(θ4)h(\theta_{1})=h(\theta_{2})=h(\theta_{3})=h(\theta_{4}) (3)

and θi\theta_{i}’s are vertices of a square inscribed in UaΣ|xU_{a}\Sigma\big{|}_{x}.

Proof.

After reordering the four-tuple ξ\xi, we can assume

ξ1=ξ3,ξ2=ξ4,\xi_{1}=-\xi_{3},\hskip 5.69054pt\xi_{2}=-\xi_{4},

and ξ1ξ2=0\xi_{1}\cdot\xi_{2}=0 since ξi\xi_{i}’s are vertices of a square centered at the origin. Thus we get

θ1θ2=ξ1ξ2h(θ1)h(θ2)=0.\theta_{1}\cdot\theta_{2}=\frac{\xi_{1}\cdot\xi_{2}}{h(\theta_{1})h(\theta_{2})}=0.

Moreover, we have

h(θ1)θ1=ξ1=ξ3=h(θ3)θ3.h(\theta_{1})\theta_{1}=\xi_{1}=-\xi_{3}=-h(\theta_{3})\theta_{3}.

Hence, we can write

θ3=h(θ1)h(θ3)θ1.\theta_{3}=\frac{-h(\theta_{1})}{h(\theta_{3})}\theta_{1}.

Since θ1\theta_{1} and θ3\theta_{3} have the same length, we deduce that h(θ1)=h(θ3)h(\theta_{1})=h(\theta_{3}) by positivity of hh and θ1=θ3\theta_{1}=-\theta_{3}. A similar argument shows h(θ2)=h(θ4)h(\theta_{2})=h(\theta_{4}) and θ2=θ4\theta_{2}=-\theta_{4}. Therefore, θi\theta_{i} ’s are vertices of a square on the circle with radius aa around the origin in TxΣT_{x}\Sigma; two vertices of this square are scaled by h(θ1)h(\theta_{1}) and the other two by h(θ2)h(\theta_{2}). Since the scaled shape is also a square by assumption, we conclude that h(θ1)=h(θ2)h(\theta_{1})=h(\theta_{2}). ∎

Definition.

Assume ff is a positive function on Σ\Sigma. We define a positive function f~\tilde{f} on Ua(Σ)U_{a}(\Sigma) by

(x,v)f(exp(x,v)).(x,v)\mapsto f(\exp(x,v)).
Proof of Theorem 1.1.

Since Σ\Sigma is compact, it suffices to prove the Theorem for positive functions. Let ff be a positive continuous function on Σ\Sigma and consider f~\tilde{f} on UaΣU_{a}\Sigma for a=d2a=\frac{d}{2}. By Corollary 4.9, we can find a sequence of functions unu_{n} in regular values of F|𝒮0F\big{|}_{\mathcal{S}_{0}} such that unu_{n} converges to f~\tilde{f} uniformly on UaΣU_{a}\Sigma. By Lemma 5.4 and Proposition 5.3, we know there is a sequence of graceful squares θn\theta_{n} inscribed inside the fibers of UaΣU_{a}\Sigma such that unu_{n} takes the same value on the four vertices of θn\theta_{n} for every nn. After passing to a subsequence, we can assume θn\theta_{n} converges to θ\theta, a square with the same side length as θn\theta_{n}’s and inscribed inside the fiber of UaΣU_{a}\Sigma over a point xx in Σ\Sigma. All the vertices of θ\theta take the same value under f~\tilde{f} by uniform convergence and the assumption on unu_{n}’s. Hence, we conclude ff admits a table determined by the four vertices of θ\theta. ∎

Proof of Corollary 1.2.

Let ff be an even function on S2S^{2} and gg a Riemannian metric invariant under the antipodal map. This gives us a function f¯\bar{f} and a Riemannian metric g¯\bar{g} on P2\mathbb{R}P^{2}. Now apply Theorem 1.1 to this Riemannian surface and the function f¯\bar{f}. The table for f¯\bar{f} on P2\mathbb{R}P^{2} lifts to two tables for ff on S2S^{2}. ∎

Appendix

Appendix A Square Peg for Star-Shaped Curves

Our goal in this appendix is to prove the square peg problem for C2C^{2} star-shaped curves (Theorem A.5). We will also prove a generic star-shaped curve has an odd number of squares; this is corollary A.6. Furthermore, we will show that if we orient a generic star-shaped curve, then the curve inscribes an odd number of squares that are consistent with this orientation. The first version of this result was proved in [12] for all smooth curves; see [5] for a modern version of this proof. We will reprove this result for C2C^{2} star-shaped curves using similar ideas to [5] in combination with modern transversality arguments.

Definition.

Let γ\gamma be an oriented curve in the plane and suppose QQ is an inscribed square inside γ\gamma. We say QQ is graceful if γ\gamma induces the same order on the vertices of QQ as the circle that inscribes this square.

We will prove every star-shaped curve inscribes a graceful square. Let hh be a positive C2C^{2} function on S1S^{1}; we can define a C2C^{2} curve in the plane via the following.

θh(θ)θ,θS1.\theta\mapsto h(\theta)\cdot\theta,\hskip 8.53581pt\forall\theta\in S^{1}.

Every C2C^{2} star-shaped curve can be parametrized by a positive C2C^{2} function on S1S^{1} in this manner. Let Δ3\Delta_{3} denote the three dimensional simplex and consider its interior Δ̊3\mathring{\Delta}_{3}. Suppose γ\gamma is a star-shaped curve in the plane and it is parametrized by a positive function hh on S1S^{1}. We can parametrize all the quadrilaterals inscribed in γ\gamma with S1×Δ̊3S^{1}\times\mathring{\Delta}_{3} in the following way.

[x,(t0,t1,t2,t3)]\displaystyle[x,(t_{0},t_{1},t_{2},t_{3})]\mapsto [f(x)x,f(eiπt0x)(eiπt0x),f(eiπ(t0+t1)x)(eiπ(t0+t1)x),\displaystyle[f(x)\cdot x,f(e^{i\pi t_{0}}\cdot x)\cdot(e^{i\pi t_{0}}\cdot x),f(e^{i\pi(t_{0}+t_{1})}x)\cdot(e^{i\pi(t_{0}+t_{1})}\cdot x),
f(eiπ(t0+t1+t2)x)(eiπ(t0+t1+t2)x)],\displaystyle f(e^{i\pi(t_{0}+t_{1}+t_{2})}\cdot x)\cdot(e^{i\pi(t_{0}+t_{1}+t_{2})}\cdot x)],

where xx is a point in S1S^{1} and tt is an interior point of Δ3\Delta_{3}. We will denote S1×Δ̊3S^{1}\times\mathring{\Delta}_{3} by P~\tilde{P} and the above equation gives us a map from P~\tilde{P} to 8\mathbb{R}^{8}. Every positive function hh on S1S^{1} gives us such a map and we denote this map by φh\varphi_{h}

Consider all the four tuples (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) in (2)48(\mathbb{R}^{2})^{4}\cong\mathbb{R}^{8} such that

x1x2=\displaystyle\|x_{1}-x_{2}\|= x2x3=x3x4=x4x1,\displaystyle\|x_{2}-x_{3}\|=\|x_{3}-x_{4}\|=\|x_{4}-x_{1}\|,
x1x3=x2x4,\displaystyle\|x_{1}-x_{3}\|=\|x_{2}-x_{4}\|,

and all the xix_{i}’s are distinct; we denote this subset of 8\mathbb{R}^{8} by A~\tilde{A}. This space is a non-compact submanifold of dimension 44 in 8\mathbb{R}^{8}.

Remark.

For a fixed positive function hh on S1S^{1}, the set φh(P~)A~\varphi_{h}(\tilde{P})\cap\tilde{A} corresponds to graceful squares inscribed inside the star-shaped curve parametrized by hh. Every graceful square of this curve corresponds to four points in this intersection.

Remark.

Note that φh\varphi_{h} is an embedding of P~\tilde{P} into 8\mathbb{R}^{8} and image of this map avoids the fat diagonal in 8(2)4\mathbb{R}^{8}\cong(\mathbb{R}^{2})^{4} for every positive function hh.

There is a free action of 4\mathbb{Z}_{4} on P~\tilde{P} generated by

[x,(t0,t1,t2,t3)][eiπt0x,(t1,t2,t3,t0)].[x,(t_{0},t_{1},t_{2},t_{3})]\mapsto[e^{i\pi t_{0}}\cdot x,(t_{1},t_{2},t_{3},t_{0})].

We denote this generator by ε\varepsilon. There is also a 4\mathbb{Z}_{4} action on 8(2)4\mathbb{R}^{8}\cong(\mathbb{R}^{2})^{4} generated by

(x1,x2,x3,x4)(x2,x3,x4,x1).(x_{1},x_{2},x_{3},x_{4})\mapsto(x_{2},x_{3},x_{4},x_{1}).

This action is free away from the fat diagonal in 8\mathbb{R}^{8}. Note that φh\varphi_{h} is equivariant with respect to the cyclic actions on its range and domain. We quotient P~\tilde{P} by this action and denote the resulting space by PP; we also quotient complement of the fat diagonal in 8(2)4\mathbb{R}^{8}\cong(\mathbb{R}^{2})^{4} by the cyclic action and denote the resulting space by VV. Since φh\varphi_{h} is equivariant, it descends to a map from PP to VV for every positive function hh; by an abuse of notation, we also denote this map by φh\varphi_{h}. Let AA be the quotient of A~\tilde{A} in VV. Now there is only one intersection point in φh(P)A\varphi_{h}(P)\cap A corresponding to each graceful square inscribed in the star-shaped curve parametrized by hh.

Define a map Φ:C+2(S1)×PV\Phi\colon C^{2}_{+}(S^{1})\times P\to V by

Φ(h,(x,t))=φh(x,t).\Phi(h,(x,t))=\varphi_{h}(x,t).

Note that C+2(S1)C^{2}_{+}(S^{1}) is an open subset of C2(S1)C^{2}(S^{1}) and in particular, it is a Banach manifold.

Lemma A.1.

The map Φ\Phi is a submersion.

Proof.

Consider a point (h,(x,t))(h,(x,t)) in C+2(S1)×PC^{2}_{+}(S^{1})\times P and suppose t=(t0,t1,t2,t3)t=(t_{0},t_{1},t_{2},t_{3}). Let δ\delta be a positive number less than t0t_{0}. We define a curve γ\gamma in PP given by

γ(s)=(e2πisx,t0s,t1,t2,t3+s)\gamma(s)=(e^{2\pi is}\cdot x,t_{0}-s,t_{1},t_{2},t_{3}+s)

for ss in [0,δ)[0,\delta). This curve moves the first vertex of the quadrilateral corresponding to (x,t)(x,t) and fixes the other three. Now if we consider the curve (h,γ(s))(h,\gamma(s)) in C+2(S1)×PC^{2}_{+}(S^{1})\times P, we have

Φ(h,γ(s))=(h(e2πisx)e2πisx,h(x2)x2,h(x3)x3,h(x4)x4)\Phi(h,\gamma(s))=(h(e^{2\pi is}\cdot x)\cdot e^{2\pi is}\cdot x,h(x_{2})\cdot x_{2},h(x_{3})\cdot x_{3},h(x_{4})\cdot x_{4})

where x2,x3,x_{2},x_{3}, and x4x_{4} are the other three vertices corresponding to (x,t)(x,t). Hence, if we let vv denote the derivative of (h,γ(s))(h,\gamma(s)) at s=0s=0, we get

dΦ(h,(x,t))[v]=(h(x)2πix+h(x)x,0,0,0),d\Phi_{(h,(x,t))}[v]=(h(x)\cdot 2\pi i\cdot x+h^{\prime}(x)\cdot x,0,0,0), (4)

where h(x)h(x) is a positive number by definition and h(x)h^{\prime}(x) is an arbitrary real number. Similarly, we can move the other vertices and get vectors of the form in equation (4) such that all the coordinates are zero except one of them and the non-zero coordinate is equal to rix+axr\cdot i\cdot x+a\cdot x for a positive rr and an arbitrary number aa. Now consider a C2C^{2} function gg on S1S^{1} so that we have g(x)=1g(x)=1 and g(x2)=g(x3)=g(x4)=0g(x_{2})=g(x_{3})=g(x_{4})=0. For small real numbers ss, all the functions h+sgh+s\cdot g will be in C+2(S1)C^{2}_{+}(S^{1}) and we have

dΦ(h,(x,t))[(g,0)]=dds|0Φ(h+sg,(x,t))=(g(x)x,g(x2)x2,g(x3)x3,g(x4)x4)=(x,0,0,0).d\Phi_{(h,(x,t))}[(g,0)]=\frac{d}{ds}\Big{|}_{0}\Phi(h+s\cdot g,(x,t))=(g(x)\cdot x,g(x_{2})\cdot x_{2},g(x_{3})\cdot x_{3},g(x_{4})\cdot x_{4})=(x,0,0,0). (5)

We conclude the proof since all the vectors of the form given in equations (4) and (5) generate 8TΦ(h,(x,t))V\mathbb{R}^{8}\cong T_{\Phi(h,(x,t))}V. ∎

Now that we know Φ\Phi is a submersion, we conclude that Φ1(A)\Phi^{-1}(A) is a codimension 44 submanifold of C+2(S1)×PC^{2}_{+}(S^{1})\times P. We denote this submanifold by 𝒬\mathcal{Q} and let π:𝒬C+2(S1)\pi\colon\mathcal{Q}\to C^{2}_{+}(S^{1}) be restriction of the first projection map to 𝒬\mathcal{Q}.

Lemma A.2.

The space 𝒬\mathcal{Q} is connected and the map π\pi is Fredholm. Moreover, π\pi has index 0.

Proof.

The connectivity of 𝒬\mathcal{Q} follows the same way we proved 𝒮0\mathcal{S}_{0} is connected in Proposition 4.3 and π\pi being Fredholm follows from the same strategy in 4.1. For the index computation, take a function gg such that the curve parametrized by gg is an ellipse. Then we know φg(P)\varphi_{g}(P) intersect AA transversely and we can prove π\pi has index zero at this point using an argument similar to the one given in Proposition 4.6. ∎

Definition.

Let hh be a positive function on S1S^{1}. We say the star-shaped curve corresponding to hh is generic if φh(P)\varphi_{h}(P) intersects AA transversely. Note that a C2C^{2} generic curve has finitely many graceful squares.

In the following, we call a positive function generic if its corresponding curve is generic.

Lemma A.3.

A positive function hh is generic if and only if it is a regular value of π\pi.

Proof.

This can be proved using a similar argument as in Lemma 4.4 and 4.5. ∎

Corollary A.4.

The set of generic functions are dense in C+2(S1)C^{2}_{+}(S^{1}).

Proof.

This follows from Sard-Smale Theorem. ∎

Theorem A.5.

Every star-shaped C2C^{2} curve inscribes at least one graceful square.

Proof.

By contradiction, assume there is a star-shaped C2C^{2} curve that does not admit a graceful square and let hh be the positive function that parametrizes this function. By assumption, hh is not in the image of π\pi so it is a regular value of π\pi by definition. Let gg be a positive function that parametrizes an ellipse; hence gg is in regular values of π\pi and π1(g)\pi^{-1}(g) is just a point corresponding to the unique square inscribed inside this ellipse. Consider a path of positive functions hsh_{s} in C+2(S1)C^{2}_{+}(S^{1}) such that

h0=g,h1=hh_{0}=g,\hskip 8.53581pth_{1}=h

and π1(hs)\pi^{-1}(h_{s}) is a one manifold. We can find such a generic path because both gg and hh are regular values of π\pi. The one manifold π1(hs)\pi^{-1}(h_{s}) is compact since we can uniformly bound the total curvature of all the curves corresponding to hsh_{s} for each ss. This compact one manifold has only one boundary point corresponding to the square inscribed inside the ellipse which is a contradiction. ∎

We get the following as a corollary of the cobordism argument given in the previous proof.

Corollary A.6.

A generic star-shaped curve inscribes an odd number of graceful squares.

References

  • [1] Jason Cantarella, Elizabeth Denne, and John McCleary, Configuration spaces, multijet transversality, and the square-peg problem, Illinois J. Math. 66 (2022), no. 3, 385–420. MR 4477422
  • [2] F. J. Dyson, Continuous functions defined on spheres, Ann. of Math. (2) 54 (1951), 534–536. MR 44620
  • [3] Roger Fenn, The table theorem, Bull. London Math. Soc. 2 (1970), 73–76. MR 271940
  • [4] Joshua Evan Greene and Andrew Lobb, The rectangular peg problem, Ann. of Math. (2) 194 (2021), no. 2, 509–517. MR 4298749
  • [5] Benjamin Matschke, Equivariant topology methods in discrete geometry, (2011).
  • [6] Benjamin Matschke, A survey on the square peg problem, Notices Amer. Math. Soc. 61 (2014), no. 4, 346–352. MR 3184501
  • [7] John W. Milnor and James D. Stasheff, Characteristic classes, Annals of Mathematics Studies, vol. No. 76, Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1974. MR 440554
  • [8] John Douglas Moore, Introduction to global analysis, Graduate Studies in Mathematics, vol. 187, American Mathematical Society, Providence, RI, 2017, Minimal surfaces in Riemannian manifolds. MR 3729450
  • [9] H. R. Morton, Symmetric products of the circle, Proc. Cambridge Philos. Soc. 63 (1967), 349–352. MR 210096
  • [10] Richard Evan Schwartz, Rectangles, curves, and Klein bottles, Bull. Amer. Math. Soc. (N.S.) 59 (2022), no. 1, 1–17. MR 4340824
  • [11] S. Smale, An infinite dimensional version of Sard’s theorem, Amer. J. Math. 87 (1965), 861–866. MR 185604
  • [12] L. G. Snirelman, On certain geometrical properties of closed curves, Uspehi Matem. Nauk 10 (1944), 34–44. MR 12531

Boston College. Massachusetts, USA.

naserisa@bc.edu