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

Rep-tiles

Ryan Blair, Patricia Cahn, Alexandra Kjuchukova,
And Hannah Schwartz
(Date: June 2025)
Abstract.

An nn-dimensional rep-tile is a compact, connected submanifold of n\mathbb{R}^{n} with non-empty interior which can be decomposed into pairwise isometric rescaled copies of itself whose interiors are disjoint. We show that every smooth compact nn-dimensional submanifold of n\mathbb{R}^{n} with connected boundary is topologically isotopic to a polycube that tiles the nn-cube, and hence is topologically isotopic to a rep-tile. Consequently, there is a rep-tile in the homotopy type of any finite CW complex. In addition to classifying rep-tiles in all dimensions up to isotopy, we also give new explicit constructions of rep-tiles in all dimensions, including examples in the homotopy type of any finite bouquet of spheres.

1. Manifolds which are rep-tiles

We prove that every compact codimension-0 smooth submanifold of n\mathbb{R}^{n} with connected boundary can be topologically isotoped to a polycube which tiles the cube [0,1]n[0,1]^{n}. As a consequence, any such manifold is topologically isotopic to a rep-tile, defined next. A rep-tile XX is a codimension-0 subset of n\mathbb{R}^{n} with non-empty interior which can be written as a finite union X=iXiX=\bigcup_{i}X_{i} of pairwise isometric sets Xi,X_{i}, each of which is similar to XX; and such that Xi,XjX_{i},X_{j} have non-intersecting interiors whenever iji\neq j. A rep-tile in n\mathbb{R}^{n} which is also homeomorphic to a compact smooth manifold will be called a nn-dimensional rep-tile.

Since every nn-dimensional rep-tile has connected boundary (Lemma 3.2), as does every nn-dimensional manifold that tiles the nn-cube, our result proves that any submanifold of n\mathbb{R}^{n} which could potentially be homeomorphic to an nn-dimensional rep-tile is in fact isotopic to one. Thus, our work completes the isotopy classification of manifolds that tile the cube, and of nn-dimensional rep-tiles, in all dimensions.

Early sightings of rep-tiles were recorded in [Gar63, Gol64]. Because nn-dimensional rep-tiles tile n\mathbb{R}^{n}, rep-tiles have been studied not only for their intrinsic beauty but also in connection with tilings of Eucidean space; see [Gar77] or [Rad21] for a discussion of the case n=2n=2. A notable achievement was a non-periodic tiling of the plane by a rep-tile, due to Conway, which was later used to create the first example of a pinwheel tiling, i.e. one in which the tile occurs in infinitely many orientations [Rad94]. The elegant 2-dimensional rep-tile portrayed in Figure 1 was the building block in one of Goodman-Strauss’s constructions of a hierarchical tiling of 2\mathbb{R}^{2} [Goo98] and is also found in [Thu89].

Refer to caption
Figure 1. The “chair” rep-tile.

The first planar rep-tile with non-trivial fundamental group was discovered by Grünbaum, settling a question of Conway [Cro91, C17]. In 1998, Gerrit van Ophuysen found the first example of a rep-tile homeomorphic to a solid torus, answering a question by Goodman-Strauss [vOp97]. Tilings of 3\mathbb{R}^{3} of higher genus were also constructed in [Sch94]. Tilings of BnB^{n} by mutually isometric knots were constructed by [Oh96]. Adams proved that any compact submanifold of n\mathbb{R}^{n} with connected boundary tiles it [Ada95]. Building on the above work, in 2021 came the homeomorphism classification of 3-dimensional rep-tiles.

Theorem 1.1.

[Bla21] A submanifold RR of 3\mathbb{R}^{3} is homeomorphic to a 3-dimensional rep-tile if and only if it is homeomorphic to the exterior of a finite connected graph in S3S^{3}.

The above implies that any 3-manifold which could potentially be homeomorphic to a rep-tile is indeed homeomorphic to one. This follows from Fox’s re-embedding theorem [Fox48], which classifies compact 3-manifolds that embed in S3S^{3}, together with Lemma 3.2, which shows that a rep-tile has connected boundary.

Our main result is Theorem 1.2, which completes the isotopy classification of manifold rep-tiles in all dimensions, without relying on a classification of codimension-0 submanifolds of n\mathbb{R}^{n}.

Theorem 1.2.

Let RnR\subset\mathbb{R}^{n} be a compact smooth nn-manifold with connected boundary. Then, RR is topologically isotopic to a rep-tile.

Corollary 1.2.1.

Let XX be a compact connected CW complex of dimension n1n\geq 1. Then XX is homotopy equivalent to a (2n+1)(2n+1)-dimensional rep-tile.

Proof.

Suppose that XX is a compact connected CW complex of dimension nn. Then, XX embeds in 2n+1\mathbb{R}^{2n+1}, by the Nöbeling-Pontryagin Theorem [Den90, p. 125, Theorem 9]. Let RR be a closed regular neighborhood of XX in 2n+1\mathbb{R}^{2n+1}. Then RR is a compact (2n+1)(2n+1)-manifold embedded in 2n+1\mathbb{R}^{2n+1}. Moreover, RR has a single boundary component. Indeed, suppose (R)\partial(R) has two or more connected components N1,N2,Nk.N_{1},N_{2},\dots N_{k}. Since NjN_{j} is a closed 2n2n manifold embedded in 2n+1\mathbb{R}^{2n+1}, it is orientable, so H2n(Nj;)H_{2n}(N_{j};\mathbb{Z})\cong\mathbb{Z}. Moreover, since k>1k>1, we have that H2n+1(R,Nj)=0H_{2n+1}(R,N_{j})=0 for each j=1,2,,k.j=1,2,\dots,k. Therefore, we see from the long exact sequence of the pair that the inclusion-induced map i:H2n(Nj)H2n(R)i_{\ast}:H_{2n}(N_{j})\to H_{2n}(R) is injective. But RR has the homotopy type of an nn-complex, which is a contradiction. Therefore, by Theorem 1.2, RR is isotopic to a rep-tile. ∎

The proof of Theorem 1.2 describes a procedure for isotoping any codimension-0 smooth submanifold RR of n\mathbb{R}^{n} with connected boundary to a rep-tile. While the proof is constructive, in effect it is done without writing down any new rep-tiles. In Section 2 we therefore also give, for any n0n\geq 0, an explicit construction of a rep-tile homeomorphic to Sn×D2S^{n}\times D^{2}. This leads to an almost equally explicit construction of a rep-tile in the homotopy type of any finite bouquet of spheres. In particular, we can build explicit rep-tiles with non-vanishing homotopy groups in arbitrarily many dimensions.

The paper is organized as follows: in Section 2 we construct a rep-tile homeomorphic to Sn×D2S^{n}\times D^{2}, presented explicitly as a union of cubes in n+2\mathbb{R}^{n+2}, introduce the technique of ball swapping, show how to construct rep-tiles homotopy equivalent to wedges of spheres, investigate suspensions of rep-tiles, and construct rep-tiles with arbitrary footprints. Section 3 is where we prove the main theorem.

1.1. Rep-tiles and tilings of Euclidean space

Rep-tiles induce self-similar tilings of Euclidean space. Thus, they can potentially be used to construct non-periodic and aperiodic tilings of the plane and higher-dimensional Euclidean spaces. Self-similar tilings have connections to combinatorial group theory [Con90], propositional logic [Wan60, Ber66, Rob71] (where some of the questions in the field originated), and dynamical systems [Thu89], among others. Our rep-tiles give new self-similar tilings of n\mathbb{R}^{n} by tiles with interesting topology. Additionally, in our proof of Theorem 1.2 we show that every compact smooth nn-manifold with connected boundary in n\mathbb{R}^{n} is topologically isotopic to a polycube that tiles a cube. In [Gol66], Golomb developed a hierarchy for polycubes that tile n\mathbb{R}^{n}, and tiling a cube is the most restrictive level of his hierarchy. Hence all compact smooth nn-manifolds with connected boundary lie in the most restrictive level of Golomb’s hierarchy, up to isotopy.

2. Explicit rep-tiles in all dimensions

In this section, an nn-dimensional polycube is a union of unit nn-cubes whose vertices lie on the integer lattice in n\mathbb{R}^{n}. We will be repeatedly using the fact that a polycube that tiles a cube is a rep-tile (see Lemma 3.1). In this section we will realize the homotopy type of certain mm-manifolds XX as rep-tiles by the following procedure. We will construct a polycube RmR\subset\mathbb{R}^{m} such that RXR\simeq X (where \simeq denotes homotopy equivalence) and such that two copies of RR, related by a rotation, tile the mm-dimensional cube. In the case where XX has the homotopy type of a sphere, XSnX\simeq S^{n}, the rep-tile RR we build is homeomorphic to a trivial 2-dimensional disk bundle on the sphere, RSn×D2R\cong S^{n}\times D^{2}. (This demonstrates that rep-tiles can have non-trivial πn\pi_{n} for all n0,n\geq 0, answering Conway’s and Goodman-Strauss’s question in dimensions three and higher.) It is a fairly straightforward consequence that finite wedges of spheres of different dimensions can be built similarly.

2.1. Stacks of cubes

A stack of nn-cubes with stacking direction xnx_{n} is an nn-dimensional polycube SnS\subset\mathbb{R}^{n} such that: (1) All nn-cubes in SS lie above the hyperplane xn=0x_{n}=0 (that is, every point in SS has non-negative xnx_{n} coordinate), and (2) for every nn-cube in SS that does not have a face contained in xn=0x_{n}=0, there is another nn-cube of SS directly below it (where height is measured by the xnx_{n}-axis).

Let the subspace of n\mathbb{R}^{n} determined by xn=0x_{n}=0 have the standard tiling by (n1)(n-1)-cubes induced by the integer lattice in n\mathbb{R}^{n}. Given SS, a stack of nn-cubes with stacking direction xnx_{n}, we consider its projection to the hyperplane xn=0x_{n}=0, which we call its footprint. By the definition of a stack of cubes, we can think of SS as consisting of columns of nn-cubes lying above each (n1)(n-1)-cube in its footprint S\mathcal{F}_{S}, which is itself an (n1)(n-1)-dimensional polycube. In other words, the homotopy type of SS is determined by S\mathcal{F}_{S}; and SS itself is determined by S\mathcal{F}_{S}, together with integer labels in each (n1)(n-1)-cube of S\mathcal{F}_{S}, specifying the height of the column of nn-cubes which lie above it. Therefore, we can describe SS by such a labeled footprint. Figure 2 illustrates a 2-dimensional stack of cubes (left) and its description via a labeling on its 1-dimensional footprint (right).

The image of such a stack of cubes under an isometry of n\mathbb{R}^{n} is also called a stack of cubes, with the image of xnx_{n} under the isometry being the stacking direction.

Refer to caption
Figure 2. A stack of cubes (left) and its labeled footprint (right). This polycube and its image under rotation by π\pi about P0P_{0} tile [0,4]2[0,4]^{2}. Thus, the polycube is a rep-tile.

2.2. Rep-tiles homotopy equivalent to SnS^{n}.

We use cube-stacking notation as above to describe a rep-tile homeomorphic to Sn×D2S^{n}\times D^{2} for all n0n\geq 0. This description is a simplification, suggested by Richard Schwartz [Sch25], of the construction given in [Bla24]. We define a stack of cubes S[0,4]n+2S\subset[0,4]^{n+2} as follows. The footprint S\mathcal{F}_{S} is a polycube in [0,4]n+1[0,4]^{n+1}.

Refer to caption
Figure 3. Top : Footprint of a 3-dimensional rep-tile homeomorphic to S1×D2S^{1}\times D^{2} (left), and its corresponding stack of cubes (right), rotated by 90 degrees for visualization. Bottom: Footprint of a 4-dimensional rep-tile homeomorphic to S2×D2S^{2}\times D^{2} (left), and space to imagine the corresponding stack of cubes (right).

Throughout the following discussion, the reader should refer to Figure 3. Define the core, denoted 𝒞\mathcal{C}, of [0,4]n+1[0,4]^{n+1} to be the union of unit cubes in the standard integer-lattice tiling of [0,4]n+1[0,4]^{n+1} containing the point (2,,2)(2,...,2). The shell of [0,4]n+1[0,4]^{n+1} is [0,4]n+1𝒞¯.\overline{[0,4]^{n+1}\setminus\mathcal{C}}. To create the labeled footprint S\mathcal{F}_{S} of our stack of cubes SS, we first partition 𝒞\mathcal{C} into two halves: 𝒞+\mathcal{C}^{+}, those containing cubes with xn+1x_{n+1}-coordinate at least 2; and 𝒞\mathcal{C}^{-}, those containing cubes with xn+1x_{n+1}-coordinate less than 2. Finally, we label each cube in 𝒞+\mathcal{C}^{+} with a 4, and each cube in 𝒞\mathcal{C}^{-} with a 0. All cubes in the shell are labeled 2. (We recall that the label of each (n+1)(n+1)-cube in the footprint indicates the height of the column of (n+2)(n+2)-cubes stacked on top of it.) Observe that S\mathcal{F}_{S}, which consists of all unit cubes in [0,4]n+1[0,4]^{n+1} with nonzero label, is homeomorphic to the shell, which is in turn homeomorphic to Sn×D1S^{n}\times D^{1}. Similarly, the stack of cubes SS determined by this labeling is homeomorphic to S×ISn×D2\mathcal{F}_{S}\times I\cong S^{n}\times D^{2}.

Next we show that SS is a rep-tile. Let rπ:n+2n+2r_{\pi}:\mathbb{R}^{n+2}\rightarrow\mathbb{R}^{n+2} denote rotation by π\pi about the nn-plane which is the intersection of xn+2=2x_{n+2}=2 and xn+1=2x_{n+1}=2. Observe that the closure of the complement of SS in [0,4]n+2[0,4]^{n+2} is also a stack of cubes, with stacking direction xn+2-x_{n+2}, is isometric to SS, and in particular, is the image of SS under rπ.r_{\pi}. As SS and rπ(S)r_{\pi}(S) tile the cube [0,4]n+2[0,4]^{n+2}, and since SS is a union of cubes, SS is a rep-tile.

2.3. Ball swapping.

We note that there is a lot of flexibility regarding the heights of columns in the construction of a rep-tile SSn×D2S\cong S^{n}\times D^{2} given above. Consider any column HH in SS of height h{1,2,3}h\in\{1,2,3\}. Let HH^{\prime} denote the column of SS which shares a footprint with rπ(H)r_{\pi}(H). Since Hrπ(H)H^{\prime}\cup r_{\pi}(H) form a column of height 4, the heights of HH and HH^{\prime} add up to 4. Moreover, unit cubes can be traded between HH and HH^{\prime} while preserving the property that the resulting polycube and its image under rπr_{\pi} tile [0,4]n+2[0,4]^{n+2}. As long as both columns remain of height strictly between 0 and 4 and their heights add up to 4, this swap preserves both the homeomorphism type of SS and the property that two copies of SS tile a cube.

More generally, let RR be any non-empty nn-dimensional polycube in n\mathbb{R}^{n}. Let GG be a group of isometries of n\mathbb{R}^{n} such that the orbit of RR under GG tiles a cube 𝒞\mathcal{C}. (As before, this implies that RR is a rep-tile.) Let uu denote any unit cube contained in RR and let gg be an arbitrary element of GG. Denote by gugu the image of uu under gg. We note that R:=R\u¯guR^{\prime}:=\overline{R\backslash u}\cup gu is also a polycube whose orbit under GG is 𝒞\mathcal{C}. Hence, RR^{\prime} is also a rep-tile. We will refer to this move as a ball swap. (Aside: the fact that performing a ball swap on a polycube that tiles an nn-cube produces another polycube that tiles an nn-cube does not depend in an essential way on the fact that uu is a nn-cube. More complicated pieces could be swapped as well, preserving the tiling property.) Ball swapping, which was inspired by work of Adams [Ada95, Ada97], turns out to be a powerful tool for building rep-tiles, as we will see in Section 3. To be precise, a version of the ball swap – one which involves an action of a group of order 2m2^{m} on [1,1]m[-1,1]^{m} and trading multiple balls simultaneously across their individual orbits – is the key idea in the Proof of Theorem 1.2. The second main ingredient in the proof is this: a priori, RR^{\prime} might not have a clear relationship to RR; to guarantee that RR^{\prime} is homeomorphic to or isotopic to RR, care must be taken in the choice of group action and the choice of uu.

2.4. Rep-tilean bouquets

Let Pn={xn+2|xn+2=xn+1=2}P_{n}=\{\textbf{x}\in\mathbb{R}^{n+2}|x_{n+2}=x_{n+1}=2\}. The construction in Section 2.2 has produced stacks of (n+2)(n+2)-dimensional cubes in [0,4]n+2[0,4]^{n+2} with the following useful properties:

  1. (1)

    Each polycube intersects x1=0x_{1}=0 in an (n+1)(n+1)-ball equal to {0}×[0,4]n×[0,2]\{0\}\times[0,4]^{n}\times[0,2] and intersects x1=4x_{1}=4 in an (n+1)(n+1)-ball equal to {4}×[0,4]n×[0,2]\{4\}\times[0,4]^{n}\times[0,2];

  2. (2)

    the polycube and its image under rotation by π\pi about PnP_{n} tile [0,4]n+2[0,4]^{n+2}.

Note that any two such polycubes of the same dimension R1R_{1} and R2R_{2} can be placed side-by-side in the x1x_{1} direction so that R1R_{1} is contained in 0x140\leq x_{1}\leq 4 and R2R_{2} is contained in 4x184\leq x_{1}\leq 8. For example, place two copies of the stack of cubes in the top of Figure 3 back-to-back. In this configuration R1R2={4}×[0,4]n×[0,2]Bn+1R_{1}\cap R_{2}=\{4\}\times[0,4]^{n}\times[0,2]\cong B^{n+1}. Thus, R1R2R_{1}\cup R_{2} has the homotopy type of the wedge R1R2R_{1}\vee R_{2}; and, after rescaling in the x1x_{1} direction and subdividing the integer lattice, it too satisfies the conditions (1) and (2) above.

Now consider SmS_{m} and SkS_{k}, two of the rep-tiles constructed in Section 2.2 of dimension mm and kk respectively. If mkm\leq k, then Sm×DkmS_{m}\times D^{k-m} can be embedded in [0,4]k+2[0,4]^{k+2} so that conditions (1) and (2) hold. By stacking SkS_{k} and this embedding of Sm×DkmS_{m}\times D^{k-m} as in the previous paragraph, we construct a rep-tile in the homotopy type of SmSkS^{m}\vee S^{k}, itself capable of becoming part of a further rep-tilean wedge. By iterating this process, rep-tiles in the homotopy type of any finite wedge of spheres can be constructed.

2.5. Suspending Rep-Tiles

Let rπr_{\pi} be an order 2 rotation about some (n2)(n-2)-subspace in n\mathbb{R}^{n}. We note that if RR is any connected nn-dimensional stack of cubes such that two copies of RR, related by rπr_{\pi}, tile an nn-cube, then RR can be used to construct an (n+1)(n+1)-dimensional rep-tile in the homotopy type of the suspension of RR. We sketch this construction with a specific choice of coordinates below. For clarity, we assume that Rrπ(R)R\cup r_{\pi}(R) tile the cube [0,4]n[0,4]^{n}.

Let RR denote any nn-dimensional stack of unit cubes which has the property that RR and its image under under rπr_{\pi} tile [0,4]n[0,4]^{n}. (For instance, RR could be one of the rep-tiles in the homotopy type of a wedge of spheres that we previously constructed.) Because Rrπ(R)=[0,4]n,R\cup r_{\pi}(R)=[0,4]^{n}, we know that rπr_{\pi} takes cubes at height 4 (with respect to the stacking direction) to holes at height zero; and vice-versa. In particular, RR contains as many cubes at height 4 as it has unit-cube-sized holes at height 0. Therefore, we may suspend RR as by the following steps

  1. (1)

    embed R×[0,4]R\times[0,4] into n+1\mathbb{R}^{n+1};

  2. (2)

    cubify in the natural way, writing R×[0,4]R\times[0,4] as a union of unit (n+1)(n+1)-cubes of the form (n-cube in R)×[i,i+1](n\text{-cube in }R)\times[i,i+1];

  3. (3)

    move all height-4 cubes in R×[0,1]R\times[0,1] to fill all holes at height zero in that slice;

  4. (4)

    repeat the last step in R×[3,4]R\times[3,4].

Crucially, steps 3 and 4 constitute ball swaps (see Section 2.3). This guarantees that the resulting polycube is still a rep-tile. Moreover, since the slice R×[i,i+1]R\times[i,i+1] is a stack of cubes, filling all cubes that correspond to “height-0 holes” in R×[i,i+1]R\times[i,i+1] (that is, those holes in R×[i,i+1]R\times[i,i+1] which are height-0 holes in RR crossed with [i,i+1][i,i+1]) turns R×[i,i+1]R\times[i,i+1] into a ball. Therefore, as before, ball swapping in the first and last slices of R×[0,4]R\times[0,4] has the effect, up to homotopy, of contracting each of the ends of R×[0,4]R\times[0,4] to a point. This completes the suspension of RR. Figures 4 and 5 illustrate the suspensions of rep-tiles homeomorphic to S0×D2S^{0}\times D^{2} and S1×D2S^{1}\times D^{2}, respectively.

Let HH and HH^{\prime} denote any pair of columns in Figures 4 or 5 which trade a cube during the ball-swapping operations. Specifically, say HH^{\prime} is height 0 and the top cube of HH is moved to HH^{\prime} during the ball swapping. Now suppose next highest cube of HH (now at height 3) is also moved to column HH^{\prime}. This would constitute a ball swap, since the unit cube remains within its orbit under the rotation. Executing this additional swap between all such pairs has the effect that all columns of heights 3 and 1 become columns of height 2. The result would be the Sn×D2S^{n}\times D^{2} rep-tile constructed in Section 2.2. Put differently, rep-tiles homeomorphic to Sn×D2S^{n}\times D^{2} can also be obtained from the S0×D2S^{0}\times D^{2} rep-tile in Figure 2 inductively, via a sequence of suspensions and ball swaps. For details on this approach, see Section 2 of [Bla24].

Refer to caption
Figure 4. Ball swap in a 3-dimensional rep-tile. The swap effectuates the suspension of a polycube representation of S0×D2S^{0}\times D^{2} to obtain a polycube representation of S1×D2S^{1}\times D^{2}. Top: S0×D2×[0,4]S0×D3S^{0}\times D^{2}\times[0,4]\cong S^{0}\times D^{3}. Middle: A cube swap which ensures that the first and last slices become disks. Bottom: the union of the four layers is a rep-tile homeomorphic to S1×D2S^{1}\times D^{2}, the result of the suspension. A further cube swap between the same pairs of columns would result in the S1×D2S^{1}\times D^{2} rep-tile given in Section 2.2.
Refer to caption
Figure 5. The left and right columns represent the labeled footprints of 44-dimensional stacks of cubes. Taken together, the three columns depict the process of suspension from S1×D2S^{1}\times D^{2} to S2×D2S^{2}\times D^{2}.
Left column: four layers of S1×D2×[i,i+1]S^{1}\times D^{2}\times[i,i+1], combining to form S1×D2×[0,4]S^{1}\times D^{2}\times[0,4]. Middle column: ball swapping occurs in the first and fourth slices. Right column: bottom slice: D3×[0,1]D^{3}\times[0,1], second slice: S1×D2×[1,2]S^{1}\times D^{2}\times[1,2]; third slice: S1×D2×[2,3]S^{1}\times D^{2}\times[2,3]; fourth slice: D3×[3,4]D^{3}\times[3,4]. The union of the four slices is the suspended rep-tile.
Note that by a further ball swap we could replace all 3’s and all 1’s by 2’s. This would produce another rep-tile homeomorphic to S2×D2S^{2}\times D^{2}, namely the one described in Section 2.2.

2.6. Rep-tiles with arbitrary footprints

The following was observed by Richard Schwartz [Sch25] while perusing the first version of our article.

Proposition 2.1.

[Sch25] There is an nn-dimensional rep-tile in the homotopy type of any compact polycube in n1\mathbb{R}^{n-1}.

This result, together with the existence of cubifications for smooth codimension-0 submanifolds of n\mathbb{R}^{n} (see Section 3.5) can be used to prove a version of Corollary 1.2.1. Specifically, we see that it is possible to realize the homotopy type of any compact nn-dimensional CW complex as a (2n+2)(2n+2)-dimensional rep-tile RR, without appealing to Theorem 1.2. The present approach uses an extra dimension; but it is rather explicit (given a polycube footprint to start with) and has the advantage that just 2 copies of RR can tile the (2n+2)(2n+2)-cube.

Proof of Proposition 2.1.

We first observe that for any compact (n1)(n-1)-polycube PP there is a positive even integer kk such that PP is isotopic to an (n1)(n-1)-polycube PP^{\prime} in [0,k+2]n1[0,k+2]^{n-1} such that PP^{\prime} contains all unit cubes in [0,k+2]n1[0,k+2]^{n-1} whose smallest xn1x_{n-1}-coordinate is equal to 0. (To see this, begin by translating PP so that it is contained in [0,k]n1[0,k]^{n-1}. Then, apply the following sequence of isotopies: shift PP at least two units away from the xn1=0x_{n-1}=0 hyperplane in the positive xn1x_{n-1} direction; then grow a (cubical) finger out of PP until it touches xn1=0x_{n-1}=0; then add the cubes whose union is [0,k+2]n2×[0,1][0,k+2]^{n-2}\times[0,1] to PP.)

Next create an nn-dimensional stack of cubes SS whose footprint is a polycube in [0,k+2]n2×[(k+2),(k+2)][0,k+2]^{n-2}\times[-(k+2),(k+2)], namely the boundary connected sum of PP^{\prime} with [0,k+2]n2×[(k+2),0][0,k+2]^{n-2}\times[-(k+2),0]. We shall label the (n1)(n-1)-cubes contained in [0,k+2]n2×[(k+2),(k+2)][0,k+2]^{n-2}\times[-(k+2),(k+2)] to indicate the height of the corresponding column. In this manner, we will obtain the desired stack of nn-cubes SS in the homotopy type of PP. The (n1)(n-1)-cubes in PP^{\prime} are labeled k+2k+2. All (n1)(n-1)-cubes which are contained in [0,k+2]n2×[0,(k+2)][0,k+2]^{n-2}\times[0,(k+2)] but not in PP^{\prime} are labeled 0. Let rr denote reflection in n1\mathbb{R}^{n-1} about the plane xn1=0x_{n-1}=0. Cubes in [0,k+2]n2×[(k+2),0][0,k+2]^{n-2}\times[-(k+2),0] that are contained in r(P)r(P^{\prime}) are labeled k+2k+2. Remaining cubes are labeled 2k+42k+4. See Figure 6.

Let ρ\rho be rotation about the (n2)(n-2)-plane in n\mathbb{R}^{n} determined by xn1=0x_{n-1}=0 and xn=k+2x_{n}=k+2. Next we observe that the sum of the labels of each unit cube in [0,k+2]n2×[(k+2),(k+2)][0,k+2]^{n-2}\times[-(k+2),(k+2)] and its reflection rr about xn1=0x_{n-1}=0 sum to 2k+42k+4. It follows that the stack of cubes SS determined by this labeling, together with ρ(S)\rho(S), tile [0,k+2]n2×[(k+2),k+2]×[0,2k+4][0,k+2]^{n-2}\times[-(k+2),k+2]\times[0,2k+4]. Then, after rescaling, two isometric copies of SS tile an nn-cube. This produces a rep-tile in the homotopy type of the original footprint, PP, as desired. ∎

Refer to caption
Figure 6. PP denotes a compact polycube in n1\mathbb{R}^{n-1}, embedded as a proper subset of a (n1)(n-1)-dimensional cube (pictured as the top 6×66\times 6 square) in the hyperplane xn=0x_{n}=0 in n\mathbb{R}^{n}. The figure is a schematic for constructing an nn-dimensional rep-tile in the homotopy type of PP. Specifically, the rep-tile’s footprint is the pictured stack of cubes, which is isotopic to PP. Each unlabeled box has k+2k+2 cubes stacked on top of it; the heights of other stacks are as written. Remark that the bottom half of the picture is a stack of cubes homeomorphic to BnB^{n}; its footprint is an (n1)(n-1)-dimensional cube. This ball is added to PP to ensure symmetry.

3. All is rep-tile

We will denote the standard integer lattice in n\mathbb{R}^{n}, consisting of all points in n\mathbb{R}^{n} with integer coordinates, by 𝒵n\mathcal{Z}^{n}. This lattice induces a cell structure 𝒞(𝒵n)\mathcal{C}(\mathcal{Z}^{n}) on n\mathbb{R}^{n}, whose kk-cells are the kk-facets of unit cubes with vertices in 𝒵n\mathcal{Z}^{n}.

We will also work with subdivisions of this lattice, and refer to the closed nn-cells in any such decomposition as atomic cubes. The size of an atomic cube will depend on the subdivision used. Precisely, suppose λ>0\lambda>0 and let fλ:nnf_{\lambda}:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n} denote the scaling function given by f(x)=λxf(x)=\lambda x. Let 𝒵λn=f(𝒵n)\mathcal{Z}_{\lambda}^{n}=f(\mathcal{Z}^{n}), and let 𝒞(𝒵λn)\mathcal{C}(\mathcal{Z}_{\lambda}^{n}) denote the corresponding cell structure.

Definition 3.1.

An nn-dimensional polycube is a submanifold of n\mathbb{R}^{n} that is isometric to a finite union of atomic cubes in 𝒞(𝒵λn)\mathcal{C}(\mathcal{Z}^{n}_{\lambda}) for some λ\lambda\in\mathbb{R}.

Definition 3.2.

A compact nn-manifold TT is said to kk-tile a subset AnA\subseteq\mathbb{R}^{n} if A=i=1kTiA=\cup_{i=1}^{k}T_{i} such that TiT_{i} is isometric to TjT_{j} for all ii and jj, and int(Ti)int(Tj)=int(T_{i})\cap int(T_{j})=\emptyset for all iji\neq j.

Lemma 3.1.

Let RR be an nn-dimensional polycube that tiles a cube CC. Then, RR is a rep-tile.

Proof.

By identifying each atomic cube in the polycube decomposition of RR with CC, we can tile each cube in RR with a finite number of pairwise isometric manifolds, each of which is similar to RR. We have thus tiled RR by rescaled copies of RR. ∎

In particular, a polycube that tiles the cube must have connected boundary, which follows from the following Lemma.

Lemma 3.2.

Let XnX^{n} be a manifold which is homeomorphic to an nn-dimensional rep-tile. Then (X)\partial(X) is non-empty and connected.

Proof.

Since XnX^{n} is a homeomorphic to a rep-tile, we have that XnX^{n} embeds in n\mathbb{R}^{n}. Hence, (X)\partial(X)\neq\emptyset. The proof that (X)\partial(X) is connected when n=3n=3 is given in [Bla21, Theorem 4.2] and works without modification in all dimensions. ∎

We recall our main theorem below.

Theorem 1.2.

Let RnR\subset\mathbb{R}^{n} be a compact smooth nn-manifold with connected boundary. Then, RR is topologically isotopic to a rep-tile.

Our main theorem is a consequence of the following.

Theorem 3.3.

Let RnR\subset\mathbb{R}^{n} be a compact smooth nn-manifold with connected boundary. Then, RR is topologically isotopic to a nn-dimensional polycube RR^{*} which 2n2^{n}-tiles a cube.

A key step in the proof that any RnR\subseteq\mathbb{R}^{n} satisfying the hypotheses of Theorem 1.2 is isotopic to a rep-tile is to decompose Cn\R¯\overline{C^{n}\backslash R}, the closure of the complement of RR in an nn-cube, into a union of closed nn-balls with non-overlapping interiors. Given a manifold XnX^{n}, the smallest number of nn-balls in such a decomposition of XX is called the ball number of XX, denoted b(X)b(X). Upper bounds on the ball number of a manifold in terms of its algebraic topology have been found by Zeeman [Zee63] and others [Luf69, Kob76, Sin79]. We rely on the following.

Theorem 3.4.

[2.11 of [Kob76]] Let MnM^{n} be a connected compact PL nn-manifold with non-empty boundary. Then b(M)nb(M)\leq n.

Refer to caption
Figure 7. The green disk RR^{*}, constructed from the top right square via ball-swapping, tiles the square [1,1]×[1,1][-1,1]\times[-1,1].

3.1. Overview of the proof of Theorem 1.2

The main ingredient is Theorem  3.3, which we prove using a strategy we refer to as a ball swap. To start, RR is smoothly embedded in Cn=[0,1]nC^{n}=[0,1]^{n} so that RCn=R\cap\partial C^{n}=\emptyset. In turn, the unit cube CnC^{n} sits inside the cube =[1,1]n\boxplus=[-1,1]^{n}. Since RR is disjoint from Cn\partial C^{n} and has a single boundary component, CnRC^{n}\setminus R is connected. By Theorem  3.4, we may decompose CnR¯\overline{C^{n}\setminus R} into nn nn-dimensional balls B1,,BnB_{1},\dots,B_{n}.111If b(CnR)<nb(C^{n}\setminus R)<n, one could use fewer balls here and tile the cube with fewer copies of RR, but we use nn balls for simplicity in the proof of the main theorem. After a homotopy of CnC^{n} which restricts to an isotopy on each piece of the decomposition {R,B1,,Bn}\{R,B_{1},\dots,B_{n}\} of CnC^{n}, we ensure that the pieces of this decomposition intersect an (n1)(n-1)-disk on Cn\partial{C}^{n} as shown in Figure 8, in what we call a taloned pattern. The defining features of taloned patterns include: there is an (n1)(n-1)-disk on Cn\partial C^{n} such that RR and each of B1,BnB_{1},\dots B_{n} intersect that disk in an (n1)(n-1)-ball and intersect the boundary of the disk in an (n2)(n-2)-ball; the (n1)(n-1)-balls BiB_{i} are disjoint inside this disk; and RR is adjacent to each ball BiB_{i} in this disk. (See Section 3.2 for the formal definition.) The homotopy used to create the taloned pattern is achieved in Lemmas 3.5 and 3.6 below. We then isotope CnC^{n} so that the (n1)(n-1)-disk which constitutes the taloned pattern of Figure 8 is identified with the union of faces of Cn=[0,1]nC^{n}=[0,1]^{n} whose interiors lie in the interior of =[1,1]n\boxplus=[-1,1]^{n}, with certain additional restrictions. These restrictions guarantee that certain rotated copies of the BiB_{i} contained in cubes adjacent to [0,1]n[0,1]^{n} in =[1,1]n\boxplus=[-1,1]^{n} are disjoint, allowing us to form the boundary connected sum of RR with these balls without changing the isotopy class of RR. Indeed, we give a family of rotations rkr_{k}, 1kn/21\leq k\leq\lfloor{n/2}\rfloor, together with one additional rotation ff if nn is odd, such that the orbit of CnC^{n} under these rotations tiles \boxplus. By taking the boundary sum of RCnR\subset C^{n} with the image of each BiB_{i} under an appropriate choice of rotation above, we obtain the desired manifold RR^{*}. By construction, RR^{*} is isotopic to RR and, moreover, the orbit of RR^{*} under the above set of rotations gives a tiling of \boxplus. A 2-dimensional tile RR^{*} created via ball swapping is shown in Figure 7. A sketch of RR^{*} in dimension n=3n=3 is shown in Figure 14. Finally, we show that this construction can be “cubified”, so that RR^{*} is a polycube tiling \boxplus, completing the proof of Theorem 3.3. Once this is established, Theorem 1.2 follows from Lemma 3.1.

3.2. Taloned Patterns

We define the desired boundary pattern described above. A kk-claw is a tree which consists of one central vertex vv and kk leaves, each connected to vv by a single edge. See Figure 8.

Refer to caption
Figure 8. Taloned boundary pattern corresponding to a kk-claw in Cn\partial C^{n}.

Our goal is to construct a boundary pattern on CnC^{n} such that there exists an embedded disk Dn1CnD^{n-1}\subset\partial C^{n} with the following properties:

  • Dn1BiD^{n-1}\cap B_{i} is a single (n1)(n-1)-disk, for all 1ik1\leq i\leq k;

  • (Dn1Bi)Dn1(D^{n-1}\cap B_{i})\cap\partial D^{n-1} is an (n2)(n-2)-disk, for all 1ik1\leq i\leq k;

  • Dn1(i=1kBiDn1)RD^{n-1}\setminus(\cup_{i=1}^{k}B_{i}\cap D^{n-1})\subset R.

  • BiBjDn1=B_{i}\cap B_{j}\cap D^{n-1}=\emptyset for iji\neq j.

We regard the boundary pattern as the regular neighborhood of a kk-claw, with the following decomposition: RR contains a neighborhood of the central vertex; and each BiB_{i} containing a neighborhood of a leaf. See Figure 8. We call this a taloned pattern of intersections.

We begin by proving Lemma 3.5, which ensures that, in the interior of CnC^{n}, the union of the boundaries of the pieces {R,B1,,Bn}\{R,B_{1},\dots,B_{n}\} in the interior of our decomposition of CnC^{n} can be assumed to be connected.

Lemma 3.5.

Let RR be a compact nn-manifold with a single boundary component embedded in the nn-cube CnC^{n} such that Cn=RB1BkC^{n}=R\cup B_{1}\cup\dots\cup B_{k}, where each BiB_{i} is an nn-ball, and such that the interiors of RR and the BiB_{i} are pairwise disjoint. Then after a homotopy of CnC^{n} which restricts to isotopies on the interiors of RR and the BiB_{i}, W=(RB1Bk)Cn¯W=\overline{\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n}} is a connected (n1)(n-1)-complex.

Proof.

Let ={R,B1,,Bk}\mathcal{B}=\{R,B_{1},\dots,B_{k}\}. We partition \mathcal{B} into layers i\mathcal{L}_{i} as follows (see Figure 9). Define the first layer as 1={L|LCn}\mathcal{L}_{1}=\{L\in\mathcal{B}\leavevmode\nobreak\ |\leavevmode\nobreak\ \partial L\cap\partial C^{n}\neq\emptyset\}. We will use the notation 1:=L1L\partial\mathcal{L}_{1}:=\bigcup_{L\in\mathcal{L}_{1}}\partial L. Next choose a minimal collection of disjoint, embedded paths α1,αl\alpha_{1},\dots\alpha_{l} on Cn\partial C^{n} such that

  • (1Cn¯)α1αl\left(\overline{\partial\mathcal{L}_{1}\setminus\partial C^{n}}\right)\cup\alpha_{1}\cup\dots\cup\alpha_{l} is connected,

  • the interior of αi\alpha_{i} is contained in a single element B(αi)B(\alpha_{i}) of \mathcal{B}; and

  • no αi\alpha_{i} has both endpoints on the same connected component of 1Cn¯\overline{\partial\mathcal{L}_{1}\setminus\partial C^{n}}.

Refer to caption
Figure 9. (Left) A partition of the decomposition {R,B1,,Bk}\{R,B_{1},\dots,B_{k}\} of CnC^{n} into layers, with the number in reach region indicating its level; (Middle) A choice of paths αi\alpha_{i} on Cn\partial C^{n} such that after performing finger moves along the αi\alpha_{i}, W=(RB1Bk)Cn¯W=\overline{\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n}} is connected (Right).
Refer to caption
Figure 10. Performing finger moves to ensure W=(RB1Bk)Cn¯W=\overline{\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n}} is a connected (n1)(n-1)-complex.

Note that any given element of the decomposition \mathcal{B} may contain the interior of more than one of the paths αi\alpha_{i}, i.e., it is possible to have B(αi)=B(αj)B(\alpha_{i})=B(\alpha_{j}) for iji\neq j. For each AA\in\mathcal{B}, we let P(A)P(A) denote the set of all ii such that A=B(αi)A=B(\alpha_{i}).

Since the αi\alpha_{i} are disjoint, for each 1il1\leq i\leq l, we can choose a disjoint regular neighborhood RiR_{i} in B(αi)B(\alpha_{i}) of αi\alpha_{i} such that RiR_{i} intersects the boundary of exactly two other elements B(αi)0B(\alpha_{i})_{0} and B(αi)1B(\alpha_{i})_{1} of \mathcal{B}, one at each of the endpoints αi(0)\alpha_{i}(0) and αi(1)\alpha_{i}(1), respectively. For each AA\in\mathcal{B}, let P0(A)P_{0}(A) denote the set of all ii such that A=B(αi)0A=B(\alpha_{i})_{0}. Next, modify the decomposition \mathcal{B} of CnC^{n} as follows (see Figure 10).

  • For each AA\in\mathcal{B}, delete all the RiR_{i} whose interiors intersect AA, replacing each AA\in\mathcal{B} by

    A=A(iP(A)Ri)¯A^{\prime}=\overline{A\setminus\left(\bigcup_{i\in P(A)}R_{i}\right)}
  • Then, attach each RiR_{i} to B(αi)0B(\alpha_{i})_{0}, replacing each AA^{\prime} (which may coincide with AA, if AA did not intersect the interior of any RiR_{i}) by

    A′′=A(iP0(A)Ri)A^{\prime\prime}=A^{\prime}\cup\left(\bigcup_{i\in P_{0}(A)}R_{i}\right)

This process can be achieved by a homotopy of CnC^{n} which restricts to isotopies on the interiors of the elements of \mathcal{B}. We imagine elements of \mathcal{B} as growing fingers along the αi\alpha_{i}. From now on, we will simply call these finger moves and will not describe them explicitly.

After performing finger moves on the elements of 1\mathcal{L}_{1} along the αi\alpha_{i}, we can assume 1Cn\partial\mathcal{L}_{1}\setminus\partial C^{n} is connected. Then inductively define i={Lj=1i1j|Li1}\mathcal{L}_{i}=\{L\in\mathcal{B}\setminus\bigcup_{j=1}^{i-1}\mathcal{L}_{j}|L\cap\partial\mathcal{L}_{i-1}\neq\emptyset\}, where i\partial\mathcal{L}_{i} is defined analogously to 1\partial\mathcal{L}_{1}. Since L\partial L is connected for each L2L\in\mathcal{L}_{2} and meets 1Cn\partial\mathcal{L}_{1}\setminus\partial C^{n}, we have that 2(1Cn)\partial{\mathcal{L}}_{2}\cup(\partial{\mathcal{L}}_{1}\setminus\partial C^{n}) is connected. Continue inductively for each 3im3\leq i\leq m, where mm is the number of layers. By construction, L\partial L is connected for each LiL\in\mathcal{L}_{i} and intersects j=1i1jCn\bigcup_{j=1}^{i-1}\partial\mathcal{L}_{j}\setminus\partial C^{n} non-trivially. Therefore i=1miCn=(RB1Bk)Cn\bigcup_{i=1}^{m}\partial\mathcal{L}_{i}\setminus\partial C^{n}=\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n} is connected, so W=(RB1Bk)Cn¯W=\overline{\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n}} is connected as well. ∎

Lemma 3.6.

Let RR be a compact nn-manifold with connected boundary embedded in the nn-cube CnC^{n} such that Cn=RB1BkC^{n}=R\cup B_{1}\cup\dots\cup B_{k}, with knk\leq n, where each BiB_{i} is a nn-ball and such that the interiors of RR and the BiB_{i} are pairwise disjoint. After applying a self-homotopy of CnC^{n} that restricts to an isotopy on the interior of each component in the above decomposition, we can find an kk-claw embedded in Cn\partial C^{n} such that its regular neighborhood in Cn\partial C^{n} is a taloned pattern.

Proof.

By Lemma 3.5, we can assume W=(RB1Bk)Cn¯W=\overline{\left(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{k}\right)\setminus\partial C^{n}} is connected, so we can perform a finger move on RR along a path in WW to ensure that RR meets Cn\partial C^{n}. Since RR has a single boundary component and (RCn)Cn(\partial R\cap\partial C^{n})\subsetneq\partial C^{n}, we can assume there exists a point pp on the interior of an (n1)(n-1) face of CnC^{n} that lies on RBi\partial R\cap\partial B_{i} for some ii. Relabeling the BiB_{i} if necessary, we assume i=1i=1.

Without loss of generality, assume pp lies on the face F1F_{1} defined by {x1=0}Cn\{x_{1}=0\}\cap C^{n}, and let p=(0,p2,,pn)p=(0,p_{2},\dots,p_{n}). After an isotopy of CnC^{n}, we can assume some ϵ\epsilon-ball Bϵ(p)B_{\epsilon}(p) satisfies the following:

RBϵ(p)={(x1,,xn)CnBϵ(p)|x2p2}R\cap B_{\epsilon}(p)=\{(x_{1},\dots,x_{n})\in C^{n}\cap B_{\epsilon}(p)|x_{2}\geq p_{2}\}
B1Bϵ(p)={(x1,,xn)CnBϵ(p)|x2p2}.B_{1}\cap B_{\epsilon}(p)=\{(x_{1},\dots,x_{n})\in C^{n}\cap B_{\epsilon}(p)|x_{2}\leq p_{2}\}.

We can further assume that RB1Bϵ(p)=WBϵ(p)R\cap B_{1}\cap B_{\epsilon}(p)=W\cap B_{\epsilon}(p) and RB1Bϵ(p)={(x1,,xn)CnBϵ(p)|x2=p2}R\cap B_{1}\cap B_{\epsilon}(p)=\{(x_{1},\dots,x_{n})\in C^{n}\cap B_{\epsilon}(p)|x_{2}=p_{2}\}.

Choose distinct points q2,,qkq_{2},\dots,q_{k} on the (n2)(n-2) disk RB1Bϵ(p)CnR\cap B_{1}\cap B_{\epsilon}(p)\cap\partial C^{n}, as in Figure 11. We claim that one can choose disjoint paths δiW\delta_{i}\subset W from a point rir_{i} in Biint(Cn)B_{i}\cap int(C^{n}) to the point qiq_{i} for each 2ik2\leq i\leq k.

To produce the δi\delta_{i}, we again apply Lemma  3.5. In dimensions 4 and higher, we can achieve disjointness of the δi\delta_{i} by a perturbation. In dimension 3, we perform an oriented resolution at each point of intersection of the δi\delta_{i}’s which can not be removed by perturbation inside WW. In dimension 2, there is only one such path, δ2\delta_{2}, since 2ik=2.2\geq i\geq k=2.

Once the paths are disjoint, we perform a finger move which pushes a neighborhood of rir_{i} in BiB_{i} along δi\delta_{i} to a neighborhood of qiq_{i} in Bϵ(p)B_{\epsilon}(p). As a result, the balls BiB_{i} intersect CnBϵ(p)\partial C^{n}\cap B_{\epsilon}(p) in the boundary pattern shown in Figure 11 (middle). We then choose a claw as shown in Figure 11 (bottom). The regular neighborhood of this claw in Cn\partial C^{n} is isotopic to a taloned pattern (Figure 8), as desired. ∎

Refer to caption
Figure 11. Three views of CnBϵ(p)\partial C^{n}\cap B_{\epsilon}(p), showing the stages of obtaining the claw. First, perform finger moves so that each ball BiB_{i}, with i2i\geq 2, meets Cn\partial C^{n} along the (n2)(n-2)-disk of intersection of RR and B1B_{1} inside Bϵ(p)B_{\epsilon}(p). One can then choose a kk-claw (bottom) which has a small regular neighborhood in Cn\partial C^{n} giving a taloned pattern of intersection.

3.3. Proof of main theorem.

We begin by setting up the necessary notation. For each i=1,,ni=1,\dots,n, let FiF_{i} be the (n1)(n-1)-dimensional face of the nn-cube CnC^{n} contained in the hyperplane xi=0x_{i}=0. For the moment, we will assume that nn is even. The case of nn odd requires an extra step, which we leave until the end of the proof.

Let ri:nnr_{i}:\mathbb{R}^{n}\to\mathbb{R}^{n} be the rotation by π2\frac{\pi}{2} about the (n2)(n-2)-plane x2i1=x2i=0x_{2i-1}=x_{2i}=0 that carries the x2i1x_{2i-1}-axis to the x2ix_{2i}-axis. Note that each rir_{i} has order four and that these rotations commute, generating a group isomorphic to (4)n/2(\mathbb{Z}_{4})^{n/2}. Given a vector 𝐲=(y1,,yn/2)(4)n/2\mathbf{y}=(y_{1},\dots,y_{n/2})\in(\mathbb{Z}_{4})^{n/2}, we define the rotation 𝐫𝐲{\bf r_{y}} as follows:

𝐫𝐲=rn/2yn/2r1y1.{\bf r_{y}}=r_{n/2}^{y_{n/2}}\circ\dots\circ r_{1}^{y_{1}}.

We set C𝐲:=𝐫𝐲(Cn).C_{\mathbf{y}}:=\mathbf{r_{y}}(C^{n}).

We claim that the orbit of a unit sub-cube under this group action is the entire nn-dimensional cube :=[1,1]n\boxplus:=[-1,1]^{n}. In other words, \boxplus is tiled by the 2n2^{n} distinct unit cubes {𝐫𝐲(Cn)|𝐲(4)n/2}.\{\mathbf{r_{y}}(C^{n})|\mathbf{y}\in(\mathbb{Z}_{4})^{n/2}\}.

To see this, first decompose \boxplus into 2n2^{n} unit sub-cubes of the form J1××JnJ_{1}\times\dots\times J_{n}, where each JiJ_{i} is either [1,0][-1,0] or [0,1][0,1]. Fixing kk, for each choice of J2k1J_{2k-1} and J2kJ_{2k} from the set {[1,0],[0,1]}\{[-1,0],[0,1]\}, the product J2k1×J2kJ_{2k-1}\times J_{2k} is a unit square in the x2k1x2kx_{2k-1}x_{2k}- plane, which we denote by k2.\mathbb{R}^{2}_{k}. Let Pk:=Cnk2P_{k}:=C^{n}\cap\mathbb{R}^{2}_{k}, i.e PkP_{k} is the unit square in the first quadrant of k2\mathbb{R}^{2}_{k}. Then J2k1×J2k=ryk(Pk)J_{2k-1}\times J_{2k}=r^{y_{k}}(P_{k}) for some yk{0,1,2,3}y_{k}\in\{0,1,2,3\}. Hence, each of the 2n2^{n} unit cubes above can be expressed as

J1××Jn=r1y1(P1)××rn/2yn/2(Pn/2)=C𝐲J_{1}\times\dots\times J_{n}=r_{1}^{y_{1}}(P_{1})\times\dots\times r_{n/2}^{y_{n/2}}(P_{n/2})=C_{\mathbf{y}}

for some 𝐲=(y1,,yn/2)(4)n/2\mathbf{y}=(y_{1},\dots,y_{n/2})\in(\mathbb{Z}_{4})^{n/2}. Moreover, for each J1××JnJ_{1}\times\dots\times J_{n}, the 𝐲\mathbf{y} such that J1××Jn=C𝐲J_{1}\times\dots\times J_{n}=C_{\mathbf{y}} is unique. To see this, note that each J1××JnJ_{1}\times\dots\times J_{n} has exactly one corner with all nonzero coordinates (and therefore with all coordinates ±1\pm 1). On the other hand, the cube C𝐲C_{\mathbf{y}} also has exactly one corner (c1,,cn)(c_{1},\dots,c_{n}) with all ci=±1c_{i}=\pm 1 (namely, the image of the point (1,1,1,,1)Cn(1,1,1,\dots,1)\in C^{n}), and its coordinates satisfy the formula yk=(c2k1)12(c2k1c2k1).y_{k}=-(c_{2k}-1)-\frac{1}{2}(c_{2k-1}c_{2k}-1). In other words, the coordinates (c1,cn)(c_{1},\dots c_{n}) uniquely determine each component yk,y_{k}, and therefore 𝐲\mathbf{y} itself.

Observe that the cube rk(Cn)r_{k}(C^{n}) intersects CnC^{n} along its face F2k1F_{2k-1}, and the cube rk1(Cn)r^{-1}_{k}(C^{n}) intersects CnC^{n} along its face F2kF_{2k}. Thus, each rotation rkr_{k} gives a pairing of the faces of CnC^{n}. We use this pairing to carry out a ball swap as previously described. This will allow us to build the rep-tile RR^{*}.

3.4. Realizing the taloned pattern on Cn\partial C^{n}

We will now describe a homotopy of CnC^{n} which restricts to an isotopy on the interiors of RR and the balls B1,,BnB_{1},\dots,B_{n}. Our goal is to use Lemma 3.6 to position RR and B1,,BnB_{1},\dots,B_{n} so that their intersections with the boundary of CnC^{n} satisfy:

  1. (1)

    For each 1in1\leq i\leq n, the only ball meeting the face FiF_{i} is BiB_{i} (and thus Fi(BiFi)RF_{i}\setminus(B_{i}\cap F_{i})\subset R),

  2. (2)

    rk(B2kF2k)F2k1r_{k}(B_{2k}\cap F_{2k})\subset F_{2k-1} is disjoint from B2k1B_{2k-1}, and

  3. (3)

    rk1(B2k1F2k1)F2kr_{k}^{-1}(B_{2k-1}\cap F_{2k-1})\subset F_{2k} is disjoint from B2kB_{2k}.

In what follows, we refer the reader to a schematic in Figure 12. Figure 13 illustrates this configuration in dimension 4.

For each k=1,,n2k=1,\dots,\frac{n}{2}, let φ2k1\varphi_{2k-1} be the (n2)(n-2)-facet in CC equal to the intersection of CC with the (n2)(n-2)-plane given by setting x2k1=0x_{2k-1}=0 and x2k=1x_{2k}=1. Likewise, let φ2k\varphi_{2k} be the (n2)(n-2)-facet in CC equal to the intersection of CC with the (n2)(n-2)-plane given by setting x2k1=1x_{2k-1}=1 and x2k=0x_{2k}=0. Note that this pair of facets are exactly those that are simultaneously parallel to the intersection F2k1F2kF_{2k-1}\cap F_{2k} and contained in F2k1F2kF_{2k-1}\cup F_{2k}.

Refer to caption
Figure 12. Intersections of B2k1B_{2k-1} and B2kB_{2k} with faces F2k1F_{2k-1} and F2kF_{2k} of Cn\partial C^{n}, and their images under the rotations rk1r_{k}^{-1} and rkr_{k}, respectively.
Refer to caption
Figure 13. Intersections of B1B_{1} and B2B_{2} with faces F1F_{1} and F2F_{2} of C4\partial C^{4}, and their images under the rotations r11r_{1}^{-1} and r1r_{1}, respectively.

Now, we fix points α2k1φ2k1\alpha_{2k-1}\in\varphi_{2k-1} and α2kφ2k\alpha_{2k}\in\varphi_{2k} by setting

α2k1=(14,,14,0,1,14,,14)\alpha_{2k-1}=\left(\frac{1}{4},\dots,\frac{1}{4},0,1,\frac{1}{4},\dots,\frac{1}{4}\right)

and

α2k=(34,,34,1,0,34,,34),\alpha_{2k}=\left(\frac{3}{4},\dots,\frac{3}{4},1,0,\frac{3}{4},\dots,\frac{3}{4}\right),

where the 0 and 11 entries are taken to be in the (2k1)st(2k-1)^{st} and 2kth2k^{th} coordinates.

Let B1,B2,,BnB_{1},B_{2},\dots,B_{n} be the nn-balls whose existence is guaranteed by Theorem 3.4. By Lemma 3.6, after an isotopy of RR and the BiB_{i}, there is an nn-claw embedded in Cn\partial C^{n} such that its regular neighborhood in Cn\partial C^{n} is a taloned pattern as shown in Figure 8. Moreover, after an isotopy of CnC^{n} supported near its boundary, we can assume that the taloned pattern is mapped homeomorphically to i=1nFi\bigcup_{i=1}^{n}F_{i} such that the intersection FiBi:=NiF_{i}\cap B_{i}:=N_{i} is a closed regular neighborhood of radius 1/81/8 of the point αi\alpha_{i} in FiF_{i}, and also that if iji\not=j, then FiBj=F_{i}\cap B_{j}=\emptyset. We do not assume any restrictions on the intersections of RR and the BiB_{i} with the remaining faces xi=1x_{i}=1 of CnC^{n}.

Note that this set-up has several convenient consequences. First, the union i=1nFi\bigcup_{i=1}^{n}F_{i} intersects R\partial R in a single (n1)(n-1)-ball, since RR meets the taloned pattern in a single (n1)(n-1)-ball. Furthermore, the center and radius of N2kN_{2k} were chosen to guarantee that the ball rk(N2k)F2k1r_{k}(N_{2k})\subset F_{2k-1} is disjoint from the neighborhood N2k1N_{2k-1}, and therefore contained in F2k1\N2k1=RF2k1F_{2k-1}\backslash N_{2k-1}=R\cap F_{2k-1}. Similarly, the ball rk1(N2k1)r_{k}^{-1}(N_{2k-1}) is contained in F2k\N2k=RF2kF_{2k}\backslash N_{2k}=R\cap F_{2k}.

3.5. Cubification of the decomposition

Recall that for any positive integer mm, by 𝒞(𝒵1mn)\mathcal{C}(\mathcal{Z}_{\frac{1}{m}}^{n}) we denote the lattice in n\mathbb{R}^{n} whose unit cubes have side length 1m\frac{1}{m}.

Let W=(RB1Bn)Cn¯W=\overline{(\partial R\cup\partial B_{1}\cup\dots\cup\partial B_{n})\setminus\partial C^{n}}. Since RR and each BiB_{i} can be assumed piecewise-smooth, WW has a closed regular neighborhood N(W)N(W). Being a codimension-0 compact submanifold of n,\mathbb{R}^{n}, it is isotopic to a polycube, also denoted N(W)N(W), in a sufficiently fine lattice 𝒞(𝒵1mn)\mathcal{C}(\mathcal{Z}_{\frac{1}{m}}^{n}). (In the course of cubification, we shall increase mm as needed without further comment.) We also assume that all cubes in N(W)N(W) which intersect RR form a regular neighborhood of R\partial R. Similarly for each BiB_{i}; and for each double intersection, BiBj\partial B_{i}\cap\partial B_{j} or RBi\partial R\cap\partial B_{i}; and each triple intersection, etc.

The closure R\N(W)¯\overline{R\backslash N(W)} is then also a polycube; similarly for each Bi\N(W)¯\overline{B_{i}\backslash N(W)}. To complete the cubification of the ensemble {R,B1,,Bn},\{R,B_{1},\dots,B_{n}\}, we assign cubes in N(W)N(W) back to the constituent pieces in an iterative fashion. Specifically, all cubes in N(W)N(W) which intersect RR are assigned to RR, and their union is denoted RcuR^{cu}; of the remaining cubes, all that intersect B1B_{1} are assigned to B1B_{1}, and the resulting polycube is denoted B1cuB_{1}^{cu}; and so on. By the above assumptions, each of the pieces {R,B1,,Bn}\{R,B_{1},\dots,B_{n}\} is isotopic to the corresponding polycube since we are only adding or removing small cubes intersecting the boundary. In addition, the union of the interiors of {R,B1,,Bn}\{R,B_{1},\dots,B_{n}\} is isotopic to the union of the interiors of {Rcu,B1cu,,Bncu}\{R^{cu},B_{1}^{cu},\dots,B_{n}^{cu}\}.

Furthermore, by selecting a sufficiently fine lattice, we can ensure that the isotopies performed, taking each of {R,B1,,Bn},\{R,B_{1},\dots,B_{n}\}, to a polycube, are arbitrarily small. Thus, they preserve properties (1), (2) and (3) from Section 3.4.

Recycling notation, we will from now on refer to RcuR^{cu}, B1cuB^{cu}_{1}, \dots, BncuB^{cu}_{n} as RR, B1B_{1}, \dots, BnB_{n} respectively.

3.6. Construction of the rep-tile.

Finally we construct our rep-tile RR^{*}\subset\boxplus:

R=R(k=1n/2rk1(B2k1)rk(B2k))R^{*}=R\cup\left(\bigcup_{k=1}^{n/2}r_{k}^{-1}(B_{2k-1})\cup r_{k}(B_{2k})\right)

We claim that (1) RR^{*} is isotopic to RR, and (2) 2n2^{n} isometric copies of RR^{*} tile the cube \boxplus. A schematic of RR^{*} in dimension n=3n=3 is shown in Figure  14 (for intuition in the case of nn even, simply ignore B3B_{3} and its rotated copy in the figure).

Refer to caption
Figure 14. The manifold RR^{*}, shown in blue, constructed via ball-swapping, which tiles [1,1]3[-1,1]^{3}.

Proof of (1). The images of the cube CnC^{n} under the rotations r1,r11,,rn/2,rn/21r_{1},r_{1}^{-1},\dots,r_{n/2},r_{n/2}^{-1} give a family of nn distinct unit cubes in \boxplus, each of which shares a unique face with CnC^{n}. More specifically, the cube rk(Cn)r_{k}(C^{n}) intersects CnC^{n} along its face F2k1F_{2k-1}, and the cube rk1(Cn)r_{k}^{-1}(C^{n}) intersects CC along its face F2k.F_{2k}. Refer to Figure 12.

It follows that the intersections of each ball rk(B2k)r_{k}(B_{2k}) and rk1(B2k1)r_{k}^{-1}(B_{2k-1}) with the cube CnC^{n} are disjoint (n1)(n-1)-balls contained in RCn\partial R\cap C^{n}. (Recall that the center and radius of the NiN_{i} were chosen carefully so that this is the case.) Therefore, RR^{*} is a boundary connected sum of RCnR\subset C^{n} with a collection of nn-balls, one in each neighboring cube. An isotopy therefore brings RR^{*} to the initial embedding of RR, as desired. This concludes the proof of (1).

Proof of (2). Let R𝐲:=𝐫𝐲(R)R_{\bf y}:={\bf r_{y}}(R), Bi𝐲:=𝐫𝐲(Bi)B_{i\,{\bf y}}:={\bf r_{y}}(B_{i}), and R𝐲:=𝐫𝐲(R)R^{*}_{\bf y}:={\bf r_{y}}(R^{*}). Note that, since Cn=R(iBi)C^{n}=R\cup(\cup_{i}B_{i}), we have that R𝐲C𝐲R_{\bf y}\subseteq C_{\bf y} and Bi𝐲C𝐲B_{i\,{\bf y}}\subseteq C_{\bf y}. In addition, the first equality on the next line clearly implies the second:

=𝐲(4)n/2C𝐲=(𝐲(4)n/2R𝐲)(𝐲(4)n/2(iBi𝐲)).\boxplus=\bigcup_{\mathbf{y}\in(\mathbb{Z}_{4})^{n/2}}C_{\bf y}=\left(\bigcup_{\mathbf{y}\in(\mathbb{Z}_{4})^{n/2}}R_{\bf y}\right)\cup\left(\bigcup_{\mathbf{y}\in(\mathbb{Z}_{4})^{n/2}}(\cup_{i}B_{i\,{\bf y}})\right).

We now show that

=𝐲(4)n/2R𝐲.\boxplus=\bigcup_{\mathbf{y}\in(\mathbb{Z}_{4})^{n/2}}R^{*}_{\mathbf{y}}.

Since \boxplus decomposes into the cubes C𝐲C_{\mathbf{y}}, it is sufficient to show that every point pC𝐲p\in C_{\mathbf{y}} is contained in R𝐯R^{*}_{\bf v} for some 𝐯(4)n/2{\mathbf{v}\in(\mathbb{Z}_{4})^{n/2}}. This is a consequence of the fact that RR^{*} is the union of RR and one ball from the orbit of BiB_{i} for each ii. However, this fact may not be self-evident, so we provide an explicit proof.

Consider a point pC𝐲p\in C_{\mathbf{y}}. If pp is in the orbit of RR, then pR𝐲C𝐲p\in R_{\mathbf{y}}\subset C_{\mathbf{y}}, so pR𝐲R𝐲p\in R_{\bf y}\subseteq R^{*}_{\mathbf{y}}. Now, suppose pBi𝐲p\in B_{i\,{\bf y}} for some i=1,ni=1,\dots n. To find which rotation of RR^{*} contains pp, consider the isometric ball BiCnB_{i}\subset C^{n}. There are two cases: if i=2k1i=2k-1, then Birk(R)B_{i}\subset r_{k}(R^{*}), and if i=2ki=2k, Birk1(R)B_{i}\subset r_{k}^{-1}(R^{*}).

Let 𝐯(4)n/2\mathbf{v}\in(\mathbb{Z}_{4})^{n/2} be the vector with r𝐯r_{\mathbf{v}} equal to r𝐲rkr_{\mathbf{y}}\circ r_{k} if i=2k1i=2k-1 and r𝐲rk1r_{\mathbf{y}}\circ r_{k}^{-1} if i=2ki=2k. In other words, the vector 𝐯\mathbf{v} is equal to the vector 𝐲\mathbf{y} modified only by shifting its kthk^{th} coordinate by ±1\pm 1. Observe that (Bi)𝐲(R)𝐯(B_{i})_{\mathbf{y}}\subset(R^{*})_{\mathbf{v}}. This shows that \boxplus indeed is equal to the union of the R𝐲R^{*}_{\mathbf{y}}.

To show that \boxplus is tiled by isometric copies of RR^{*}, we need to check that the R𝐲R^{*}_{\mathbf{y}} have non-overlapping interiors. First observe that RR^{*} has nn-volume 1, and that \boxplus has nn-volume 2n2^{n}. Since exactly 2n2^{n} isometric copies of RR^{*} make up \boxplus, they must have disjoint interiors. This concludes the proof of (2).

3.7. Constructing the rep-tile in odd dimensions

We have yet to handle the case where nn is odd, i.e. n=2m+1n=2m+1 for some integer m>0m>0. As before, let FiF_{i} denote the face of CnC^{n} intersecting the (n1)(n-1)-plane where xi=0x_{i}=0. In this case, in addition to the rotations r1,,rmr_{1},\dots,r_{m} defined above, we require an additional rotation f:nnf:\mathbb{R}^{n}\to\mathbb{R}^{n} by an angle of π\pi about the (n2)(n-2)-plane where xn1=0=xnx_{n-1}=0=x_{n}. Note that by definition, Fn=fr(n1)/21(Fn)F_{n}=f\circ r_{(n-1)/2}^{-1}(F_{n}), and so fr(n1)/21f\circ r_{(n-1)/2}^{-1} carries CnC^{n} to its nthn^{th} neighboring cube in \boxplus.

For i=1,,n1i=1,\dots,n-1, choose points αiFi\alpha_{i}\in F_{i} as before. Choose the point αn\alpha_{n} on the (n2)(n-2)-facet of FnF_{n} where FnF_{n} intersects the (n1)(n-1)-plane xn1=1x_{n-1}=1. More specifically, we let

αn=(12,,12,1,0)\alpha_{n}=\left(\frac{1}{2},\dots,\frac{1}{2},1,0\right)

and NnN_{n} be a neighborhood of αn\alpha_{n} in the face FnF_{n} with radius 1/81/8. This guarantees that fr(n1)/21(Nn)f\circ r_{(n-1)/2}^{-1}(N_{n}) is disjoint from NnN_{n}. Therefore, we can again define the boundary sum:

R=Rfr(n1)/21(Bn)k=1mrk1(B2k1)rk(B2k),R^{*}=R\cup f\circ r_{(n-1)/2}^{-1}(B_{n})\cup\bigcup_{k=1}^{m}r_{k}^{-1}(B_{2k-1})\cup r_{k}(B_{2k}),

which is isotopic to RR and tiles =[1,1]n\boxplus=[-1,1]^{n} as before. To complete our proof that RR^{*} is a rep-tile for any nn, we appeal to Lemma 3.1. ∎

Acknowledgments:

This paper is the product of a SQuaRE. We are indebted to AIM, whose generous support and hospitality made this work possible. AK is partially supported by NSF grant DMS-2204349, PC by NSF grant DMS-2145384, RB by NSF grant DMS-2424734, and HS by NSF grant DMS-1502525. We thank Kent Orr for many helpful discussions. We are grateful to Richard Schwartz for his feedback on the first version of this paper; and for numerous valuable suggestions, notably a simplification of our original construction of spherical rep-tiles.

Ball Number

Let RR be a frog with a cube for a bride

Place RR in a box with some balls beside

Set free, the balls

Dance through walls

Out plops a Rep-tile with frogs inside

References

  • [Ada95] Colin C. Adams. Tilings of space by knotted tiles. Math. Intelligencer, 17(2):41–51, 1995.
  • [Ada97] Colin C. Adams. Knotted tilings. In The Mathematics of Long-Range Aperiodic Order (Waterloo, ON, 1995), volume 489 of NATO Adv. Sci. Inst. Ser. C: Math. and Phys. Sci., pages 1–8. Kluwer Academic Publishers, Dordrecht, 1997.
  • [Ber66] Robert Berger. The Undecidability of the Domino Problem. Number 66. American Mathematical Soc., 1966.
  • [Bla24] Ryan Blair, Patricia Cahn, Alexandra Kjuchukova, and Hannah Schwartz. Rep-Tiles. arXiv preprint arXiv:2412.19986v1, 2024.
  • [Bla21] Ryan Blair, Zoe Marley, and Ilianna Richards. Three-dimensional rep-tiles. Proceedings of the AMS, To appear; arXiv 2107.10216.
  • [Cro91] HT Croft, KJ Falconer, and RK Guy. Unsolved Problems in Geometry. Springer, 1991.
  • [Con90] John H Conway and Jeffrey C Lagarias. Tiling with polyominoes and combinatorial group theory. Journal of Combinatorial Theory, Series A, 53(2):183–208, 1990.
  • [Den90] BH Denton. General Topology I, edited by AV Arkhangel’skii and LS Pontryagin. Springer, 1990.
  • [Fox48] Ralph H Fox. On the imbedding of polyhedra in 3-space. Annals of Mathematics, pages 462–470, 1948.
  • [Gar63] Martin Gardner. On rep-tiles, polygons that can make larger and smaller copies of themselves. Scientific American, 208:154–164, 1963.
  • [Gar77] Martin Gardner. Extraordinary nonperiodic tiling that enriches the theory of tiles. Scientific American, 236(1):110–121, 1977.
  • [Gol64] Solomon W Golomb. Replicating figures in the plane. The Mathematical Gazette, 48(366):403–412, 1964.
  • [Gol66] Solomon W Golomb. Tiling with polyominoes. J. Combinatorial Theory, 1:280–296, 1966.
  • [Goo98] Chaim Goodman-Strauss. Matching rules and substitution tilings. Annals of Mathematics, pages 181–223, 1998.
  • [Kob76] Kazuaki Kobayashi and Yasuyuki Tsukui. The ball coverings of manifolds. J. Math. Soc. Japan, 28(1):133–143, 1976.
  • [Luf69] Ehard Luft. Covering of manifolds with open cells. Illinois Journal of Mathematics, 13(2):321–326, 1969.
  • [Oh96] Seungsang Oh. Knotted solid tori decompositions of B3B^{3} and S3S^{3}. Journal of Knot Theory and Its Ramifications, 5(03):405–416, 1996.
  • [Rad94] Charles Radin. The pinwheel tilings of the plane. Annals of Mathematics, 139(3):661–702, 1994.
  • [Rad21] Charles Radin. Conway and aperiodic tilings. The Mathematical Intelligencer, 43(2):15–20, 2021.
  • [Rob71] Raphael M Robinson. Undecidability and nonperiodicity for tilings of the plane. Inventiones Mathematicae, 12:177–209, 1971.
  • [Sch94] Egon Schulte. Space fillers of higher genus. Journal of Combinatorial Theory, Series A, 68(2):438–453, 1994.
  • [Sch25] Richard Schwartz. Personal communication, January 2025.
  • [Sin79] Wilhelm Singhof. Minimal coverings of manifolds with balls. Manuscripta Mathematica, 29(2):385–415, 1979.
  • [Thu89] W. P. Thurston. Tilings and finite state automata. AMS Colloquim Lectures, 1989.
  • [vOp97] G. van Ophuysen. Problem 19. Tagungsbericht, 20, 1997.
  • [Wan60] Hao Wang. Proving theorems by pattern recognition i. Communications of the ACM, 3(4):220–234, 1960.
  • [Zee63] Eric Christopher Zeeman. Seminar on combinatorial topology. (No Title), 1963.