Ferromagnetically ordered states in the Hubbard model on the hexagonal golden-mean tiling
Abstract
We study magnetic properties of the half-filled Hubbard model on the two-dimensional hexagonal golden-mean quasiperiodic tiling. The tiling is composed of large and small hexagons, and parallelograms, and its vertex model is bipartite with a sublattice imbalance. The tight-binding model on the tiling has macroscopically degenerate states at . We find the existence of two extended states in one of the sublattices, in addition to confined states in the other. This property is distinct from that of the well-known two-dimensional quasiperiodic tilings such as the Penrose and Ammann-Beenker tilings. Applying the Lieb theorem to the Hubbard model on the tiling, we obtain the exact fraction of the confined states as , where is the golden mean. This leads to a ferromagnetically ordered state in the weak coupling limit. Increasing the Coulomb interaction, the staggered magnetic moments are induced and gradually increase. Crossover behaviour in the magnetically ordered states is also addressed in terms of perpendicular space analysis.
I Introduction
Quasiperiodic systems have attracted considerable interest since the discovery of the Al-Mn quasicrystal Shechtman et al. (1984). Their properties are of equal interest, in part driven by the observation of behaviour traditionally observed in periodic systems. For example, electron correlations in quasicrystals have been actively studied after quantum critical behavior was observed in the Au-Al-Yb quasicrystal Deguchi et al. (2012). Similarly, long-range correlative states have been reported despite the lack of periodicity inherent in these materials: such as superconductivity in the Al-Zn-Mg quasicrystal Kamiya et al. (2018), and ferromagnetically ordered states in the Au-Ga-X (X = Gd, Tb, Dy) quasicrystals Tamura et al. (2021); Takeuchi et al. (2023). These studies have stimulated, and continue to motivate theoretical investigations on electron correlations and the spontaneously symmetry breaking states in quasicrystals Okabe and Niizeki (1988); Wessel et al. (2003); Jagannathan et al. (2007); Watanabe and Miyake (2013); Takemori and Koga (2015); Takemura et al. (2015); Andrade et al. (2015); Fulga et al. (2016); Otsuki and Kusunose (2016); Sakai et al. (2017); Araújo and Andrade (2019); Sakai and Arita (2019); Varjas et al. (2019); Duncan et al. (2020); Cao et al. (2020); Takemori et al. (2020); Hauck et al. (2021); Ghadimi et al. (2021); Sakai and Koga (2021). For example, magnetically ordered states in the Hubbard model on quasiperiodic bipartite tilings have been studied, including the Penrose Jagannathan et al. (2007); Koga and Tsunetsugu (2017), Ammann-Beenker Wessel et al. (2003); Jagannathan and Schulz (1997); Koga (2020); Oktel (2021), and Socolar dodecagonal Koga (2021). One of the common properties among the majority of these studies is the existence of strictly localized states with (i.e., confined states) in the non-interacting case Kohmoto and Sutherland (1986); Arai et al. (1988); Koga and Tsunetsugu (2017); Koga (2020); Oktel (2021); Koga (2021); Keskiner and Oktel (2022); Matsubara et al. (2023). This leads to interesting magnetic properties in the weak coupling limit.
Recently, we introduced a family of golden–mean hexagonal and trigonal aperiodic tilings produced using a generalization of de Bruijn’s grid method Coates et al. (2023). In this work, we showcased the structural properties and substitution rules of two ‘special’ cases of this family. These are the and tilings, where the subscript refers to the tunable grid-shift parameters used in their construction (for more details, see Coates et al. (2023)). These tilings hold distinct structural properties compared to the Penrose, Ammann-Beenker, and Socolar tilings – not only do they share rotational symmetries associated with periodic systems, but, they also possess a sublattice imbalance due to their vertex structure. However, they are still rooted in the ‘physical’ world of experimentally observed trigonal and hexagonal quasiperiodic systems Woods et al. (2014); Uri et al. (2023); Oka and Koshino (2021); Coates et al. (2020).
It is therefore desirable to study magnetic properties on quasiperiodic systems with sublattice imbalances, in order to systematically understand and compare correlated electron behavior across the widest range of relevant quasiperiodic tilings. In fact, we have already shown the effect which an imbalance has on the magnetic states on one of the special cases from the hexagonal family; the hexagonal golden-mean tiling realizes a ferrimagnetically ordered state in the ground state Koga and Coates (2022), which is in contrast to that in the Penrose Koga and Tsunetsugu (2017), Ammann-Beenker Koga (2020), and Socolar dodecagonal tilings Koga (2021) where antiferromagnetically ordered states are realized without a uniform magnetization.

In this paper, we discuss the relevant properties of the tiling structure and then study the macroscopically degenerate states with in the tight-binding model, which should play an important role for finding magnetic properties in the weak coupling limit. We clarify that two extended states appear in one of the sublattices, while confined states appear in the other. Furthermore, we obtain the exact fraction of the confined states in terms of Lieb’s theorem Lieb (1989), considering magnetism in the weak coupling limit. We also discuss how magnetic properties are affected by electron correlations in the half-filled Hubbard model.
The paper is organized as follows. In Sec. II, we briefly describe the properties of the hexagonal golden-mean tiling needed for our work. In Sec. III, we introduce the half-filled Hubbard model on the hexagonal golden-mean tiling. Then, we study the macroscopically degenerate states with in Sec. IV. By means of the real-space Hartree approximations, we clarify how a magnetically ordered state is realized in the Hubbard model in Sec. V. Finally, crossover behavior in the ordered state is addressed by mapping the spatial distribution of the magnetization to perpendicular space. A summary is given in the last section.
II Properties of the hexagonal golden-mean tiling
Here, we will give an overview and describe the relevant properties of the hexagonal golden-mean tiling which we need for our calculations. The tiling is composed of large hexagons (LH), parallelograms (P), and small hexagons (SH). A section of the tiling, and the schematics of its proto-tiles are shown in Figure 1. The vertex system of the tiling is bipartite, since it is composed of polygons with even edges (hexagons and parallelograms). For our work, we require the exact fractions of tile and vertex frequencies across the tiling, which we take directly from Coates et al. (2023), in which we explicitly explain our methods of derivation.
In the thermodynamic limit, the fractions of the LH, P, and SH tiles are given as:

(1) | |||||
(2) | |||||
(3) |
where is the golden-mean . Similarly, there are seven types of vertices: A, B, C, D, E, F, and G vertices, which are explicitly shown in Figure 2. Their fractions across the tiling are given as:
(4) | |||||
(5) | |||||
(6) | |||||
(7) | |||||
(8) | |||||
(9) | |||||
(10) |
As the tiling is bipartite, trivially, we have two distinct sublattices of vertices. These sublattices can be distinguished either by their occupation of distinct sub-planes in perpendicular space Coates et al. (2023), or by grouping by their coordination number. For example, one sublattice consists of A, B, and C vertices (coordination number of 3), while the other consists of D, E, F, and G vertices (coordination numbers of 4, 5, 6, and 4, respectively). From here on, the sublattice including A, B, and C vertices is denoted as the sublattice and the other is denoted as the sublattice.
As we previously mentioned, this sublattice structure is in contrast to that of the well-known bipartite tilings such as the Penrose and Ammann-Beenker tilings, where half of the vertices for each type exist in both sublattices. The sublattice structure inherent in the tiling, however, leads to the sublattice imbalance , such that Coates et al. (2023):
(11) | |||||
(12) | |||||
(13) |
where and are the fractions of the and sublattices, respectively.
We note the following property which is convenient for reducing the computational cost of mean-field calculations. In the hexagonal tiling, certain tiles or vertices have a local threefold rotational symmetry, e.g. the LH and SH tiles, and the A and F vertices, as seen in Figure 2. Following the substitution rules in Coates et al. (2023), this threefold ‘group’ is changed in a cyclical manner as: LH tile SH tile A vertex F vertex LH tile . Therefore, the system belongs to the point group when one generates the tiling by iteratively applying the deflation rule to an LH or SH tile as its seed, allowing us to save computational time by applying symmetry operations. However, in the thermodynamic limit, the entire system has sixfold rotational symmetry, which is seen in Fourier space Coates et al. (2023).
III Model and hamiltonian
We study the Hubbard model on the hexagonal golden-mean tiling, which is given by the following Hamiltonian
(14) |
where annihilates (creates) an electron with spin at the th site and . is the nearest-neighbor transfer integral and is the on-site Coulomb interaction. For simplicity, we have assumed that the magnitude of the hopping integral is uniform in the system. The chemical potential is always when the electron density is fixed to be half filling.
To discuss magnetic properties in the Hubbard model, we make use of the real-space mean-field theory. This method has an advantage in treating large clusters, which is crucial to clarify magnetic properties inherent in the quasiperiodic systems. Here, we introduce the site-dependent mean-field and the mean-field Hamiltonian is then given as
(15) |
For given values of , we numerically diagonalize the Hamiltonian , update , and repeat this procedure until the result converges. The uniform and staggered magnetizations are given as
(16) | |||||
(17) | |||||
(18) |
where () is the number of the sites (the average of the magnetization) in the sublattice and is the local magnetization at the th site.
Here, we discuss electronic properties in the noninteracting case (), where the model Hamiltonian is reduced to the tightbinding model. Diagonalizing the Hamiltonian for the system with , we obtain the density of states as
(19) |
where () is the number of the sites in the whole system and is the th eigenenergy. The results are shown in Figure 3.

We find the delta-function peak at , suggesting the existence of macroscopically degenerate states. In fact, the clear jump singularity appears at in the integrated density of states. These states should be important for magnetic properties in the weak coupling limit. In the next section, we discuss the macroscopically degenerate states with .
IV Macroscopically degenerate states
Here, we focus on the degenerate states with in the tightbinding model. Since the hexagonal golden-mean tiling is bipartite, these states should exist in both sublattices. First, we focus on the sublattice composed of D, E, F, and G vertices. It is clarified that the number of the degenerate states in the sublattice is at most two, which will be proven in Appendix A. This proof is based on the fact that there exist no tiles with zero amplitudes. Since all tiles have finite amplitudes in either corner site, this should indicate the existence of extended states. To clarify this, we consider the detail of the sublattice. Figure 1(a) shows that the sublattice can be divided into three groups , which are shown as the numbers in the open circles. Each site in the sublattice connects to three nearest-neighbor sites belonging to each of the , , and sublattices. Again, this is proven in Appendix B. Therefore, two states are the exact eigenstates with in the sublattice, where the amplitudes are given as
(20) |
where is the site index in the sublattice and is a solution of the equation . Since finite amplitudes appear in the whole system, these states can be regarded as the extended states. Therefore, we can say that there exist only two extended states with in the sublattice.
By contrast, there are the other macroscopically degenerate states in the sublattice. We construct a simple form, considering their linear combinations, as discussed in previous papers Koga and Tsunetsugu (2017); Kohmoto and Sutherland (1986); Arai et al. (1988). The states can be represented to be exactly localized in a certain region and can be regarded as confined states. These are in contrast to the extended states in the sublattice. Five simple examples of the confined states and are explicitly shown in Figure 4.

According to Conway’s theorem, a certain diagram appears repeatedly in the quasiperiodic tiling, in general. This means that each confined state exists with a finite fraction in the tiling. The diagram for the site structures of and , which is shown in Figure 4, always appears around the F vertex due to the matching rule of the tiles. Therefore, the fractions for these confined states are given by the fraction of the F vertex, . On the other hand, the site structures for , , and do not always appear around the LH and SH tiles, and A vertices, respectively. Taking the tiling structure into account, we obtain the fractions of , , and as , , and , respectively. In the tightbinding model on the hexagonal golden-mean tiling, there are many kinds of confined states (not shown) and therefore the fraction of the confined states is bounded by .
Next, we try to directly obtain the exact fraction of the confined states, making use of magnetic properties at half filling Koga and Coates (2022). According to Lieb’s theorem, the ground state of the half-filled Hubbard model has a total spin for arbitrary . In the weak coupling limit, the magnetically ordered state originates only from the macroscopically degenerate states with . Two extended states in the sublattice should be negligible in the thermodynamic limit. Therefore, magnetic properties little depend on these states and mainly depend on the confined states in the sublattice. Thus, the uniform magnetization can be given as , where is the fraction of the confined states. From these two equations, we obtain the exact fraction of the confined states as
(21) |
This is consistent with the numerical results for the finite cluster with .
We wish to note that in the hexagonal golden-mean tiling the extended states appear in addition to the confined states. The extended states are also found in the tight-binding model on the hexagonal tiling, although this was not discussed in our previous work Koga and Coates (2022).
V Magnetic properties
Here, we discuss magnetic properties in the half-filled Hubbard model on the hexagonal golden-mean tiling. We mainly treat the system with by means of real-space mean-field approximations. When the system is non-interacting, the macroscopically degenerate states appear at the Fermi level, as shown in Figure 3. The introduction of interaction leads to a magnetically ordered state with finite magnetizations: the magnetization profile for the case with is shown in Figure 5(a), where red circles indicate positive magnetizations, and its size is proportional to the magnitude.

We find finite magnetizations only in the sublattice, as discussed above. In particular, the magnetizations in the A vertices are smaller than those in the B and C vertices. This quantitative difference is clearly found in the distribution of the magnetization in Figure 6(a), where the magnetizations on the A vertices are , while those on the B and C vertices are .

This behaviour can be explained by the spatial distribution of the confined states. When one considers the local tiling structure for the confined states (see Figure 4), the A vertices have an amplitude only in the wave function . On the other hand, multiple confined states have amplitudes at B and C vertices. This should lead to a difference in the magnetizations, namely, the confined states in the larger regions have amplitudes in many sites and therefore have a minimal effect on the magnetizations at the A vertices. By contrast, we find no magnetization in the sublattice, which is consistent with the fact that two extended states little affect magnetic properties in the weak coupling limit. From these results, we can say that, in the weak coupling limit, the ferromagnetically ordered state is realized with the total uniform moment .
Increasing the interaction strength, the local magnetization in the sublattice monotonically increases and the magnetizations in the other sublattice are induced. The spatial structure in the magnetization for is shown in Figure 5(b). The magnetization is induced in the sublattice, as shown in Figure 6(b). Further increasing the interaction strength changes the distribution of the local magnetizations: when , the magnetization is almost , as shown in Figure 6(c).

Figure 7 shows the change in the distribution of the local moments. When , the distribution is similar to that in the weak coupling limit . Namely, a sharp peak appears at (the sublattice), while some peaks appear at (the sublattice). When , distinct behavior appears in the magnetic distribution. In the strong coupling case, the local magnetization should be classified into some groups. In the sublattice with , the magnetization at the A vertices is distinct from that at the B and C vertices, and this behavior appears in the whole parameter space. On the other hand, in the sublattice with , the magnetization is classified into three groups characteristic of the coordination number . Namely, for D and G vertices with , for the E vertices, and for the F vertices when , as shown in Figure 6(c). This is distinct from the weak coupling case.
The crossover between the weak and strong coupling regimes occurs around . In the strong coupling limit , the Hubbard model is reduced to the antiferromagnetic Heisenberg model with nearest-neighbor couplings . The mean-field ground state is described by the staggered moment . This means that the mean-field approach cannot correctly describe the reduction of the magnetic moment due to quantum fluctuations. Therefore, an alternative method is necessary to clarify magnetic properties in this regime, which is beyond the scope of the present study. Nevertheless, interesting magnetic properties inherent in the hexagonal golden-mean tiling, eg. the ferromagnetically ordered state in the weak coupling limit, can be captured even in our simple mean-field method.
Finally, we wish to demonstrate the spatial profile of the magnetizations characteristic of the hexagonal golden-mean tiling. To this purpose, we map the tiling to perpendicular space , where the positions in perpendicular space have one-to-one correspondence with the positions in physical space. We have previously shown that there are four windows, which can be labelled by pairs of integer heights Coates et al. (2023), which we re-label here. These heights correspond to where each vertex of the tiling projects onto the body-diagonals of the two 3–dimensional cubes which can be formed by the 6–dimensional super-space basis vectors Coates et al. (2023). Thus, the four planes of the hexagonal golden-mean tiling are described by these heights as where and . The A, B, and C vertices uniquely occupy the (0, 0) and (1, 1) planes, with the remaining vertices occupying the remaining planes, which is schematically shown in Figure 8.


The magnetization profiles of the (0, 0) and (0, 1) planes in perpendicular space are shown in Figure 9, where we show the absolute values of the local magnetizations. As the (1, 0) and (1, 1) planes are equivalent, it is unnecessary to show them. When , the system is essentially the same as that with , where no magnetization appears in the planes (0, 1) and (1, 0) for the sublattice. This is consistent with the fact that the extended states have little effect on the magnetic properties in the sublattice. By contrast, finite magnetization appears across the entirety of the (0, 0) and (1, 1) planes, implying that the spontaneous magnetizations appear in the sublattice. Therefore, we can say that the ferromagnetically ordered state is realized in the weak coupling limit.

We also find a spatial pattern in the B and C vertex regions in the (0, 0) and (1, 1) planes, and, a spatial pattern with a tiny difference appears in the magnetic moments in the A region, as shown in Figure 10. These suggest the existence of many kinds of confined states in relatively large regions. This is because the overlapping structure in the confined states should classify the vertices into hierarchical groups, which yields a detailed structure in perpendicular space, distinct from the simple pattern for the vertices (see Figure 8).
Upon increasing the interaction strength, all vertex sites have magnetizations, as shown in Figure 9(b). In the strong coupling case, the Coulomb interactions become crucial to stabilize the ferrimagnetically ordered states with staggered moments. When , the local magnetization takes large values. In this case, the magnitude of local magnetizations can be classified into two groups in the sublattice and three groups in the sublattice, discussed above.
Before concluding, we would like to summarize and compare the magnetic properties in the Hubbard models on the Penrose, Ammann-Beenker, Socolar dodecagonal, hexagonal golden-mean, and hexagonal golden-mean tilings. One of the common features is the existence of confined states at in the noninteracting case (), which play a crucial role in stabilizing the magnetically ordered states in the weak coupling limit. Nevertheless, their confined state properties are distinct from each other. The number of types of confined states are six in the Penrose case Kohmoto and Sutherland (1986); Arai et al. (1988), while it should be infinite in the others. As for sublattice structures, the hexagonal golden-mean tiling has a sublattice imbalance, leading to a ferrimagnetically ordered state even in the weak coupling limit Koga and Coates (2022). In our tiling, however, there exists a sublattice imbalance such that the confined states appear in one of the sublattices, leading to a ferromagnetically ordered state in the weak coupling limit. The other tilings have an equivalent sublattice structure, and the corresponding Hubbard model shows the antiferromagnetically ordered state without a uniform magnetization.
VI Summary
We have studied magnetic properties in the half-filled Hubbard model on the hexagonal golden-mean tiling by means of the real-space mean-field approach. We have found the delta-function peak in the density of states of the tight-binding model, implying the existence of macroscopically degenerate confined states at . We have then clarified that two extended states exist in the sublattice and the confined states appear only in the sublattice. For the above properties, we have obtained the exact fraction of the confined states as . The introduction of the Coulomb interaction lifts the macroscopic degeneracy at the Fermi level and drives the system to a ferromagnetically ordered state. We have clarified how the spatial distribution of the magnetizations continuously changes with increasing interaction strength. Crossover behaviour in the magnetically ordered states has been discussed by applying perpendicular space analysis to the local magnetizations.
Acknowledgements.
We would like to thank T. Dotera for valuable discussions. Parts of the numerical calculations were performed on the supercomputing systems at ISSP, The University of Tokyo. This work was supported by Grant-in-Aid for Scientific Research from JSPS, KAKENHI Grant Nos. JP22K03525, JP21H01025, JP19H05821 (A.K.), and the EPSRC grant EP/X011984/1 (S.C.).Appendix A Upper bound of the number of the states with in the sublattice
We examine the number of the states with in the sublattice for the noninteracting Hamiltonian . The states with in the sublattice can be described as follows,
(22) |
where is the local state at the th site and is its coefficient. The equation is reduced to the following simultaneous equation,
(23) |
where is the site index in the sublattice and the summation runs to the nearest-neighbor sites of the th site. The number of the equations is given by and the number of coefficients is given by . Although , the solutions of eq. (23) and their number should not be trivial due to the quasiperiodic structure in the tiling.
To clarify the upper bound of the number of solutions, we consider a certain domain which is composed of finite tiles connected by the shared edges. Then, we define a "forbidden domain" so that for the vertices inside and on its boundary. By taking the matching rule of tiles into account, we sometimes find that, on a certain tile outside of the forbidden domain and adjacent to its boundary, the amplitudes of the vertices are zero. This allows us to redefine the forbidden domain to include the tile. In the other words, the forbidden domain can be regarded as to be expanded. In the following, we demonstrate that the forbidden domain can be expanded to the whole system and clarify the upper bound of the number of the degenerate states with .

First, we focus on a P tile outside of the forbidden domain and adjacent to its boundary, as shown in Figure 11(a). Here, we have labeled three sites in the sublattice as , , and . The site is located on the shared edge, and the site is located on the other corner of the P tile. The site on the shared edge in the sublattice connects to the nearest-neighbor sites , , and . Since the definition of the forbidden domain, , and the site must be on the boundary of the forbidden domain, we obtain . We then obtain since [eq. (23)]. Therefore, each site on the P tile has no amplitude, meaning that the forbidden domain is expanded to include the P tile, as shown in Figure 11(b). By taking into account the above rule, the forbidden domain can be expanded so that no P tiles touch outside it. In the hexagonal golden-mean tiling, the P tiles densely exist with their fraction and some of them are connected to each other (see Figure 2). Therefore, the forbidden domain should be expanded according to the above rules.
Next, we focus on a certain LH tile outside of the forbidden domain and adjacent to its boundary, as shown in Figure 12(a). We have assumed that the LH tile and forbidden domain share the edge with the sites and , which belong to the and sublattices, respectively.

When the A vertex is located at the site , the local tiling structure is shown in Figure 12(b). The LH tile in the forbidden domain is adjacent to four P tiles and some P tiles are also connected to each other. Therefore, the forbidden domain should be expanded, which is shown as the shaded area in Figure 12(b). Furthermore, according to eq. (23). Therefore, we conclude that the amplitudes of all corner sites of three LH tiles are zero and the forbidden domain is expanded to include two LH tiles.
When the B vertex is located at the site , it is necessary to consider three cases according to the type of vertex at the site ; D, E, and G vertices:
-
(i)
the E vertex is located at the site : the local structure around the site is shown in Figure 12(c). The amplitudes of the corner sites of the LH tile are zero since three sites belonging to the sublattice share the connected P tiles. Therefore, the forbidden domain can be spatially expanded to include the LH tile.
-
(ii)
the D vertex is located at the site : the local structure is symmetric, as shown in Figure 12(d). At the sites and the D vertex can not be found due to the matching rule of tiles, however, E or G vertices can be. In the case of the E vertex being located at either or then we essentially find the same case as (i) and thereby the forbidden domain is expanded to include these two LH tiles. When G vertices are located at both sites and , the local structure is shown in Figure 13(a). In this case, either E or G vertex is located at the site . When it is the G vertex, the local structure is shown in Figure 13(b). In this case, the forbidden domain is expanded to include the LH tiles, which are shown as the hatched hexagons, and the P tiles adjacent to them. Furthermore, the forbidden domain is expanded to include the S tiles sharing the sites and . Therefore, finally, the forbidden domain is expanded to include all the tiles shown in Figure 13(b). On the other hand, when the E vertex is located at the site , the vertex structure is shown in Figure 13(c). In this case, one may not expand the forbidden domain, which is shown as the shaded area, to a larger domain in terms of our simple rule. Now, we must consider the fifteen vertices in the sublattice outside the area, which is denoted as , , , . The equations (23) are explicitly given as
(24) (26) since the amplitudes of the vertices on the shaded area are zero. Thus, we obtain . This means that the amplitudes of the vertices on the LH and SH tiles outside of the area must be zero and the forbidden domain can be expanded to include these tiles.
-
(iii)
the G vertex is located at the site : the local structure is shown in Figure 14(a). When the E (D) vertex sits at site , the local vertex structure is the same as the case (i) [(ii)]. Therefore, the forbidden domain is expanded. When the G vertices are located at both sites and , the F vertex is located at the site due to the matching rule of the tiles. The forbidden region is shown as the shaded area in Figure 14(b). Here, we focus on five vertices in the sublattice outside the shaded area, which is denoted as , to examine their amplitudes. According to eq. (23), the state with is satisfied by the following equations,
(27) (28) (29) (30) (31) Thus, we obtain . This means that the amplitudes of the vertices on the LH tiles are zero and the forbidden domain can be expanded to the LH tiles.


From these results, we can say that the forbidden domains can be expanded to the whole system since the S tiles are always isolated in the hexagonal golden-mean tiling, as shown in Figure 2. Namely, the amplitude of the wave function is zero in the whole system and no degenerate states with appear in the sublattice under the assumption of the existence of the forbidden domain. The assumption is equivalent to two conditions imposed on the wave function, where the th site belongs to the sublattice on a certain P tile. Therefore, we can prove that the number of the degenerate states with in the sublattice is at most two.
Appendix B Three groups in the sublattice
Here, we will prove that the sublattice can be divided into three groups , and , so that each site in the sublattice connects to three nearest-neighbor sites belonging to each of these groups. Similar to appendix A, we introduce a domain so that, inside and on its boundary, the groups for these vertices are determined. In the following, we demonstrate that the domain can be expanded to the whole system. We first focus on a P tile outside of the domain and adjacent to its boundary, as shown in Figure 11(a). When the groups for the sites and are determined, the group of the site is uniquely determined. Therefore, the domain can be expanded to include the P tiles according to this rule.
Next, we consider the LH tile outside of the domain and adjacent to its boundary, as shown in Figure 12(a). By considering some cases according to the type of vertex at the site , , etc., the group for each site is uniquely and trivially determined except for two cases, which are shown in Figure 13(c) and Figure 14(b). In the former (latter) case, the group for the site () is uniquely determined to belong to the groups for the site , since the two of next nearest-neighbor sites of () belong to the groups for the sites . And the groups for the site () are uniquely determined by the above rules. Then, we can expand the domain to include the LH tiles. From these results, we can say that the sublattice are divided into three groups and each site in the sublattice connects to three nearest-neighbor sites belonging to each of these groups.
References
- Shechtman et al. (1984) D. Shechtman, I. Blech, D. Gratias, and J. W. Cahn, Phys. Rev. Lett. 53, 1951 (1984).
- Deguchi et al. (2012) K. Deguchi, S. Matsukawa, N. K. Sato, T. Hattori, K. Ishida, H. Takakura, and T. Ishimasa, Nat. Mat. 11, 1013 (2012).
- Kamiya et al. (2018) K. Kamiya, T. Takeuchi, N. Kabeya, N. Wada, T. Ishimasa, A. Ochiai, K. Deguchi, K. Imura, and N. K. Sato, Nat. Commun. 9, 154 (2018).
- Tamura et al. (2021) R. Tamura, A. Ishikawa, S. Suzuki, T. Kotajima, Y. Tanaka, T. Seki, N. Shibata, T. Yamada, T. Fujii, C.-W. Wang, M. Avdeev, K. Nawa, D. Okuyama, and T. J. Sato, J. Ame. Chem. Soc. 143, 19938 (2021).
- Takeuchi et al. (2023) R. Takeuchi, F. Labib, T. Tsugawa, Y. Akai, A. Ishikawa, S. Suzuki, T. Fujii, and R. Tamura, Phys. Rev. Lett. 130, 176701 (2023).
- Okabe and Niizeki (1988) Y. Okabe and K. Niizeki, J. Phys. Soc. Jpn. 57, 16 (1988).
- Wessel et al. (2003) S. Wessel, A. Jagannathan, and S. Haas, Phys. Rev. Lett. 90, 177205 (2003).
- Jagannathan et al. (2007) A. Jagannathan, A. Szallas, S. Wessel, and M. Duneau, Phys. Rev. B 75, 212407 (2007).
- Watanabe and Miyake (2013) S. Watanabe and K. Miyake, J. Phys. Soc. Jpn. 82, 083704 (2013).
- Takemori and Koga (2015) N. Takemori and A. Koga, J. Phys. Soc. Jpn. 84, 023701 (2015).
- Takemura et al. (2015) S. Takemura, N. Takemori, and A. Koga, Phys. Rev. B 91, 165114 (2015).
- Andrade et al. (2015) E. C. Andrade, A. Jagannathan, E. Miranda, M. Vojta, and V. Dobrosavljević, Phys. Rev. Lett. 115, 036403 (2015).
- Fulga et al. (2016) I. C. Fulga, D. I. Pikulin, and T. A. Loring, Phys. Rev. Lett. 116, 257002 (2016).
- Otsuki and Kusunose (2016) J. Otsuki and H. Kusunose, J. Phys. Soc. Jpn. 85, 073712 (2016).
- Sakai et al. (2017) S. Sakai, N. Takemori, A. Koga, and R. Arita, Phys. Rev. B 95, 024509 (2017).
- Araújo and Andrade (2019) R. N. Araújo and E. C. Andrade, Phys. Rev. B 100, 014510 (2019).
- Sakai and Arita (2019) S. Sakai and R. Arita, Phys. Rev. Research 1, 022002(R) (2019).
- Varjas et al. (2019) D. Varjas, A. Lau, K. Pöyhönen, A. R. Akhmerov, D. I. Pikulin, and I. C. Fulga, Phys. Rev. Lett. 123, 196401 (2019).
- Duncan et al. (2020) C. W. Duncan, S. Manna, and A. E. B. Nielsen, Phys. Rev. B 101, 115413 (2020).
- Cao et al. (2020) Y. Cao, Y. Zhang, Y.-B. Liu, C.-C. Liu, W.-Q. Chen, and F. Yang, Phys. Rev. Lett. 125, 017002 (2020).
- Takemori et al. (2020) N. Takemori, R. Arita, and S. Sakai, Phys. Rev. B 102, 115108 (2020).
- Hauck et al. (2021) J. B. Hauck, C. Honerkamp, S. Achilles, and D. M. Kennes, Phys. Rev. Res. 3, 023180 (2021).
- Ghadimi et al. (2021) R. Ghadimi, T. Sugimoto, K. Tanaka, and T. Tohyama, Phys. Rev. B 104, 144511 (2021).
- Sakai and Koga (2021) S. Sakai and A. Koga, Mater. Trans. 62, 380 (2021).
- Koga and Tsunetsugu (2017) A. Koga and H. Tsunetsugu, Phys. Rev. B 96, 214402 (2017).
- Jagannathan and Schulz (1997) A. Jagannathan and H. J. Schulz, Phys. Rev. B 55, 8045 (1997).
- Koga (2020) A. Koga, Phys. Rev. B 102, 115125 (2020).
- Oktel (2021) M. O. Oktel, Phys. Rev. B 104, 014204 (2021).
- Koga (2021) A. Koga, Mater. Trans. 62, 360 (2021).
- Kohmoto and Sutherland (1986) M. Kohmoto and B. Sutherland, Phys. Rev. Lett. 56, 2740 (1986).
- Arai et al. (1988) M. Arai, T. Tokihiro, T. Fujiwara, and M. Kohmoto, Phys. Rev. B 38, 1621 (1988).
- Keskiner and Oktel (2022) M. A. Keskiner and M. O. Oktel, Phys. Rev. B 106, 064207 (2022).
- Matsubara et al. (2023) T. Matsubara, A. Koga, and S. Coates, J. Phys. Conf. Ser. 2461, 012003 (2023).
- Coates et al. (2023) S. Coates, A. Koga, T. Matsubara, R. Tamura, H. R. Sharma, R. McGrath, and R. Lifshitz, “Hexagonal and trigonal quasiperiodic tilings,” (2023), arXiv:2201.11848 [cond-mat.soft] .
- Woods et al. (2014) C. Woods, L. Britnell, A. Eckmann, R. Ma, J. Lu, H. Guo, X. Lin, G. Yu, Y. Cao, R. Gorbachev, A. Kretinin, J. Park, L. Ponomarenko, M. Katsnelson, Y. Gornostyrev, K. Watanabe, T. Taniguchi, C. Casiraghi, H.-J. Gao, A. Geim, and K. Novoselov, Nat. Phys. 10, 451 (2014).
- Uri et al. (2023) A. Uri, S. C. de la Barrera, M. T. Randeria, D. Rodan-Legrain, T. Devakul, P. J. D. Crowley, N. Paul, K. Watanabe, T. Taniguchi, R. Lifshitz, L. Fu, R. C. Ashoori, and P. Jarillo-Herrero, “Superconductivity and strong interactions in a tunable moiré quasiperiodic crystal,” (2023), arXiv:2302.00686 [cond-mat.mes-hall] .
- Oka and Koshino (2021) H. Oka and M. Koshino, Phys. Rev. B 104, 035306 (2021).
- Coates et al. (2020) S. Coates, S. Thorn, R. McGrath, H. R. Sharma, and A. P. Tsai, Phys. Rev. Mater. 4, 026003 (2020).
- Koga and Coates (2022) A. Koga and S. Coates, Phys. Rev. B 105, 104410 (2022).
- Lieb (1989) E. H. Lieb, Phys. Rev. Lett. 62, 1201 (1989).