Non-Abelian topological defects and strain mapping in 2D moiré materials
Abstract
We present a general method to analyze the topological nature of the domain boundary connectivity that appeared in relaxed moiré superlattice patterns at the interface of 2-dimensional (2D) van der Waals (vdW) materials. At large enough moiré lengths, all moiré systems relax into commensurated 2D domains separated by networks of dislocation lines. The nodes of the 2D dislocation line network can be considered as vortex-like topological defects. We find that a simple analogy to common topological systems with an order parameter, such as a superconductor or planar ferromagnet, cannot correctly capture the topological nature of these defects. For example, in twisted bilayer graphene, the order parameter space for the relaxed moiré system is homotopy equivalent to a punctured torus. Here, the nodes of the 2D dislocation network can be characterized as elements of the fundamental group of the punctured torus, the free group on two generators, endowing these network nodes with non-Abelian properties. Extending this analysis to consider moiré patterns generated from any relative strain, we find that antivortices occur in the presence of anisotropic heterostrain, such as shear or anisotropic expansion, while arrays of vortices appear under twist or isotropic expansion between vdW materials. Experimentally, utilizing the dark field imaging capability of transmission electron microscopy (TEM), we demonstrate the existence of vortex and antivortex pair formation in a moiré system, caused by competition between different types of heterostrains in the vdW interfaces. We also present a methodology for mapping the underlying heterostrain of a moiré structure from experimental TEM data, which provides a quantitative relation between the various components of heterostrain and vortex-antivortex density in moiré systems.
I Introduction
Moiré patterns are quasi-periodic in-plane projections of two similar stacked 2-dimensional (2D) periodic lattices. Atomic scale moiré superlattices can be formed by stacking atomically thin van der Waals (vdW) materials; one such example is twisted bilayer graphene. Moiré patterns formed by incommensurately stacking 2D materials have been used to manipulate a system’s electronic structure, from Hofstadter’s butterfly [1, 2, 3] to the valley Hall effect [4, 5] to magic angle strongly correlated physics [6, 7]. As the number and type of layers in experimentally relevant systems proliferates, including twisted double bilayer [8], twisted trilayer [9, 10, 11] and twisted quadrilayer graphene [12], as well as hexagonal boron nitride [13] and transition metal dichacolgenides (TMDs) [14], it is important to be able to predict the structure in vdW stacked combinations of atomic layers.
Increasing attention has been paid to the effects of strain disorder on the structure and properties of such systems [15]. The effect and extent of twist angle disorder in magic angle graphene is an active area of research [16, 17]. Strained moiré patterns in excitonic systems have been proposed as a way to create 1D quantum wires [18]. In this paper, we present a generalizable topological interpretation of the structure of moiré interfaces that allows for the characterization of arbitrary strain and the proposition of new types of moiré patterns.
A topological description of the moiré structure is appealing in part because some of the major features of the structure seem to be fixed once certain boundary conditions, such as total twist angle and strain, are pinned. For large enough moiré length, moiré systems are known to relax into domains of nearly commensurate alignment, separated by domain walls which can be characterized as dislocation lines [19]. The topological connectivity of the network of dislocation lines remains fixed even as the domain lengths become distorted by local strain fields.
The nodes of the network where dislocation lines meet in the relaxed moiré system (sometimes known as AA points in graphene or TMD moiré) have been referred to as topological defects by Alden et al. [20], and again by Turkel et al. [21], who also emphasize their role in transport as tunable, local concentrations of twist angle. They have also been shown to play the role of defects in electrochemistry [22]. However, we find that the order parameter describing the topology of these defects is rather different from that of the conventional vortex descriptions in the planar magnet or superfluid, where the order parameter can be described by a complex number of unit length. The unit length complex order parameter space can be mapped to a circle (), hence the fundamental group is . The fundamental group is a useful tool in the analysis of topological defects, as defects and their collections can each be mapped to an element of the fundamental group [23]. For defects with an order parameter, this manifests as an integer winding number [24]. Interestingly, we find that the order parameter space describing the relaxation structure of a moiré superlattice is not homotopy equivalent to due to the periodic boundary condition of the order parameter imposed by the moiré superlattice. We present a mathematical framework, similarly based on the fundamental group, by which the 2D dislocation network nodes can be characterized as topological defects.
Throughout the paper, we refer to the space of order parameters as the configuration space, in keeping with the convention adopted in earlier works [25]. For most of this work, we focus on graphene-like moiré superlattices, where two different domain types, AB and BA, are separated by dislocation lines meeting at the AA site. The starting point for our new vortex description of the AA nodes is the realization that the configuration space for the graphene-like moiré superlattice is the punctured torus, which is homotopy equivalent to theta space (to be defined later). The fundamental group of the punctured torus is , the free group on two generators and . The generators and are noncommuting, so is non-Abelian. The commutator corresponds to a closed loop around a vortex centered at the AA node of the moiré superlattice (its inverse corresponds to a path around an antivortex). The non-commutativity of the generators is a consequence of the removed point at the AA node, which is physically motivated by the node’s high energy barrier in the generalized stacking fault energy [25].
The domain walls that emerge in the moiré superlattice will be color-coded as , which correspond to the three distinct ways in which the AB stacking order makes a transition to the BA stacking order as the domain wall is crossed. The aforementioned generators and can be related to the color coding of the domain walls, which are experimentally accessible quantities. Examination of the color distribution of domain walls crossing an arbitrary closed boundary reveals the vortex content enclosed within, according to the non-Abelian vortex theory developed here. The use of the free group language to characterize the topological structure of a moiré superlattice is likely to find application in other kinds of superlattices formed from incommensurate stacking of non-hexagonal crystals, with details of the group determined by the material-dependent stacking energy profiles and lattice symmetry.
As mentioned before, the non-Abelian vortices in a moiré superlattice have a counterpart in non-Abelian antivortices. We find that the relative strain tensor between two constituent layers determines the vortex/antivortex distribution of the sample. The mathematical tools to understand configuration space combined with experimental information on the configuration distribution in real space enables us to estimate the strain distribution underlying a moiré pattern, allowing for the characterization or engineering of strain distributions in van der Waals heterostructures.
The remainder of the paper is organized as follows. In Sec. II we go over various tools and concepts used in analyzing the strain patterns and the nature of topological defects in the moiré superlattice. The notion of theta space as the proper configuration space of the graphene-like moiré superlattice is introduced and justified by energetic consideration. Sec. III discusses the formal theory of the vortex and antivortex in a moiré superlattice using the language of the free group and its generators. The mathematical discussion is followed by Sec. IV in which the new algebraic formulation of vortices and anti-vortices is employed to identify antivortex formation in a real moiré superlattice, by way of a novel method of strain mapping. Sec. V gives a summary and discussion. Technical details of the theory of vortex algebra are included in Appendix A and Appendix B with the hope that future investigations of moiré superlattices, including lattices other than sublattice-symmetric honeycomb lattices, can make use of the type of vorticity formulation presented here. Appendix C includes details on the image processing to experimentally measure the lattice displacement.
II Order parameter and configuration space for moiré superlattice
The natural choice of order parameter in a bilayer system is the local shift vector, , determined as the in-plane vector that points from a lattice site in one layer to the equivalent lattice site in the other layer. Fig. 1 illustrates of how we define the shift vector in a graphene-like honeycomb lattice. Because of the periodicity of the lattice, a shift vector larger than a unit cell, as shown in Fig. 1(b), is equivalent to the shorter vector folded into the first unit cell. In other words, the configuration space in which this order parameter exists is a torus [25].

After labeling the two honeycomb lattice sites as A and B, the standard naming convention for the bilayer honeycomb stackings is obtained by listing the pair of vertically aligned sites. The condition , when every atom is on top of an equivalent atom in the other layer, is known as AA stacking. Shifting the top layer along one of the three atomic bonds from an A site to a B site of the other layer (Fig. 1(d)), or along the negative of those three vectors (Fig. 1(c)) yields a structure where only half the atoms in the top and bottom layers coincide in the 2D projection. The latter two stacking configurations are often termed AB and BA stacking, respectively. In graphene, the AA stacking is energetically unfavorable, while the AB and BA stackings are symmetry-related lowest-energy layer stacking configurations, called Bernal stacking. Note that for each of the two graphene Bernal stacking configurations, three different shifting directions result in the same stacking configuration, represented by a single point on the toroidal configuration space. AB and BA stacking are connected by spatial inversion, leading to for the corresponding shift vectors in the configuration space (Fig. 1(e)). The high-energy nature of the AA stacking can be reflected by removing from the configuration space altogether, rendering the topology from that of a torus to that of a punctured torus. This will play a crucial role in the theory of vorticity we develop in Sec. III.
Fig. 1(e) shows the points in configuration space corresponding to the real-space configurations in (a-d). The high symmetry stackings, BA and AB, are represented by the dark gray and light gray points in Fig. 1(e), respectively. As the unit cell can be defined in various ways, it is equally valid to use the parallelogram definition of the unit cell shown in Fig. 1, where AA is the corner, or the hexagonal unit cell shown in dotted lines, in which AB and BA are corners of the hexagon.
II.1 Experimental measurement of order parameter
Transmission electron microscopy (TEM) provides an experimental route to characterizing local atomic configurations in real space, including detecting the change in across a domain wall, known as the Burgers vector. Changes in the stacking order are distinguishable by a dark field imaging technique that consists of inserting an aperture into the diffraction plane around a single Bragg position and recording the resulting filtered real space image. Depending on the choice of diffraction peak, contrast between domains (see Fig. 2(a)) or the partial dislocations that form the domain walls (see Fig. 2(b)) in the bilayer stacking are visible [20]. The Burgers vectors of the dislocations are exactly determinable, as the vector perpendicular to the diffraction peak for which their contrast vanishes in the dark field image [20, 26]. The Burgers vector of a dislocation in the bilayer is equal to , the change in shift vector across the boundary.
We note that the dislocations form a network with distinct topology. Twisted bilayer graphene (Fig. 2(b)) has a structure where six dislocation lines meet at a node. Despite being a different material with different symmetries, MoS2 slightly twisted from a 3R-like stacking (close to 0∘ twisting) appears to share the same dislocation network topology as graphene (Fig. 2(c)). In contrast, MoS2 twisted from a 2H configuration (close to 180∘ twisting) has a structure where three lines meet at a node rather than six (Fig. 2(d)). As we will explore later, the topology of the graphene and 3R-like case is defined by a punctured torus, originating from the single energy maximum in the stacking fault energy as a function of configuration (and two degenerate minima). However, for twisted 2H MoS2 there are two energy maxima and one minimum, leading the configuration space to be described as a twice-punctured torus and generating a different topology. While in this paper we focus on the topology generated by the energy profile of bilayer graphene, we are motivated by the realization that a topological description of each system’s configuration space should distinguish the different possible connectivities of the network, including those arising from materials of different lattice symmetry, such as a square lattice.

While constructing the tools to analyze the topological defects of the moiré superlattice in configuration space, we also attain the ability to analyze the configuration distribution and extract the overall strain profile in the twisted bilayer. In order to do these analyses, we need to consider several important quantities of the moiré lattice: the displacement gradient matrix, the moiré vectors, and the Burgers vectors. All of these quantities give important information that allows us to reconstruct .
II.2 Displacement gradient matrix
A displacement vector field which describes the shift in positions of a lattice compared to a reference lattice such that , is closely related to strain. Typically, is a 2D vector corresponding to the positions in the strained lattice, and is a 2D vector corresponding to the positions in the intrinsic lattice, prior to application of forces. In the case of moiré materials, presented in Fig. 1, the reference lattice (with coordinates ) is the bottom layer of the material, and the displaced lattice () is the top layer. Thus the vector field that produces a moiré pattern is in a sense analogous to the relative strain (heterostrain) between the two layers.
The linear strain matrix, used in the modeling of strain energies, is defined by taking derivatives of the field and symmetrizing:
(1) |
Note that this linear strain tensor removes rotation contributions to first order in angle, since rotating the lattice does not cost energy and thus should not be counted towards the total strain energy.
Without symmetrizing, the gradient of the field is known as the displacement gradient matrix, , obtained from
(2) |
The vector can be recovered by integrating over distance. If we assume that and that the strain is spatially uniform from to , we obtain
(3) |
Thus, or
(4) |
Any arbitrary displacement of layer 2 relative to layer 1, formed by a combination of twist and strain, can be written in terms of a displacement gradient matrix, , with four independent components. It is useful to define the four components as follows:
(5) |
where is a linear approximate of the twist (measured in radians), is isotropic strain, and and are uniaxial and shear strain, respectively [27].
II.3 Moiré length
The moiré length is the distance over which the lattice has been shifted by a unit vector with respect to lattice , creating a local return to the starting configuration.
Consider the shift vector , as defined in Fig. 1. If we twist our lattices about a point where two atoms are stacked on top of each other, the origin has Traveling away from the origin in a direction , the displacement increases until we reach the point where the lattices have diverged by a whole unit cell ( or , where and are two lattice vectors of the unstrained lattice) and thus are aligned ( again, resulting in moiré periodicity.
A general interface of 2D lattices with heterostrain, , can be characterized by a pair of moiré vectors, (1, 2), which describes the moiré periodicity in the 2D space. Considering the corresponding pair of two coincident lattice points in the upper and lower layers, and , respectively, the two moiré vectors can be expressed
(6) |
where the constant vectors can each be or , depending on the coincident lattice condition for the moiré superlattice.
If and are collinear (without and being collinear), then the matrix has determinant zero and the moiré pattern is 1D rather than 2D. We will ignore this case for now, and assume that and are each used once.
We can relate the lattice constants , the matrix and the moiré vectors, by Thus if is invertible, and
(7) |
Putting Eq. 7 together with the corresponding equation for the other moiré vector, , we obtain a matrix equation that can be solved for if we have measured and :
(8) |
where and are matrices formed by horizontally concatenating the and column vectors, respectively. Note that if the two ’s are not linearly independent, that is again the case.
II.4 Burgers vector
The shift vector is also related to the Burgers vector of the dislocations that form in the relaxed moiré system. To define the Burgers vector, first one considers a closed path along the lattice points of a perfect crystal, sometimes called a Burgers circuit. When a dislocation is introduced within this path the circuit becomes broken, and the vector connecting the now separated start and end points of the circuit is called the Burgers vector. In a moiré superlattice, the twisted interface acts as a dislocation, and we can obtain Burgers vectors by considering circuits that traverse the interface. The simplest closed circuits in the aligned, untwisted structure take the following form: travel from a lattice point to along the top layer, then down vertically into the bottom layer, then along to in the bottom layer, and finally return to in the top layer by moving vertically upwards. To connect the Burgers vector to the shift vector , it is easiest to think of obtaining the twisted geometry by keeping the bottom layer fixed and introducing a dislocation via rotation of the top layer. In this case, the failure of the circuit to close after applying the twist is equal to the relative change in positions of the lattice points that corresponded to and in the top layer. However this is simply the change in the local shift between the two points, e.g.
(9) |
where the non-integral form is only true if one does not map into the compact unit-cell torus.
While this expression was obtained under the assumption of a uniformly twisted system, the result is quite general. For the relaxed systems that occur in experimental devices, one finds that the Burgers vector is only non-zero if a “dislocation line” (that is, a domain wall) is enclosed in the circuit. Each dislocation line between a pair of AA nodes can be associated with a specific Burgers vector, as was done experimentally using DF TEM in Fig. 2. The pair of AA nodes also provides a moiré vector , which can be linked to obtained from the Burgers vectors. Therefore, knowledge of the moiré vectors and the Burgers vectors across the dislocations associated with is sufficient to solve for the displacement gradient matrix using Eq. 8. As the moiré vectors and Burgers vectors can be measured from experimental dark field and diffraction images, it is possible to obtain a map of from the information provided by DF TEM images, as we will discuss in Sec. IV.2.
II.5 Configuration space
Although a torus is topologically nontrivial, a small loop around the AA site () on the torus can be contracted to a point and thus does not inherit any nontrivial topological properties from the torus. The topological defects associated with winding around the holes of the torus are the dislocations [24], but this description alone provides no constraint on the manner in which the dislocations meet at the nodes. In the case of a graphene moiré superlattice, the three dislocation lines with different Burgers vector, colored red(), green(), and blue(), converge at the AA defect and diverge out again. What we need is a proper mathematical description of the topological nature of the AA nodes consistent with the experimental findings.
We start by investigating the distribution in configuration space of the order parameter for unrelaxed and relaxed moiré systems. Fig. 3(a) shows an unrelaxed twisted bilayer of a honeycomb lattice, and Fig. 3(d) shows the configuration space, colored via a scheme that is used consistently throughout this work. A Gaussian intensity distribution of a chosen color is centered around each key point in configuration space. The region of configuration space centered around the AA point is colored white, and those centered around AB (BA) are dark gray (light gray). Red, green, and blue regions are placed on the midpoint of the three equidistant shortest paths between AB and BA, two of which require crossing the unit cell boundaries. The coloring in real space in Fig. 3(a-c) is determined for each point, by determining the local shift vector, finding it as a point in configuration space, and adopting the color corresponding to that point.

Fig. 3(b-c) depict real space structures that have been modulated by a periodic lattice distortion to mimic relaxation, with two different amplitudes. As the amplitude of relaxation is increased from Fig. 3(a) to (b) to (c), light or dark gray regions, corresponding to nearly AB or BA stacking, take up increasing areas in real space, while AA regions shrink, and red, green, and blue regions evolve into lines. Because the red, green, and blue color tell us which path on the torus was used to get between BA and AB, i.e. , the color tells us the Burgers vector of that line. The corresponding distributions in configuration space are shown in Fig. 3(d-f). As relaxation strength increases, decreasing point density around the AA configuration reveals that fewer points in real space correspond to AA stacking, while configurations on the red, green, and blue lines, and especially AB and BA sites, become more numerous. Thus, the atomic lattice relaxation process in real space can be viewed as emptying out most of configuration space and populating only the AB and BA points and dislocation lines connecting them.
A similar emptying of AA and concentration at AB, BA, and the colored lines between AB and BA, occurs for the three strain types: isotropic, uniaxial, and shear. Fig. 4(a-c) show the configurations formed by with a spatially constant displacement gradient matrix , corresponding to only one nonzero component or in Eq. 5. Fig. 4(d-f) show corresponding structures after applying the modulation function to mimic atomic scale relaxation. We note that the order and orientation of red, green, and blue lines differs in real space for each strain component, as well as twist. In this sense, the colors (encoding local order parameter) provide information about which strain components are present.

III Theory of dislocation network nodes in moiré superlattices
As emphasized previously, the proper order parameter space for a moiré superlattice and the various network structures observed in it is the space of shift vectors with the boundary conditions of a torus and one point removed due to energy considerations (and the ensuing atomic relaxation), as shown in Fig. 5(b). But the TEM images shown in Fig. 2 suggest an even more constrained space for the order parameter. For moiré regions with a sufficiently large moiré length scale, the order parameter is locked to either the AB () or BA () point in the configuration space. There are three equivalent ways to make a transition from BA to AB stacking orders, designated by red, blue, and green arrows in Fig. 5(d) with corresponding label , , and . The AB to BA transition is accomplished by their inverses, shown as , and in Fig. 5(d). Since the , and the three lines connecting the two points span the entirety of the relevant order parameters, one can “gouge out” the unnecessary portions of the punctured torus. The result is the theta space shown in Fig. 5(c). This is the relevant configuration on which to make a proper definition of vorticity, not the circle () where the usual homotopic classification of vorticity takes place [24].
Before developing the formal theory of vorticity in the next subsection we complete the phenomenological classification of possible vortex patterns around the AA node. According to the TEM data, a path around a single AA node in real space simultaneously implies encircling the AA spot in configuration space by . Closed paths in real space that encircle a node of the dislocation network now correspond to non-contractible paths in configuration space. The vortex winding number can be intuitively defined as or if the configuration space loop cycles the same or opposite direction as the real space loop, respectively (Fig. 5(d)). Recalling that the red, green, and blue paths in configuration space correspond to Burgers vectors of dislocations in real space, there are four distinct orderings of Burgers vectors in real space, two corresponding to vortices (, Fig. 5(e-f)), and two corresponding to antivortices (, Fig. 5(g-h)). Comparing the arrangement of Burgers vectors around the loop with the configurations in Fig. 4, it can be concluded that twist and isotropic strain produce vortex-type defects at the nodes, and uniaxial or shear strain produce antivortex-type defects at the nodes.

III.1 Algebraic formulation of vorticity

To understand the algebraic structure of paths in the theta space (Fig. 5(c)), it is useful to consider the same path through configuration space in both the unit-cells shown in Fig. 1(e). Let us consider the full encirclement of the AA node shown in Fig. 5(d). We begin by observing that the BA AB BA transition through the and dislocations in Fig. 5(d) becomes, in the parallelogram scheme shown in Fig. 6(a), a non-contractible loop about the torus’ periodic boundary conditions. The other BA AB BA transition through the and arrows in the hexagonal scheme becomes another non-contractible loop in the parallelogram unit cell shown in Fig. 6(b). We label the first and the second transitions as and , which will soon be identified with two generators of the free group [28, 29]. The third BA AB BA move through and arrows of the hexagon can be decomposed as the inverse of followed by the inverse of , or [Fig. 6(c)]. A complete loop around the edges of the hexagonal unit cell is equivalent to the algebraic operation . Our convention is to perform the operation appearing on the left side of the product first.
The commutator in the language of the free group with two generators represents a “vortex” centered about the AA defect. The vorticity of this topological defect can naturally defined as the commutator (more detailed discussion in Appendix A). Similar to the conventional vortex defined in , this vorticity defined for the AA node is non-trivial in the sense that it is not contractible to an identity, as illustrated in Fig. 6(d). After cancelling out the paths that are traversed both ways, the overall path for becomes equivalent to four partial loops around the four corners of the parallelogram, equal to a full loop round . On a torus such a loop can be contracted to zero and become trivial, but not for a punctured torus. The anti-vortex has the algebraic representation . Geometrically, this amounts to starting from the same BA point on the upper left corner of the hexagonal unit cell in Fig. 5(d) and making a complete counter-clockwise loop. Appendix A gives a more complete account of the vortex structures in the language of free groups with two generators . One can find group-theoretic representations for vortex dipoles (vortex + antivortex) and vortex quadrupoles (two vortices and two antivortices) as well.
III.2 RGB formulation of vorticity
As we described in Section IIA, dark field TEM imaging can identify the dislocation lines with given Burgers vectors and the AB and BA domains separated by them. Each dislocation line converging on a given AA node can then be color-coded as , , or one of their inverses , considering Burgers vector and the neighboring AB/BA domains in the TEM measurement. The free group language of the previous subsection gives a mathematically complete account of the vortex and antivortex structures, but it is helpful to translate the same statement to the more tangible and experimentally measurable scheme according to
(10) |
By direct substitution we obtain the commutator
(11) |
which is a product of transition vectors over the six domain walls in succession. In the same scheme we have the antivortex commutator
(12) |
Any cyclic permutation of the six letters gives rise to the equivalent vortex or antivortex.
While the sign of the exponent in , , and operators can be obtained considering the order of the neighboring AB and BA domains by combining the DF TEM images with the first and second order Bragg peaks as shown in Fig. 2, there is a simpler scheme to assign the sign of the operators, considering AB/BA domains are always complementary. For example, if the six dislocation lines converging on an AA node appear, for instance, in the order of while going clockwise around it, it ought to be interpreted as given in Eq. (11) and classified as a vortex. If the colors appear as , it is an anti-vortex according to Eq. (12). One only needs to keep in mind that the sequence of colors is to be understood as one color letter followed by the inverse of another color letter, and vice versa. Generalizations of the RGB scheme to vortex-antivortex dipole and/or vortex quadrupole structure are discussed in the Appendix B.
IV Experimental observation of antivortices and strain
IV.1 Detection of antivortices
While moiré patterns and commensurated domain systems with vortex-type nodes have been studied extensively, those with antivortex-type nodes have not been demonstrated. We postulate this is due to the energy required to maintain sufficient strain, whereas twist and lattice constant mismatch can create vortex-type moiré without global strain. Nonetheless, we observe a line of antivortex nodes along a boundary of non-uniformly strained moiré superlattice.
Fig. 7(a) shows combined DF TEM images of a twisted bilayer graphene sample that contains a 1 m sized bubble formed underneath the sample. Near the boundary of the bubble, non-uniform relative strain builds up in the moiré superlattice, which in turn creates the various strain components discussed in Eq. 5. The antivortices (vortices) can be identified by the () order in which the dislocations occur in a clockwise loop. The antivortices, which form along the top edge of a closed-loop dislocation, are each capable of annihilating with a nearby vortex, keeping the net winding number constant within a fixed-boundary region. In Fig. 7(b), loop A surrounds a vortex-antivortex pair, with . Loop D, which surrounds the entire closed-loop dislocation, also has net as the entire feature could annihilate if the local strain were removed, in which case it would become similar to loop E. When circling a single antivortex (B) or vortex (C), the winding number is nonzero.

IV.2 Strain mapping
Existence of antivortices is a measure of the fact that anisotropic strains (uniaxial and shear) are dominating over the isotropic and twist components. Anisotropic strains alter the band structure and can produce pseudomagnetic fields [30]. We can quantify the various strain components from the displacement gradient matrix. Computation of the displacement gradient matrix from a DF TEM image is discussed in Appendix C. In brief, one component of the order parameter is known at every colored line, and an elastic model is used to interpolate in between, obtaining an estimate of the order parameter in the continuum. The displacement gradient matrix can be obtained by differentiating.
The density of vortices or antivortices is computed in Fig. 7(c), by taking the determinant of the displacement gradient matrix, , which by the definition in Eq. 5 is equal to in terms of the strain and twist components. If , vortices are present and if antivortices are present (see Appendix C). Thus, if the isotropic components (twist and isotropic scaling) outweigh the anisotropic components (shear and uniaxial strain), vortices form, and if the opposite, antivortices form.
We further use our estimated displacement gradient matrix to create strain maps of the three strain components, plus twist. Note that the moiré pattern only provides information on the heterostrain, or difference in strain between the two lattices. Unlike other methods to estimate heterostrain from the spatial structure of a moiré pattern [31], this DF TEM method includes knowledge of the lattice orientation and Burgers vector information, avoiding the need to make additional assumptions. Still, the need to interpolate within the moiré cell means that information smaller than the moiré scale is not deterministic from the data. In Fig. 8, strain maps are shown for a heterostructure of MoSe2 and WSe2. The intrinsic lattice constant mismatch of 0.3% should show up in the isotropic component. However, isotropic mismatch smaller than 0.3% is measured, indicating that the lattice globally strains to achieve closer to epitaxial matching, whereas global lattice mismatch is often assumed to be fixed when calculating moiré lengths. This type of strain mapping can reveal useful information about the phenomenology of moiré materials.

V Conclusion

In conclusion, we have presented a general and rigorous approach to describing the topology of nodes formed in moiré materials. Vortex and antivortex are described as the commutator and its inverse of the free group on generators and . The two generators have an intuitive geometric interpretation as two distinct ways by which to make a transition from the AB to BA stacking order, and then back to AB. High-quality TEM measurements are then utilized to represent the abstract generators in terms of colors of domain walls, leading to a dictionary by which to infer the vortex content inside a given boundary. This dictionary relies upon the order in which the colored domain walls cross the boundary. The idea is schematically illustrated in Fig. 9.
We discover an antivortex-type node and present a DF TEM based method for characterizing the type of node and strain field, which does not rely on the usual assumption that the dominant component creating the moiré pattern is twist. This opens the door for the design and characterization of moiré materials based on anisotropic strain fields.
VI Acknowledgement
We thank Frans Spaepen, David Nelson, S.-W. Cheong for important discussions. P.K. acknowledges support from ARO MURI (W911NF-21-2-0147). R.E. acknowledges support from NSF DMR-1922172 for TEM analysis. H.Y. acknowledges the support by the National Research Foundation (NRF) grant funded by the Korean government (MSIT) (No.2021R1C1C1010924). S.C. acknowledges support from NSF Grant No. OIA-1921199. P.C. acknowledges support from NSF DMS Award No. 1819220. ML acknowledges support from NSF DMREF Award No. 1922165. J.H. was supported by NRF-2019R1A6A1A10073079 and by EPIQS Moore theory centers at MIT and Harvard. He acknowledges the help of Manhyung Han for drawing the figures in the Appendix and the discussion with Harry Baik on free group.
Appendix A Vortex algebra
We have avoided explicit use of the language of group theory in the main body of the paper. The two generators and their commutators , were introduced through physical motivation. Here we provide more in-depth discussion and generalizations based on the theory of the free group.
The fundamental group of the punctured torus is , the free group on two generators [28, 29]. The and generators correspond to the two independent ways in which one can encircle the torus. In an ordinary torus the two operations and do commute (Abelian), and the only elements of the fundamental group of the torus are , which count the number of loops in both directions. For a punctured torus such commutativity is lost, and consequently the group structure becomes non-Abelian.
Elements of the free group consist of every conceivable sequence of “letters” such as called “words”. Keep in mind that both letters and have specific geometric moves associated with them. At this point it is helpful to go over well-established theorems in free groups to guide our thinking.
We now use to denote the original free group . Given two elements , the commutator is denoted . The “lower central series” of the free group can be defined as follows. One begins with which is the original free group, then is the subgroup of consisting of all commutators and their products, i.e., all elements of the form
The subgroup is also a normal subgroup, meaning that the quotient space is a group.
Now one can proceed inductively and define
the subgroup generated by all elements of the form where and . It is easy to check from the definition that
Much like the study of van der Waals materials, such “filtration” gives a nice way to study a free group structure ‘one layer at a time’!
Some facts that are worth noting about the lower central series are summarized:
-
1.
Any element of can be uniquely written with and .
-
2.
Any element of can be uniquely written where and .
-
3.
Any element of can be uniquely written with and .
-
4.
Any element of can be uniquely written with and .
By putting all of the above statements together, one sees that any element in the free group can be uniquely written as
(13) |
and so on. In general, each is a normal subgroup, and , meaning the quotient group is isomorphic to a product of integer groups . The number of generators is for a given quotient group . Although the free group itself is non-Abelian, the quotient group is Abelian, characterized by a set of integers. These integers then go on to play the role of topological quantum numbers in physical contexts.

Elements of the free group (13) for which refer to closed loops in real-space graphical representation. It is clear that these are the only elements of that we are interested in. Elements for which (an identity) and are with nonzero . These are the elements of the quotient group and represent the vortices () and antivortices in physical contexts.
To consider higher-order topological defects, consider elements for which and all integers in Eq. (13) equal to zero except :
(14) |
Pictorial representations for the double commutators are easily obtained by tracing out paths according to definitions of and given in Fig. 6. We encourage readers to perform such exercises themselves and arrive at their graphical representations shown in Fig. 10. They are precisely the graphical representation of vortex-antivortex pairs (vortex dipoles) lying along the two crystallographic directions of the triangular lattice.

Next in line is the description of vortex quadrupole structure as triple commutators. According to Eq. (13), there are only three generators of the quotient group . How does one know there are only three generators at this level of filtration?
There is a theorem that gives the number of generators ) at each level through the formula
(15) |
In this formula the sum runs over all divisors of the given integer . For a free group with only two generators we have on the left side of the equation. To see how the formula works with , first set to find . It means that the quotient group has two generators, namely and . At we have or , hence there is only one generator of which is the commutator . At we have , and is the number of generators for , namely and . Finally, at we get , and is the number of generators of given by . It is an arduous, but fun exercise to draw the real-space paths corresponding to each of the triple commutator. The results are the three distinct vortex quadrupole configurations in real space shown in Fig. 11.
Appendix B RGB scheme for higher-order vortices
In Sec. III we discussed ways to characterize a vorticity in terms of the RGB color scheme. A similar RGB scheme to characterize various higher-order vortex structures can be developed.

Fig. 12 shows the vortex dipole and quadrupole configurations in terms of intersecting RGB loops. A small circle drawn around each intersection can determine the vorticity of that point. For instance the filled (empty) circle round the top (bottom) intersection in Fig. 12(a) reads the product of letters () going counter-clockwise, corresponding to a vortex (an antivortex). Vortex quadrupole construction is done by having the three RGB loops intersect at four different points, as shown in Fig. 12(b). In both cases, a large circle drawn far away from the loops fails to cross any of the RGB lines.

It is possible to construct examples of vortex dipole configurations with loops extending out to infinity (hence crossing an arbitrary large circle) as in Fig. 13(a) and (b). The group elements assigned to each configuration can be calculated straightforwardly, leading to and for the left and right configurations, respectively.
This kind of scheme is applicable to experimental situations. Draw a large loop enclosing a given TEM image in the manner shown in Fig. 9, then start counting the dislocation lines according to the scheme. The RGB-based words can be converted to the -letter scheme with the help of the dictionary given in Eq. (10). The -based word can subsequently be converted to various commutators and higher-order commutators until it is cast in the general form given in Eq. (13), from which the total vorticity, dipole numbers, quadrupole numbers and so on can be read off.
Appendix C TEM image processing
TEM DF images were taken on a JEOL 2010F microscope, with 80kV accelerating voltage. A m objective aperture was used to form the dark field images. RGB-colored composite images were formed using Adobe Photoshop blending modes from three distinct second order images of a given sample region.
To create strain maps in Fig. 8, we performed image processing using Python programming codes on images of red, blue and green lines, such as in Fig. 7(b), to prepare to interpolate shift vector values. Ideally an algorithm could be created to extract the lines straight from TEM images but we skip that step for now, and manually trace the red, green, and blue lines to create an “ideal” (i.e. noiseless), but still raster, image. It is then necessary to split all lines into individual line segments and points of intersections, which we call nodes. First, using the connectedComponents function in Python’s OpenCV library, nodes are found by searching for places where red, green, and blue pixels coincide. In a similar manner, pixels that belong to each red, green, or blue line are grouped into lists after dilating them to ensure continuity. We include each line in a dictionary that describes the color of the line and what nodes are part of it. Next, a circle centered around each node is removed from the image (set to R,G,B = 0,0,0), effectively breaking the lines into line segments. Again, the pixels of each segment are found and a dictionary is created for each line segment. The parent line of each segment is identified, as well as the nodes that are its endpoints.
Next, each node must be assigned three integer values corresponding to the coefficients of and , where is the lattice vector associated with the red line, and so on. An origin node is picked arbitrarily and assigned the coefficients . Then, a red segment adjacent to the starting node is chosen to reach a second node, which is assigned . Now that a second node is assigned, the choices are not arbitrary because the direction of increase for each vector has been determined. Note that it is the case that at every node, because . We also know that is constant on red lines, on green lines, etc. From each node, we move to its neighboring nodes (those connected by a segment), and use these properties to fill in the rest of the coefficients. A nice property of this manner of assigning coefficients is it does not need to be told whether a given point is a vortex or an antivortex. Lastly, using the csap and scipy libraries, we fit B-splines to each line segment while forcing the spline to pass through the nodes of the segment, and then use the knots of the splines to generate a Gmsh mesh file.
The next part of the computation is done in the Julia programming language. The vector values of and (measured from a diffraction pattern) must be input ( can be found from the other two). Recall that the vectors in the image plane should appear in the order of RGB while going clockwise to correctly distinguish vortices and antivortices. Coefficients are then attributed to points on lines of the mesh and interpolated via an elastic model using the Gridap library [32] and its interface with Gmsh [33]. The elastic model applies a cost to a large derivative in the -field, as well as a cost to deviating from the known values on the lines. The resulting -field is differentiated to get strain components. The values are defined on a mesh that is small compared to the moiré length. For plotting, the mesh values are interpolated onto a grid in the Matlab software package.
In addition to spatially mapping each strain component, knowledge of the displacement gradient matrix can be used to map the density of vortices and antivortices. Antivortices are distinguished from vortices by the chirality of the rotation in configuration space as you make a loop in real space. To quantify the chirality, we can compare the sign of the cross product of a pair of real-space vectors to their corresponding vectors in configuration space. Consider the real space cartesian vectors and where is positive. They correspond to and in configuration space, by Eq. 3. If the sign of the cross product in configuration space is also positive, it is a vortex. If negative, it is an antivortex.
Given that and are basis vectors and the matrix
. Thus the condition for a vortex is det and for antivortex is det.
Writing in terms of the strain components, the condition is
(16) |
References
- Ponomarenko et al. [2013] L. A. Ponomarenko, R. V. Gorbachev, G. L. Yu, D. C. Elias, R. Jalil, A. A. Patel, A. Mishchenko, A. S. Mayorov, C. R. Woods, J. R. Wallbank, M. Mucha-Kruczynski, B. A. Piot, M. Potemski, I. V. Grigorieva, K. S. Novoselov, F. Guinea, V. I. Falko, and A. K. Geim, Cloning of dirac fermions in graphene superlattices, Nature 497, 594 (2013).
- Dean et al. [2013] C. R. Dean, L. Wang, P. Maher, C. Forsythe, F. Ghahari, Y.Gao, J. Katoch, M. Ishigami, P. Moon, M. Koshino, T. Taniguchi, K. Watanabe, K. L. Shepard, J.Hone, and P. Kim, Hofstadter’s butterfly and the fractal quantum hall effect in moire superlattices, Nature 497, 598 (2013).
- Hunt et al. [2013] B. Hunt, J. D. Sanchez-Yamagishi, A. F. Young, M. Yankowitz, B. J. LeRoy, K. Watanabe, T. Taniguchi, P. Moon, M. Koshino, P. Jarillo-Herrero, and R. C. Ashoori, Massive dirac fermions and hofstadter butterfly in a van der waals heterostructure, Science 340, 1427 (2013).
- Gorbachev et al. [2014] R. V. Gorbachev, J. C. W. Song, G. L. Yu, A. V. Kretinin, F. Withers, Y. Cao, A. Mishchenko, I. V. Grigorieva, K. S. Novoselov, L. S. Levitov, and A. K. Geim, Detecting topological currents in graphene superlattices, Science 346, 448 (2014).
- Endo et al. [2019] K. Endo, K. Komatsu, T. Iwasaki, E. Watanabe, D. Tsuya, K. Watanabe, T. Taniguchi, Y. Noguchi, Y. Wakayama, Y. Morita, and S. Moriyama, Topological valley currents in bilayer graphene/hexagonal boron nitride superlattices, Appl. Phys. Lett. 114 (2019).
- Cao et al. [2018a] Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, R. C. Ashoori, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 80 (2018a).
- Cao et al. [2018b] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Correlated insulator behaviour at half-filling in magic-angle graphene superlattices, Nature 556, 43 (2018b).
- Liu et al. [2020] X. Liu, Z. Hao, E. Khalaf, J. Y. Lee, Y. Ronen, H. Yoo, D. Haei Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath, and P. Kim, Tunable spin-polarized correlated states in twisted double bilayer graphene, Nature 583, 221 (2020).
- Zhang et al. [2021] X. Zhang, K.-T. Tsai, Z. Zhu, W. Ren, Y. Luo, S. Carr, M. Luskin, E. Kaxiras, and K. Wang, Correlated insulating states and transport signature of superconductivity in twisted trilayer graphene superlattices, Physical Review Letters 127 (2021).
- Hao et al. [2021] Z. Hao, A. M. Zimmerman, P. Ledwith, E. Khalaf, D. H. Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath, and P. Kim, Electric field tunable superconductivity in alternating-twist magic-angle trilayer graphene, Science 371, 1133 (2021).
- Park et al. [2021] J. M. Park, Y. Cao, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Tunable strongly coupled superconductivity in magic-angle twisted trilayer graphene, Nature 590, 249 (2021).
- Park et al. [2022] J. M. Park, Y. Cao, L.-Q. Xia, S. Sun, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Robust superconductivity in magic-angle multilayer graphene family, Nature Materials , https://doi.org/10.1038/s41563 (2022).
- Woods et al. [2021] C. R. Woods, P. Ares, H. Nevison-Andrews, M. J. Holwill, R. Fabregas, F. Guinea, A. K. Geim, K. S. Novoselov, N. R. Walet, and L. Fumagalli, Charge-polarized interfacial superlattices in marginally twisted hexagonal boron nitride, Nature Communications 12 (2021).
- Wang et al. [2022] X. Wang, K. Yasuda, Y. Zhang, S. Liu, K. Watanabe, T. Taniguchi, J. H. ad Liang Fu, and P. Jarillo-Herrero, Interfacial ferroelectricity in rhombohedral stacked bilayer transition metal dichalcogenides, Nature Nanotechnology 17, 367 (2022).
- Lau et al. [2022] C. N. Lau, M. W. Bockrath, K. F. Mak, and F. Zhang, Reproducibility in the fabrication and physics of moiré materials, Nature 602, 41 (2022).
- Uri et al. [2020] A. Uri, S. Grover, Y. Cao, J. A. Crosse, K. Bagani, D. Rodan-Legrain, Y. Myasoedov, K. Watanabe, T. Taniguchi, P. Moon, M. Koshino, P. Jarillo-Herrero, and E. Zeldov, Mapping the twist-angle disorder and landau levels in magic-angle graphene, Nature 581, 47 (2020).
- Wilson et al. [2020] J. H. Wilson, Y. Fu, S. Das Sarma, and J. H. Pixley, Disorder in twisted bilayer graphene, Phys. Rev. Research 2 (2020).
- Bai et al. [2020] Y. Bai, L. Zhou, J. Wang, W. Wu, L. J. McGilly, D. Halbertal, C. F. B. Lo, F. Liu, J. Ardelean, P. Rivera, N. R. Finney, X.-C. Yang, D. N. Basov, W. Yao, X. Xu, J. Hone, A. N. Pasupathy, and X.-Y. Zhu, Excitons in strain-induced one-dimensional moiré potentials at transition metal dichalcogenide heterojunctions, Nat. Mater. 10, 1124 (2020).
- Yoo et al. [2019] H. Yoo, R. Engelke, S. Carr, S. Fang, K. Zhang, P. Cazeaux, S. H. Sung, R. Hovden, A. W. Tsen, T. Taniguchi, K. Watanabe, G.-C. Yi, M. Kim, M. Luskin, E. B. Tadmor, E. Kaxiras, and P. Kim, Atomic and electronic reconstruction at the van der waals interface in twisted bilayer graphene, Nature Materials 18, 448 (2019).
- Alden et al. [2013] J. S. Alden, A. W. Tsen, P. Y. Huang, R. Hovden, L. Brown, J. Park, D. A. Muller, and P. L. McEuen, Strain solitons and topological defects in bilayer graphene, Proceedings of the National Academy of Sciences of the USA 110, 11256 (2013).
- Turkel et al. [2022] S. Turkel, J. Swann, Z. Zhu, M. Christos, K. Watanabe, T. Taniguchi, S. Sachdev, M. S. Scheurer, E. Kaxiras, C. R. Dean, and A. N. Pasupathy, Orderly disorder in magic-angle twisted trilayer graphene, Science 376, 193 (2022).
- Yu et al. [2020] Y. Yu, K. Zhang, H. Parks, M. Babar, S. C. andIsaac M. Craig, M. V. Winkle, A. Lyssenko, T. Taniguchi, K. Watanabe, V. Viswanathan, and D. K. Bediako, Tunable angle-dependent electrochemistry at twisted bilayer graphene with moiré flat bands, Nature Chemistry 14, 267 (2020).
- Thouless [1998] D. Thouless, Topological quantum numbers in nonrelativistic physics (World Scientific, 1998).
- Mermin [1979] N. D. Mermin, The topological theory of defects in ordered media, Rev. Mod. Phys. 51, 591 (1979).
- Carr et al. [2018] S. Carr, D. Massatt, S. B. Torrisi, P. Cazeaux, M. Luskin, and E. Kaxiras, Relaxation and domain formation in incommensurate two-dimensional heterostructures, Physical Review B 98 (2018).
- Lin et al. [2013] J. Lin, W. Fang, W. Zhou, A. R. Lupin, J. C. Idrobo, J. Kong, S. J. Pennycook, , and S. T. Pantelide, AC/AB stacking boundaries in bilayer graphene, Nano Letters 13, 3262 (2013).
- Friedrichs [1947] K. O. Friedrichs, On the boundary-value problems of the theory of elasticity and korn’s inequality, Annals of Mathematics 48, 441 (1947).
- Cohen [1989] D. E. Cohen, Combinatorial Group Theory: A Topological Approach (London Mathematical Society Student Texts, Series Number 14) (Cambridge University Press, 1989).
- Kharlampovich and Myasnikov [2006] O. Kharlampovich and A. Myasnikov, Elementary theory of free non-abelian groups, Journal of Algebra 302, 451 (2006).
- Guinea et al. [2010] F. Guinea, M. I. Katsnelson, and A. K. Geim, Energy gaps and a zero-field quantum hall effect in graphene by strain engineering, Nature Physics 6, 30 (2010).
- Halbertal et al. [2022] D. Halbertal, S. Shabani, A. N. Passupathy, , and D. N. Basov, Moiré metrology of energy landscapes in van der waals heterostructures, ACS Nano 16, 1471 (2022).
- Badia and Verdugo [2020] S. Badia and F. Verdugo, Gridap: An extensible finite element toolbox in julia, Journal of Open Source Software 5, 2520 (2020).
- Geuzaine and Remacle [2009] C. Geuzaine and J.-F. Remacle, Gmsh: A 3-D finite element mesh generator with built-in pre- and post-processing facilities: THE GMSH PAPER, International Journal on Numerical Methods in Engineering 79, 1309 (2009).