Topological Dynamics of Synthetic Molecules

Yuming Zhu Bronx High School of Science,
Bronx, NY 10016, USA
[email protected]
 and  Emil Prodan Department of Physics and
Department of Mathematical Sciences
Yeshiva University
New York, NY 10016, USA
[email protected]
(Date: November 18, 2024)
Abstract.

We study the dynamics of synthetic molecules whose architectures are generated by space transformations from a point group acting on seed resonators. We show that the dynamical matrix of any such molecule can be reproduced as the left regular representation of a self-adjoint element from the stabilized group’s algebra. Furthermore, we use elements of representation theory and K-theory to rationalize the dynamical features supported by such synthetic molecules up to topological equivalences. These tools enable us to identify a set of fundamental models which generate by superposition all possible dynamical matrices up to homotopy equivalences. Interpolations between these fundamental models give rise to topological spectral flows.

This work was supported by the U.S. National Science Foundation through the grant CMMI-2131760 and by U.S. Army Research Office through contract W911NF-23-1-0127.

1. Introduction

The discovery of topological insulators [1, 2, 3, 4, 5, 6, 7, 8] and of topological photonic and mechanical systems [9, 10, 11, 12, 13, 14, 15, 16], together with their complete classification [17, 18, 19, 20], have permanently changed how research and discovery are conducted in materials science. For example, the focus has shifted from enhancing the properties of a material to making it different. Furthermore, many applications rely now on robust effects that do not require fine tuning, such as generating wave channels at the interface between two topologically distinct materials, edge-to-edge topological pumping, or closing and opening of the resonant gaps by interpolating between two distinct material configurations. The latter application is called topological spectral engineering and we will see it at work in the present study. Besides supplying the means for important applications, observing a topological spectral flow is the simplest way to assess if two metamaterials are topologically distinct (see e.g [21]). These aforementioned robust effects are applicable to information and sensing technologies, where the main function of a device is to switch between on and off states as robustly as possible.

Physical models in condensed matter physics are developed on periodic lattices {\mathcal{L}}caligraphic_L and most of the time they take the form of tight-binding Hamiltonians

H=x,ytxy|xy|+txy|yx|,𝐻subscript𝑥𝑦tensor-productsubscript𝑡𝑥𝑦ket𝑥bra𝑦tensor-productsuperscriptsubscript𝑡𝑥𝑦ket𝑦bra𝑥H=\sum_{x,y\in{\mathcal{L}}}t_{xy}\otimes|x\rangle\langle y|+t_{xy}^{\ast}% \otimes|y\rangle\langle x|,italic_H = ∑ start_POSTSUBSCRIPT italic_x , italic_y ∈ caligraphic_L end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ⊗ | italic_x ⟩ ⟨ italic_y | + italic_t start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ⊗ | italic_y ⟩ ⟨ italic_x | , (1.1)

where the intuition is that an electron hops between the orbitals supported by the sites of the lattice, with probabilities encoded in the matrices txysubscript𝑡𝑥𝑦t_{xy}italic_t start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT. In particular, the Hamiltonians of the established topological insulators are of this form (see [20] for complete list). If one moves away from this context, unless specialized mathematical tools are utilized, it is not clear what a topological classification might be, how to write all representative Hamiltonians, and how to compute topological invariants. For example, statements like “two spectrally gappped Hamiltonians are topologically equivalent if we can deform one into the other without closing the gap” are without value if the space in which the deformations occur is not specified. Imposing symmetry constraints on the Hamiltonians is still not enough because locality conditions on the hopping amplitudes are also needed. In a remarkable paper [22], Jean Bellissard realized that the full set of constraints can be communicated by simply stating that the Hamiltonians belong to a particular Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra. The isomorphism class of this Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra identifies a specific class of materials and a concrete realization of it supplies the topological space for the deformations. These statements of Bellissard are not purely formal because, in the same paper, he showed how to construct such Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebras (see also [23]). Once the deformation space is identified, then various flavors of stable homotopy theories, aka K-theories, supply the devices that answer the classification problems.

To the experts, the above statements and their implications are very clear, but this is not at all the case for a broader community of materials scientists, despite of many worked examples available in the literature. The authors believe that this current state of affairs is due to the fact that since these mentioned applications targeted challenging situations, there was no room for pedagogical expositions. As a result, the connections between the K-theoretic methods and other more familiar methods have not been revealed explicitly enough. The purpose of this expository paper is to fill in some of the mentioned shortcomings. The plan is to engage simple synthetic systems built from a finite number of self-coupling resonators and whose architecture is generated using the action of a finite subgroup of the full Euclidean group, hence, a point group (see section 2). The dynamical matrices determining the dynamical linear regimes of such systems all fall into the group algebra of the point group. As we shall see, we are in a situation where the representation theory of this group and the K-theory of its algebra overlap quite strongly (see section 4). Thus, we are in a special situation where the K-theoretic tools can be seen at work through the prism of representation theory. For a community familiar with the latter, this class of physical systems and their analysis can be used as a door into the world of the K-theoretic ideas, which divert from representation theory as soon as the number of degrees of freedom become infinite.

The exposition is organized as follows. Section 2 describes quantum and classical materials generated from actions of the Euclidean group on seeding synthetic atoms (such as quantum dots) or self-coupling classical resonators (such as solid shapes fitted with magnets), respectively. We use the terms synthetic molecules to refer to both these systems. Using simple physical reasoning, we calculate the algebra of the Hamiltonians. In the case where all the actions come from a subgroup of the Euclidean group, we find that this algebra drops to the group algebra of the space transformations. This, together with a natural topology on the algebra, completes the task of finding the deformation space for this particular class of synthetic molecules. Section 4 explains the interplay between representation theory and K-theory and demonstrates how numerical invariants are computed in general and in particular for setting of synthetic molecules. Section 5 builds a set of Hamiltonians derived from K-theoretic data. Any other Hamiltonian can be deformed to a direct sum of Hamiltonians from this list. Furthermore, we demonstrate that interpolations between pairs of Hamiltonians from the list result in topological spectral flows, hence confirming that they are topologically distinct.

2. Synthetic molecules

2.1. Building architectures with space transformations

Refer to caption
Figure 2.1. A resonator design that leads to a stable equilibrium configuration symmetric to the icosahedron group.

For simplicity, we describe these synthetic materials in a classical mechanical setting, but all our predictions apply in the acoustic, photonic and quantum settings as well. The synthetic mechanical materials that we have in mind are built from many copies of a single seed resonator. The seed resonator consists of moving parts, a physical frame and sources of force fields, such that:

  • The frame of the seed resonator can be decorated with sources of force fields, e.g. magnets, in order to adjust its internal resonant modes and frequencies;

  • The frame of the seed resonator can also be fitted with sensors in order to quantify the motion of its moving parts.

  • The frames of the copies111A copy of the seed resonator will include the frame, the moving parts and all the other fittings. can be rigidly anchored once their positions and orientations are decided;

  • The moving parts of the identical copies of the seed resonators self-couple via the force fields generated by the attached sources of force fields.

A schematic example of such synthetic material is shown in Fig. 2.1, and an actual laboratory model is shown and described in Fig. 2.2. We carefully listed our assumptions because they have several important consequences:

  • The physical frames and their fitted sensors supply local reference frames, marked by the red arrows in Fig. 2.1, from where the motion is quantified. Thus, if {qα}subscript𝑞𝛼\{q_{\alpha}\}{ italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT } are the generalized degrees of freedom of the seed resonator and qi(t)subscript𝑞𝑖𝑡q_{i}(t)italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) are their recorded values as function of time in an experiment, then same recordings will be reported if same experiment is carried out with the seed resonator placed at a different location and with different orientation.

  • The collective motion of the self-coupled moving parts is fully determined by the positions and orientations of the frames.

  • The motion, as quantified, or better said reported from the local frames, is unaffected by rigid translations and rotations of the whole structure, which is a consequence of Galilean invariance of the physical laws involved in the couplings.

We will be dealing only with linear dynamical regimes, which requires the existence of stable equilibrium configurations. Thus, we assume:

  • For each spatial configuration of the frames of the resonators, there is a unique lowest energy configuration.

This assumption needs to be true only for the set of configurations sampled during an experiment or for the configurations engaged during an application.

Refer to caption
Figure 2.2. Example of a fully assembled and functional synthetic molecule.

A somewhat trivial observation supplies the key to the operator algebraic approach we want to describe. Specifically, the position and orientation of the frame of a resonator can be set by a rotation followed by a translation of the seed resonator. These two space transformations define an element of the Euclidean group, which transforms the local reference frame of the seed resonator222All identical copies are generated from a seed resonator that is fixed once and for all at a specific location and orientation in space. into the local reference frame of an actual resonator. As such, there is a one-to-one correspondence between a resonator and an element of the Euclidean group. We reached our first important conclusion:

Proposition 2.1 ([23]).

Let us use the terms “synthetic molecules” for clusters of self-coupled resonators and let us define the architecture of the synthetic molecule to be the information contained in the positions and orientations of the frames of the resonators. Then the architecture can be conceptualized as a discrete subset or better said a lattice {\mathcal{L}}caligraphic_L of the Euclidean group of space transformations.

Remark 2.2.

The Euclidean group and its subgroup of rotations will play active and passive roles in our discussion. For example, these groups will supply space transformations on {\mathcal{L}}caligraphic_L, but will also supply the labels (or the coordinates) for the points in {\mathcal{L}}caligraphic_L. When appearing with an active role, the elements of the groups are indicated by g𝑔gitalic_g, and otherwise by x𝑥xitalic_x. \Diamond

We recall that the Euclidean group 𝑬𝑬\bm{E}bold_italic_E is a topological group and that, for the Euclidean space with 3-dimensions, the underlying topological space is

𝑬3×3×{1,1}.similar-to-or-equals𝑬superscript3superscript311\bm{E}\simeq{\mathbb{R}}^{3}\times{\mathbb{R}}\mathbb{P}^{3}\times\{-1,1\}.bold_italic_E ≃ blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT × blackboard_R blackboard_P start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT × { - 1 , 1 } . (2.1)

Note that 3×{1,1}superscript311{\mathbb{R}}\mathbb{P}^{3}\times\{-1,1\}blackboard_R blackboard_P start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT × { - 1 , 1 } is the topological space underlying the subgroup O(3)𝑂3O(3)italic_O ( 3 ) of rotations, with ±1plus-or-minus1\pm 1± 1 indicating if a rotation is proper or improper, respectively. According to our statement from above, the architecture of a synthetic molecule containing N𝑁Nitalic_N resonators can be quantified as a subset of N𝑁Nitalic_N distinct points on the space (2.1).

Remark 2.3.

Since we will put strong emphasis on synthetic molecules produced by a finite subgroup of O(3)𝑂3O(3)italic_O ( 3 ), we supply further details about this group. If a point of the 3-dimensional space is specified by a column vector with three entries encoding its coordinates, then the rotation transformations take the form of a matrix. Any proper rotation takes the form of a 3×3333\times 33 × 3 matrix of the form

R(n^,θ)=exp(θn^L)𝑅^𝑛𝜃𝜃^𝑛𝐿R(\hat{n},\theta)=\exp(\theta\,\hat{n}\cdot\vec{L})italic_R ( over^ start_ARG italic_n end_ARG , italic_θ ) = roman_exp ( italic_θ over^ start_ARG italic_n end_ARG ⋅ over→ start_ARG italic_L end_ARG ) (2.2)

where θ𝜃\thetaitalic_θ is an angle, n^=(nx,ny,nz)^𝑛subscript𝑛𝑥subscript𝑛𝑦subscript𝑛𝑧\hat{n}=(n_{x},n_{y},n_{z})over^ start_ARG italic_n end_ARG = ( italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) is 3-component vector of norm one and L𝐿\vec{L}over→ start_ARG italic_L end_ARG is the 3-component vector with entries

Lx=(000001010),Ly=(001000100),Lz=(010100000).formulae-sequencesubscript𝐿𝑥matrix000001010formulae-sequencesubscript𝐿𝑦matrix001000100subscript𝐿𝑧matrix010100000L_{x}={\small\begin{pmatrix}0&0&0\\ 0&0&-1\\ 0&1&0\end{pmatrix}},\ L_{y}={\small\begin{pmatrix}0&0&1\\ 0&0&0\\ -1&0&0\end{pmatrix}},\ L_{z}={\small\begin{pmatrix}0&-1&0\\ 1&0&0\\ 0&0&0\end{pmatrix}}.italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - 1 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ) , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL - 1 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ) , italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL - 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ) . (2.3)

Up to a factor, L𝐿\vec{L}over→ start_ARG italic_L end_ARG coincides with the angular momentum in quantum mechanics for a particle with spin 1. The somewhat complicated space 3superscript3{\mathbb{R}}\mathbb{P}^{3}blackboard_R blackboard_P start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is a result of the allowed ranges for the pair (θ,n^)𝜃^𝑛(\theta,\hat{n})( italic_θ , over^ start_ARG italic_n end_ARG ) or, equivalently, for the vector nθ=θn^subscript𝑛𝜃𝜃^𝑛\vec{n}_{\theta}=\theta\hat{n}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = italic_θ over^ start_ARG italic_n end_ARG. For nθ<πnormsubscript𝑛𝜃𝜋\|\vec{n}_{\theta}\|<\pi∥ over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∥ < italic_π, (2.2) supplies distinct rotations, but for nθ=πnormsubscript𝑛𝜃𝜋\|\vec{n}_{\theta}\|=\pi∥ over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∥ = italic_π, (2.2) supplies the same rotation if plug in nθsubscript𝑛𝜃\vec{n}_{\theta}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT and nθsubscript𝑛𝜃-\vec{n}_{\theta}- over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT. Thus, rotations are fully parameterized by the points of a 3-dimensional ball of radius π𝜋\piitalic_π if the opposite points at its surface are identified. The resulting space is diffeomorphic to 3superscript3{\mathbb{R}}\mathbb{P}^{3}blackboard_R blackboard_P start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, the space of lines in 4superscript4{\mathbb{R}}^{4}blackboard_R start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT passing through the origin. \Diamond

Refer to caption
Figure 2.3. Architectures generated by acting with the proper icosahedral group on a seeding shape. The difference between panels (a) and (b) is the location of the seeding shape.

So far, the discussion was about formalizing the architecture of given material. However, we can reverse the arrows and use Proposition 2.1 to generate interesting architectures. In Fig. 2.3, we show two architectures generated by acting with the space transformations contained in the proper icosahedral point group on a seed resonator placed at two different locations in space. These and other similar examples will be discussed in detail in subsection 2.3.

2.2. Dynamics

We denote by G𝐺Gitalic_G the underlying group which labels the resonators of a synthetic molecule. Thus, G𝐺Gitalic_G can be the full Euclidean group or just its O(3)𝑂3O(3)italic_O ( 3 ) subgroup. The seed resonator is labeled by the identity eG𝑒𝐺e\in Gitalic_e ∈ italic_G and its degrees of freedom are q(e,α)𝑞𝑒𝛼q(e,\alpha)italic_q ( italic_e , italic_α ), α=1,,D𝛼1𝐷\alpha=1,\ldots,Ditalic_α = 1 , … , italic_D, with D𝐷Ditalic_D specifying the number of internal degrees of freedom. Then the degrees of freedom of a resonator labeled by xG𝑥𝐺x\in{\mathcal{L}}\subset Gitalic_x ∈ caligraphic_L ⊂ italic_G in the synthetic molecule are q(x,α)𝑞𝑥𝛼q(x,\alpha)italic_q ( italic_x , italic_α ), α=1,,D𝛼1𝐷\alpha=1,\ldots,Ditalic_α = 1 , … , italic_D. When the system is driven harmonically with pulsation ω𝜔\omegaitalic_ω, under the assumption of very small but finite dissipation, the degrees of freedom display an oscillatory behavior as functions of time t𝑡titalic_t,

q(x,α;t)=QR(x,α)cos(ωt+ϕx,α)=Re[Q(x,α)eıωt],𝑞𝑥𝛼𝑡subscript𝑄𝑅𝑥𝛼𝜔𝑡subscriptitalic-ϕ𝑥𝛼Redelimited-[]𝑄𝑥𝛼superscript𝑒italic-ı𝜔𝑡q(x,\alpha;t)=Q_{R}(x,\alpha)\cos(\omega t+\phi_{x,\alpha})={\rm Re}[Q(x,% \alpha)e^{\imath\omega t}],italic_q ( italic_x , italic_α ; italic_t ) = italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_x , italic_α ) roman_cos ( italic_ω italic_t + italic_ϕ start_POSTSUBSCRIPT italic_x , italic_α end_POSTSUBSCRIPT ) = roman_Re [ italic_Q ( italic_x , italic_α ) italic_e start_POSTSUPERSCRIPT italic_ı italic_ω italic_t end_POSTSUPERSCRIPT ] , (2.4)

where QRsubscript𝑄𝑅Q_{R}\in{\mathbb{R}}italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ∈ blackboard_R are real-valued amplitudes, and Q𝑄Q\in{\mathbb{C}}italic_Q ∈ blackboard_C are complex-valued amplitudes which efficiently incorporate the phases ϕitalic-ϕ\phiitalic_ϕ. We can place the complex amplitudes in one single vector

|Q=xα=1DQ(x,α)|x,α,ket𝑄subscript𝑥superscriptsubscript𝛼1𝐷𝑄𝑥𝛼ket𝑥𝛼|Q\rangle=\sum_{x\in{\mathcal{L}}}\sum_{\alpha=1}^{D}Q(x,\alpha)|x,\alpha\rangle,| italic_Q ⟩ = ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_L end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT italic_Q ( italic_x , italic_α ) | italic_x , italic_α ⟩ , (2.5)

where |x,αket𝑥𝛼|x,\alpha\rangle| italic_x , italic_α ⟩ can be thought of as an abstract basis for the state space, or as a concrete column vector with entry 1 at one place and 0 entry at all other places.333The latter requires an (un-natural) ordering of (x,α)𝑥𝛼(x,\alpha)( italic_x , italic_α )’s and this is why the former viewpoint is preferred. The vectors (2.5) span the Hilbert space 2(,d)D2()similar-to-or-equalssuperscript2superscript𝑑tensor-productsuperscript𝐷superscript2\ell^{2}({\mathcal{L}},{\mathbb{C}}^{d})\simeq{\mathbb{C}}^{D}\otimes\ell^{2}(% {\mathcal{L}})roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L , blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) ≃ blackboard_C start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ⊗ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L ). As a vector in D2()tensor-productsuperscript𝐷superscript2{\mathbb{C}}^{D}\otimes\ell^{2}({\mathcal{L}})blackboard_C start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ⊗ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L ), the dynamical state of the system takes the form |Q=x𝑸x|xket𝑄subscript𝑥tensor-productsubscript𝑸𝑥ket𝑥|Q\rangle=\sum_{x\in{\mathcal{L}}}\bm{Q}_{x}\otimes|x\rangle| italic_Q ⟩ = ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_L end_POSTSUBSCRIPT bold_italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⊗ | italic_x ⟩, with 𝑸xdsubscript𝑸𝑥superscript𝑑\bm{Q}_{x}\in{\mathbb{C}}^{d}bold_italic_Q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

External generalized forces {fg,αeıωt}subscript𝑓𝑔𝛼superscript𝑒italic-ı𝜔𝑡\{f_{g,\alpha}e^{\imath\omega t}\}{ italic_f start_POSTSUBSCRIPT italic_g , italic_α end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_ı italic_ω italic_t end_POSTSUPERSCRIPT } driving the synthetic molecule can be also encoded as a vector from this Hilbert space,

|F=xfx,α|x,α=x𝒇x|x.ket𝐹subscript𝑥subscript𝑓𝑥𝛼ket𝑥𝛼subscript𝑥tensor-productsubscript𝒇𝑥ket𝑥|F\rangle=\sum_{x\in{\mathcal{L}}}f_{x,\alpha}|x,\alpha\rangle=\sum_{x\in{% \mathcal{L}}}\bm{f}_{x}\otimes|x\rangle.| italic_F ⟩ = ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_L end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_x , italic_α end_POSTSUBSCRIPT | italic_x , italic_α ⟩ = ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_L end_POSTSUBSCRIPT bold_italic_f start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⊗ | italic_x ⟩ . (2.6)

If the quadratic form of the Lagrange function is

L(QR,Q˙R)=12Q˙RT𝑻Q˙R+12QRT𝑽QR,𝐿subscript𝑄𝑅subscript˙𝑄𝑅12superscriptsubscript˙𝑄𝑅𝑇𝑻subscript˙𝑄𝑅12superscriptsubscript𝑄𝑅𝑇𝑽subscript𝑄𝑅L(Q_{R},\dot{Q}_{R})=\tfrac{1}{2}\dot{Q}_{R}^{T}\bm{T}\dot{Q}_{R}+\tfrac{1}{2}% Q_{R}^{T}\bm{V}Q_{R},italic_L ( italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , over˙ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG over˙ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_italic_T over˙ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_italic_V italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , (2.7)

where 𝑻𝑻\bm{T}bold_italic_T and 𝑽𝑽\bm{V}bold_italic_V are positive operators on D2()tensor-productsuperscript𝐷superscript2{\mathbb{C}}^{D}\otimes\ell^{2}({\mathcal{L}})blackboard_C start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ⊗ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L ), then the response of the synthetic molecule to the driving forces is given by the solution to the equation

ω2𝑻|Q+𝑽|Q=|F𝑻12(𝑫ω2)𝑻12|Q=|F,formulae-sequencesuperscript𝜔2𝑻ket𝑄𝑽ket𝑄ket𝐹superscript𝑻12𝑫superscript𝜔2superscript𝑻12ket𝑄ket𝐹-\omega^{2}\bm{T}|Q\rangle+\bm{V}|Q\rangle=|F\rangle\ \ \Leftrightarrow\ \ \bm% {T}^{\frac{1}{2}}(\bm{D}-\omega^{2})\bm{T}^{\frac{1}{2}}|Q\rangle=|F\rangle,- italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_italic_T | italic_Q ⟩ + bold_italic_V | italic_Q ⟩ = | italic_F ⟩ ⇔ bold_italic_T start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( bold_italic_D - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) bold_italic_T start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT | italic_Q ⟩ = | italic_F ⟩ , (2.8)

where

𝑫:=𝑻12𝑽𝑻12assign𝑫superscript𝑻12𝑽superscript𝑻12\bm{D}:=\bm{T}^{-\frac{1}{2}}\bm{V}\bm{T}^{-\frac{1}{2}}bold_italic_D := bold_italic_T start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT bold_italic_V bold_italic_T start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT (2.9)

is call the dynamical matrix. The resonant frequencies and the corresponding self-oscillating modes can be obtain from the spectral properties of 𝑫𝑫\bm{D}bold_italic_D, which, as any linear operator on the Hilbert space D2()tensor-productsuperscript𝐷superscript2{\mathbb{C}}^{D}\otimes\ell^{2}({\mathcal{L}})blackboard_C start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ⊗ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L ), it takes the form

𝑫=x,xwx,x()|xx|,𝑫subscript𝑥superscript𝑥tensor-productsubscript𝑤superscript𝑥𝑥ketsuperscript𝑥bra𝑥\bm{D}=\sum_{x,x^{\prime}\in{\mathcal{L}}}w_{x^{\prime},x}({\mathcal{L}})% \otimes|x^{\prime}\rangle\langle x|,bold_italic_D = ∑ start_POSTSUBSCRIPT italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_L end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x end_POSTSUBSCRIPT ( caligraphic_L ) ⊗ | italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ⟨ italic_x | , (2.10)

where wx,xsubscript𝑤superscript𝑥𝑥w_{x^{\prime},x}italic_w start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x end_POSTSUBSCRIPT are D×D𝐷𝐷D\times Ditalic_D × italic_D matrices with complex entries. Often, these are referred to as coupling matrices. Let us point out the similarity between the expressions (2.10) and (1.1), where the only difference is that the lattices live in different topological groups, Euclidean group in the former and 3superscript3{\mathbb{R}}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT in the latter.

The group G𝐺Gitalic_G can act on itself from the left or from the right and we will be interested in both. The left action is supplied by multiplication to the left, while the right action goes as gx=xg1𝑔𝑥𝑥superscript𝑔1g\cdot x=xg^{-1}italic_g ⋅ italic_x = italic_x italic_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, for all g,xG𝑔𝑥𝐺g,x\in Gitalic_g , italic_x ∈ italic_G. These actions are also well defined on subsets of G𝐺Gitalic_G, in particular, on {\mathcal{L}}caligraphic_L. Now, in Eq. (2.10), we were careful to specify that the coupling matrices of the synthetic molecule are entirely determined by the architecture of the molecule, hence by the lattice {\mathcal{L}}caligraphic_L, as it was already explained in the previous subsection. From the principles stated there, we can also infer that the coupling matrices satisfy the following covariant relation,

wgx,gx(g)=wx,x(),subscript𝑤𝑔superscript𝑥𝑔𝑥𝑔subscript𝑤superscript𝑥𝑥w_{g\cdot x^{\prime},g\cdot x}(g\cdot{\mathcal{L}})=w_{x^{\prime},x}({\mathcal% {L}}),italic_w start_POSTSUBSCRIPT italic_g ⋅ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_g ⋅ italic_x end_POSTSUBSCRIPT ( italic_g ⋅ caligraphic_L ) = italic_w start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x end_POSTSUBSCRIPT ( caligraphic_L ) , (2.11)

which is a direct consequence of the Galilean invariance of the physical laws underlining the coupling of the resonators. Taking g=x𝑔𝑥g=xitalic_g = italic_x, one finds

wx,x()=wxx,e(x),subscript𝑤superscript𝑥𝑥subscript𝑤𝑥superscript𝑥𝑒𝑥w_{x^{\prime},x}({\mathcal{L}})=w_{x\cdot x^{\prime},e}(x\cdot{\mathcal{L}}),italic_w start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x end_POSTSUBSCRIPT ( caligraphic_L ) = italic_w start_POSTSUBSCRIPT italic_x ⋅ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_e end_POSTSUBSCRIPT ( italic_x ⋅ caligraphic_L ) , (2.12)

or equivalently,

𝑫=x,xwxx(x)|xx|,𝑫subscript𝑥superscript𝑥tensor-productsubscript𝑤𝑥superscript𝑥𝑥ketsuperscript𝑥bra𝑥\bm{D}=\sum_{x,x^{\prime}\in{\mathcal{L}}}w_{x\cdot x^{\prime}}(x\cdot{% \mathcal{L}})\otimes|x^{\prime}\rangle\langle x|,bold_italic_D = ∑ start_POSTSUBSCRIPT italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_L end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_x ⋅ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_x ⋅ caligraphic_L ) ⊗ | italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ⟨ italic_x | , (2.13)

where we removed the redundant lower index. As shown in [23], this particular form of the dynamical matrices, together with the assumption that the coupling matrices depend continuously on the architecture, enable one to construct the Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra where the models live. More precisely, if one keeps the architecture fixed but changes the internal structure of the seed resonator, the dynamical matrices densely sample a groupoid Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra that is entirely determined by the lattice {\mathcal{L}}caligraphic_L. We discuss next how this Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra emerges in the particular setting.

2.3. Synthetic molecules generated with point groups

The finite subgroups of SO(3)𝑆𝑂3SO(3)italic_S italic_O ( 3 ) are all known and classified [24]. In this subsection, we consider a seed resonator placed at a point in space and with an orientation described by x0𝑬subscript𝑥0𝑬x_{0}\in\bm{E}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ bold_italic_E. Copies of this seed resonator are then acted from the left with space transformations contained in a finite subgroup ΓSO(3)Γ𝑆𝑂3\Gamma\subset SO(3)roman_Γ ⊂ italic_S italic_O ( 3 ). As such, we generate a finite lattice

={γx0,γΓ}.𝛾subscript𝑥0𝛾Γ{\mathcal{L}}=\{\gamma x_{0},\ \gamma\in\Gamma\}.caligraphic_L = { italic_γ italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_γ ∈ roman_Γ } . (2.14)

The seed resonator corresponds to the neutral element eΓ𝑒Γe\in\Gammaitalic_e ∈ roman_Γ and is part of lattice {\mathcal{L}}caligraphic_L.

Remark 2.4.

One important assumption of ours is that the cardinals of {\mathcal{L}}caligraphic_L and ΓΓ\Gammaroman_Γ always coincide: ||=|Γ|Γ|{\mathcal{L}}|=|\Gamma|| caligraphic_L | = | roman_Γ |. This forbids x0subscript𝑥0x_{0}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for being part of the set 𝑬Γsuperscript𝑬Γ\bm{E}^{\Gamma}bold_italic_E start_POSTSUPERSCRIPT roman_Γ end_POSTSUPERSCRIPT of points of 𝑬𝑬\bm{E}bold_italic_E fixed by ΓΓ\Gammaroman_Γ.\Diamond

We want to stress that resulting lattice depends quite sensitively on x0subscript𝑥0x_{0}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, a fact that was already visible in Fig. 2.3. However, regardless of the initial choice of x0𝑬𝑬Γsubscript𝑥0𝑬superscript𝑬Γx_{0}\in\bm{E}\setminus\bm{E}^{\Gamma}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ bold_italic_E ∖ bold_italic_E start_POSTSUPERSCRIPT roman_Γ end_POSTSUPERSCRIPT, we have:

Proposition 2.5.

The sets x𝑥x\cdot{\mathcal{L}}italic_x ⋅ caligraphic_L appearing in Eq. (2.13) are all identical.

Proof.

Before giving the arguments, let us point out that x𝑥x\cdot{\mathcal{L}}italic_x ⋅ caligraphic_L for x𝑥x\in{\mathcal{L}}italic_x ∈ caligraphic_L represents the lattice {\mathcal{L}}caligraphic_L as seen from the local reference frame of the resonator located at x𝑥xitalic_x. Then the statement says that synthetic molecule appears identical when looked at from any of the local reference frames. Now, if x𝑥xitalic_x and y𝑦yitalic_y belong to {\mathcal{L}}caligraphic_L, they must be of the form x=γx0𝑥𝛾subscript𝑥0x=\gamma x_{0}italic_x = italic_γ italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and y=γx0𝑦superscript𝛾subscript𝑥0y=\gamma^{\prime}x_{0}italic_y = italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, for some γ,γΓ𝛾superscript𝛾Γ\gamma,\gamma^{\prime}\in\Gammaitalic_γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_Γ. As such,

xy=yx1=(γx0)(γx0)1=γγ.𝑥𝑦𝑦superscript𝑥1superscript𝛾subscript𝑥0superscript𝛾subscript𝑥01𝛾superscript𝛾x\cdot y=yx^{-1}=(\gamma^{\prime}x_{0})(\gamma x_{0})^{-1}=\gamma\cdot\gamma^{% \prime}.italic_x ⋅ italic_y = italic_y italic_x start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = ( italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ( italic_γ italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = italic_γ ⋅ italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . (2.15)

Given Eq. (2.14), we have

x=γΓ=Γ,𝑥𝛾ΓΓx\cdot{\mathcal{L}}=\gamma\cdot\Gamma=\Gamma,italic_x ⋅ caligraphic_L = italic_γ ⋅ roman_Γ = roman_Γ , (2.16)

where we used the invariance of a group against its own right action. ∎

The above facts have important consequences: Since the coupling matrices are entirely determined by the local environment as experienced from the local frames of the resonators, we have wx,x()=wxxsubscript𝑤superscript𝑥𝑥subscript𝑤𝑥superscript𝑥w_{x^{\prime},x}({\mathcal{L}})=w_{x\cdot x^{\prime}}italic_w start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x end_POSTSUBSCRIPT ( caligraphic_L ) = italic_w start_POSTSUBSCRIPT italic_x ⋅ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT and, by setting xx=γ𝑥superscript𝑥𝛾x\cdot x^{\prime}=\gammaitalic_x ⋅ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_γ, we have

𝑫=γΓwγx|γxx|.𝑫subscript𝛾Γtensor-productsubscript𝑤𝛾subscript𝑥ket𝛾𝑥bra𝑥\bm{D}=\sum_{\gamma\in\Gamma}w_{\gamma}\otimes\sum_{x\in{\mathcal{L}}}|\gamma x% \rangle\langle x|.bold_italic_D = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ⊗ ∑ start_POSTSUBSCRIPT italic_x ∈ caligraphic_L end_POSTSUBSCRIPT | italic_γ italic_x ⟩ ⟨ italic_x | . (2.17)

We will see in the next section that any such 𝑫𝑫\bm{D}bold_italic_D can be generated from the left-regular representations of the (stabilized) group Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra of ΓΓ\Gammaroman_Γ. As explained in our introductory remarks, this is an effective though somewhat abstract way to communicate the constraints we want to impose on our synthetic molecules. Under these constraints, one is free to modify the position x0subscript𝑥0x_{0}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of the seed resonator in 𝑬𝑬Γ𝑬superscript𝑬Γ\bm{E}\setminus\bm{E}^{\Gamma}bold_italic_E ∖ bold_italic_E start_POSTSUPERSCRIPT roman_Γ end_POSTSUPERSCRIPT. One is also free to change the internal structure of the seed resonator, including adding or removing degrees of freedom. For example, we can add a second distinct seed resonator, because the pair can be seen as one seed resonator with a more complex structure. All these mentioned actions will essentially result in an alteration of the coupling matrices wγsubscript𝑤𝛾w_{\gamma}italic_w start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT and they lead to a well defined class of synthetic molecules associated with the group ΓΓ\Gammaroman_Γ. The common thing about these molecules is that their dynamical matrices are all generated from the same group Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra.

Since there is no cap on the number of degrees of freedom carried by the seed resonator, one can create quite complicated molecules with the algorithm just described, with layers upon layers of resonators, each having their own spatial distributions. Due to this untamed complexity, one can be mislead into thinking that there are no limits on the eigenmodes or dynamical patterns one can produce with such molecules. Quite the contrary, once the point group ΓΓ\Gammaroman_Γ is fixed, there are only a finite number of truly distinct dynamical patterns that one can squeeze out of the synthetic molecule (see section 5).

Refer to caption
Figure 2.4. The icosahedron and its proper symmetries.
Example 2.6.

The icosahedral group Ihsubscript𝐼I_{h}italic_I start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is the group of symmetries of the icosahedron, which is one of the existing platonic shapes. The subgroup Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT of Ihsubscript𝐼I_{h}italic_I start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT containing all proper space transformations has exactly 60 elements. They can be described as follows. There are 5-fold rotations around the six axes passing through pairs of opposing vertices, shown as yellow bullets in Fig. 2.4. One of the mentioned rotation axes is shown in yellow in Fig. 2.4. Since these axes also pass through the center of the icosahedron, we can specify them by using the coordinates r=(x,y,z)𝑟𝑥𝑦𝑧\vec{r}=(x,y,z)over→ start_ARG italic_r end_ARG = ( italic_x , italic_y , italic_z ) of six vertices in a rectangular coordinate system with origin at the center of the icosahedron. Then the axes themselves are encoded in the norm one vectors n^=r/|r|^𝑛𝑟𝑟\hat{n}=\vec{r}/|\vec{r}|over^ start_ARG italic_n end_ARG = over→ start_ARG italic_r end_ARG / | over→ start_ARG italic_r end_ARG |. The mentioned coordinates are (φ=𝜑absent\varphi=italic_φ = golden ratio, ϵ=±1italic-ϵplus-or-minus1\epsilon=\pm 1italic_ϵ = ± 1)

C5:(ϵ,φ,0),(0,ϵ,φ),(φ,0,ϵ).C_{5}:\quad(\epsilon,\varphi,0),(0,\epsilon,\varphi),(\varphi,0,\epsilon).italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT : ( italic_ϵ , italic_φ , 0 ) , ( 0 , italic_ϵ , italic_φ ) , ( italic_φ , 0 , italic_ϵ ) .

Rotations by 2nπ52𝑛𝜋5\frac{2n\pi}{5}divide start_ARG 2 italic_n italic_π end_ARG start_ARG 5 end_ARG, n=1,4¯𝑛¯14n=\overline{1,4}italic_n = over¯ start_ARG 1 , 4 end_ARG, around these six axes supply a total of 24 space transformations. Additionally, there are 3-fold rotations about the ten axes passing through the centers of opposing facets shown as blue bullets in Fig. 2.4, and one of the axes is shown in blue in Fig. 2.4. Again, we specify these axes by giving the coordinates of the centers of ten facets, which are

C3:(φ,0,ϵ(1+2φ)),(ϵ(1+φ),1+φ,1+φ),(0,1+2φ,ϵφ),(12φ,ϵφ,0),(1φ,ϵ(1+φ),ϵ(1+φ)).C_{3}:\quad\begin{array}[]{c}(-\varphi,0,\epsilon(1+2\varphi)),(\epsilon(1+% \varphi),1+\varphi,1+\varphi),(0,1+2\varphi,\epsilon\varphi),\\ (-1-2\varphi,\epsilon\varphi,0),(-1-\varphi,\epsilon(1+\varphi),-\epsilon(1+% \varphi)).\end{array}italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT : start_ARRAY start_ROW start_CELL ( - italic_φ , 0 , italic_ϵ ( 1 + 2 italic_φ ) ) , ( italic_ϵ ( 1 + italic_φ ) , 1 + italic_φ , 1 + italic_φ ) , ( 0 , 1 + 2 italic_φ , italic_ϵ italic_φ ) , end_CELL end_ROW start_ROW start_CELL ( - 1 - 2 italic_φ , italic_ϵ italic_φ , 0 ) , ( - 1 - italic_φ , italic_ϵ ( 1 + italic_φ ) , - italic_ϵ ( 1 + italic_φ ) ) . end_CELL end_ROW end_ARRAY

Rotations by 2nπ32𝑛𝜋3\frac{2n\pi}{3}divide start_ARG 2 italic_n italic_π end_ARG start_ARG 3 end_ARG, n=1,2𝑛12n=1,2italic_n = 1 , 2, around the ten axes supply a total of 20 space transformations. Lastly, there are 2-fold rotations about the 15 axes passing through the centers of opposing edges, shown as red bullets in Fig. 2.4. One such rotation axis is shown in red in Fig. 2.4. We specify these axes by giving the coordinates of the centers of 15 edges, which are

C2:(φ,0,0),(0,φ,0),(0,0,φ),12(ϵ,ϵ(1+φ),φ),12(ϵ(φ+1),ϵφ,1),12(ϵφ,ϵ,1+φ),12(ϵ,ϵ(1+φ),φ),12(ϵ(φ+1),ϵφ,1),12(ϵφ,ϵ,1+φ).C_{2}:\quad\begin{array}[]{c}(\varphi,0,0),(0,\varphi,0),(0,0,\varphi),\\ \tfrac{1}{2}(-\epsilon,\epsilon(1+\varphi),\varphi),\tfrac{1}{2}(\epsilon(% \varphi+1),-\epsilon\varphi,1),\tfrac{1}{2}(\epsilon\varphi,-\epsilon,1+% \varphi),\\ \tfrac{1}{2}(\epsilon,\epsilon(1+\varphi),\varphi),\tfrac{1}{2}(\epsilon(% \varphi+1),\epsilon\varphi,1),\tfrac{1}{2}(\epsilon\varphi,\epsilon,1+\varphi)% .\end{array}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : start_ARRAY start_ROW start_CELL ( italic_φ , 0 , 0 ) , ( 0 , italic_φ , 0 ) , ( 0 , 0 , italic_φ ) , end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( - italic_ϵ , italic_ϵ ( 1 + italic_φ ) , italic_φ ) , divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ϵ ( italic_φ + 1 ) , - italic_ϵ italic_φ , 1 ) , divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ϵ italic_φ , - italic_ϵ , 1 + italic_φ ) , end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ϵ , italic_ϵ ( 1 + italic_φ ) , italic_φ ) , divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ϵ ( italic_φ + 1 ) , italic_ϵ italic_φ , 1 ) , divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ϵ italic_φ , italic_ϵ , 1 + italic_φ ) . end_CELL end_ROW end_ARRAY

Rotations by π𝜋\piitalic_π around these 15 axes supply a total of 15 space transformations. Together with the identity transformation, we have listed 60 space transformations, hence, all the elements of the proper icosahedral group. \Diamond

Remark 2.7.

The procedures used in Example 2.6 apply to all platonic shapes, in the sense that all their proper point symmetries can be found by identifying the pairs of opposing vertices, opposing edge centers and opposing face centers, together with the appropriate angles of rotations. The latter are easy to determine once the axes of rotations are identified by the pairs, as explained.\Diamond

The ensembles seen in Fig. 2.3 were produced by applying the transformations listed above on a seeding shape, a cone in this instance. If this seeding shape is fitted with magnets as in Fig. 2.1, then this procedure produces genuine synthetic molecules from the class associated with the Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT group. On the contrary, the synthetic molecule shown in Fig. 2.2 has 12 resonators placed exactly above the vertices of an icosahedron. Thus, this is one of the cases where the seed resonator sits at a forbidden location. Another observation is that the molecule in Fig. 2.3(b) is not as symmetric as the platonic shape, since some pairs of resonators are closer spaced than others. Nevertheless, its lattice is still symmetric under the full Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT group.

Refer to caption
Figure 2.5. The C60 molecule contains 60 carbon atoms arranged in a spatial configuration that matches the vertices of a truncated icosahedron.

Fig. 2.5 illustrates the C60 molecule [25], where the 60 carbon atoms sit at the vertices of a truncated icosahedron. The nodes seen in this picture represent carbon atoms. Their outer-shell electrons are sp2𝑠superscript𝑝2sp^{2}italic_s italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-hybridized as in graphene, hence 3 of the valence electrons of a carbon atom are locked in strong covalent chemical bonds, seen as edges in Fig. 2.5, while the remaining valence electrons (one per carbon atom) hop between the π𝜋\piitalic_π-orbitals carried by each of the atoms. If |aket𝑎|a\rangle| italic_a ⟩ is the π𝜋\piitalic_π-orbital carried by atom a𝑎aitalic_a of the molecule, as defined in a local frame attached to the atom, then we can write the discrete Hamiltonian

HC60=E0a,a|aa|,E0>0,formulae-sequencesubscript𝐻subscript𝐶60subscript𝐸0subscript𝑎superscript𝑎ket𝑎brasuperscript𝑎subscript𝐸00H_{C_{60}}=-E_{0}\sum_{\langle a,a^{\prime}\rangle}|a\rangle\langle a^{\prime}% |,\quad E_{0}>0,italic_H start_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT 60 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT ⟨ italic_a , italic_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ end_POSTSUBSCRIPT | italic_a ⟩ ⟨ italic_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | , italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 , (2.18)

where ,\langle\,,\,\rangle⟨ , ⟩ means near neighboring atoms. This Hamiltonian accounts well for the low energy dynamics of the 60 valence electrons of the molecule.

2.4. The Cayley graph picture

We mentioned at one point the similarity between the expressions (2.17) and (1.1), which is a nice feature of the formalism. Still, in Eq. (2.17), the lattice needs to be imagined inside an abstract group, but this can be corrected using the notion of Cayley graph.

The isometry class of all space groups can be abstractly presented in terms of generators and relations [26]. In particular, if 𝔽2=X,Ysubscript𝔽2𝑋𝑌{\mathbb{F}}_{2}=\langle X,Y\rangleblackboard_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ⟨ italic_X , italic_Y ⟩ is the free nonabelian group in two generators and N𝑁Nitalic_N is its normal subgroup generated by the set {X2,Y2,(XY)3}superscript𝑋2superscript𝑌2superscript𝑋𝑌3\{X^{2},Y^{2},(XY)^{3}\}{ italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_Y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , ( italic_X italic_Y ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT }, then 𝔽2/Nsubscript𝔽2𝑁{\mathbb{F}}_{2}/Nblackboard_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_N is isomorphic to Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. Explicitly, the isomorphism is given by the identifications XC5𝑋subscript𝐶5X\leftrightarrow C_{5}italic_X ↔ italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and YC2𝑌subscript𝐶2Y\leftrightarrow C_{2}italic_Y ↔ italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, where C5subscript𝐶5C_{5}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the rotations displayed in Fig. 2.4. Thus, Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT can be presented as

Ip=C5,C2|C55,C22,(C5C2)3.subscript𝐼𝑝inner-productsubscript𝐶5subscript𝐶2superscriptsubscript𝐶55superscriptsubscript𝐶22superscriptsubscript𝐶5subscript𝐶23I_{p}=\big{\langle}C_{5},C_{2}|C_{5}^{5},C_{2}^{2},(C_{5}C_{2})^{3}\big{% \rangle}.italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = ⟨ italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT , italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT , italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , ( italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ . (2.19)

As such, any element of Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT can be written as a word made up from the two letters C5subscript𝐶5C_{5}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, but this writing is not unique due to the presence of relations. One should also be aware that other generating sets and relations are possible, and, to distinguish the presentation of Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT displayed above from others, we will refer to Eq. (2.19) as the standard presentation of Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. Such facts apply to any finite subgroup of SO(3)𝑆𝑂3SO(3)italic_S italic_O ( 3 ).

Definition 2.8 ([27]).

Given a discrete group ΓΓ\Gammaroman_Γ and a finite subset S𝑆Sitalic_S of ΓΓ\Gammaroman_Γ, the Cayley graph 𝒞(Γ,S)𝒞Γ𝑆{\mathcal{C}}(\Gamma,S)caligraphic_C ( roman_Γ , italic_S ) is the un-directed graph with vertex set ΓΓ\Gammaroman_Γ and edge set containing an edge between γ𝛾\gammaitalic_γ and sγ𝑠𝛾s\gammaitalic_s italic_γ whenever γΓ𝛾Γ\gamma\in\Gammaitalic_γ ∈ roman_Γ and sS𝑠𝑆s\in Sitalic_s ∈ italic_S.

If S𝑆Sitalic_S is a standard generating set of ΓΓ\Gammaroman_Γ, then we call 𝒞(Γ,S)𝒞Γ𝑆{\mathcal{C}}(\Gamma,S)caligraphic_C ( roman_Γ , italic_S ) the standard Cayley graph of ΓΓ\Gammaroman_Γ. A more refined and more useful geometric object is the Cayley digraph:

Definition 2.9 ([27]).

Given a discrete group ΓΓ\Gammaroman_Γ and a subset S𝑆Sitalic_S of ΓΓ\Gammaroman_Γ, let c:SColor:𝑐𝑆Colorc:S\to{\rm Color}italic_c : italic_S → roman_Color assign a distinct color to each sS𝑠𝑆s\in Sitalic_s ∈ italic_S. Then the Cayley digraph 𝒞(Γ,S,c)𝒞Γ𝑆𝑐\vec{\mathcal{C}}(\Gamma,S,c)over→ start_ARG caligraphic_C end_ARG ( roman_Γ , italic_S , italic_c ) is the colored graph with vertex set ΓΓ\Gammaroman_Γ and directed edges from γ𝛾\gammaitalic_γ to sγ𝑠𝛾s\gammaitalic_s italic_γ for γΓ𝛾Γ\gamma\in\Gammaitalic_γ ∈ roman_Γ and sS𝑠𝑆s\in Sitalic_s ∈ italic_S. All directed edges produced by sS𝑠𝑆s\in Sitalic_s ∈ italic_S are assigned the color c(s)𝑐𝑠c(s)italic_c ( italic_s ).

Refer to caption
Figure 2.6. Cayley digraph of the proper icosahedral group. Red arrows represent the C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT generator and blue arrows represent the C5subscript𝐶5C_{5}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT generator.

The standard Cayley digraph of Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is illustrated in Fig. 2.6, which renders the group in a geometric fashion. We can select any vertex of this graph to represent the neutral element e𝑒eitalic_e, and various paths between e𝑒eitalic_e and a vertex γΓ𝛾Γ\gamma\in\Gammaitalic_γ ∈ roman_Γ supply a valid word for γ𝛾\gammaitalic_γ made up from the letters C5subscript𝐶5C_{5}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT and C2subscript𝐶2C_{2}italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and their inverses. The Hamiltonian (2.18) coincides, up to a factor, with the adjacency operator of this graph

Δ=γIp(12|C5γγ|+12|C51γγ|+|C2γγ|).Δsubscript𝛾subscript𝐼𝑝12ketsubscript𝐶5𝛾quantum-operator-product𝛾12superscriptsubscript𝐶51𝛾bra𝛾ketsubscript𝐶2𝛾bra𝛾\Delta=-\sum_{\gamma\in I_{p}}\big{(}\tfrac{1}{2}|C_{5}\gamma\rangle\langle% \gamma|+\tfrac{1}{2}|C_{5}^{-1}\gamma\rangle\langle\gamma|+|C_{2}\gamma\rangle% \langle\gamma|\big{)}.roman_Δ = - ∑ start_POSTSUBSCRIPT italic_γ ∈ italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG | italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT italic_γ ⟩ ⟨ italic_γ | + divide start_ARG 1 end_ARG start_ARG 2 end_ARG | italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_γ ⟩ ⟨ italic_γ | + | italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_γ ⟩ ⟨ italic_γ | ) . (2.20)

We will see in subsection 5.2 how such Hamiltonians can be graphically represented on the Cayley graph.

As advertised in [28], Cayley graphs can be used to produce real synthetic molecules from abstract finitely generated groups. Furthermore, Cayley graphs can be used very effectively to render the resonant modes of molecules.

3. The Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra of dynamical matrices

This section establishes the connection between the abstract Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra of a point group and the algebra of dynamical matrices for the class of synthetic molecules introduced in the previous section.

3.1. Group Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebras

The material of this subsection is standard and it can be found, for example, in [29][Ch. VIII]. Given a generic discrete group ΓΓ\Gammaroman_Γ, its group algebra ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ consists of formal series

q=γΓαγγ,αγ,formulae-sequence𝑞subscript𝛾Γsubscript𝛼𝛾𝛾subscript𝛼𝛾q=\sum_{\gamma\in\Gamma}\alpha_{\gamma}\,\gamma,\quad\alpha_{\gamma}\in{% \mathbb{C}},italic_q = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_γ , italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∈ blackboard_C , (3.1)

where all but a finite number of terms are zero.444Of course, this is of no concern if ΓΓ\Gammaroman_Γ is finite. The addition and multiplication of such formal series are defined in the obvious way, using the group and algebraic structures of ΓΓ\Gammaroman_Γ and {\mathbb{C}}blackboard_C, respectively. In addition, there exists a natural \ast-operation

q=γΓαγγ1,(q)=q,(αq)=αq,α.formulae-sequencesuperscript𝑞subscript𝛾Γsuperscriptsubscript𝛼𝛾superscript𝛾1formulae-sequencesuperscriptsuperscript𝑞𝑞formulae-sequencesuperscript𝛼𝑞superscript𝛼superscript𝑞𝛼q^{\ast}=\sum_{\gamma\in\Gamma}\alpha_{\gamma}^{\ast}\,\gamma^{-1},\quad(q^{% \ast})^{\ast}=q,\quad(\alpha q)^{\ast}=\alpha^{\ast}q^{\ast},\ \alpha\in{% \mathbb{C}}.italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , ( italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_q , ( italic_α italic_q ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_α ∈ blackboard_C . (3.2)

Hence, ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ is a \ast-algebra in a natural way.

We denote by e𝑒eitalic_e the neutral element of ΓΓ\Gammaroman_Γ. Then the map

𝒯:Γ,𝒯(q)=αe:𝒯formulae-sequenceΓ𝒯𝑞subscript𝛼𝑒{\mathcal{T}}:{\mathbb{C}}\Gamma\to{\mathbb{C}},\quad{\mathcal{T}}(q)=\alpha_{e}caligraphic_T : blackboard_C roman_Γ → blackboard_C , caligraphic_T ( italic_q ) = italic_α start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT (3.3)

defines a faithful positive trace on ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ and a pre-Hilbert structure on ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ via

q,q:=𝒯(qq),q,qΓ.formulae-sequenceassign𝑞superscript𝑞𝒯superscript𝑞superscript𝑞𝑞superscript𝑞Γ\langle q,q^{\prime}\rangle:={\mathcal{T}}(q^{\ast}q^{\prime}),\quad q,q^{% \prime}\in{\mathbb{C}}\Gamma.⟨ italic_q , italic_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ := caligraphic_T ( italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , italic_q , italic_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_C roman_Γ . (3.4)

The completion of the linear space G𝐺{\mathbb{C}}Gblackboard_C italic_G under this pre-Hilbert structure supplies the Hilbert space 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ), spanned by |γket𝛾|\gamma\rangle| italic_γ ⟩, γΓ𝛾Γ\gamma\in\Gammaitalic_γ ∈ roman_Γ, which form an orthonormal basis

γ,γ=δγ,γ,γ,γΓ.formulae-sequence𝛾superscript𝛾subscript𝛿𝛾superscript𝛾𝛾superscript𝛾Γ\langle\gamma,\gamma^{\prime}\rangle=\delta_{\gamma,\gamma^{\prime}},\quad% \gamma,\gamma^{\prime}\in\Gamma.⟨ italic_γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = italic_δ start_POSTSUBSCRIPT italic_γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , italic_γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_Γ . (3.5)

The left action of ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ on itself can be extended to the action of a bounded operator on 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ), and this supplies the left regular representation πLsubscript𝜋𝐿\pi_{L}italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT of ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ inside the algebra 𝔹(2(Γ))𝔹superscript2Γ{\mathbb{B}}\big{(}\ell^{2}(\Gamma)\big{)}blackboard_B ( roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ) ) of bounded operators on 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ). Specifically,

πL(q)|γ=γΓαγ|γγ,qΓ,γΓ,formulae-sequencesubscript𝜋𝐿𝑞ketsuperscript𝛾subscript𝛾Γsubscript𝛼𝛾ket𝛾superscript𝛾formulae-sequence𝑞Γsuperscript𝛾Γ\pi_{L}(q)|\gamma^{\prime}\rangle=\sum_{\gamma\in\Gamma}\alpha_{\gamma}|\gamma% \gamma^{\prime}\rangle,\quad q\in{\mathbb{C}}\Gamma,\quad\gamma^{\prime}\in\Gamma,italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT | italic_γ italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ , italic_q ∈ blackboard_C roman_Γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_Γ , (3.6)

where q=γΓαγγ𝑞subscript𝛾Γsubscript𝛼𝛾𝛾q=\sum_{\gamma\in\Gamma}\alpha_{\gamma}\gammaitalic_q = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_γ with αγsubscript𝛼𝛾\alpha_{\gamma}\in{\mathbb{C}}italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∈ blackboard_C. The completion of ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ with respect to the norm

q:=πL(q)𝔹(2(Γ))assignnorm𝑞subscriptnormsubscript𝜋𝐿𝑞𝔹superscript2Γ\|q\|:=\|\pi_{L}(q)\|_{{\mathbb{B}}(\ell^{2}(\Gamma))}∥ italic_q ∥ := ∥ italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) ∥ start_POSTSUBSCRIPT blackboard_B ( roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ) ) end_POSTSUBSCRIPT (3.7)

supplies the reduced group Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra CrΓsubscriptsuperscript𝐶𝑟ΓC^{\ast}_{r}\Gammaitalic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ of ΓΓ\Gammaroman_Γ.

Remark 3.1.

It may seems that, for finite groups, we can equally well work with the group algebra. However, the norm (3.7) puts a topology on the algebra, which will then enable us to speak of “continuous deformations” of the models. \Diamond

The right-regular representation of the group algebra acts on 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ) as

πR(q)|γ:=γΓαγ|γγ1,qCrΓ,γΓ.formulae-sequenceassignsubscript𝜋𝑅𝑞ketsuperscript𝛾subscript𝛾Γsubscript𝛼𝛾ketsuperscript𝛾superscript𝛾1formulae-sequence𝑞subscriptsuperscript𝐶𝑟Γsuperscript𝛾Γ\pi_{R}(q)|\gamma^{\prime}\rangle:=\sum_{\gamma\in\Gamma}\alpha_{\gamma}|% \gamma^{\prime}\gamma^{-1}\rangle,\quad q\in C^{\ast}_{r}\Gamma,\quad\gamma^{% \prime}\in\Gamma.italic_π start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_q ) | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ := ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ⟩ , italic_q ∈ italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ , italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_Γ . (3.8)
Remark 3.2.

We note that, if the group ΓΓ\Gammaroman_Γ is finite, then

𝒯(q)=1|Γ|Tr(πL(q))=1|Γ|Tr(πR(q)),𝒯𝑞1ΓTrsubscript𝜋𝐿𝑞1ΓTrsubscript𝜋𝑅𝑞{\mathcal{T}}(q)=\frac{1}{|\Gamma|}{\rm Tr}\big{(}\pi_{L}(q)\big{)}=\frac{1}{|% \Gamma|}{\rm Tr}\big{(}\pi_{R}(q)\big{)},caligraphic_T ( italic_q ) = divide start_ARG 1 end_ARG start_ARG | roman_Γ | end_ARG roman_Tr ( italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) ) = divide start_ARG 1 end_ARG start_ARG | roman_Γ | end_ARG roman_Tr ( italic_π start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_q ) ) , (3.9)

which can be seen straight from (3.6) and (3.8).\Diamond

3.2. Symmetry and structure of dynamical matrices

The right-regular representation induces a unitary representation of ΓΓ\Gammaroman_Γ on 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ):

U:Γ𝔹(2(Γ)),Uγ|γ:=πR(γ)|γ=|γγ1.:𝑈formulae-sequenceΓ𝔹superscript2Γassignsubscript𝑈𝛾ketsuperscript𝛾subscript𝜋𝑅𝛾ketsuperscript𝛾ketsuperscript𝛾superscript𝛾1U:\Gamma\to{\mathbb{B}}\big{(}\ell^{2}(\Gamma)\big{)},\quad U_{\gamma}|\gamma^% {\prime}\rangle:=\pi_{R}(\gamma)|\gamma^{\prime}\rangle=|\gamma^{\prime}\gamma% ^{-1}\rangle.italic_U : roman_Γ → blackboard_B ( roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ) ) , italic_U start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ := italic_π start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_γ ) | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ = | italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ⟩ . (3.10)

Since the left and right actions commute, we automatically have that

UγπL(q)Uγ=πL(q),qCrΓ,γΓ.formulae-sequencesuperscriptsubscript𝑈𝛾subscript𝜋𝐿𝑞subscript𝑈𝛾subscript𝜋𝐿𝑞formulae-sequence𝑞subscriptsuperscript𝐶𝑟Γ𝛾ΓU_{\gamma}^{\ast}\,\pi_{L}(q)\,U_{\gamma}=\pi_{L}(q),\quad q\in C^{\ast}_{r}% \Gamma,\quad\gamma\in\Gamma.italic_U start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) italic_U start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT = italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) , italic_q ∈ italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ , italic_γ ∈ roman_Γ . (3.11)

Hence, the group Csuperscript𝐶C^{\ast}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-algebra supplies models that are naturally ΓΓ\Gammaroman_Γ-symmetric.

Now, given the action (3.6) of left-regular representation on the basis of 2(Γ)superscript2Γ\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ), we can write

πL(q)=γΓαγγΓ|γγγ|.subscript𝜋𝐿𝑞subscript𝛾Γsubscript𝛼𝛾subscriptsuperscript𝛾Γket𝛾superscript𝛾brasuperscript𝛾\pi_{L}(q)=\sum_{\gamma\in\Gamma}\alpha_{\gamma}\sum_{\gamma^{\prime}\in\Gamma% }|\gamma\gamma^{\prime}\rangle\langle\gamma^{\prime}|.italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_q ) = ∑ start_POSTSUBSCRIPT italic_γ ∈ roman_Γ end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_Γ end_POSTSUBSCRIPT | italic_γ italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ⟨ italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | . (3.12)

Using the isomorphism of Hilbert spaces 2()2(Γ)similar-to-or-equalssuperscript2superscript2Γ\ell^{2}({\mathcal{L}})\simeq\ell^{2}(\Gamma)roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_L ) ≃ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ), |x=|γx0|γket𝑥ket𝛾subscript𝑥0maps-toket𝛾|x\rangle=|\gamma x_{0}\rangle\mapsto|\gamma\rangle| italic_x ⟩ = | italic_γ italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ ↦ | italic_γ ⟩, we can compare (3.12) with (2.17) to conclude the 𝑫𝑫\bm{D}bold_italic_D is just the left-regular representation of the element h=γwγγCrΓsubscript𝛾subscript𝑤𝛾𝛾subscriptsuperscript𝐶𝑟Γh=\sum_{\gamma}w_{\gamma}\gamma\in C^{\ast}_{r}\Gammaitalic_h = ∑ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_γ ∈ italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ. The only difference is that wγsubscript𝑤𝛾w_{\gamma}italic_w start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT is a matrix in general and we can correct for it by tensoring CrΓsubscriptsuperscript𝐶𝑟ΓC^{\ast}_{r}\Gammaitalic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ with 𝕂𝕂{\mathbb{K}}blackboard_K, the algebra of compact operators over 2(𝒟)superscript2𝒟\ell^{2}({\mathcal{D}})roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_D ), where 𝒟𝒟{\mathcal{D}}caligraphic_D is the set of internal degrees of freedom.555To allow unrestricted turning off and on of the degrees of freedom, 𝒟𝒟{\mathcal{D}}caligraphic_D is assumed to be infinite. This operation is commonly referred to as the stabilization of the algebra. At this point, we reached one of our main points:

Proposition 3.3.

All dynamical matrices for the class of synthetic molecules associated with the group ΓΓ\Gammaroman_Γ can be generated as left-regular representations of self-adjoint elements from 𝕂CrΓtensor-product𝕂subscriptsuperscript𝐶𝑟Γ{\mathbb{K}}\otimes C^{\ast}_{r}\Gammablackboard_K ⊗ italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_Γ.

Example 3.4.

The adjacency operator (2.20) is the left-regular representation of the self-adjoint element δ=12(C5+C51)+C2𝛿12subscript𝐶5superscriptsubscript𝐶51subscript𝐶2\delta=\frac{1}{2}(C_{5}+C_{5}^{-1})+C_{2}italic_δ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) + italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT from CrIpsubscriptsuperscript𝐶𝑟subscript𝐼𝑝C^{\ast}_{r}I_{p}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT.\Diamond

4. Algebraic and Topological Aspects of the Dynamics

The subject of this section is on the decomposition of the linear space of resonant modes in the smallest possible subspaces that remain invariant under dynamics, as well as on the classification of these subspaces from algebraic and topological point of views. The former one engages the ordinary representation theory while the latter one requires more specialized tools, specifically, Kasparov’s bi-variant K-theory. While an in-depth mastery of the latter requires substantial effort, the concepts and the working principles can be communicated in very few words. For a reader unfamiliar with the theory, seeing them in action in the simplest but non-trivial context of finite groups and algebras can serve as a door to the world of these ideas. This is one of the main reasons KK-theory appears in our exposition. Another reason is that there is no other natural alternative to the topological point of view. In our opinion, KK-theory is a must-have tool for anyone planning to enter the business of topological dynamics. We hope that the patient reader can sense that from the exposition below.

4.1. The algebraic representation ring

A unitary representation of a group ΓΓ\Gammaroman_Γ is a group morphism between the group and the group 𝕌()𝕌{\mathbb{U}}({\mathcal{H}})blackboard_U ( caligraphic_H ) of unitary operators over a Hilbert space {\mathcal{H}}caligraphic_H. Thus, a unitary representation π:Γ𝕌():𝜋Γ𝕌\pi:\Gamma\to{\mathbb{U}}({\mathcal{H}})italic_π : roman_Γ → blackboard_U ( caligraphic_H ), typically shorthanded to just π𝜋\piitalic_π, specifies both the Hilbert space {\mathcal{H}}caligraphic_H and the morphism itself. If the Hilbert space has a finite dimension N𝑁Nitalic_N, then we are talking about finite dimensional representations, which are nothing but group morphisms from ΓΓ\Gammaroman_Γ to U(N)𝑈𝑁U(N)italic_U ( italic_N ), the group of unitary matrices of rank N𝑁Nitalic_N. Two representations π1:Γ𝕌(1):subscript𝜋1Γ𝕌subscript1\pi_{1}:\Gamma\to{\mathbb{U}}({\mathcal{H}}_{1})italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT : roman_Γ → blackboard_U ( caligraphic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) and π2:G𝕌(2):subscript𝜋2𝐺𝕌subscript2\pi_{2}:G\to{\mathbb{U}}({\mathcal{H}}_{2})italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : italic_G → blackboard_U ( caligraphic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are said to be unitarily equivalent if there exists a unitary map U:12:𝑈subscript1subscript2U:{\mathcal{H}}_{1}\to{\mathcal{H}}_{2}italic_U : caligraphic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT → caligraphic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT such that π1=Uπ2Usubscript𝜋1superscript𝑈subscript𝜋2𝑈\pi_{1}=U^{\ast}\pi_{2}Uitalic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_U start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_U. Representation theory is the art of classifying the representations of a group up to unitary equivalence. Since the relation between finite groups and their group algebras is functorial, any finite group representation of a finite group extends by linearity to a finite representation of its algebra as bounded linear operators over the same Hilbert space. Reciprocally, any finite representation of a group’s algebra supplies, by restriction to the group elements, a representation of the group.

Given two finite group representations π1subscript𝜋1\pi_{1}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and π2subscript𝜋2\pi_{2}italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, we can define a new finite representation by using their direct sum π1π2direct-sumsubscript𝜋1subscript𝜋2\pi_{1}\oplus\pi_{2}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⊕ italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. This binary operation is compatible with the unitary equivalence, hence it descends on the set of unitarily equivalent classes of finite representations, where it becomes an abelian binary operation. Thus, the set {\mathcal{R}}caligraphic_R of classes of unitarily equivalent finite group representations has a natural structure of an abelian semigroup. It can be closed to a group (,)direct-sum({\mathcal{R}},\oplus)( caligraphic_R , ⊕ ), which is called the group of virtual representations.

𝐈𝐩subscript𝐈𝐩\mathbf{I_{p}}bold_I start_POSTSUBSCRIPT bold_p end_POSTSUBSCRIPT E𝐸Eitalic_E 12C512subscript𝐶512C_{5}12 italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT 12C5212superscriptsubscript𝐶5212C_{5}^{2}12 italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 20C320subscript𝐶320C_{3}20 italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT 15C215subscript𝐶215C_{2}15 italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
Agsubscript𝐴𝑔A_{g}italic_A start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT 1 1 1 1 1
T1gsubscript𝑇1𝑔T_{1g}italic_T start_POSTSUBSCRIPT 1 italic_g end_POSTSUBSCRIPT 3 φ𝜑\varphiitalic_φ 1φ1𝜑1-\varphi1 - italic_φ 0 -1
T2gsubscript𝑇2𝑔T_{2g}italic_T start_POSTSUBSCRIPT 2 italic_g end_POSTSUBSCRIPT 3 1φ1𝜑1-\varphi1 - italic_φ φ𝜑\varphiitalic_φ 0 -1
Ggsubscript𝐺𝑔G_{g}italic_G start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT 4 -1 -1 1 0
Hgsubscript𝐻𝑔H_{g}italic_H start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT 5 0 0 -1 1
Table 1. The character table of Ipsubscript𝐼𝑝I_{p}italic_I start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT group, where φ𝜑\varphiitalic_φ is the golden ratio.

A representation π:Γ𝕌():𝜋Γ𝕌\pi:\Gamma\to{\mathbb{U}}({\mathcal{H}})italic_π : roman_Γ → blackboard_U ( caligraphic_H ) is said to be irreducible if there is no linear subspace of {\mathcal{H}}caligraphic_H that is left invariant by the actions of all π(γ)𝜋𝛾\pi(\gamma)italic_π ( italic_γ ), γΓ𝛾Γ\gamma\in\Gammaitalic_γ ∈ roman_Γ. Up to unitary equivalences, the irreducible representations of a finite group ΓΓ\Gammaroman_Γ can be enumerated using the characters of the group, which are readily available in the literature. For example, the table of characters of the proper icosahedral group can be found in [30] and is reproduced in Table 1. There is a simple and practical way to generate an irreducible representation πχsubscript𝜋𝜒\pi_{\chi}italic_π start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT from a character χ𝜒\chiitalic_χ (see section 5.1). Any finite group representation π𝜋\piitalic_π is unitarily equivalent to a finite direct sum of irreducible representations

πχ(πχπχ)similar-to-or-equals𝜋subscriptdirect-sum𝜒direct-sumsubscript𝜋𝜒subscript𝜋𝜒\pi\simeq\bigoplus\nolimits_{\chi}(\pi_{\chi}\oplus\cdots\oplus\pi_{\chi})italic_π ≃ ⨁ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ( italic_π start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ⊕ ⋯ ⊕ italic_π start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) (4.1)

The above can be written more compactly as πχnχπχsimilar-to-or-equals𝜋subscriptdirect-sum𝜒subscript𝑛𝜒subscript𝜋𝜒\pi\simeq\oplus_{\chi}\,n_{\chi}\,\pi_{\chi}italic_π ≃ ⊕ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT italic_π start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT, nχsubscript𝑛𝜒n_{\chi}\in{\mathbb{N}}italic_n start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ∈ blackboard_N, which shows that (,)direct-sum({\mathcal{R}},\oplus)( caligraphic_R , ⊕ ) is freely and commutatively generated over the classes of irreducible representations.

There is additional algebraic structure on {\mathcal{R}}caligraphic_R. Indeed, given two finite group representations π1subscript𝜋1\pi_{1}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and π2subscript𝜋2\pi_{2}italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, we can generate a new finite representation using their tensor product π1π2tensor-productsubscript𝜋1subscript𝜋2\pi_{1}\otimes\pi_{2}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⊗ italic_π start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. This binary operation is also compatible with the unitary equivalence, hence it descends on {\mathcal{R}}caligraphic_R, where it becomes an associative commutative binary operation, which distributes over direct-sum\oplus. Thus, the set {\mathcal{R}}caligraphic_R of classes of unitarily equivalent finite group virtual representations has a natural structure of a commutative ring, (,,)direct-sumtensor-product({\mathcal{R}},\oplus,\otimes)( caligraphic_R , ⊕ , ⊗ ), called the ring of virtual representations [31].

4.2. Dynamics vs group representations

When analyzing the dynamics of a synthetic molecule, the ring of representations comes into play in the following way. Let 𝑫𝑫\bm{D}bold_italic_D be a dynamical matrix as in (2.17). As we have already seen, 𝑫𝑫\bm{D}bold_italic_D is a linear operator over the Hilbert space =D2(Γ)tensor-productsuperscript𝐷superscript2Γ{\mathcal{H}}={\mathbb{C}}^{D}\otimes\ell^{2}(\Gamma)caligraphic_H = blackboard_C start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT ⊗ roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Γ ), which is necessarily the left-regular representation of an element hhitalic_h from the group algebra ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ. Suppose we compute or measure the resonant spectrum of 𝑫𝑫\bm{D}bold_italic_D and we find an eigenvalue λ𝜆\lambdaitalic_λ. By standard methods, we can associate a spectral projection Pλsubscript𝑃𝜆P_{\lambda}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT such that Pλsubscript𝑃𝜆P_{\lambda}{\mathcal{H}}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT caligraphic_H is the linear space spanned by all the eigenvectors of 𝑫𝑫\bm{D}bold_italic_D corresponding to λ𝜆\lambdaitalic_λ. One finds that this Pλsubscript𝑃𝜆P_{\lambda}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is also the left-regular representation of a projection pλsubscript𝑝𝜆p_{\lambda}italic_p start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT from ΓΓ{\mathbb{C}}\Gammablackboard_C roman_Γ. For example, consider the so-called interpolating polynomials defined by the relations Fλ(λ)=δλ,λsubscript𝐹𝜆superscript𝜆subscript𝛿𝜆superscript𝜆F_{\lambda}(\lambda^{\prime})=\delta_{\lambda,\lambda^{\prime}}italic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_δ start_POSTSUBSCRIPT italic_λ , italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT for any λ𝜆\lambdaitalic_λ and λsuperscript𝜆\lambda^{\prime}italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT from the discrete spectrum of 𝑫=πL(h)𝑫subscript𝜋𝐿\bm{D}=\pi_{L}(h)bold_italic_D = italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_h ). Recalling that the finite representations commute with polynomial calculus, we have

Pλ=Fλ(𝑫)=Fλ(πL(h))=πL(Fλ(h)),subscript𝑃𝜆subscript𝐹𝜆𝑫subscript𝐹𝜆subscript𝜋𝐿subscript𝜋𝐿subscript𝐹𝜆P_{\lambda}=F_{\lambda}(\bm{D})=F_{\lambda}(\pi_{L}(h))=\pi_{L}(F_{\lambda}(h)),italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = italic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( bold_italic_D ) = italic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_h ) ) = italic_π start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_h ) ) , (4.2)

which shows that Pλsubscript𝑃𝜆P_{\lambda}italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is the left-regular representation of pλ=Fλ(h)Γsubscript𝑝𝜆subscript𝐹𝜆Γp_{\lambda}=F_{\lambda}(h)\in{\mathbb{C}}\Gammaitalic_p start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = italic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_h ) ∈ blackboard_C roman_Γ. Lastly, since the left and right regular representations commute, we see that

πλ:Γ𝔹(Pλ),πλ(γ):=PλπR(γ)Pλ,:subscript𝜋𝜆formulae-sequenceΓ𝔹subscript𝑃𝜆assignsubscript𝜋𝜆𝛾subscript𝑃𝜆subscript𝜋𝑅𝛾subscript𝑃𝜆\pi_{\lambda}:\Gamma\to{\mathbb{B}}(P_{\lambda}{\mathcal{H}}),\quad\pi_{% \lambda}(\gamma):=P_{\lambda}\pi_{R}(\gamma)P_{\lambda},italic_π start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT : roman_Γ → blackboard_B ( italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT caligraphic_H ) , italic_π start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_γ ) := italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_π start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_γ ) italic_P start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT , (4.3)

defines a finite representation of ΓΓ\Gammaroman_Γ. We learned in the previous sub-section that each πλsubscript𝜋𝜆\pi_{\lambda}italic_π start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is unitarily equivalent to a direct sum of irreducible representations. Thus, we have the following inescapable fact:

Proposition 4.1.

Up to a unitary transformation, all resonant modes, without exception, coincide with a vector from the representation space of one of the πχsubscript𝜋𝜒\pi_{\chi}italic_π start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT’s, or linear combinations of such vectors.

The above fact shows that the dynamical patterns supported by a synthetic molecule from the class associated to the group ΓΓ\Gammaroman_Γ can be all generated as linear combinations of vectors drawn from a finite number of spaces, which can be enumerated by the character of the group. The statement is purely algebraic and, as such, it cannot reveal anything about the stability of those dynamical patterns against continuous deformations of the molecule. In the light of this remark, it is then natural to approach the representation theory and the classification of dynamical patterns from a topological point of view. This is the subject of the following two subsections.

4.3. The topological representation ring