Structure preserving discretization:
A Berezin-Toeplitz Quantization viewpoint


Abstract

In this paper, we introduce a comprehensive axiomatization of structure-preserving discretization through the framework of commutative diagrams. By establishing a formal language that captures the essential properties of discretization processes, we provide a rigorous foundation for analyzing how various structures—such as algebraic, geometric, and topological features—are maintained during the transition from continuous to discrete settings. Specifically, we establish that the transition from continuous to discrete differential settings invariably leads to noncommutative structures, reinforcing previous observations on the interplay between discretization and noncommutativity.
We demonstrate the applicability of our axiomatization by applying it to the Berezin-Toeplitz quantization, showing that this quantization method adheres to our proposed criteria for structure-preserving discretization. We establish in this setting a precise limit theorem for the approximation of the Laplacian by a sequence of matrix approximations. This work enriches the theoretical understanding of discretization and sets the stage for its broader applications to various discretization methods.

1 Introduction↩︎

The discretization of continuous structures plays a pivotal role in both theoretical and applied mathematics, particularly in fields such as numerical analysis, computational physics, and data science. Traditional approaches to discretization often rely on heuristic or ad hoc methods, which can lead to inconsistencies and inefficiencies. Recent advances suggest that a more rigorous approach to discretization lies in preserving the essential features of the continuous structures while transitioning to a discrete framework. A key aspect of effective discretization is ensuring that the discrete model retains the underlying geometric and algebraic properties of the continuous one.
Structure-preserving discretization methods have gained significant attention due to their ability to retain key properties of continuous systems, such as symmetries, conservation laws, and geometric structures, when transitioning to discrete models. State-of-the-art approaches include finite element exterior calculus (FEEC) [1], [2], which ensures stability and convergence in numerical approximations of partial differential equations by preserving the de Rham complex. Geometric integration methods, such as symplectic integrators, maintain the symplectic structure in Hamiltonian systems, ensuring long-term accuracy in energy conservation [3], [4]. Discrete exterior calculus (DEC) extends these ideas to computational geometry by discretizing differential forms while preserving topological properties like divergence and curl [5], [6]. Recent advances have also focused on the relation between noncommutative geometries and discrete differential structures [7], [8]. This is in particular exemplified by the Berezin-Toeplitz quantization. This quantization discretizes certain types of Poisson algebras while maintaining algebraic structures in the transition from continuous to discrete systems [9]. Hence, quantization provides a robust framework for discrete geometry applications.
Although there are similarities among the different approaches of structure-preserving discretization, it is yet not clear how one can describe under a unified theory, eclectic methods ranging from finite exterior calculus to quantization; we identify this as a gap in the theory.
In this paper, we present an axiomatization of structure-preserving discretization by leveraging the language of commutative diagrams. Our primary objective is to formalize the process of discretization such that it maintains the integrity of underlying geometric and algebraic structures. We argue that by employing commutative diagrams, one can systematically capture and enforce the preservation of essential relationships between continuous and discrete models. A subsequent objective is to demonstrate that the structure-preserving discretization of differential structures, following the previous axioms, lead to noncommutative differential structure on operator algebras.
Commutative diagrams provide a visual and conceptual tool from category theory that allow us to represent the interrelations between various mathematical structures and their discretizations. By constructing and analyzing these diagrams, we can ensure that the discretization process respects the morphisms and relationships inherent in the continuous setting. This approach not only facilitates a clearer understanding of the discretization process but also helps in developing algorithms that are both theoretically sound and practically effective.
We demonstrate, by applying our axiomatization, that the discretization of a continuous differential structure is inherently given by a noncommutative geometry. This means that the resulting discrete structure is represented by operator algebras and the differentiation is realized by a commutator with a selfadjoint operator \(D\). The results presented in this work corroborate previous observations and constructions [10], [11]. The dictionary between continuous and discrete settings is identical to the dictionary between classical geometry and noncommutative one. This dictionary is summarized in the following table.

Table 1: No caption
Structure Continuous Discrete
Space Smooth manifold \(M\) simplex or graph
Algebra Algebra of functions \(C^\infty(M)\) Operator algebra \(\mathfrak{B}\)
Differential Differential form \(df\) Commutator \([D,F]\)
Integration \(\int f(x)dx\) \(\mathrm{Tr}_{\omega}(F)\)

The Berezin-Toeplitz quantization is an archetypal example of deformation of a classical geometry given by a compact Kähler manifold \((M,g,\lbrace\cdot,\cdot\rbrace)\) to a noncommutative finite dimensional operator algebra \((M_n(\mathbb{C}),\left[ \cdot, \cdot \right] )\). Thus, it serves as a fundamental example of transition from a continuous to a discrete space. We use it in this work as a central application of our axiomatization in order to show that it satisfies the conditions of a structure preserving discretization.
The paper is organized as follows: Section 2 introduces the key concepts and axioms that form the foundation of our framework. Section 3 demonstrates how commutative diagrams can effectively formalize the preservation of structures and explains how noncommutative geometry emerges from the structure-preserving discretization of a differential algebra. Section 4 provides a comprehensive review of diverse examples of structure-preserving discretizations, showcasing the framework’s flexibility. Finally, Sections 5 and 6 apply these findings to the Berezin-Toeplitz quantization, highlighting its adherence to the proposed criteria; Section 6 being a walk through example on the sphere.
Through this work, we hope to contribute to a deeper understanding of discretization methods and to provide a foundation for future research in this area.

2 Axiomatization of the theory of discretizations↩︎

In the first section, we introduce our axiomatization of structure-preserving discretization. The axioms rely on Category theory tools that we introduce briefly.

2.1 Category theory↩︎

Let us recall that a category \(\mathcal{C}\) is defined by a class \(\mathrm{ob}(\mathcal{C})\) whose elements are called objects and a class \(\mathrm{hom}(\mathcal{C})\) whose elements are called morphisms or arrows. Each arrow \(f\) has a source object \(X\) and a target object \(Y\). A morphism \(f\) from \(X\) to \(Y\) is then denoted \(f:X\rightarrow Y\). The class of arrows \(\mathrm{hom}(\mathcal{C})\) is equipped with a binary operation called composition of arrows between two morphism \(f\) and \(g\): \[f:X\rightarrow Y, \quad g:Y\rightarrow Z \quad \text{define}\quad g\circ f:X\rightarrow Z\] such that \(f\circ g\) is an element of \(\mathrm{hom}(\mathcal{C})\). The composition is associative and there is exactly one identity morphism denoted \(1_X\) for every object \(X\) such that for every arrow \(f:X\rightarrow Y\): \[1_Y\circ f =f =f\circ f_X.\] One says that a morphism \(f:X\rightarrow Y\) is invertible and called an isomorphism (or more simply an invertible arrow) if there exists a morphism \(g:Y\rightarrow X\) such that \(f\circ g=1_Y\) and \(g\circ f=1_X\).

2.2 Axiomatization↩︎

In what will follow, we will always, unless stated otherwise, consider objects in the category of Banach spaces. Nevertheless, one should have in mind that many of the constructions can be extended to merely topological spaces.

Definition 1. A discretization \(\mathfrak{D}(f)\) of an arrow \((f:C_1\rightarrow C_2)\) is a sequence of arrows \((f_n:C_1^n\rightarrow C_2^n)_{n\in \mathbb{N}}\) producing the following diagram \[\begin{tikzcd} C_1 \arrow[r,"f"] \arrow[d,"\pi^n_1"'] & C_2 \arrow[d,"\pi^n_2"] \\ C^n_1 \arrow[r,"f_n"] & C^n_2 \end{tikzcd} \label{diagram}\tag{1}\] for each \(n\in \mathbb{N}\). The maps \(\pi^n_i:C_i\rightarrow C^n_i\), \(i=1,2\) are surjective and contractive linear maps for all \(n\in \mathbb{N}\). In addition, these maps satisfy the limit conditions: \[\lim_{n\rightarrow \infty}\|\pi_i^nx\|_{C^n_i}=\|x\|_{C_i}, \quad \text{for all x\in C_i where i=1,2.}\] A discretization \(\mathfrak{D}(f)\) is faithful if: \[(\text{f is an invertible arrow) \Longrightarrow (f_n\in \mathfrak{D}(f) is invertible for all n\in \mathbb{N}}).\] In the special case where the discretization \(\mathfrak{D}(f:C\rightarrow C)\) does not depend on the map \(f\), i.e. the maps \(\pi^n\) are independent of \(f\), then we say that we have a discretization of the object \(C\) and we write \(\mathfrak{D}(C)\).

Remark 2. The previous definitions generalize mutatis mutandis to the discretization \(\mathfrak{D}(\lbrace f_n\rbrace_{n\in \mathbb{N}})\) of a sequence of arrows \(\lbrace f_n\rbrace_{n\in \mathbb{N}}\) in the category \(\mathcal{C}\).

Definition 3 (Structure preservation). Let \(C_1\) and \(C_2\) be objects in a category \(\mathcal{C}\). Consider \(f\in \mathrm{hom}(C_1,C_2)\) an arrow in \(\mathcal{C}\) such that \((C_1,f)\) defines an object in a category \(\mathcal{B}\). A discretization \(\mathfrak{D}(f)\) is structure preserving if it satisfies the following conditions:

  • \((C^n_1,f_n)\in \mathrm{ob}(\mathcal{B})\) and \(C_1^n\in \mathrm{ob}(\mathcal{C})\) for all \(n\in \mathbb{N}\).

  • the diagram 1 commutes asymptotically: \[\|f_n\circ\pi_1^n(x)-\pi_2^n\circ f(x)\|\longrightarrow 0 \quad \text{as n\rightarrow \infty and for all x\in C_1}\] In such case, we will say that the discretization \(\mathfrak{D}(f)\) is consistent.

Example 4 (Euler method). Consider the algebra \(C^\infty(\mathbb{S}^1)\) of smooth functions over the circle. For every \(n\in \mathbb{N}\), let \(X_n=\lbrace x_1,x_2,\dots,x_n\rbrace\) be a finite collection of points on \(\mathbb{S}^1\) such that \(d(x_i,x_{i+1})=\frac{2\pi}{n},\) for \(i=1,2,\dots,n-1\) and where \(d\) is the Euclidean distance on the circle. Consider now the right-shift map on \(\mathbb{C}^n\) \[S:\mathbb{C}^n\rightarrow \mathbb{C}^n \quad S(y_1,y_2,\dots,y_n)=(y_n,y_1,y_2,\dots,y_{n-1})\] and define the family Euler operators by \[E_n=\frac{n}{2\pi}(S-1)\] Moreover, we define the maps \[\pi^n:C^\infty(\mathbb{S}^1)\rightarrow \mathrm{Hom}(X_n,\mathbb{C}^n)\qquad \pi^n(f)=(f(x_1),f(x_2),\cdots,f(x_n)).\] Consider now the differential operator \[\frac{d}{d\theta}:C^\infty(\mathbb{S}^1)\rightarrow C^\infty(\mathbb{S}^1) \qquad f\mapsto \frac{df}{d\theta}\] such that \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\) is an object of the category of differential algebra.
We can now look at the Euler discretization method of the pair \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\) defined by the following diagram: \[\begin{tikzcd} C^\infty(\mathbb{S}^1) \arrow[r,"\frac{d}{d\theta}"] \arrow[d,"\pi^n"'] & C^\infty(\mathbb{S}^1) \arrow[d,"\pi^n"] \\ \mathrm{Hom}(X_n,\mathbb{C}^n) \arrow[r,"E_n"] & \mathrm{Hom}(X_n,\mathbb{C}^n) \end{tikzcd}\] We readily verify \[\lim_{n\rightarrow \infty}\|\pi^n(f)\|_\infty = \|f\|_\infty \quad \text{and} \quad \|E_n\circ\pi^n(f)-\pi^n\circ \frac{d}{d\theta}(f)\| \longrightarrow 0\] and thus, the Euler method defines a discretization. Therefore, it would be tempting to conclude that the Euler discretization is a structure preserving discretization of the object \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\). However, this is not the case since the pair \((\mathrm{Hom}(X_n,\mathbb{C}^n),E_n)\) does not define a differential algebra, since \(E_n\) does not satisfy the Leibniz rule.

Remark 5. In fact, a structure-preserving discretization of the differential algebra \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\) necessitates the use of a Lie bracket in a finite dimensional context, which indeed satisfies the product rule and defines a derivation. This will be explored in the next section.

We can also require stronger conservation properties between the continuous object \(C_1\) and its discretization \(C_1^n\).

Definition 6 (Strongly structure preserving). Let \(C_1\) and \(C_2\) be objects in a category \(\mathcal{C}\). Consider \(f\in \mathrm{hom}(C_1,C_2)\) an arrow in \(\mathcal{C}\) such that \((C_1,f)\) defines an object in a category \(\mathcal{B}\). A discretization \(\mathfrak{D}(f)\) is strongly structure preserving if it satisfies the following conditions:

  • \((C^n_1,f_n)\in \mathrm{ob}(\mathcal{B})\) and \(C_1^n\in \mathrm{ob}(\mathcal{C})\) for all \(n\in \mathbb{N}\).

  • the diagram 1 commutes for all \(n\in \mathbb{N}\);

  • the maps \(\pi_i^n:C_i\rightarrow C_i^n\) are surjective homomorphisms for all \(n\in \mathbb{N}\) and for \(i=1,2\) ( i.e. \(\pi_i^n\in \mathrm{hom}(\mathcal{C})\) for all \(n\in \mathbb{N}\)).

Example 7. Let \(\mathcal{C}\) the category of prehilbert space and \(\mathcal{B}\) be the category of differential algebra. Then, for instance, the pair \((C^\infty(\mathbb{S}^1),\left\langle \cdot ,\cdot \right\rangle )\) consisting of the space smooth function on the circle with its usual scalar product, is an object of \(\mathcal{C}\); and the pair \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\) defining a differential algebra is an object of \(\mathcal{B}\).
A strongly structure preserving discretization of the differential algebra \((C^\infty(\mathbb{S}^1),\frac{d}{d\theta})\) (if it exists) is a commutative diagram: \[\begin{tikzcd} (C^\infty(\mathbb{S}^1),\left\langle \cdot ,\cdot \right\rangle ) \arrow[r,"\frac{d}{d\theta}"] \arrow[d,"\pi^n"'] & (C^\infty(\mathbb{S}^1),\left\langle \cdot ,\cdot \right\rangle ) \arrow[d,"\pi^n"] \\ (\mathcal{A}_n ,\left\langle \cdot ,\cdot \right\rangle_n )\arrow[r,"d_n"] & (\mathcal{A}_n ,\left\langle \cdot ,\cdot \right\rangle_n ) \end{tikzcd}\] such that:

  • \((\mathcal{A}_n,d_n)\) is a differential algebra;

  • \(\frac{d}{d\theta}\circ \pi^n = \pi^n\circ d_n\) for all \(n\);

  • \((\mathcal{A}_n ,\left\langle \cdot ,\cdot \right\rangle_n )\) is a prehilbert space and \(\pi^n\) a partial isometry (i.e. an isometry on its support) that is: \[\left\langle \pi^nx,\pi^ny\right\rangle_n = \left\langle x,y\right\rangle\] for \(x,y\) in the support.

2.3 Convergence↩︎

Consider a discretization \(\mathfrak{D}(f)\) of an arrow \(f\). Heuristically, the convergence of the discretization means that \((C_1^n,f_n)\) converges to \((C_1,f)\). Hence, to properly define a notion of convergence, one needs to compare elements between different levels of dicretization.
Let us recall that, the graph of an arrow \(f\), denoted \(\mathrm{Gr}(f)\), is defined as the subset of \(C_1\times C_2\) such that: \[\mathrm{Gr}(f)=\left\lbrace (x,y)\in C_1\times C_2 : y=f(x) \right\rbrace\] We denote by \(p_i:\mathrm{Gr}(f)\rightarrow C_i\) for \(i=1,2\) the obvious coordinate projections.

Definition 8 (Convergence). Consider a discretization \(\mathfrak{D}(f)\) of an arrow \(f\). We say that \(\mathfrak{D}(f)\) is convergent if to any of the projection maps \(\pi^n_i:C_i\rightarrow C^n_i\), we can associate an injective contractive linear map \(s_i^n:C_i^n\rightarrow C_i\) such that \[\lim_{n\rightarrow +\infty}\|x-s_i^n\circ \pi^n_i(x)\|=0, \quad \text{for all x\in p_i(\mathrm{Gr}(f)), and for i=1,2.}\] The maps \(s_i^n\) will be called section map.

For a structure preserving discretization \(\mathfrak{D}(f)\), one can obtain an equivalent definition of convergence.

Proposition 9. Consider a structure preserving discretization \(\mathfrak{D}(f)\) of an arrow \(f\). The discretization \(\mathfrak{D}(f)\) is convergent if and only if to any of the projection maps \(\pi^n_i:C_i\rightarrow C^n_i\), we can associate an injective contractive linear map \(s_i^n:C_i^n\rightarrow C_i\) such that \[\lim_{n\rightarrow +\infty}\|x-s_1^n\circ \pi^n_1(x)\|=0, \qquad \text{and} \qquad \lim_{n\rightarrow +\infty}\|f(x)-s_2^n\circ f_n (\pi^n_1(x))\|=0\]

Proof. We prove the if part of the proof, the only if can be proven identically. Assume that section maps exist, then one get the following inequality \[\|f(x)-s_{2}^n\circ \pi_{2}^n(f(x)))\|\leq\|f(x)-s_{2}^n\circ f_n(\pi_1 ^n(x))\| +\|s_2^n(\pi_2^n\circ f(x)- f_n\circ\pi_1^n(x))\|\] where the first term in the right-hand-side goes to zero by assumption and the second term also vanishes from the structure preserving condition. ◻

If we assume that the arrow \(f\) is also invertible, then one can deduce the following corollary relating convergence of \(\mathfrak{D}(f)\) and \(\mathfrak{D}(f^{-1})\).

Corollary 10. Let \(f\) be an invertible arrow, such that the discretizations \(\mathfrak{D}(f)\) and \(\mathfrak{D}(f^{-1})\) are structure preserving and faithful. Then, \(\mathfrak{D}(f)\), respectively \(\mathfrak{D}(f^{-1})\), converges if \[\lim_{n\rightarrow +\infty}\|f(x)-s_2^n\circ f^{-1}_n (\pi^n_1(x))\|=0, \qquad \text{and} \qquad \lim_{n\rightarrow +\infty}\|f(x)-s_2^n\circ f_n (\pi^n_1(x))\|=0\]

3 Derivations and Noncommutative geometry↩︎

3.1 Derivations↩︎

The central focus of discretization theory is the study of differential equations derived from the differential structure of a manifold. In the following paragraph, we aim to demonstrate how the structure-preserving discretization of a smooth differential operator inevitably results in a noncommutative differential structure.
A differential structure means here the data of a unital algebra of smooth functions \(A\) and a derivation; we recall that a derivation \(d:A\rightarrow M\) is map such that:

  • \(d\) is linear,

  • \(d(1)=0\),

  • \(d(ab)=d(a)b + ad(b)\).

where \(M\) is a \(A\)-bimodule. The data \((A,d)\) is called a differential algebra and constitute one of the fundamental structure that one may want to preserve under discretization. Consider a structure preserving discretization \((A_n,d_n)\) of a given differential algebra \((A,d)\) given by the following diagram:

Proposition 11. If \(A_n\) is isomorphic to a matrix algebra then there exists a self-adjoint element \(D_n\) such that \[d_n(a)=\left[ D_n,a\right]\] In addition, if the discretization \((A_n,d_n)\) is structure preserving, then \[\lim_{n\rightarrow \infty}\|\pi^n\circ d(a)-\left[ D_n,\pi^n(a)\right] \| = 0\]

Proof. The claim immediately follows from Theorem 4.1.6 in [12] and the axiom of a structure preserving discretization. ◻

Remark 12. Structure preserving discretizations \((A_n,d_n)\) of differential algebra are also called approximately inner derivations in the theory of Operator Algebras, see for instance [13]. The last proposition provides a first bridge between application of \(C^*\)-algebra theory (or operator theory) and structure preserving discretizations.

We have established that one approximate an differential algebra \((A,d)\) of functions by a matrix algebra, then the discrete differential structure is inherently noncommutative. We now show a concrete case where such approximation exists.

Definition 13. A filtration of a Hilbert space \(\mathcal{H}\) is a sequence \(\mathcal{F}=\left\lbrace \mathcal{H}_1,\mathcal{H}_2,\dots \right\rbrace\) of finite dimensional subspaces of \(\mathcal{H}\) such that \(\mathcal{H}_n\subset\mathcal{H}_{n+1}\) and \[\mathcal{H}=\overline{\bigcup_{n=1}^\infty \mathcal{H}_{n}}\] Let \(\mathcal{F}=\left\lbrace \mathcal{H}_n\right\rbrace\) be a filtration of \(\mathcal{H}\) and let \(P_n\) be the projection onto \(\mathcal{H}_n\). The degree of an operator \(D\) is defined by \[\mathrm{deg}(D)=\sup_{n\geq 1}\mathrm{rank}(P_nD-DP_n)\]

Therefore, one sees that if the domain \(\mathcal{H}\) of a continuous differential operator \(D\) admits a filtration family \(\lbrace\mathcal{H}_n\rbrace\), then we can construct the following diagram of discretization: \[\begin{tikzcd} \mathcal{H} \arrow[r,"D"] \arrow[d,"P_n"'] & \mathcal{H} \arrow[d,"P_n"] \\ \mathcal{H}_n \arrow[r,"D_n"] & \mathcal{H}_n \end{tikzcd} \label{diagram2}\tag{2}\] Thus, the lack of commutativity of the previous diagram i.e. the existence of a structure preserving discretization, in this case is measured by the degree \(\mathrm{deg}(D)\) of the operator \(D\). We can then restrict ourselves to subclass of operators with vanishing degrees.

Definition 14 (Block diagonal operator). A linear operator \(D\) on a Hilbert space \(H\) is called block diagonal if there exists an increasing sequence of finite rank projections, \(P_1\leq P_2\leq P_3\leq \cdots\) such \(\|[D,P_n]\|=\|DP_n-P_nD\|=0\) for all \(n\in \mathbb{N}\) and \(P_n\rightarrow 1_{\mathcal{H}}\) (in the strong operator topology) as \(n\rightarrow \infty\).

Recall that two (possibly unbounded) self-adjoint operators \(A\) and \(B\) are said to commute if and only if all the projections in their associated projection-valued measures commute. Hence, a trivial example of increasing sequence of commuting projections is given by spectral projectors.
Similarly to the axioms of structure preserving discretization, the block diagonal condition might be too restrictive; one can then relax the commutativity condition.

Definition 15. A linear operator \(D\) on a Hilbert space \(H\) is called quasidiagonal if there exists an increasing sequence of finite rank projections, \(P_1\leq P_2\leq P_3\leq \cdots\) such \(\|[D,P_n]\|=\|DP_n-P_nD\|\rightarrow 0\) and \(P_n\rightarrow 1_{\mathcal{H}}\) (in the strong operator topology) as \(n\rightarrow \infty\).

We can now state a theorem on existence of structure preserving discretization.

Theorem 16 (First Existence Theorem). Consider a differential algebra of functions \((A,d)\) such that \(da=Da-aD\) such that \(D\) is a quasidiagonal (not necessarily bounded) operator, then \((A,d)\) admits a structure preserving discretization. If in addition \(D\) is block diagonal, then \((A,d)\) admits a strongly structure preserving discretization.

Proof. Since \(d\) is a quasidiagonal derivation in an algebra of functions \(A\), there exists a self-adjoint operator \(D\) such that \(da=[D,a]:=Da-aD\) and a family of finite rank projections such that \(\|[D,p_n]\|\rightarrow 0\). Then, we can define the surjective contractions: \[\pi^n:A\rightarrow A_n, \qquad \pi^n(a)=p_nap_n\] and thus satisfy \[\pi^n(da)=p_n[D,a]p_n=[p_nDp_n,\pi^n(a)]+(p_n[D,p_n]ap_n-p_na[D,p_n]p_n).\] Therefore, we deduce the following upper-bound \[\|\pi^n(da)-[D_n,\pi^n(a)]\| \leq \|[D,p_n]\| \|a\|\rightarrow 0\] which implies the structure-preserving condition.
If \(D\) is block-diagonal, then it admits a family \(\lbrace p_n\rbrace\) of finite rank projectors such that \(p_nD=p_nD\). Therefore, one has \[\pi^n(da)=p_n[D,a]p_n=[D_n,\pi^n(a)]=d_n\pi_n(a).\] i.e. the discretization is strongly structure preserving. ◻

Remark 17. From the Existence Theorem, we see that a structure preserving discretization of a differential algebra, when it exists, is given by the data of \((A_n,H_n,D_n)\) where \(A_n\) is a matrix algebra, \(H_n\) is its representation space and \(D_n\) is a selfadjoint operator defining the differential structure. This is the general setting of noncommutative geometry [14].

3.2 \(C^*\)-algebras as a general setting for structure preserving discretizations↩︎

In fact, most cases of discretization and existence of projection maps \(\pi^n\) and section maps \(s^n\) fit it in the general framework of \(C^*\)-algebras. Let us briefly recall that a \(C^*\)-algebra is a Banach algebra equipped with an involution \(*\), compatible with the norm: \(\|x^*x\|=\|x\|^2\).

Definition 18 (Nuclear \(C^*\)-algebra). A \(C^*\)-algebra \(A\) is said to be nuclear if there exists a net of natural numbers \((n_\lambda)_{\mathrm{\Lambda}}\) and nets of contractive, completely positive maps \(\varphi_\lambda :A\rightarrow M_{n_\lambda}(\mathbb{C})\) and \(\psi_\lambda : M_{n_\lambda}(\mathbb{C}) \rightarrow A\) such that \(\lim_{\mathrm{\Lambda}}\|a-\psi_\lambda(\varphi_\lambda(a))\|=0\) for every \(a\in A\).

Theorem 19 (Second Existence Theorem). Consider a pair \((A,f)\) such that \(A\) is also a nuclear \(C^*\)-algebra. There always exists a discretization \(\mathfrak{D}(A,f)\) of \((A,f)\). Moreover, if \((A,f)\) is considered as an object of the category of linear spaces, then a strongly structure preserving discretization always exists.

Proof. Since \(A\) is a nuclear \(C^*\)-algebra, there exist discretization maps \(\varphi_n:A\rightarrow M_{n}(\mathbb{C})\) and section maps \(\psi_n:M_{n}(\mathbb{C}) \rightarrow A\). We need to verify that \(\lim_{n\rightarrow \infty}\|\varphi(a)\|=\|a\|\); the limit follows from the fact that \(\lim_{n\rightarrow \infty}\|a-\psi_n\circ\phi_n(a)\|=0\) and the contraction property and: \[\|\psi_n\circ\varphi_n(a)\|\leq \|\varphi_n(a)\|\leq \|a\|\] Therefore, by taking the limit we obtain that \(\lim_{n\rightarrow\infty}\|\varphi_n(a)\|=\|a\|\). We can construct the following discretization diagram: \[\begin{tikzcd} A \arrow[r,"f"] \arrow[d,"\varphi^n"'] & A \arrow[d,"\varphi^n"] \\ M_n(\mathbb{C}) \arrow[r,"f_n"] & M_n(\mathbb{C}) \end{tikzcd}\] where the discrete arrow \((f_n)\) is defined by: \[f_n(\pi^n(a))=\pi^n(f(a)).\] ◻

Remark 20. It turns out that all commutative \(C^*\)-algebra \(C(X)\), of continuous functions on a compact space \(X\), are nuclear. More generally, if \(A\) and \(B\) are nuclear \(C^*\)-algebras then \(A\otimes B\) is a nuclear \(C^*\)-algebra. In particular, if \(A=C(X)\) then \(A\otimes B\) can be identified with \(C(X,B)\) i.e. the \(C^*\)-algebra of continuous functions with values in \(B\), under the map \((f\otimes b)(x)=f(x)b\).

Example 21. A fundamental example is given by the space of continuous differential forms on a manifold \(X\) of dimension \(n\). Recall that a continuous differential \(r\)-form is a map \[\omega:X\rightarrow \mathrm{\Lambda}^r\mathbb{R}^n, \quad x\mapsto \sum_{|I|=r}\omega_I(x)dx^I\] where the functions \(\omega_I\) are continuous. If we identify the exterior algebra \(\mathrm{\Lambda}\mathbb{R}^n\) with the Clifford algebra \(\mathrm{Cl}(\mathbb{R}^n)\) then it acquires a \(C^*\)-algebra structre. The space of continuous differential forms is then isomorphic as a \(C^*\)-algebra to \(C(X,\mathrm{Cl}(\mathbb{R}^n))\). Since the Clifford algebras are nuclear \(C^*\)-algebras, then the space of continuous differential forms is also a nuclear \(C^*\)-algebra and therefore admits a discretization according to the previous theorem.

4 Zoology of structure preserving discretizations↩︎

In the following section, we review archetypical examples of structure preserving discretizations through the lens of our axiomatization.

4.1 Finite element exterior calculus↩︎

We start by recalling some of the basic definitions of finite element exterior calculus.

Definition 22. Let \(W^1,W^2,\dots\) be Hilbert spaces and let \(d^k:W^k\rightarrow W^{k+1}\) be densely-defined closed linear operators satisfying

  • \(\mathrm{Im}(d^k)\subseteq \mathrm{Dom}(d^{k+1})\) and

  • \(d^{k+1}d^k=0\)

for all \(k\). Then, we say that the data \(\lbrace (W^k,d^k)\rbrace_{k}\) define a Hilbert complex. The operators \(d^k\) are called the differentials of the complex. If all the differentials are bounded, the Hilbert complex is said to be bounded.

Definition 23. A morphism of bounded Hilbert complexes \(\Phi:(V,d)\rightarrow (\tilde{V},\tilde{d})\) is a set of linear maps \(\Phi^k:V^k\rightarrow \tilde{V}^k\) such that \(\tilde{d}\Phi^k=\Phi^{k+1}d^k\) for all \(k\). We indicate that \(\Phi\) is a morphism by writing \(\Phi \in \mathrm{hom}((V,d),(\tilde{V},\tilde{d}))\).

The framework of the finite element exterior calculus is given by the \(L^2\)-de Rham complex, represented by the following complex \[0\rightarrow \mathrm{H\Lambda}^0(\mathrm{\Omega})\xrightarrow[]{d}\mathrm{H\Lambda}^1(\mathrm{\Omega})\xrightarrow[]{d}\cdots \xrightarrow[]{d}\mathrm{H\Lambda}^n(\mathrm{\Omega})\rightarrow 0\] where \(\mathrm{\Omega}\subseteq \mathbb{R}^n\) is a Lipschitz domain. In order to give a numerical approximation of a PDE, the method builds a finite-dimensional subcomplex \[0\rightarrow \mathrm{\Lambda}_\hbar^1(\mathrm{\Omega})\xrightarrow[]{d}\cdots \xrightarrow[]{d}\mathrm{\Lambda}_\hbar^n(\mathrm{\Omega})\rightarrow 0\] of the de Rham complex. In order to construct this subcomplex, one shows that there exists morphism \(\pi^\hbar\) projecting the de Rham complex down to the appropriate subcomplex; so that each map \(\pi^\hbar_k:\mathrm{H\Lambda}^k\rightarrow \mathrm{\Lambda}^k_\hbar\) is a projection onto the subspace \(\mathrm{\Lambda}_\hbar^k\) and induces the following commuting diagram \[\begin{tikzcd} H\mathrm{\Lambda}^k \arrow[r,"d"] \arrow[d,"\pi^\hbar_k"'] & H\mathrm{\Lambda}^{k+1} \arrow[d,"\pi^\hbar_{k+1}"] \\ \mathrm{\Lambda}_\hbar^k \arrow[r,"d"] & \mathrm{\Lambda}_\hbar^{k+1} \end{tikzcd}\] In this case, the section maps \(s_k^{\hbar}\) are simply identity maps. Therefore, if we consider \(\lbrace H\mathrm{\Lambda}^k,d^k\rbrace\) as an object of the category of Hilbert complexes, then the data \((\lbrace \mathrm{\Lambda}^k_{\hbar},d\rbrace, \pi^\hbar_k)\) gives a strongly structure preserving discretization.

Proposition 24. The Finite Element Exterior Calculus is a strongly structure preserving discretization.

Proof. We verify that the discretization \((\lbrace \mathrm{\Lambda}^k_{\hbar},d\rbrace, \pi^\hbar_k)\) satisfies the axiom of a structure preserving discretization of the continuous data \(\lbrace H\mathrm{\Lambda}^k,d^k\rbrace\) seen as an object of the category of Hilbert complexes. The following three conditions are fulfilled:

  • \(\lbrace \mathrm{\Lambda}^k_{\hbar},d\rbrace\) is a differential complex

  • the diagram commutes for all \(n\): \[d\circ \pi^{\hbar}_k=\pi^{\hbar}_{k+1}\circ d\]

  • the maps \(\pi^\hbar_{k}\) are bounded maps between normed spaces.

from which the statement follows. ◻

4.2 Transfer Operator and diffeomorphism group↩︎

In the following paragraph, we would like to illustrate how our axiomatization allows us to describe structure-preseving discretizations of diffeomorphism groups; we prove on practical examples how infinite group of symmetries are related to matrix groups.
Consider a smooth compact manifold \(X\) and its diffeomorphism group \(\mathrm{Diff}(X)\). In this section, we identify structure preserving discretization of diffeomorphism groups which play a central role in PDE theory.
The group \(\mathrm{Diff}(X)\) admits a natural action on the Hilbert space \(L^2(X)\), the space of square-summable half-densities on \(X\); by choosing a smooth measure \(m\) on \(X\) one can identify \(L^2(X)\) with the usual space of functions \(f\) on \(X\) which are square-summable with respect to \(m\). Hence, \(\mathrm{Diff}(X)\) admits an obvious unitary representation on \(H\) induced by its action on \(X\): for any \(\psi \in \mathrm{Diff}(X)\), define \[\widehat{\psi}:H\rightarrow H \qquad \widehat{\psi}(f)(x):= J_\psi(x)^{\frac{1}{2}}f(\psi^{-1}x)\] where \(J_\psi(x)=dm(\psi^{-1}x)/dm(x)\). We can restrict this representation to an operator \(\widehat{\psi}:C^\infty(X)\rightarrow C^\infty(X)\) and study the possible structure preserving discretization of \(\mathrm{GL}(C^\infty(X))\).
Fix a diffeomorphism \(\psi\), in order to present a structure preserving discretization of the pair \((C^\infty(X), \widehat{\psi})\), we need to specify in which category the object \((C^\infty(X), \widehat{\psi})\) belongs to: we will consider it as an object of the category of linear spaces. Therefore, following the definition, a structure preserving discretization is given by a commuting diagram \[\begin{tikzcd} C^\infty(X) \arrow[r,"\widehat{\psi}"] \arrow[d,"\pi^n"'] & C^\infty(X) \arrow[d,"\pi^n"] \\ A_N\arrow[r,"\widehat{\psi}_N"] & A_N \end{tikzcd} \label{diagram3}\tag{3}\] such that \(\lim_{n\rightarrow \infty}\|\pi^n(f)\|_\infty = \|f\|_\infty\) and \((A_N,\widehat{\psi}_N)\) is a pair of a vector space \(A_N\) and a linear map \(\widehat{\psi}_N\).

Lemma 25. There exists a family of surjective maps \(\pi^n:C^\infty(X)\rightarrow \mathbb{C}^{N}\) such that \(\lim_{n\rightarrow \infty}\|\pi^n(f)\|_\infty = \|f\|_\infty\) for all \(f \in C^\infty(X)\).

Proof. The algebra \(C(X)\) of continuous functions is a nuclear abelian \(C^*\)-algebra, therefore it admits a discretization map \(\pi^n:C(X)\rightarrow \mathbb{C}^{N}\); we restrict this map to \(C^\infty(X)\). Moreover, the limit \(\lim_{n\rightarrow \infty}\|\pi^n(f)\|=\|f\|\) follows from the Second Existence Theorem. ◻

Therefore, diagram 3 becomes \[\begin{tikzcd} C^\infty(X) \arrow[r,"\widehat{\psi}"] \arrow[d,"\pi^n"'] & C^\infty(X) \arrow[d,"\pi^n"] \\ \mathbb{C}^N\arrow[r,"\widehat{\psi}_n"] & \mathbb{C}^N \end{tikzcd}\] and the existence of \(\widehat{\psi}_n\) is given by the Second Existence Theorem. We can then state the following theorem.

Theorem 26. For any diffeomorphism \(\psi\in \mathrm{Diff}(X)\), there exists a faithful, strongly structure preserving discretization of \(\mathfrak{D}(C^\infty(X),\widehat{\psi})\).

Proof. Following the proof of Theorem [2nd], we define \(\widehat{\psi}_n\) by \[\widehat{\psi}_n\circ \pi^n(f)=\pi^n\circ \widehat{\psi}(f) \label{identity}\tag{4}\] which gives us a linear surjection \(\widehat{\psi}_n:\mathbb{C}^N\rightarrow \mathbb{C}^N\), since \(\pi^n\) and \(\widehat{\psi}\) are surjective; and thus \(\widehat{\psi}_n\) is an isomorphism. ◻

Remark 27. Since the previous discretization does not depend on the choice of diffeomorphism, we will abuse notation and denote by \[\mathfrak{D}(\mathrm{Diff}(X))=\mathrm{GL}_N(\mathbb{C}).\] to say that a discretization of the diffeomeorphism group \(\mathrm{Diff}(X)\) is given by the finite dimensional group of invertible matrices \(\mathrm{GL}_N(\mathbb{C})\).

One could also be interested in subgroups of the diffeomorphism group. For instance, if the manifold \(X\) is equipped with a volume form \(d\omega\), we can define the scalar product between two function: \[\left\langle f,g \right\rangle =\int_X f\overline{g} \;d\omega\] Then, one can define the subgroup of diffeomorphisms preserving this inner product and denoted it by \(\mathrm{Diff}^+_0(X)\).
One can also define the subgroup of markovian diffeomorphism, i.e. the group of diffeomorphism \(\psi\) satisfying \(\psi(1)=1\). We denote this subgroup by \(\mathfrak{D}(\mathrm{Diff}_{\mathrm{m}}(X))\).
Let us denote by \(\mathrm{GL}^{\mathrm{st}}_N(\mathbb{C})\), the group of invertible stochastic matrices.

Proposition 28. Let \((A_N,\widehat{\psi}_N)\) be a faithful strongly structure preserving discretization of the pair \((C^\infty(X), \psi)\) independent of the choice of \(\psi\). If

  • \(C^\infty(X)\) is merely a vector space, then \(\mathfrak{D}(\mathrm{Diff}(X))=\mathrm{GL}_N(\mathbb{C}).\)

  • \((C^\infty(X),\left\langle \cdot,\cdot \right\rangle )\) is a prehilbert space, then \(\mathfrak{D}(\mathrm{Diff}_0^{+}(X))=\mathrm{SO}_N(\mathbb{C}).\)

  • \((C^\infty(X),1)\) is a vector space with a unit element (i.e. containing its scalar field), then \(\mathfrak{D}(\mathrm{Diff}_{\mathrm{m}}(X))=\mathrm{GL}^{\mathrm{st}}_N(\mathbb{C}).\)

Proof. i) follows from Theorem [thmdiff].
ii) If \((A_N,\widehat{\psi}_N)\) is a strongly structure preserving discretization of the pair \((C^\infty(X), \psi)\), then \((A_N, \left\langle \cdot , \cdot \right\rangle_N )\) has a prehilbert structure as well. Moreover, the map \(\pi^N:C^\infty(X)\rightarrow A_N\) is an isometry on its support. Therefore, we have that for any \(f,g\in C^\infty(X)\) \[\left\langle \widehat{\psi}_N\circ \pi^N(f), \widehat{\psi}_N\circ \pi^N(g) \right\rangle_N = \left\langle \pi^N\circ \widehat{\psi}(f) ,\pi^N\circ \widehat{\psi}(g) \right\rangle_N= \left\langle f,g \right\rangle .\] The statement follows from the surjectivity of \(\widehat{\psi}\) and \(\pi^n\).
iii) If \((A_N,\widehat{\psi}_N)\) is a strongly structure preserving discretization of the pair \((C^\infty(X), 1)\), then \((A_N, 1_N)\) has also a unit element. Moreover, the map \(\pi^N:C^\infty(X)\rightarrow A_N\) is unital, i.e. \(\pi^n(1)=1_N\). Using the commutativity axiom, we get, \[\widehat{\psi}_N(1_N)=\widehat{\psi}_N\circ \pi^N(1)=\pi^N\circ \widehat{\psi}(1)=1_N.\] Therefore, the discretization \(\widehat{\psi}_N\) preserves the unit. Because \(\widehat{\psi}_N\) is invertible (from statement i)) then it is a left and right stochastic linear map. ◻

In addition, we notice that a structure preserving diffeomorphism discretization of a diffeomorphism group induces a structure preserving discretization of the associated Lie algebra. Let us recall that the pullback representation gives the map \[\mathrm{Diff}(X)\rightarrow \mathrm{Lin}(C^\infty(X)) \quad \psi\mapsto (f\mapsto \widehat{\psi}(f))\] Differentiating it at the identity \(\psi=\mathrm{id}\) gives a representation of the algebra of vector fields \(\mathrm{Vect}(X)\) over the space \(X\) into the space of derivations: \[\mathrm{Vect}(X)\rightarrow \mathrm{Der}(C^\infty(X)) \quad \Psi\mapsto \mathcal{L}_\Psi\] Here the operator \(\mathcal{L}_\psi:C^\infty(X)\rightarrow C^\infty(X)\) is given by the derivative of a function \(f\) in the direction of the vector field \(X\), i.e. \[\mathcal{L}_\Psi f:=\left. \frac{d}{dt}\right|_{t=0}\widehat{\psi}_t(f),\] where \(\psi^t\) denotes the flow of \(\Psi\). Moreover, one can show that the map is a Lie-algebra anti-homomorphism: \[\mathcal{L}_{[\Psi,\Phi]}=\mathcal{L}_\Phi\mathcal{L}_\Psi-\mathcal{L}_\Psi\mathcal{L}_\Phi\] for any vector fields \(\Psi,\Phi\in \mathrm{Vect}(X)\).
Hence, using the structure preserving condition 4 and differentiating Diagram 3 at the identity we obtain the new diagram: \[\begin{tikzcd} C^\infty(X) \arrow[r,"\mathcal{L}_\Psi"] \arrow[d,"\pi^n"'] & C^\infty(X) \arrow[d,"\pi^n"] \\ A_N\arrow[r,"\mathcal{L}_{\Psi_N}"] & A_N \end{tikzcd}\] where \(\mathcal{L}_\psi\) and \(\mathcal{L}_{\psi_N}\) are derivations.

Proposition 29. A structure preserving discretization of the diffeomorphism group \(\mathrm{Diff}(X)\) induces a structure preserving discretization of the Lie algebra \((\mathrm{Vect}(X),\mathcal{L}_\psi)\).

  • \(C^\infty(X)\) is merely a vector space, \(\mathfrak{D}(\mathrm{Vect}(X))=\mathfrak{gl}_N(\mathbb{C}).\)

  • \((C^\infty(X),\left\langle \cdot,\cdot \right\rangle )\) is a prehilbert space, then \(\mathfrak{D}(\mathrm{Vect}_0^{+}(X))=\mathfrak{so}_N(\mathbb{C}).\)

  • \((C^\infty(X),1)\) is a vector space with a unit element (i.e. containing its scalar field), then \(\mathfrak{D}(\mathrm{Vect}_{\mathrm{m}}(X))=\mathfrak{gl}_N(\mathbb{C}).\)

4.3 Berezin-Toeplitz Quantization↩︎

The Berezin-Toeplitz quantization bridges the gap between classical and quantum mechanics by transitioning from the Poisson algebra of smooth functions on a Kähler manifold to a finite-dimensional, noncommutative matrix algebra. This quantization method maintains essential geometric and algebraic properties, making it an ideal framework for discretizing continuous structures while ensuring structural integrity. In this section, the detailed process of Berezin-Toeplitz quantization is presented, and we demonstrate its role in the discretization of compact manifolds. We prove from first principle that it is a structure preserving method. It also provides a significant example to understand how continuous spaces are represented discretely in noncommutative settings.
The natural setting of the Berezin-Toeplitz quantization is a compact Kähler manifold \((M,\omega)\). Whereas the basic objects can be defined on an arbitrary Kähler manifold, some proofs rely on compactness to avoid technicalities. For this reason, we restrict ourselves to the compact setting.
A Kähler manifold \((M,\omega)\) is called quantizable if there exists an associated quantum line bundle \((L,h,\nabla)\), where \(L\) is a holomorphic line bundle \(L\) over \(M\) (i.e. locally a one-dimensional vector space), \(h\) a Hermitian metric on \(L\), and \(\nabla\) a connection compatible with the metric \(h\) and the complex structure.
We consider the space of holomorphic sections \(\mathrm{\Gamma}_{hol}(M,L)\) with value in the line bundle \(L\). The bundle \(L\) will be the stage for a finite dimensional quantization. For this, one needs to define the appropriate Hilbert space from \(\mathrm{\Gamma}_{hol}(M,L)\). the scalar product on the space of sections and the norm are given by \[\left\langle \varphi,\psi \right\rangle :=\int_Mh(\varphi,\psi)\mathrm{\Omega}, \quad \|\varphi\|:=\sqrt{\left\langle \varphi,\varphi \right\rangle },\] where \(\mathrm{\Omega}=\frac{1}{n!}\omega^{\wedge n}\) is the Liouville form, used as a volume form on \(M\). Let \(\mathrm{L}^2(M,L)\) be the \(\mathrm{L}^2\)-completion of \(\mathrm{\Gamma}_\infty(M,L)\), and \(\mathrm{\Gamma}_{hol}(M,L)\) be its finite dimensional (due to compactness) subspace of holomorphic sections. Let \(\mathrm{\Pi}^{(1)}:\mathrm{L}^2(M,L)\rightarrow \mathrm{\Gamma}_{hol}(M,L)\) be the projection. For \(f\in C^\infty(M)\), the Toeplitz operator \(T^{(1)}_f\) is defined to be \[T^{(1)}_f:=\mathrm{\Pi}^{(1)}(f\cdot): \quad \mathrm{\Gamma}_{hol}(M,L) \rightarrow \mathrm{\Gamma}_{hol}(M,L)\] The linear map \[T:C^\infty(M)\rightarrow \mathrm{End}\left(\mathrm{\Gamma}_{hol}(M,L) \right), \qquad f,\mapsto T_f\] is the Berezin-Toeplitz quantization map. In general, because \[T_fT_g\neq T_{fg}\] the quantization map is neither a Lie algebra homomorphism nor an associative algebra homomorphism.
Moreover, the Berezin-Toeplitz quantization map from the commutative algebra of functions to a noncommutative finite-dimensional matrix algebra. The finite dimensionality implies the loss of a lot of classical information. In order to recover this information, one should consider not just the bundle \((L,\nabla, h)\) alone but all its tensor powers \((L^{\otimes m},\nabla^{(m)},h^{(m)})\) and apply the construction for every \(m\in \mathbb{N}_0\). In this way, the direct sum of the spaces of holomorphic sections of \((L^{\otimes m},\nabla^{(m)},h^{(m)})\) gets identified with a Hilbert subspace \(\mathcal{H}\) called generalized Hardy space.
Thus, we have the following decomposition: \[\mathcal{H}=\bigoplus_m(\mathcal{H}_m,\left\langle \cdot,\cdot \right\rangle_m),\qquad \mathcal{H}_m=\mathrm{\Gamma}_{hol}(M,L^m).\] where we have defined the modified measure \[\left\langle \varphi,\psi \right\rangle_m :=\int_Mh(\varphi,\psi)\mathrm{\Omega}^{(m)}_\epsilon,\qquad \text{where}\quad\mathrm{\Omega}^{(m)}_\epsilon:=\epsilon^{(m)}(x)\mathrm{\Omega}(x)\] with \(\epsilon^{(m)}\) the Rawnsley’s epsilon function.

Definition 30 (Toeplitz operator). For \(f\in C^\infty(M)\) the Toeplitz operator \(T_f^{(m)}\) (of level \(m\)) is defined by \[T^{(m)}_f:=\mathrm{\Pi}^{(m)}f\mathrm{\Pi}^{(m)}: \mathrm{\Gamma}_{hol}(M,L^m)\rightarrow \mathrm{\Gamma}_{hol}(M,L^m)\]

This infinite family should in some sense approximate the algebra \(C^\infty(M)\); one obtains a family of matrix algebras and maps \[T^{(m)}:C^\infty(M) \rightarrow \mathrm{End}(\mathrm{\Gamma}_{hol}(M,L^m)), \qquad f\mapsto T_f^{(m)}=\mathrm{\Pi}^{(m)}(f\cdot), m\in \mathbb{N}_0\]

Proposition 31 ([15]). The Toeplitz map \[C^\infty(M)\rightarrow End(\mathcal{H}_m), \quad f\mapsto T^{(m)}(f)\] is surjective.

Definition 32 (Berezin-Toeplitz quantization map). The Berezin-Toeplitz quantization map is the map \[\mathrm{\Lambda} : C^\infty(M)\rightarrow \prod_{m\in \mathbb{N}_0}\text{End}\left( \mathrm{\Gamma}_{hol}\left( M,L^{(m)}\right) \right) , \quad f\rightarrow T_f:=\left( T_f^{(m)}\right)_{m\in \mathbb{N}_0}\]

In other words, for \(f\in C^\infty(M)\) we have the decomposition \[T_f=\prod_{m=0}^\infty T_f^{(m)}\] where \(T_f^{(m)}\) are exactly the restriction of \(T_f\) to \(\mathcal{H}^{(m)}\).

Theorem 33 (Bordemann, Meinrenken, Schlichenmaier, [15]).  
(a) For every \(f\in C^\infty(M)\) there exists a \(C>0\) such that \[\|f\|_\infty -\frac{C}{m}\leq \|T^{(m)}_f\|\leq \|f\|_\infty\] In particular, \(\lim_{m\rightarrow +\infty}\|T_f^{(m)}\|=\|f\|_\infty\).
(b) For every, \(f,g\in C^\infty(M)\) \[\|im[T_f^{(m)},T_g^{(m)}]-T^{(m)}_{\left\lbrace f,g\right\rbrace }\|=O(m^{-1})\]

Definition 34. The covariant Berezin symbol \(\sigma^{(m)}\) (of level \(m\)) is defined by \[\sigma^{(m)}:\mathrm{End}(\mathrm{\Gamma}_{hol}(M,L^{(m)}))\rightarrow C^\infty(M), \quad A\mapsto x\mapsto \sigma^{(m)}(A)(x):=\frac{\left\langle e_q^{(m)},Ae_q^{(m)} \right\rangle }{\left\langle e_q^{(m)},e_q^{(m)} \right\rangle }\]

We introduce on \(\mathrm{End}(\mathrm{\Gamma}_{hol}(M,L^{(m)}))\) the Hilbert-Schmidt norm \[<A,C>_{HS} =\mathrm{Tr}(A^*C)\]

Proposition 35 ([15]). The following statements hold:
i) The Toeplitz map \(f\mapsto T_f^{(m)}\) and the covariant symbol map \(A\mapsto \sigma^{(m)}(A)\) are adjoint \[\left\langle A,T_f^{(m)}\right\rangle_{HS}=\left\langle \sigma^{(m)}(A),f\right\rangle_m\] ii) The covariant symbol map \(\sigma^{(m)}\) is injective.
iii) The
Berezin transform* map defined \[C^\infty(M)\rightarrow C^\infty(M), \qquad f\mapsto I^{(m)}(f):=\sigma^{(m)}\circ T_f^{(m)}\] satisfies the limit \[\|f-I^{(m)}(f)\|_\infty \rightarrow 0 \quad \text{as m\rightarrow \infty}.\]*

In this section, we are going to draw a bridge between the classical problem of structure preserving discretization and the Berezin-Toeplitz quantization. We are going to show that the Berezin-Toeplitz quantization is in fact a structure preserving discretization of the Poisson algebra \((C^\infty(M), \left\lbrace \cdot, \cdot \right\rbrace )\), when \(M\) is a compact Kähler manifold.
Now that we have presented the background material on the Berezin-Toeplitz quantization, we can be described it by the following diagram. \[\begin{tikzcd} C^\infty(M)\otimes C^\infty(M) \arrow[r,"{\left\lbrace \cdot, \cdot \right\rbrace} "] \arrow[d,"T^{(m)}"] & C^\infty(M) \arrow[d,"T^{(m)}"] \\ M_m(\mathbb{C})\otimes M_m(\mathbb{C}) \arrow[r,"{\left[ \cdot, \cdot\right] } "] & M_m(\mathbb{C}) \end{tikzcd}\] Indeed, we prove that the Berezin-Toeplitz quantization satisfies the axiom of a structure preserving discretization of the pair \((C^\infty(M), \left\lbrace \cdot, \cdot \right\rbrace )\).

Proposition 36. The Berezin-Toeplitz quantization of the Poisson algebra of smooth functions \((C^\infty(M), \left\lbrace \cdot, \cdot \right\rbrace )\) is a structure preserving discretization.

Proof. We start by proving that the quantization is a discretization method in the sense of Definition [def1]. The Toeplitz maps \(T^{(m)}\) are contractive by Theorem [BMS] and surjective by Proposition [surj]. Moreover, we have the limit: \[\lim_{m\rightarrow \infty}\|T^{(m)}(f)\|=\|f\|\] again following from Theorem [BMS]. In addition, to any projection \(T^{(m)}\), we can associate an injective section map \(\sigma^{(m)}\). Therefore, the Berezin-Toeplitz quantization is a discretization.
Now, if we consider the pair \((C^\infty(M),\left\lbrace \cdot, \cdot\right\rbrace )\) as an object in the category of Lie algebra, then the pair \((M_m(\mathbb{C}),[\cdot, \cdot])\) is a structure preserving discretization since :

  • \((M_m(\mathbb{C}),[\cdot, \cdot])\) is a Lie algebra

  • \(\|T^{(m)}\left\lbrace f,g \right\rbrace - [T^{(m)}(f),T^{(m)}(g)]\|\rightarrow 0\) as \(m\rightarrow \infty\).

Therefore, the Berezin-Toeplitz quantization provides a structure preserving discretization of the Lie algebra structure on \(C^\infty(M)\). ◻

5 Noncommutative Laplacian as a discrete Laplacian↩︎

The Laplacian, which traditionally arises in the context of smooth Riemannian manifolds, is critical for understanding various physical phenomena, such as heat diffusion and wave propagation. However, when transitioning to discrete spaces, particularly in curved spaces, the formulation of the Laplacian requires a novel approach.
This section focuses on how the Berezin-Toeplitz quantization framework can be used to discretize the classical Laplacian, thereby creating a discrete version that operates on matrix algebras rather than smooth functions. The key idea is to show that the noncommutative Laplacian emerges naturally, following the axioms, as a discrete counterpart of the classical Laplace operator on compact Kähler manifolds. By leveraging the structure-preserving properties of the Berezin-Toeplitz quantization, the resulting discrete Laplacian retains important geometric and analytic features, offering a powerful tool for studying differential equations in noncommutative settings.
Consider again a compact Kähler manifold \((M,\omega)\). The Nash embedding theorem states that any Riemannian manifold can be isometrically embedded in the Euclidean space \(\mathbb{R}^d\) for sufficiently large \(d\). Thus, for a closed Poisson manifold \((M,\pi )\) with a metric \(g\), there exists an isometry embedding \[X:M\rightarrow \mathbb{R}^d\] for sufficiently large \(d\). We denote the embedding coordinate functions as \(\left\lbrace X^k \right\rbrace_{k=1,\dots, d}\). The word isometric means that the induced metric of the embedding is equal to the intrinsic metric \(g\) on \(M\): \[g_{ij}(x)=g_x(\partial_i,\partial_j)=\left\langle d_xX(\partial_i),d_xX(\partial_j) \right\rangle = \sum_{k=1}^d(\partial_iX^k)(\partial_jX^k) \label{coord}\tag{5}\] Using the smooth bivector \(\pi\in \mathfrak{X}(M)\), we can associate to any function \(f\) a derivation through the map defined by: \[\xi_f:C^\infty(M)\rightarrow \mathrm{Der}(M), \quad \xi_f(g)=\pi(df\wedge dg)\] We can now define the following family of self-adjoint differential operators \[\partial_k:C^\infty(M)\rightarrow C^\infty(M), \quad g\mapsto \partial_k(g):=\xi_{X^k}(g) =\left\lbrace X^k,g\right\rbrace\] for every \(k=1,\dots ,d\).

Remark 37. One can then construct \(\mathfrak{g}:=\mathrm{Lie}(<\partial_k>)\) the Lie algebra generated by the set of vector fields \((\partial_k)\) and define the so-called quantized Weil-algebra by \[\mathcal{W}(\mathfrak{g})=\mathrm{U}(\mathfrak{g})\otimes \mathrm{Cl}(\mathfrak{g})\] where \(\mathrm{U}(\mathfrak{g})\), respectively \(\mathrm{Cl}(\mathfrak{g})\), is the enveloping Lie algebra, respectively the Clifford algebra, associated to \(\mathfrak{g}\). One can then define a first order elliptic differential operator called the Dirac operator \[D=\sum_k\partial_ke^k + \sum_{ijk}\gamma_{ijk}e^ie^je^k\] which plays a central role in the Attiyah-Singer theory. In this work however, we will simply consider the first order operator \(\mathcal{D}\) given by \[\mathcal{D}= \sum_k \left\lbrace X^k,\cdot \right\rbrace .\]

We also define the Bochner-Laplace operator for the metric \(g\) on \(M\) defined by \[\mathrm{\Delta}\varphi := -g^{ij}\nabla_i\nabla_j\varphi.\] using the covariant derivatives. There is a relation between the operator \(D\) and the Laplacian given by \[D^2=\mathrm{\Delta} +\text{lower order terms}\]

Therefore, we will derive structure preserving discretization of the differential algebra \((C^\infty(M),\mathcal{D})\) which will induce a discretization of the Laplacian.

Theorem 38. The Berezin-Toeplitz quantization induces a structure preserving discretization of the differential algebra \((C^\infty(M),\mathcal{D})\) given by the following commuting diagram \[\begin{tikzcd} C^\infty(M) \arrow[r,"\mathcal{D}"] \arrow[d,"T^{(m)}"'] & C^\infty(M) \arrow[d,"T^{(m)}"] \\ \mathrm{End}(\mathcal{H}_m)\arrow[r,"d_m"] & \mathrm{End}(\mathcal{H}_m) \end{tikzcd}\] where the differential \(d_m\) is given by the following commutator: \[A\mapsto d_m(A)=[D_m,A], \qquad D_m:=\sum_{k=1}^d\partial_m^k:=\sum_{k=1}^d \mathrm{\Pi}^{(m)}X_k\mathrm{\Pi}^{(m)}\]

Proof. We consider the pair \((C^\infty(M),\mathcal{D} )\) as an object in the category of differentiable algebra, then the pair \((M_m(\mathbb{C}),d_m)\) is a structure preserving discretization since :

  • \((M_m(\mathbb{C}),[\cdot, \cdot])\) is a differential algebra.

  • \(\|T^{(m)}\left\lbrace X^k,f \right\rbrace - [T^{(m)}(X^k),T^{(m)}(f)]\|\rightarrow 0\) as \(m\rightarrow \infty\).

where the last statement follows from the quantization of the Poisson bracket. ◻

Corollary 39. The Berezin-Toeplitz discretization \(\mathfrak{D}(C^\infty(M),\mathcal{D})\) defines a discretization \[\mathrm{\Delta}_m: \mathrm{End}(\mathcal{H}_m)\rightarrow \mathrm{End}(\mathcal{H}_m),\qquad A\mapsto \mathrm{\Delta}_m(A)=\sum_{k=1}^d\left[ \partial_m^k\left[ \partial_m^k,A\right] \right]\] of the Bochner-Laplace operator \(\mathrm{\Delta}\) such that \[D_m^2=\mathrm{\Delta}_m+\text{lower order terms}\]

Proof. Recall the two identities that determine the quantization of the Poisson algebra: \[\begin{align} T_m(fg)=T_m(f)T_m(g) + O\left(m^{-1}\right) \quad \text{and} \quad m\left[ T_m(f),T_m(g)\right] =T_m(\left\lbrace f,g \right\rbrace )+O(m^{-1}). \end{align}\] Then, applying successively the identities to the iterated Poisson brackets defining the Laplace operator, we get: \[\begin{align} T_m\left( \left\lbrace X^k , \left\lbrace X^k,\varphi \right\rbrace \right\rbrace\right) &= m\left[ T_m(X^k),T_m\left( \left\lbrace X^k,\varphi \right\rbrace \right)\right] + O(m^{-1}) \\ &= m^2\left[ T_m(X^k),\left[ T_m(X^k),T_m(\varphi)\right] + O(m^{-2})1\right] + O(m^{-1})\\ &=m^2\left[ T_m(X^k),\left[ T_m(X^k),T_m(\varphi)\right] \right] + \left[ T_m(X^k), O(1)1\right] + O(m^{-1})\\ &=\Delta_m(\varphi) + O(m^{-1}) \end{align}\] The statement then follows by take the norm on both sides of the last equality. ◻

Corollary 40. The discrete Bochner-Laplace* operator \[\widetilde{\mathrm{\Delta}}_m:C^\infty(M)\rightarrow C^\infty(M), \qquad \widetilde{\mathrm{\Delta}}_m(f)=\sigma^{(m)}\circ \mathrm{\Delta}_m\circ T^{(m)}\] is a self-adjoint operator on \((\mathcal{H}_m,<\cdot,\cdot>_m)\) with discrete spectrum.*

Proof. (Recall that for \(f\in C^\infty\), the operator \(M_f\) is multiplication with \(\tau^*f\)).
Using the spectral theorem applied to \(\mathrm{\Delta}_m\) we get the decomposition of \(\widetilde{\mathrm{\Delta}}_m\) as \[\widetilde{\mathrm{\Delta}}_m:\mathcal{H}_m\rightarrow \mathcal{H}_m,\qquad \widetilde{\mathrm{\Delta}}_m(f)=\sum_{i=1}^m\lambda_i\sigma^{(m)}P_{\lambda_i}T^{(m)}\] where \((\lambda_i,P_{\lambda_i})\) are the pairs of eigenvalues and associated spectral projections.
The operator \(\sigma^{(m)}P_{\lambda}T^{(m)}\) is a self-adjoint projection for every \(\lambda\in \sigma(\mathrm{\Delta}_m)\). ◻

An important question that remains to address is the convergence of the discrete Bochner-Laplacian to the continuous counterpart. In order to do so, we need to show, following Definition [conv], the limit \[\lim_{m\rightarrow \infty}\|\mathrm{\Delta}\varphi-\sigma^{(m)}\circ \mathrm{\Delta}_m\circ T^{(m)}\|=0\] We will need the following definition to tackle the convergence of unbounded linear operators.

Definition 41 (Strong graph limit). Let \(A_n\) be a sequence of operators on a Hilbert space \(\mathcal{H}\). We say that \((\psi,\varphi)\in \mathcal{H}\times \mathcal{H}\) is in the strong graph limit of \(A_n\) if we can find \(\psi_n\in D(A_n)\) so that \(\psi_n\rightarrow \psi\), \(A_n\psi_n\rightarrow \varphi\). We denote the set of pairs in the strong graph limit by \(\mathrm{\Gamma}^s\). If \(\mathrm{\Gamma}^s\) is the graph of an operator \(A\) we say that \(A\) is the strong graph limit of \(A_n\) and write \(A=\mathrm{sgr-lim}A_n\).

Theorem 42. The discrete Bochner-Laplace operator \(\widetilde{\mathrm{\Delta}}_m\) converges in the strong-graph limit sense to the Bochner-Lapalce operator \(\mathrm{\Delta}\): \[\mathrm{sgr}-\lim \widetilde{\mathrm{\Delta}}_m = \mathrm{\Delta}\]

In order to prove the theorem, we need the following lemma relating the Laplacian to the Poisson bracket.

Lemma 43. Let \((M,g,\omega)\) be a Kähler manifold, with its induced Poisson structure \((M,\pi)\), such that \((M,g)\) which admits an isometric embedding \(X:M\rightarrow \mathbb{R}^d\). Then, the Laplacian operator can be expressed as: \[\mathrm{\Delta}_g\varphi = -\sum_{k=1}^d\left\lbrace X^k , \left\lbrace X^k,\varphi \right\rbrace \right\rbrace\]

Proof. We start by expressing the Poisson bracket using the bivector \(\pi\) and the connection operator: \[\begin{align} -\left\lbrace X^k , \left\lbrace X^k,\varphi \right\rbrace \right\rbrace &=- \pi\left(dX^k\wedge \pi\left( dX^k\wedge d \varphi\right) \right) =-\pi^{ij}\pi^{rs}\partial_iX^k\nabla_j\left(\partial_rX^k \nabla_s\varphi \right) \end{align}\] Using the product rule, we obtain: \[\begin{align} &-\pi^{ij}\pi^{rs}\partial_iX^k\nabla_j\left(\partial_rX^k \nabla_s\varphi \right)=-\pi^{ij}\pi^{rs}\partial_iX^k\left[ \nabla_j\left(\partial_rX^k\right)\nabla_s\varphi + \partial_rX^k\nabla_j\nabla_s\varphi \right]\\ &=-\pi^{ij}\pi^{rs}\left[\nabla_j\left(\partial_iX^k\partial_rX^k\right)\nabla_s\varphi - (\nabla_j\partial_iX^k)(\partial_rX^k)\nabla_s\varphi + \partial_iX^k\partial_rX^k\nabla_j\nabla_s\varphi\right] . \end{align}\] We have used the fact that \(X\) are coordinate embedding and satisfy Equation 5 . We use now that \(\nabla_j\left(g_{ir}\right)=0\) (metric compatible connection) and \[\pi^{ij}\nabla_j\partial_iX^k=\pi^{ij}\left(\partial_j\partial_i X^k - \Gamma_{ij}^\ell\partial_\ell X^k \right)= \pi^{ij}\partial_j\partial_i X^k - \pi^{ij}\Gamma_{ij}^\ell\partial_\ell X^k=0\] since \(\pi^{ij}=-\pi^{ji}\) and also \(\Gamma^k_{ij}=\Gamma^k_{ji}\). We deduce then that \[\begin{align} -\left\lbrace X^k , \left\lbrace X^k,\varphi \right\rbrace \right\rbrace=-\pi^{ij}\pi^{rs}\left[\nabla_j\left(g_{ir}\right)\nabla_s\varphi - (\nabla_j\partial_iX^k)(\partial_rX^k)\nabla_s\varphi + g_{ir}\nabla_j\nabla_s\varphi\right] \end{align}\] Important, here we have used \(\nabla \pi =0\) which follows from the general properties of the Kähler structure \(\nabla g =\nabla J=\nabla \omega =0\). Therefore, we are left with \[-\left\lbrace X^k , \left\lbrace X^k,\varphi \right\rbrace \right\rbrace = -\pi^{ij}\pi^{rs} g_{ir}\left(\nabla_j\nabla_s\varphi\right)=-g^{js} \left(\nabla_j\nabla_s\varphi\right).\] ◻

Lemma 44. Let \((M,g,\omega)\) be a Kähler manifold, with its induced Poisson structure \((M,\pi)\), such that \((M,g)\) which admits an isometric embedding \(X:M\rightarrow \mathbb{R}^d\). Then, the bi-Laplacian operator can be expressed as: \[\sum_{k}\left\lbrace X^k, \mathrm{\Delta}\left\lbrace X^k ,\varphi \right\rbrace \right\rbrace =\mathrm{\Delta}^2(\varphi)\] for every \(\varphi\in C^\infty(M)\).

Proof. We apply the Laplacian to the Poisson bracket with \(X^k\) \[\mathrm{\Delta}\left\lbrace X^k ,\varphi \right\rbrace = \left\lbrace \mathrm{\Delta}X^k ,\varphi \right\rbrace + \left\lbrace \nabla_i X^k ,\nabla_i\varphi \right\rbrace +\left\lbrace X^k ,\mathrm{\Delta}\varphi \right\rbrace = \left\lbrace X^k ,\mathrm{\Delta}\varphi \right\rbrace\] using the fact that \(\pi^{ij}\nabla_j\partial_iX^k=0\). ◻

Lemma 45. The section maps \(\sigma^{(m)}\) are contractives i.e. \[\|\sigma^{(m)}(A)\|\leq \|A\|\] for every \(A\in M_m(\mathbb{C})\).

Lemma 46. (a) For every \(f\in C^\infty(M)\), we have the first estimate \[I^{(m)}(f)=f+\frac{1}{m}\mathrm{\Delta}(f)+O(m^{-2}) \label{eq-I}\qquad{(1)}\] (b) For every, \(f,g\in C^\infty(M)\), we have the second estimate \[\mathrm{i}m[T^{(m)}(\varphi),T^{(m)}(\psi)]=T^{(m)}\left( \left\lbrace f,g \right\rbrace \right)+O(m^{-1})T^{(m)}(\mathrm{\Delta}\left\lbrace f,g \right\rbrace )\]

Proof. Theorem 42
We start by applying a triangular inequality \[\begin{align} &\|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty} \\ &\leq \|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ T^{(m)}(\mathrm{\Delta}(\varphi))\|_{\infty} + \|\sigma^{(m)}\circ T^{(m)}(\mathrm{\Delta}(\varphi)) - \sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty} \end{align}\] and then we apply Lemma 45 to bound the second term in the right-hand-side: \[\begin{align} \|\mathrm{\Delta}(\varphi)-&\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty} \\ &\leq \|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ T^{(m)}(\mathrm{\Delta}(\varphi))\|_{\infty} + \| T^{(m)}(\mathrm{\Delta}(\varphi)) - \mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty} \end{align}\] We then combine ?? and Lemma 44 to bound the first term in the right-hand-side and Corollary 39 to bound the second term to get \[\|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty} \leq a_m\|\mathrm{\Delta}^2(\varphi)\|_{\infty}\] Since we work on a compact smooth manifold, we have one hand that \[\|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{2}\leq c_0\|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{\infty}\] and on the other using Sobolev embedding we have \[\|\mathrm{\Delta}^2(\varphi)\|_{\infty}\leq c_1\|\varphi\|_{H^k}\] In conclusion, we get \[\|\mathrm{\Delta}(\varphi)-\sigma^{(m)}\circ\mathrm{\Delta}_m\circ T^{(m)}(\varphi)\|_{2} \leq a_m\|\varphi\|_{H^k}\] ◻

Remark 47. It is important to notice that the structure preserving property obtained in Corollary 39 plays a central role in the proof of convergence; hence the last theorem is an example of results showing how structure preservation might imply convergence.

We can also answer the question about the convergence of the spectrum of the Laplacian, which is crucial in many applications.

Theorem 48. Let \(A_n\) \((n\in \mathbb{N})\) and \(A\) be (unbounded) self-adjoint operators and assume that \(D(A)=D(A_n)\). Assume, furthermore, that there are null-sequences \((a_n)\) and \((b_n)\) from \(\mathbb{R}\) for which \[\|(A-A_n)f\|\leq a_n\|f\|+b_n\|Af\| \quad \text{for all f\in D(A)}.\] Then \(\sigma(A)=\lim_{n\rightarrow \infty}\sigma(A_n)\).

Corollary 49. The spectrum discrete Bochner-Laplace operator \(\widetilde{\mathrm{\Delta}}_m\) converges in the spectrum of the true Bochner-Lapalce operator \(\mathrm{\Delta}\) i.e. \(\sigma(\mathrm{\Delta})=\lim_{n\rightarrow \infty}\sigma(\mathrm{\Delta}_n)\).

6 Example of the sphere \(\mathbb{S}^2\)↩︎

We show in the concrete example of the sphere \((\mathbb{S}^2,\lbrace\cdot, \cdot\rbrace)\), seen as a compact Poisson manifold, how a structure-preserving discretization of the differential structure of the space \(\mathbb{S}^2\) can be derived from first principles, including the discretization of the Laplacian operator. This is achieved following the set of axioms introduced in this work; the projection maps are provided by the Berezin-Toeplitz quantization. As a result, we recover the discrete Laplacian used in Zeitlin’s model for Euler’s equations on the sphere [9], [16]; hence proving that such model is structure-preserving in the sense of the present work. In a follow-up work [17], we extend this construction to compact surfaces with higher genus.
Let us recall that there exists an isometric embedding: \[\rho : \mathbb{S}^2\rightarrow \mathbb{R}^3, \qquad \rho(s):=(x,y,z)\] of the sphere onto \(\mathbb{R}^3\), such that \(x^2+y^2+z^2=1\). We consider the differential operator \(\mathcal{D}\) defined by \[\mathcal{D}(f)=\lbrace x,f\rbrace + \lbrace y,f\rbrace + \lbrace z,f\rbrace = - \left(\partial_xf + \partial_yf + \partial_zf \right)\] Finding a structure-preserving discretization of the differential algebra \((C^\infty(\mathbb{S}^2),\mathcal{D})\) is equivalent by our axiomatization to the existence of the following diagram \[\begin{tikzcd} C^\infty(\mathbb{S}^2) \arrow[r,"\mathcal{D}"] \arrow[d,"T^{(m)}"'] & C^\infty(\mathbb{S}^2) \arrow[d,"T^{(m)}"] \\ \mathrm{End}(\mathcal{H}_m)\arrow[r,"d_m"] & \mathrm{End}(\mathcal{H}_m) \end{tikzcd}\] such that \((\mathrm{End}(\mathcal{H}_m),d_m)\) is a differential algebra and \(\|T^{(m)}\circ\mathcal{D} - d_m\circ T^{(m)}\|\rightarrow 0\).
If we rewrite the spherical polynomial the spherical harmonics \(\left\lbrace Y_{\ell m}(\theta, \varphi)\right\rbrace\) as the space \(\mathcal{P}_{\mathrm{h}}[x_1,x_2,x_3]\) of harmonic homogeneous polynomial in \(3\) variables \(x_1,x_2,x_3\), then \(T^{(m)}\) can be taken as the projection onto the subspace \(\mathcal{P}^m_{\mathrm{h}}[x_1,x_2,x_3]\) of degree \(m\) homogeneous polynomials. In this case, \(T^{(m)}\) corresponds to the Berezin-Toeplitz quantization of \(\mathbb{S}^2\) and it satisfies the condition of a structure preserving discretization.
Following the procedure of the previous section, the continuous operators \((x,y,z)\) are replaced by \(m\)-dimensional matrices: \[T^{(m)}(x)=X_m, \quad T^{(m)}(y)=Y_m, \quad T^{(m)}(z)=Z_m,\] and the Lie algebra generated by \((X_m,Y_m,Z_m)\) is the Spin algebra \(\mathfrak{so}(m)\). Then, the differential operator \(\mathcal{D}\) is discretized using the Lie bracket of matrices and gives the derivation \(d_m\): \[d_m(A)=[X_m,A]+[Y_m,A]+[Z_m,A]\] Similarly, the Laplace operator on the sphere is replaced by \[\Delta_m(A)=[X_m,[X_m,A]]+[Y_m,[Y_m,A]]+[Z_m,[Z_m,A]],\] which is identified as the non-commutative Laplacian.

References↩︎

[1]
Snorre H. Christiansen, A Construction of Spaces of Compatible Differential Forms on Cellular Complexes, Mathematical Models and Methods in Applied Sciences 18(2008), no. 05, 739–757 (en).
[2]
D. N. Arnold, R. S. Falk, and R. Winther, Finite element exterior calculus: from Hodge theory to numerical stability, Bulletin of the American Mathematical Society 47(2010), no. 2, 281–354 (en), arXiv: 0906.4325.
[3]
S. H. Christiansen, H. Z. Munthe-Kaas, and B. Owren, Topics in structure-preserving discretization, Acta Numerica 20(2011), 1–119 (en).
[4]
E.S. Gawlik, P. Mullen, D. Pavlov, J.E. Marsden, and M. Desbrun, Geometric, variational discretization of continuum theories, Physica D: Nonlinear Phenomena 240(2011), no. 21, 1724–1760 (en).
[5]
A. N. Hirani, Discrete Exterior Calculus, Ph.D. thesis, Caltech, Pasadena, California, 2003.
[6]
M. Desbrun, A. N. Hirani, M. Leok, and J. E. Marsden, Discrete Exterior Calculus, arXiv:math/0508341 (2005) (en), arXiv: math/0508341.
[7]
Alexander Lundervold and Hans Munthe-Kaas, Hopf algebras of formal diffeomorphisms and numerical integration on manifolds, September 2010, arXiv:0905.0087 [math].
[8]
Robert I. McLachlan, Klas Modin, Hans Munthe-Kaas, and Olivier Verdier, Butcher series: A story of rooted trees and numerical methods for evolution equations, February 2017, arXiv:1512.00906 [math].
[9]
Klas Modin and Manolis Perrot, Eulerian and Lagrangian Stability in Zeitlin’s Model of Hydrodynamics, Communications in Mathematical Physics 405(2024), no. 8, 177 (en).
[10]
Damien Tageddine and Jean-Christophe Nave, Noncommutative Differential Geometry on Infinitesimal Spaces, accepted for publication in Annales Mathématiques du Québec (2025), arXiv:2209.12929.
[11]
, Statistical Fluctuation of Infinitesimal Spaces, passed first round of revisions in Annales Mathématiques du Québec (2025), arXiv:2304.10617.
[12]
S. Sakai, C*-algebras and W*-algebras, Classics in mathematics, Springer, Berlin ; New York, 1998.
[13]
Shôichirô Sakai, Operator algebras in dynamical systems: the theory of unbounded derivations in C*-algebras, Cambridge University Press, Cambridge, 2008 (eng), OCLC: 190967415.
[14]
A. Connes, Noncommutative geometry, Academic Press, San Diego, 1994 (eng).
[15]
Martin Schlichenmaier, Berezin-Toeplitz quantization for compact Kaehler manifolds. A Review of Results, Advances in Mathematical Physics 2010(2010), 1–38 (en), arXiv:1003.2523 [math-ph].
[16]
Klas Modin and Milo Viviani, Lie–Poisson Methods for Isospectral Flows, Foundations of Computational Mathematics 20(2020), no. 4, 889–921 (en).
[17]
, Noncommutative Laplacian and numerical approximation of Laplace-Beltrami spectrum of compact Riemann surfaces, in preparation (2025).