← Back to home

Weyl Algebra

Oh, you want to delve into the Weyl algebra, do you? Fine. Don't expect a warm welcome. This isn't a tea party. It's an exploration of structures that are, frankly, more interesting than most conversations I'm forced to endure. Just try not to break anything.

Differential algebra

In the realm of abstract algebra, the Weyl algebras emerge as abstract echoes of differential operators that have been coaxed into behaving nicely, specifically those with polynomial coefficients. They bear the name of Hermann Weyl, a man who apparently found the intricacies of the Heisenberg uncertainty principle in quantum mechanics sufficiently compelling to warrant such an abstraction. A quaint notion, if you ask me.

Consider the most rudimentary form. Let there be a field, which we'll call FF, and the ring of polynomials in a single variable, F[x]F[x]. The corresponding Weyl algebra, the first of its kind, A1A_1, is composed of differential operators of this particular flavor:

fm(x)xm+fm1(x)xm1++f1(x)x+f0(x)f_{m}(x)\partial _{x}^{m}+f_{m-1}(x)\partial _{x}^{m-1}+\cdots +f_{1}(x)\partial _{x}+f_{0}(x)

where each fi(x)f_{i}(x) is, as previously stated, an element of F[x]F[x]. It's a rather straightforward construction, really. The nn-th Weyl algebra, denoted AnA_n, follows a similar, albeit scaled-up, blueprint.

Alternatively, A1A_1 can be conjured into existence by taking the free algebra on two generators, let's call them qq and pp, and then crushing it down by introducing the ideal generated by the relation ([p,q]1)([p,q]-1). This [p,q][p,q] business is, of course, the commutator, a fundamental source of discord in these algebraic worlds. For AnA_n, we simply expand this to the free algebra on 2n2n generators and impose the relations ([pi,qj]δi,j)([p_i,q_j]-\delta_{i,j}) for all i,ji,j from 1 to nn, where δi,j\delta_{i,j} is the ever-present Kronecker delta. It’s a way of forcing a specific kind of non-commutativity.

More generally, if you're presented with a partial differential ring, (R,Δ)(R, \Delta), where Δ={1,,m}\Delta = \{\partial_1, \ldots, \partial_m\} is a set of commuting derivatives, the associated Weyl algebra is the noncommutative ring R[1,,m]R[\partial_1, \ldots, \partial_m]. This ring adheres to the rule ir=ri+i(r)\partial_i r = r\partial_i + \partial_i(r) for any rRr \in R. The initial case we discussed is a specific instance where R=F[x1,,xn]R = F[x_1, \ldots, x_n] and Δ={x1,,xn}\Delta = \{\partial_{x_1}, \ldots, \partial_{x_n}\}, with FF being a field, naturally.

This article, for the most part, will confine itself to the AnA_n constructed over a field FF of characteristic zero, unless some peculiar circumstance dictates otherwise. It's worth noting that the Weyl algebra stands as a peculiar example of a simple ring that stubbornly refuses to be a matrix ring over a division ring. It's also a non-commutative specimen of a domain and a prime illustration of an Ore extension.

Motivation

The genesis of the Weyl algebra is intimately tied to the abstract machinations of quantum mechanics and the rather ambitious process of canonical quantization. Imagine a classical phase space, a rather abstract landscape defined by canonical coordinates (q1,p1,,qn,pn)(q_1, p_1, \ldots, q_n, p_n). These coordinates are bound by the Poisson bracket relations:

{qi,qj}=0,{pi,pj}=0,{qi,pj}=δij.\{q_i, q_j\} = 0, \quad \{p_i, p_j\} = 0, \quad \{q_i, p_j\} = \delta_{ij}.

Now, in the grand scheme of canonical quantization, the objective is to construct a Hilbert space of states and then map these classical observables—essentially functions on this phase space—to self-adjoint operators inhabiting this Hilbert space. The crucial step is to impose the canonical commutation relations:

[q^i,q^j]=0,[p^i,p^j]=0,[q^i,p^j]=iδij,[\hat{q}_i, \hat{q}_j] = 0, \quad [\hat{p}_i, \hat{p}_j] = 0, \quad [\hat{q}_i, \hat{p}_j] = i\hbar \delta_{ij},

where [,][\cdot, \cdot] denotes the commutator and q^i,p^i\hat{q}_i, \hat{p}_i are the operators that represent the classical qiq_i and pip_i. It was Erwin Schrödinger who, back in 1926, proposed a rather elegant identification:

  • q^j\hat{q}_j is to be represented by simple multiplication by xjx_j.
  • p^j\hat{p}_j is to be represented by the derivative ixj-i\hbar \partial_{x_j}.

With this mapping, the canonical commutation relations, the very bedrock of quantum mechanics, are satisfied. It's a rather neat trick, turning abstract concepts into tangible operators.

Constructions

The Weyl algebras, in their essence, can be constructed through various means, each offering a different vantage point, from the concrete to the profoundly abstract.

Representation

The Weyl algebra AnA_n can be materialized as a concrete representation. In the differential operator representation, much like Schrödinger's approach to canonical quantization, we assign:

  • qjq_j is mapped to multiplication by xjx_j.
  • pjp_j is mapped to differentiation by xj\partial_{x_j}.

Then, the commutation relation [qi,pj]=δij[q_i, p_j] = \delta_{ij} translates directly to [x^i,p^j]=δij[\hat{x}_i, \hat{p}_j] = \delta_{ij} in this operational setting.

In the realm of matrix representations, reminiscent of matrix mechanics, A1A_1 can be embodied by the following infinite matrices:

P=[010000200003],Q=[000010000100]P = \begin{bmatrix} 0 & 1 & 0 & 0 & \cdots \\ 0 & 0 & 2 & 0 & \cdots \\ 0 & 0 & 0 & 3 & \cdots \\ \vdots & \vdots & \vdots & \vdots & \ddots \end{bmatrix}, \quad Q = \begin{bmatrix} 0 & 0 & 0 & 0 & \cdots \\ 1 & 0 & 0 & 0 & \cdots \\ 0 & 1 & 0 & 0 & \cdots \\ \vdots & \vdots & \vdots & \vdots & \ddots \end{bmatrix}

These matrices, PP and QQ, when subjected to the commutator [P,Q][P, Q], yield the identity matrix, thereby satisfying the fundamental relation of A1A_1.

Generator

AnA_n can also be constructed as a quotient of a free algebra, defined by its generators and the relations that bind them. One method involves starting with an abstract vector space VV of dimension 2n2n, endowed with a symplectic form ω\omega. The Weyl algebra W(V)W(V) is then defined as:

W(V):=T(V)/((vuuvω(v,u)) for v,uV)W(V) := T(V) / ((v \otimes u - u \otimes v - \omega(v, u)) \text{ for } v, u \in V)

Here, T(V)T(V) signifies the tensor algebra on VV, and the notation (())(( \cdot )) denotes the ideal generated by the enclosed elements. In simpler terms, W(V)W(V) is the algebra generated by VV with the sole constraint that vuuv=ω(v,u)vu - uv = \omega(v, u). When VV is equipped with a Darboux basis for ω\omega, W(V)W(V) becomes isomorphic to AnA_n.

Furthermore, AnA_n can be viewed as a quotient of the universal enveloping algebra of the Heisenberg algebra. The Heisenberg algebra, itself the Lie algebra of the Heisenberg group, has a central element, [q,p][q, p]. By setting this central element equal to 1, the unit element in the universal enveloping algebra, we arrive at AnA_n.

Quantization

The algebra W(V)W(V), as defined earlier, is a quantization of the symmetric algebra Sym(V)\text{Sym}(V). When VV is over a field of characteristic zero, W(V)W(V) bears a natural isomorphism to Sym(V)\text{Sym}(V) itself, but with a subtly altered product – the Groenewold–Moyal product. This perspective views the symmetric algebra as polynomial functions on VV^*, where the variables span VV, and the Moyal product is obtained by replacing ii\hbar with 1.

The isomorphism is realized through a symmetrization map from Sym(V)\text{Sym}(V) to W(V)W(V):

a1an1n!σSnaσ(1)aσ(n)a_1 \cdots a_n \mapsto \frac{1}{n!} \sum_{\sigma \in S_n} a_{\sigma(1)} \otimes \cdots \otimes a_{\sigma(n)}

If one prefers to retain the ii\hbar and work over the complex numbers, the Weyl algebra could be defined from the outset using generators qiq_i and iqii\hbar\partial_{q_i}, aligning more directly with the conventions of quantum mechanics.

Thus, the Weyl algebra serves as a quantization of the symmetric algebra. It's closely related to Moyal quantization, particularly when restricted to polynomial functions. The key distinction lies in their formulation: the Weyl algebra is presented via generators and relations (as differential operators), while Moyal quantization is defined through a deformed multiplication.

To put it another way, if we denote the Moyal star product by fgf \star g, the Weyl algebra is isomorphic to (C[x1,,xn],)(\mathbb{C}[x_1, \dots, x_n], \star).

For comparison, the analogous quantization to the Weyl algebra in the context of exterior algebras is the Clifford algebra, sometimes referred to as the orthogonal Clifford algebra. The Weyl algebra itself is also known as the symplectic Clifford algebra. It essentially mirrors the role that Clifford algebras play for non-degenerate symmetric bilinear forms, but for symplectic bilinear forms.

D-module

The Weyl algebra can also be understood as a D-module. Specifically, the Weyl algebra associated with the polynomial ring R[x1,,xn]R[x_1, \dots, x_n], equipped with its standard differential structure, precisely corresponds to Grothendieck's ring of differential operations, DARn/RD_{\mathbb{A}^n_R/R}.

More broadly, consider a smooth scheme XX over a ring RR. Locally, the map XRX \to R can be seen as an étale cover over some ARn\mathbb{A}^n_R, with the standard projection. Since "étale" implies flatness and a null cotangent sheaf, this means that any D-module over such a scheme can be locally viewed as a module over the nthn^{\text{th}} Weyl algebra.

Let RR be a commutative algebra over a subring SS. The ring of differential operators DR/SD_{R/S} (or simply DRD_R when SS is implicit) is inductively defined as a graded subalgebra of EndS(R)\text{End}_S(R):

  • DR0=RD_R^0 = R
  • DRk={dEndS(R):[d,a]DRk1 for all aR}D_R^k = \{d \in \text{End}_S(R) : [d, a] \in D_R^{k-1} \text{ for all } a \in R\}.

The entire ring DRD_R is the union of all DRkD_R^k for k0k \geq 0, forming a subalgebra of EndS(R)\text{End}_S(R).

In the specific case where R=S[x1,,xn]R = S[x_1, \dots, x_n], the ring of differential operators of order n\leq n presents similarly to the characteristic zero case, but with the addition of "divided power operators." These are operators that correspond to those in the complex case but cannot be expressed as integral combinations of higher-order operators. An example is the operator x1[p]\partial_{x_1}^{[p]}, which maps x1Nx_1^N to (Np)x1Np\binom{N}{p}x_1^{N-p}.

Explicitly, a presentation is given by:

DS[x1,,x]/Sn=Sx1,,x,{xi,xi[2],,xi[n]}1iD_{S[x_1, \dots, x_\ell]/S}^n = S\langle x_1, \dots, x_\ell, \{\partial_{x_i}, \partial_{x_i}^{[2]}, \dots, \partial_{x_i}^{[n]}\}_{1 \leq i \leq \ell}\rangle

subject to the relations:

[xi,xj]=[xi[k],xj[m]]=0[x_i, x_j] = [\partial_{x_i}^{[k]}, \partial_{x_j}^{[m]}] = 0 [xi[k],xj]={xi[k1]if i=j0if ij[\partial_{x_i}^{[k]}, x_j] = \begin{cases} \partial_{x_i}^{[k-1]} & \text{if } i=j \\ 0 & \text{if } i \neq j \end{cases} xi[k]xi[m]=(k+mk)xi[k+m]when k+mn\partial_{x_i}^{[k]}\partial_{x_i}^{[m]} = \binom{k+m}{k}\partial_{x_i}^{[k+m]} \quad \text{when } k+m \leq n

where xi[0]=1\partial_{x_i}^{[0]} = 1 by convention. The Weyl algebra then emerges as the limit of these algebras as nn \to \infty.

When SS is a field of characteristic 0, DR1D_R^1 is generated as an RR-module by 1 and the SS-derivations of RR. Furthermore, DRD_R is generated as a ring by DR1D_R^1. Specifically, if S=CS=\mathbb{C} and R=C[x1,,xn]R=\mathbb{C}[x_1, \dots, x_n], then DR1=R+iRxiD_R^1 = R + \sum_i R\partial_{x_i}. As previously noted, An=DRA_n = D_R.

Properties of AnA_n

Many of the properties observed in A1A_1 extend to AnA_n with minimal fuss, as the higher dimensions commute in a rather predictable fashion.

General Leibniz rule

  • Theorem (general Leibniz rule): pkqm=l=0k(kl)m!(ml)!qmlpkl=qmpk+mkqm1pk1+p^k q^m = \sum_{l=0}^{k} \binom{k}{l} \frac{m!}{(m-l)!} q^{m-l} p^{k-l} = q^m p^k + mk q^{m-1} p^{k-1} + \cdots

    Proof: When viewed through the lens of the px,qxp \mapsto \partial_x, q \mapsto x representation, this identity arises directly from the general Leibniz rule. Since this rule can be proven through sheer algebraic manipulation, it holds sway over A1A_1 as well.

    Specifically, we find that: [q,qmpn]=nqmpn1[q, q^m p^n] = -nq^m p^{n-1} [p,qmpn]=mqm1pn[p, q^m p^n] = mq^{m-1} p^n

  • Corollary: The center of the Weyl algebra AnA_n is simply its underlying field of scalars, FF.

    Proof: If the commutator of an element ff with either pp or qq is zero, then, based on the preceding statement, ff cannot contain any monomial of the form pnqmp^n q^m where n>0n > 0 or m>0m > 0. This leaves only scalar multiples of the identity, thus placing ff within the center FF.

Degree

  • Theorem: AnA_n possesses a basis given by {qmpn:m,n0}\{q^m p^n : m, n \geq 0\}.

    Proof: Through repeated application of the commutator relations, any monomial can be expressed as a linear combination of these basis elements. The challenge then becomes demonstrating their linear independence. This can be verified in the differential operator representation. For any linear sum m,ncm,nxmxn\sum_{m,n} c_{m,n} x^m \partial_x^n with non-zero coefficients, we can group it by descending order:

    pN(x)xN+pN1(x)xN1++pM(x)xMp_N(x)\partial_x^N + p_{N-1}(x)\partial_x^{N-1} + \cdots + p_M(x)\partial_x^M

    where pM(x)p_M(x) is a non-zero polynomial. Applying this operator to xMx^M yields M!pM(x)M! p_M(x), which is demonstrably non-zero.

    This basis structure allows A1A_1 to be a graded algebra, where the degree of m,ncm,nqmpn\sum_{m,n} c_{m,n} q^m p^n is determined by the maximum value of m+nm+n among its non-zero monomials. A similar definition of degree applies to AnA_n.

  • Theorem: For AnA_n:

    • deg(g+h)max(deg(g),deg(h))\deg(g+h) \leq \max(\deg(g), \deg(h))
    • deg([g,h])deg(g)+deg(h)2\deg([g,h]) \leq \deg(g) + \deg(h) - 2
    • deg(gh)=deg(g)+deg(h)\deg(gh) = \deg(g) + \deg(h)

    Proof: We'll prove this for A1A_1, as the AnA_n case follows a similar logic. The first relation is inherent in the definition of degree. The second relation stems directly from the general Leibniz rule. For the third relation, we know that deg(gh)deg(g)+deg(h)\deg(gh) \leq \deg(g) + \deg(h). Thus, it suffices to show that ghgh contains at least one non-zero monomial with the degree deg(g)+deg(h)\deg(g) + \deg(h). To find such a monomial, we select the term with the highest degree in gg. If there are multiple such terms, we choose the one with the highest power of qq. We do the same for hh. The product of these two selected monomials yields a unique monomial within ghgh, ensuring it remains non-zero.

  • Theorem: AnA_n is a simple domain. That is, it possesses no non-trivial two-sided ideals and no zero divisors.

    Proof: The property deg(gh)=deg(g)+deg(h)\deg(gh) = \deg(g) + \deg(h) directly implies that there are no zero divisors.

    Now, suppose, for the sake of contradiction, that II is a non-zero two-sided ideal of A1A_1, with IA1I \neq A_1. Let's pick a non-zero element fIf \in I that has the minimum possible degree.

    If ff contains a non-zero monomial of the form xxmn=xm+1nx \cdot x^m \partial^n = x^{m+1} \partial^n, then the commutator [,f]=ff[\partial, f] = \partial f - f\partial will contain a non-zero monomial of the form xm+1nxm+1n=(m+1)xmn\partial x^{m+1} \partial^n - x^{m+1} \partial^n \partial = (m+1)x^m \partial^n. Thus, [,f][\partial, f] is non-zero and has a degree strictly less than deg(f)\deg(f). Since II is a two-sided ideal, [,f]I[\partial, f] \in I, which contradicts the assumption that ff had the minimal degree.

    Similarly, if ff contains a non-zero monomial of the form xmnx^m \partial^n \partial, then the commutator [x,f]=xffx[x, f] = xf - fx will be non-zero and have a lower degree, again leading to a contradiction. These arguments, when applied across all possible monomials, demonstrate that no such non-zero ideal II can exist.

Derivation

  • Further information: Derivation (differential algebra)

  • Theorem: The derivations of AnA_n are in a one-to-one correspondence with the elements of AnA_n, up to an additive scalar. In other words, any derivation DD can be expressed as [,f][ \cdot, f ] for some fAnf \in A_n. Conversely, any fAnf \in A_n yields a derivation [,f][ \cdot, f ]. If f,fAnf, f' \in A_n yield the same derivation, i.e., [,f]=[,f][ \cdot, f ] = [ \cdot, f' ], then fff - f' must be an element of the scalar field FF.

    The proof for this theorem mirrors the process of finding a potential function for a conservative polynomial vector field in the plane.

    Proof: Since the commutator is a derivation in both of its arguments, [,f][ \cdot, f ] is indeed a derivation for any fAnf \in A_n. The uniqueness up to an additive scalar is guaranteed by the fact that the center of AnA_n is precisely the field of scalars FF.

    The core of the proof involves demonstrating that any derivation is an inner derivation, achieved through induction on nn.

    Base case (n=1): Let D:A1A1D: A_1 \to A_1 be a linear map that is a derivation. We aim to construct an element rr such that [p,r]=D(p)[p, r] = D(p) and [q,r]=D(q)[q, r] = D(q). Since both DD and [,r][ \cdot, r ] are derivations, these relations imply that [g,r]=D(g)[g, r] = D(g) for all gA1g \in A_1. Given that [p,qmpn]=mqm1pn[p, q^m p^n] = mq^{m-1}p^n, there exists an element f=m,ncm,nqmpnf = \sum_{m,n} c_{m,n} q^m p^n such that [p,f]=m,nmcm,nqmpn=D(p)[p, f] = \sum_{m,n} mc_{m,n} q^m p^n = D(p). The Jacobi identity, combined with the derivation property of DD, leads to: 0=D([p,q])=[p,D(q)]+[D(p),q]=[p,D(q)]+[[p,f],q]0 = D([p, q]) = [p, D(q)] + [D(p), q] = [p, D(q)] + [[p, f], q] Rearranging, we get: [p,D(q)[q,f]]=0[p, D(q) - [q, f]] = 0 This implies that D(q)[q,f]D(q) - [q, f] must commute with pp. Since [p,qmpn]=nqm1pn[p, q^m p^n] = -nq^{m-1}p^n, we can find a polynomial h(p)h(p) such that [q,h(p)]=D(q)[q,f][q, h(p)] = D(q) - [q, f]. Crucially, [p,h(p)]=0[p, h(p)] = 0. Thus, r=f+h(p)r = f + h(p) serves as the desired element.

    Inductive step: For the inductive step, we similarly establish the existence of an element rAnr \in A_n such that [q1,r]=D(q1)[q_1, r] = D(q_1) and [p1,r]=D(p1)[p_1, r] = D(p_1). The commutation relations imply that [x,D(y)[y,r]]=0[x, D(y) - [y, r]] = 0 for all x{p1,q1}x \in \{p_1, q_1\} and y{p2,,pn,q2,,qn}y \in \{p_2, \dots, p_n, q_2, \dots, q_n\}. Since [x,D(y)[y,r]][x, D(y) - [y, r]] is a derivation in both xx and yy, it follows that [x,D(y)[y,r]]=0[x, D(y) - [y, r]] = 0 for all xp1,q1x \in \langle p_1, q_1 \rangle and all yp2,,pn,q2,,qny \in \langle p_2, \dots, p_n, q_2, \dots, q_n \rangle. Here, \langle \cdot \rangle denotes the subalgebra generated by the elements. Therefore, for any yp2,,pn,q2,,qny \in \langle p_2, \dots, p_n, q_2, \dots, q_n \rangle, we have D(y)[y,r]p2,,pn,q2,,qnD(y) - [y, r] \in \langle p_2, \dots, p_n, q_2, \dots, q_n \rangle. Since D[,r]D - [ \cdot, r ] is also a derivation, by induction, there exists an rp2,,pn,q2,,qnr' \in \langle p_2, \dots, p_n, q_2, \dots, q_n \rangle such that D(y)[y,r]=[y,r]D(y) - [y, r] = [y, r'] for all yp2,,pn,q2,,qny \in \langle p_2, \dots, p_n, q_2, \dots, q_n \rangle. As p1,q1p_1, q_1 commute with p2,,pn,q2,,qn\langle p_2, \dots, p_n, q_2, \dots, q_n \rangle, we find that D(y)=[y,r+r]D(y) = [y, r + r'] for all y{p1,,pn,q1,,qn}y \in \{p_1, \dots, p_n, q_1, \dots, q_n\}, and consequently, for all of AnA_n.

Representation theory

Zero characteristic

When the ground field FF is of characteristic zero, the nthn^{\text{th}} Weyl algebra AnA_n is a simple, Noetherian domain. Its global dimension is nn, a notable contrast to its deformed counterpart, Sym(V)\text{Sym}(V), which has a global dimension of 2n2n.

AnA_n possesses no finite-dimensional representations. While this is a consequence of its simplicity, it can be more directly demonstrated by considering the trace of σ(q)\sigma(q) and σ(p)\sigma(p) for a hypothetical finite-dimensional representation σ\sigma (where [q,p]=1[q, p] = 1).

tr([σ(q),σ(p)])=tr(1)\text{tr}([\sigma(q), \sigma(p)]) = \text{tr}(1)

Since the trace of a commutator is always zero, and the trace of the identity matrix is the dimension of the representation, this equation implies that the dimension of the representation must be zero.

In fact, the absence of finite-dimensional representations is just the tip of the iceberg. For any finitely generated AnA_n-module MM, there exists a corresponding subvariety Char(M)\text{Char}(M) of V×VV \times V^*—termed the 'characteristic variety'—whose "size" is roughly correlated with the "size" of MM. A finite-dimensional module would, by definition, have a characteristic variety of zero dimension. Bernstein's inequality states that for a non-zero module MM,

dim(char(M))n\dim(\text{char}(M)) \geq n

An even more profound result is Gabber's theorem, which asserts that Char(M)\text{Char}(M) is a co-isotropic subvariety of V×VV \times V^* with respect to the natural symplectic form.

Positive characteristic

The landscape shifts dramatically when the Weyl algebra is defined over a field of characteristic p>0p > 0.

In this scenario, for any element DD in the Weyl algebra, the element DpD^p becomes central. Consequently, the Weyl algebra exhibits a remarkably large center. It functions as a finitely generated module over its center and, even more strikingly, as an Azumaya algebra over its center. This leads to a plethora of finite-dimensional representations, all constructed from simple representations of dimension pp.

Generalizations

The ideals and automorphisms of A1A_1 have been meticulously studied, and the moduli space for its right ideals is well-understood. However, the case for AnA_n presents a far greater challenge and is intimately connected to the notoriously difficult Jacobian conjecture.

For a more in-depth exploration of this quantization, particularly for n=1n=1 and its extension to a class of integrable functions beyond polynomials using the Fourier transform, consult the Wigner–Weyl transform.

Weyl algebras and Clifford algebras can be further endowed with the structure of a *-algebra. They can be unified within the framework of a superalgebra, as discussed in CCR and CAR algebras.

Affine varieties

Weyl algebras also find a natural generalization in the context of algebraic varieties. Consider a polynomial ring:

R=C[x1,,xn]IR = \frac{\mathbb{C}[x_1, \dots, x_n]}{I}

Here, a differential operator is defined as a composition of C\mathbb{C}-linear derivations of RR. This can be explicitly described as the quotient ring:

Diff(R)={DAn:D(I)I}IAn\text{Diff}(R) = \frac{\{D \in A_n : D(I) \subseteq I\}}{I \cdot A_n}

This construction allows us to study differential operators acting on functions defined on more complex geometric objects than simple affine spaces.