Fact-checked by Grok 2 weeks ago

Differential operator

A differential operator is a mathematical that applies to functions, typically represented by the symbol D = \frac{d}{dx} for the first in one , and extended to higher-order derivatives via powers like D^n = \frac{d^n}{dx^n}. These operators are linear, satisfying D(af + bg) = aD(f) + bD(g) for constants a, b and functions f, g, and can be combined into polynomials such as L = a_n D^n + \cdots + a_1 D + a_0, where the coefficients a_k may be functions of the independent . In multiple variables, they generalize to partial differential operators, like the Laplacian \Delta = \sum \frac{\partial^2}{\partial x_i^2}, which measures the of the . Differential operators form the foundation of the of differential equations, where equations of the form L(u) = f describe how functions evolve under , enabling the modeling of dynamic systems. For instance, linear constant-coefficient operators facilitate the solution of ordinary differential equations by factoring into characteristic equations, revealing exponential solutions. In partial differential equations, they underpin fundamental laws in physics, such as the \frac{\partial u}{\partial t} = k \Delta u for diffusion processes and the wave equation \frac{\partial^2 u}{\partial t^2} = c^2 \Delta u for . Applications extend to and , including electromagnetic fields via and through the Navier-Stokes equations, where these operators capture spatial and temporal changes. Advanced concepts, such as pseudodifferential operators, arise in to handle singularities and wavefronts in solutions.

Basic Concepts

Definition

A differential operator is a linear map D: C^\infty(\Omega) \to C^\infty(\Omega), where \Omega \subset \mathbb{R}^n is an and C^\infty(\Omega) denotes the of smooth real-valued functions on \Omega, that satisfies a generalized Leibniz rule characterizing its finite order. Specifically, D is a differential operator of order at most k if, for every smooth function g \in C^\infty(\Omega), the commutator [D, m_g]: m \mapsto D(g m) - g D(m) (where m_g denotes multiplication by g) is a differential operator of order at most k-1, with order 0 operators being precisely the continuous linear maps (i.e., multiplication operators). The order of D is the minimal such nonnegative integer k for which the (k+1)-fold iterated commutator with multiplication operators vanishes identically. This inductive definition via commutators ensures that D locally behaves like a finite of partial derivatives, up to smooth multiplication factors. In local coordinates x = (x_1, \dots, x_n) on \Omega, any differential operator D of order at most k admits the explicit expression D = \sum_{|\alpha| \leq k} a_\alpha(x) \partial^\alpha, where \alpha = (\alpha_1, \dots, \alpha_n) \in \mathbb{N}^n is a multi-index with |\alpha| = \alpha_1 + \dots + \alpha_n, \partial^\alpha = \partial_{x_1}^{\alpha_1} \cdots \partial_{x_n}^{\alpha_n} denotes the corresponding operator of order |\alpha|, and the coefficients a_\alpha: \Omega \to \mathbb{R} are functions. The highest-order terms (with |\alpha| = k) determine the principal symbol of D, which plays a key role in analyzing its global properties. More generally, differential operators extend to smooth manifolds M of dimension n, acting as linear maps D: C^\infty(M) \to C^\infty(M) that are locally of the above form in any coordinate chart. On a manifold, the order k is independent of the choice of coordinates, preserving the commutator characterization relative to multiplication by functions on M. This framework captures extensions of classical differentiation while ensuring consistency across overlapping charts via the partition of unity theorem.

Historical Development

The concept of differential operators traces its origins to the , amid the development of , particularly in the . Leonhard Euler and began treating higher-order derivatives as successive applications of a basic operation acting on functions, facilitating the analysis of variational problems. Euler's foundational contributions in the 1740s, developed further through collaboration with Lagrange in the 1760s, marked an early recognition of derivatives as operator-like entities that could be composed and applied systematically to functionals. This intuitive approach gained formal structure in the early 19th century with Louis François Antoine Arbogast's introduction of the standalone differential operator notation D, which separated operational symbols from quantities and enabled algebraic manipulation of derivatives. Published in 1800 in Du calcul des dérivations et ses usages dans la théorie des suites et dans la géométrie, Arbogast's work represented a pivotal conceptual shift, allowing differential operators to be viewed independently of specific functions. Subsequent 19th-century advancements by and further embedded differential operators within the rigorous theory of partial differential equations (PDEs). Cauchy's 1827 analysis of the Cauchy-Riemann equations exemplified first-order differential operators in , while his 1840 power series methods for nonlinear PDE initial value problems highlighted their role in solution existence and uniqueness. Weierstrass's mid-1870s emphasis on analytical rigor critiqued earlier informal approaches, influencing the study of elliptic operators and boundary value problems in PDEs. In the early 20th century, Élie Cartan's development of exterior systems from onward provided a geometric framework for higher-order differential operators, integrating them with Lie groups and moving frames for solving systems of PDEs. The mid-20th century brought abstract generalizations: Laurent Schwartz's 1950-1951 theory of distributions extended differential operators to act on generalized functions, enabling solutions to PDEs beyond classical smoothness. Pseudodifferential operators, building on singular integral techniques, were advanced by Alberto Calderón in 1959 to address Cauchy problems for broad classes of PDEs. Algebraically, the ring of differential operators on smooth manifolds was formalized by in the 1960s, revealing deep ties to sheaf theory and D-modules. During the 1940s-1950s, connections to s emerged, with the ring of constant-coefficient differential operators identified as the universal enveloping algebra of the translation , influencing and quantization. profoundly shaped this evolution, as in introduced differential operators like the momentum operator -i\hbar \frac{d}{dx} to represent observables in wave mechanics, bridging classical analysis with operator algebras.

Examples and Notation

Examples

A fundamental example of a first-order differential operator in one variable is the differentiation operator D = \frac{d}{dx}, which maps a function f(x) to its Df = f'(x). This has order 1, as its principal is the nonzero \xi \mapsto i\xi in the cotangent \xi. It satisfies the Leibniz rule for derivations: for functions f and g, D(fg) = f \, Dg + g \, Df = f g' + g f'. A typical second-order differential operator in one variable is P = \frac{d^2}{dx^2} + x \frac{d}{dx}, which arises in contexts such as the Airy differential equation. Its order is 2, determined by the highest-order term \frac{d^2}{dx^2}, with principal symbol \xi \mapsto -\xi^2. To verify it qualifies as a differential operator of order at most 2, consider its action under multiplication: for a smooth function f and variable h, compute P(f h) - f \, P h = f'' h + 2 f' h' + x f' h. This remainder equals f'' \cdot h + (2 f' + x f') \cdot h', which is a first-order differential operator in h (order reduced by 1), confirming the Leibniz condition recursively. In multiple variables, the divergence operator \operatorname{div} = \sum_{i=1}^n \frac{\partial}{\partial x_i} acts on a V = (V_1, \dots, V_n) by \operatorname{div} V = \sum_{i=1}^n \frac{\partial V_i}{\partial x_i}. This is a differential operator, with principal \xi \mapsto i \sum_{i=1}^n \xi_i. It obeys the multivariable Leibniz rule: for a vector field V and scalar function f, \operatorname{div}(f V) = f \, \operatorname{div} V + \nabla f \cdot V = f \sum_{i=1}^n \frac{\partial V_i}{\partial x_i} + \sum_{i=1}^n \frac{\partial f}{\partial x_i} V_i. The Laplacian \Delta = \sum_{i=1}^n \frac{\partial^2}{\partial x_i^2} is a second-order operator on scalar functions, with principal \xi \mapsto -\lVert \xi \rVert^2, and satisfies the corresponding higher-order Leibniz condition, such as \Delta(f h) - f \, \Delta h = 2 \sum_{i=1}^n \frac{\partial f}{\partial x_i} \frac{\partial h}{\partial x_i} + h \Delta f, where the remainder is order at most 1 in h. On Riemannian manifolds, the \nabla provides a differential operator that generalizes partial to tensor fields while respecting the manifold's . For a Y along a with X, \nabla_X Y measures the rate of change of Y parallel to the , satisfying \nabla_X (f Y) = f \nabla_X Y + (X f) Y as the Leibniz rule. Its is 1, with the principal symbol determined by the metric and . Similarly, the D on a over a is a elliptic differential operator, locally expressed as D = \sum_{j=1}^n e_j \cdot \nabla_{e_j} for an orthonormal \{e_j\}, where \cdot denotes Clifford . As a differential operator, it satisfies the Leibniz rule D(f s) = f D s + c(df) s for a function f and section s, where c denotes Clifford . Its principal symbol is \sigma_D(X) = i \sum_{j=1}^n X(e_j) e_j \cdot, which is invertible for X \neq 0. An example of a differential operator with variable coefficients and mixed derivatives is the operator L = \frac{\partial}{\partial t} - \Delta, acting on functions u(t, x) in \mathbb{R} \times \mathbb{R}^n. This has order 2, dominated by the second-order spatial Laplacian \Delta, with principal symbol i \tau + \lVert \xi \rVert^2 in variables (\tau, \xi). It satisfies the Leibniz condition for order 2; for instance, in one spatial dimension, L(f u) = \partial_t (f u) - \partial_{xx} (f u) = (\partial_t f) u + f \partial_t u - [f_{xx} u + 2 f_x u_x + f u_{xx}], and f L u = f (\partial_t u - u_{xx}), so the difference is (\partial_t f) u - f_{xx} u - 2 f_x u_x, which rearranges to multiplication by (\partial_t f - f_{xx}) times u plus multiplication by -2 f_x times u_x, a operator in u.

Notations

In the univariate case, the differential operator corresponding to the first is commonly denoted by D, representing \frac{d}{dx}, with higher powers D^k indicating the k-th \frac{d^k}{dx^k}. In the multivariate setting, partial derivatives with respect to variables x_1, \dots, x_n are denoted by \partial_i = \frac{\partial}{\partial x_i} for i = 1, \dots, n. To compactly express higher-order partial derivatives, is standard: a multi-index \alpha = (\alpha_1, \dots, \alpha_n) is an n- of nonnegative integers, with |\alpha| = \sum_{i=1}^n \alpha_i denoting its order, and \partial^\alpha = \prod_{i=1}^n \partial_i^{\alpha_i} representing the corresponding mixed . This notation facilitates the over all partial derivatives of order at most k, as in \sum_{|\alpha| \leq k} a_\alpha(x) \partial^\alpha f(x). Polynomial differential operators are often symbolized as P(x, D), where D = (\partial_1, \dots, \partial_n) and P is a polynomial in the variables x and the formal symbols D_j, such as P(x, D) = \sum_{|\alpha| \leq m} a_\alpha(x) D^\alpha with D^\alpha = (-i)^{|\alpha|} \partial^\alpha in some conventions to align with Fourier analysis. On a smooth manifold M, the space of differential operators of order at most k acting on smooth sections of a vector bundle is denoted by \mathrm{Diff}^k(M), forming a filtered algebra under composition. Notational conventions vary between and physics: mathematicians typically use italicized or \mathcal{D} for formal differential operators and emphasize formal adjoints, while physicists often employ boldface or upright \mathbf{D} and prioritize Hermitian adjoints in contexts. For instance, the Laplacian operator may appear as \Delta = \sum_i \partial_i^2 in mathematical texts but as -\hbar^2 \nabla^2 in , highlighting domain-specific adjustments. Composition of differential operators can be left or right ordered, affecting the symbol in quantization schemes; in Weyl quantization, the symbol of the product Op(a) Op(b) corresponds to a symmetric (Weyl) ordering where multiplication operators act midway between left and right derivatives, given by the oscillatory integral formula for the composed symbol.

Fundamental Properties

General Properties

Differential operators are linear maps, meaning that for a differential operator P of order m, and functions u, v and scalars a, b, P(au + bv) = a P u + b P v. In appropriate function spaces, such as Sobolev spaces H^s(\mathbb{R}^n), these operators are continuous: a differential operator P(D) maps H^{s+m}_{\mathrm{loc}}(\Omega) continuously to H^s_{\mathrm{loc}}(\Omega) for all s \in \mathbb{R}. This continuity holds in the topology induced by the Sobolev norms, ensuring well-defined behavior on spaces of functions with controlled derivatives. The of two operators exhibits a Leibniz-type structure. If P has k and Q has m, then P \circ Q has k + m, and the leading terms in the composition arise from the product of the leading coefficients via a generalized Leibniz rule. Specifically, for operators P = \sum_{|\alpha| \leq k} p_\alpha(x) \partial^\alpha and Q = \sum_{|\beta| \leq m} q_\beta(x) \partial^\beta, the highest-order part is simply the product of the principal parts. Commutators with operators further illustrate this: for a operator D = \partial_j and smooth function f, [D, f] g = D(f g) - f D g = (\partial_j f) g, which is a zero-order operator, and in general, [D, f] reduces the order by at least one. Each differential operator P of k has a principal \sigma_k(P)(x, \xi) = \sum_{|\alpha| = k} a_\alpha(x) (i \xi)^\alpha, a of degree k in the cotangent variable \xi. This captures the highest-order behavior and is independent of lower-order terms. An is elliptic if its principal satisfies |\sigma_k(P)(x, \xi)| \geq c |\xi|^k for some c > 0 and all \xi \neq 0, ensuring the operator is "invertible" in the high-frequency regime and leading to improved regularity properties for solutions. For systems, ellipticity requires the matrix to be invertible for \xi \neq 0.

Fourier Interpretation

The Fourier transform provides a powerful of differential operators by transforming them into multiplication operators in the . For a u on \mathbb{R}^n, the Fourier transform \hat{u}(\xi) = \int_{\mathbb{R}^n} u(x) e^{-i x \cdot \xi} \, dx (up to normalization constants) converts partial derivatives into multiplications: the operator \partial_j acts as \widehat{\partial_j u}(\xi) = i \xi_j \hat{u}(\xi). More generally, for a multi-index \alpha, \partial^\alpha corresponds to (i \xi)^\alpha, so a linear differential operator P = \sum_{|\alpha| \leq m} a_\alpha(x) \partial^\alpha with coefficients a_\alpha has \sigma_P(x, \xi) = \sum_{|\alpha| \leq m} a_\alpha(x) (i \xi)^\alpha, a in \xi of degree at most m. For constant coefficients, \widehat{P u}(\xi) = \sigma_P(\xi) \hat{u}(\xi). For variable coefficients, P is a whose action involves an oscillatory integral with this , but \widehat{P u}(\xi) is not pointwise multiplication by \sigma_P(x, \xi) \hat{u}(\xi). A concrete example illustrates this correspondence: consider the first-order operator D = -i \frac{d}{dx} on \mathbb{R}, which is the momentum operator in quantum mechanics. Its Fourier transform yields \hat{D} f(\xi) = \xi \hat{f}(\xi), directly associating the operator with the frequency variable \xi as a multiplier. This multiplier property extends to higher-order operators, such as the Laplacian \Delta = \sum_{j=1}^n \partial_j^2, where \widehat{\Delta u}(\xi) = -|\xi|^2 \hat{u}(\xi), facilitating the solution of elliptic equations like \Delta u = f via \hat{u}(\xi) = -\hat{f}(\xi) / |\xi|^2 for \xi \neq 0. Such transformations underscore the role of differential operators as special cases of pseudodifferential operators, where the symbol is precisely a polynomial in \xi. Beyond basic , the interpretation connects to the of singularities in solutions to partial differential equations. Singularities in the set of a propagate along bicharacteristic curves defined by the flow of the principal \sigma_m(P)(x, \xi), the homogeneous leading term of degree m. This is analyzed using integral operators, which generalize pseudodifferential operators to handle phase shifts and oscillatory integrals, ensuring that singularities neither appear nor disappear except along these flows for hyperbolic or properly supported operators. The theorem, established through , applies to solutions of P u = f, where the set of u is contained in that of f union the flow-out from characteristic sets. Finally, the framework links operators to symbol quantization in Weyl calculus, a symmetric quantization where the associated to a a(x, \xi) is \mathrm{Op}_w(a) f(x) = (2\pi)^{-n} \iint e^{i (x-y) \cdot \eta} a\left(\frac{x+y}{2}, \xi\right) f(y) \, dy \, d\eta with \eta = \xi adjusted via oscillatory integrals. For symbols of operators, Weyl quantization coincides with the standard left or right quantizations due to the exact structure, providing a bridge to semiclassical and deformation quantization in . This connection preserves the total and enables precise control over composition via the .

Adjoint Operators

Formal Adjoint in One Variable

In the context of linear operators acting on functions in one variable, the formal D^* of an operator D is defined such that for suitable test functions f and g with compact support, the formula holds: \int_{-\infty}^{\infty} (D f) g \, dx = \int_{-\infty}^{\infty} f (D^* g) \, dx, where boundary terms vanish due to the compact support assumption. This definition ensures that the captures the duality between the operator and its action under the L^2 inner product, up to boundary contributions that are controlled in appropriate function spaces. For a polynomial differential operator D = \sum_{k=0}^m a_k(x) \frac{d^k}{dx^k} with coefficients a_k(x), the explicit form of the formal is D^* g = \sum_{k=0}^m (-1)^k \frac{d^k}{dx^k} \left( a_k(x) g \right). This formula arises from applying the (Leibniz rule) repeatedly during to transfer all derivatives from f to g. The derivation begins with the first-order case and proceeds inductively: for the differentiation \frac{d}{dx}, gives \int (f') g \, dx = -\int f g' \, dx, so \left( \frac{d}{dx} \right)^* = -\frac{d}{dx}. For higher orders, the Leibniz rule for adjoints follows: if D = P \frac{d^k}{dx^k} with multiplication by P(x), then (D)^* = (-1)^k \frac{d^k}{dx^k} (P \cdot), and the full sums these terms. A key example is the first-order operator, where D = \frac{d}{dx} yields D^* = -\frac{d}{dx}, as noted above. For a second-order operator D = \frac{d^2}{dx^2} + b(x) \frac{d}{dx} + c(x), repeated produces D^* = \frac{d^2}{dx^2} - b \frac{d}{dx} + (c - b'), where b' denotes the of b with respect to x. This reflects the sign flip for odd-order terms and the adjustment for variable coefficients via the . An operator D is formally self-adjoint if D = D^*, a condition that simplifies the analysis of symmetric problems in partial differential equations and ensures real eigenvalues under suitable boundary conditions. For instance, the pure second derivative \frac{d^2}{dx^2} satisfies this property, as its adjoint is itself.

Formal Adjoint in Several Variables

In several variables, the formal adjoint of a linear partial differential operator extends the concept from the one-dimensional case by incorporating the multivariable integration by parts formula, often derived via the divergence theorem. For a domain \Omega \subset \mathbb{R}^n with smooth boundary, consider smooth functions f, g with appropriate support or boundary conditions such that boundary integrals vanish. The formal adjoint D^* of a linear partial differential operator D is defined by the relation \int_\Omega (D f) g \, dx = \int_\Omega f (D^* g) \, dx, where the integrals are taken with respect to the Lebesgue measure on \Omega. This definition ensures that D^* is the unique differential operator satisfying the bilinear pairing identity for test functions, ignoring boundary terms in the formal sense. For a general linear partial differential operator of at most m, D f = \sum_{|\alpha| \leq m} a_\alpha(x) \partial^\alpha f, where \alpha = (\alpha_1, \dots, \alpha_n) is a multi-index with |\alpha| = \sum_{i=1}^n \alpha_i, \partial^\alpha = \prod_{i=1}^n \frac{\partial^{\alpha_i}}{\partial x_i^{\alpha_i}}, and the coefficients a_\alpha(x) are smooth functions on \Omega, the formal is given by D^* g = \sum_{|\alpha| \leq m} (-1)^{|\alpha|} \partial^\alpha (a_\alpha(x) g). This expression preserves the m of the operator and transforms the leading terms accordingly. The formula for D^* arises from repeated applications of the multivariable rule, which leverages the for derivatives and the . For a single first-order , integration by parts yields \int_\Omega (\partial_i f) g \, dx = -\int_\Omega f (\partial_i g) \, dx + \int_{\partial \Omega} f g n_i \, dS, where n_i is the i-th component of the outward unit normal, showing that the formal of \partial_i is -\partial_i when terms are neglected. For a term of higher order \partial^\alpha f, integration by parts is applied successively to each \partial_i^{\alpha_i}, introducing a factor of (-1)^{\alpha_i} per variable and pulling the coefficient a_\alpha inside the derivatives via the Leibniz rule: \partial^\alpha (a_\alpha g) = \sum_{\beta \leq \alpha} \binom{\alpha}{\beta} (\partial^\beta a_\alpha) (\partial^{\alpha - \beta} g). The overall sign (-1)^{|\alpha|} accounts for the total number of integrations by parts across all variables. A fundamental illustration is the divergence operator \operatorname{div}: C^\infty(\Omega; \mathbb{R}^n) \to C^\infty(\Omega), defined by \operatorname{div} V = \sum_{i=1}^n \partial_i V_i for a V = (V_1, \dots, V_n). Its formal is the negative (\operatorname{div})^* \phi = -\nabla \phi = (-\partial_1 \phi, \dots, -\partial_n \phi), satisfying \int_\Omega (\operatorname{div} V) \phi \, dx = -\sum_{i=1}^n \int_\Omega V_i (\partial_i \phi) \, dx + \int_{\partial \Omega} \phi (V \cdot n) \, dS. This relation follows directly from applying to each component, confirming the structure for vector-valued operators. The distinction between the formal adjoint and the L^2 adjoint lies in their settings: the formal D^* is a purely differential expression without specified , applicable to functions, whereas the L^2 adjoint is the densely defined on the L^2(\Omega) whose ensures the holds for functions in its maximal , typically requiring D^* g \in L^2(\Omega). Under assumptions that compactly supported functions are dense in L^2(\Omega) and the coefficients a_\alpha are sufficiently regular (e.g., bounded and continuous), the L^2 adjoint coincides with the formal on this dense , enabling extension by continuity.

Examples of Adjoints

A classic example in one variable is the Euler operator E = x \frac{d}{dx}, acting on smooth functions on (0, \infty). Its formal with respect to the L^2 inner product \langle f, g \rangle = \int_0^\infty f(x) g(x) \, dx is E^* = -x \frac{d}{dx} - 1. To verify this, consider \langle E f, g \rangle = \int_0^\infty g(x) \left( x f'(x) \right) dx. yields [g(x) x f(x)]_0^\infty - \int_0^\infty f(x) \frac{d}{dx} \left( x g(x) \right) dx = B - \int_0^\infty f(x) \left( x g'(x) + g(x) \right) dx, where B denotes boundary terms that vanish for suitable test functions with compact support. Thus, \langle E f, g \rangle = B - \langle f, x g' + g \rangle, so E^* g = - x g' - g = -x \frac{d}{dx} g - g. In several variables, the gradient operator \nabla: C_c^\infty(\mathbb{R}^n) \to C_c^\infty(\mathbb{R}^n, \mathbb{R}^n) has formal adjoint \nabla^* = -\operatorname{div}, where \operatorname{div} \mathbf{v} = \sum_{i=1}^n \frac{\partial v_i}{\partial x_i}. This follows from : \langle \nabla u, \mathbf{v} \rangle = \int_{\mathbb{R}^n} \nabla u \cdot \mathbf{v} \, dx = -\int_{\mathbb{R}^n} u \operatorname{div} \mathbf{v} \, dx + B, with boundary terms B=0 for compactly supported functions. Consequently, the Laplacian \Delta = \operatorname{div} \circ \nabla is formally , \Delta^* = \Delta, as \langle \Delta u, v \rangle = \langle \nabla u, \nabla v \rangle = \langle u, \Delta v \rangle by applying the adjoint twice. In physics, the momentum operator p = -i \frac{d}{dx} on L^2(\mathbb{R}), restricted to the dense domain C_c^\infty(\mathbb{R}) of smooth compactly supported functions, is essentially self-adjoint. This means its closure is self-adjoint, ensuring a unique self-adjoint extension used in quantum mechanics for the free particle Hamiltonian. Essential self-adjointness follows from the fact that the deficiency indices are zero, verified via solutions to p^* \psi = \pm i \psi, which lie outside L^2(\mathbb{R}). For a variable coefficient operator in \mathbb{R}^2, consider P = \frac{\partial}{\partial x} + x \frac{\partial}{\partial y}. The formal adjoint is P^* = -\frac{\partial}{\partial x} - x \frac{\partial}{\partial y} - 1. Verification proceeds by integration by parts in the inner product \langle P u, v \rangle = \int_{\mathbb{R}^2} v \left( u_x + x u_y \right) dx dy = B - \int_{\mathbb{R}^2} u \left( v_x + (x v)_y \right) dx dy = B - \int_{\mathbb{R}^2} u \left( v_x + x v_y + v \right) dx dy, where the extra -1 arises from \frac{\partial x}{\partial y} = 1. Thus, P^* v = - v_x - x v_y - v. The formal adjoint ignores boundary terms and is defined locally on smooth functions, but the actual adjoint in a Hilbert space like L^2(\Omega) depends on the domain, incorporating boundary conditions to ensure \langle L u, v \rangle = \langle u, [L^*](/page/Adjoint) v \rangle without boundary contributions. For instance, on a bounded [a,b], the \frac{d}{dx} requires boundary conditions (e.g., Dirichlet u(a)=u(b)=0) for self-adjointness, altering the domain of the adjoint relative to the formal version.

Algebraic Structures

Ring of Univariate Polynomial Differential Operators

The ring of univariate polynomial differential operators, often denoted \operatorname{Diff}(R) for a R, is the generated by the polynomials R and the differentiation \frac{d}{dx}, where elements are finite sums \sum_{i=0}^n f_i \left(\frac{d}{dx}\right)^i with f_i \in R. The multiplication in \operatorname{Diff}(R) is defined via composition, incorporating the Leibniz rule: for f, g \in R, (f \frac{d}{dx})(g) = f \frac{d g}{dx} + f' g, where f' denotes the formal of f with respect to x. This structure ensures that \operatorname{Diff}(R) acts naturally on R as a ring of endomorphisms. \operatorname{Diff}(R) admits an Ore extension presentation: \operatorname{Diff}(R) \cong R[\partial; \delta], where \delta is the standard on R given by \delta(x) = 1 and extended R-linearly, with the multiplication rule \partial \cdot a = \sigma(a) \partial + \delta(a) for a \in R and \sigma = \mathrm{id}. This construction highlights the non-commutative nature of the , arising from the commutation [\frac{d}{dx}, x] = \frac{d}{dx} \cdot x - x \cdot \frac{d}{dx} = 1. The algebra \operatorname{Diff}(R) is the first Weyl algebra when R is a of characteristic zero, generated by x and \partial (with \partial corresponding to \frac{d}{dx}) subject to the key relation \partial x - x \partial = 1. This relation generates the entire non-commutative structure, distinguishing the Weyl algebra from commutative polynomial rings. For R = \mathbb{C}, the Weyl algebra is simple, possessing no nontrivial two-sided ideals.

Ring of Multivariate Polynomial Differential Operators

The ring of multivariate polynomial differential operators, often denoted as \operatorname{Diff}(\mathbb{R}^n) or D(\mathbb{R}^n), is the noncommutative over \mathbb{R} (or \mathbb{C}) generated by the coordinate multiplication operators x_1, \dots, x_n and the partial differentiation operators \partial_1 = \frac{\partial}{\partial x_1}, \dots, \partial_n = \frac{\partial}{\partial x_n}, subject to the commutation relations [\partial_i, x_j] = \delta_{ij}, [\partial_i, \partial_j] = 0, and [x_i, x_j] = 0 for all i, j = 1, \dots, n, where \delta_{ij} is the . These relations ensure that the algebra faithfully represents the action of linear partial differential operators with coefficients on the space of smooth functions on \mathbb{R}^n. This algebra is known as the n-th Weyl algebra, denoted A_n, and can be formally constructed as the quotient A_n = \mathbb{C}\langle x_1, \dots, x_n, \partial_1, \dots, \partial_n \rangle / I, where I is the two-sided ideal generated by the specified commutators. The multivariate structure generalizes the univariate case, introducing additional generators and relations that capture interactions across multiple variables. The Weyl algebra A_n carries a natural filtration by operator order, where the j-th filtered component F^j A_n consists of elements of total at most j in the \partial_i's (with x_i's having order 0). The associated \operatorname{gr} A_n = \bigoplus_j F^j A_n / F^{j-1} A_n is isomorphic to the commutative \mathbb{C}[x_1, \dots, x_n, \partial_1, \dots, \partial_n] in $2n variables, reflecting the commutative approximation of the noncommutative structure. In the global setting, A_n acts on the entire space \mathbb{R}^n or \mathbb{C}^n, whereas local versions arise as stalks of the sheaf of differential operators on algebraic varieties, such as the sheaf \mathcal{D}_X over a smooth variety X. Over an of characteristic zero, such as \mathbb{C}, the Weyl algebra A_n(\mathbb{C}) is simple, possessing no nontrivial two-sided ideals. A key application arises in quantum mechanics, where the commutation relations [\partial_i, x_j] = \delta_{ij} (up to scaling by i\hbar) encode the canonical commutation relations of the Heisenberg algebra, governing the position and momentum operators in n-dimensional phase space.

Coordinate-Independent Description

In the coordinate-independent setting, differential operators on a smooth manifold M act between smooth sections of vector bundles E \to M and F \to M. The space \mathrm{Diff}^k(E, F) consists of all linear maps P: C^\infty(M, E) \to C^\infty(M, F) of order at most k, defined such that for any point x \in M, there exists a neighborhood U of x where the iterated commutator [\cdots [P, m_{s_1}], m_{s_2}] \cdots, m_{s_{k+1}}] vanishes for all smooth sections s_1, \dots, s_{k+1} of E, with m_s denoting pointwise multiplication by the section s. This characterization ensures the operators satisfy a generalized Leibniz rule, extending the product rule to higher orders via tensor products of sections: for an order-k operator, P(f \cdot s) = \sum_{j=0}^k \binom{k}{j} (P_j f) \cdot \nabla^j s, where \nabla is a connection and P_j are lower-order terms, though the precise form depends on the bundle structure. Locally, in a trivialization of E and F over a (U, \phi) on M, such operators reduce to the classical coordinate form \sum_{|\alpha| \leq k} a_\alpha(x) \partial^\alpha, where a_\alpha are sections and \partial^\alpha are partial derivatives. Globally, however, the coordinate-free description employs bundles J^k(E), which parametrize the k-th order of sections of E—equivalence classes of sections agreeing up to k-th order derivatives at a point—allowing differential operators to be viewed as morphisms between jet bundles and F. Covariant derivatives induced by a on E provide prototypical order-one differential operators in \mathrm{Diff}^1(E, E \otimes T^*M), mapping sections to covector-valued sections while preserving the Leibniz rule \nabla_X (f s) = (\nabla_X f) s + f \nabla_X s for vector fields X. Higher-order operators arise naturally from compositions of such covariant derivatives, generating the full space \mathrm{Diff}^*(E, F) in a manner independent of local coordinates. The collection of all differential operators \bigcup_k \mathrm{Diff}^k(E, F) forms a filtered ring under composition, with the filtration \mathrm{Diff}^k(E, F) \subseteq \mathrm{Diff}^{k+1}(E, F) preserved such that the product of order-k and order-l operators has order at most k+l. The associated graded ring is isomorphic to the ring of symbols via the principal symbol map \sigma_k: \mathrm{Diff}^k(E, F)/\mathrm{Diff}^{k-1}(E, F) \to \Gamma(S^k(T^*M) \otimes \mathrm{Hom}(E, F)), where S^k(T^*M) denotes symmetric k-th powers of the , rendering the symbol sequence exact and facilitating algebraic analysis. A canonical example is the de Rham complex on M, a chain complex of differential forms \Omega^\bullet(M) where the exterior derivative d: \Omega^p(M) \to \Omega^{p+1}(M) serves as a first-order differential operator satisfying d^2 = 0 and the graded Leibniz rule d(\alpha \wedge \beta) = d\alpha \wedge \beta + (-1)^{\deg \alpha} \alpha \wedge d\beta.

Advanced Variants

Differential Operators of Infinite Order

Differential operators of infinite order generalize finite-order differential operators by allowing formal power series expansions in the , typically expressed in form to ensure on suitable spaces. Specifically, such an D on f is defined as Df(x) = \sum_{k=0}^\infty \frac{a_k}{k!} \partial_x^k f(x), where the coefficients \{a_k\} form a with positive , making the series converge for analytic or in appropriate topologies. This form arises naturally from composing the with the \phi(z) = \sum_{k=0}^\infty a_k \frac{z^k}{k!}, which is entire, ensuring the \phi(\partial_x) acts continuously on spaces beyond . The operator satisfies a Leibniz rule derived from that of the derivative: D(fg)(x) = \sum_{k=0}^\infty \frac{a_k}{k!} \sum_{j=0}^k \binom{k}{j} \partial_x^j f(x) \cdot \partial_x^{k-j} g(x). Prominent examples include the translation operator e^{h \partial_x}, which shifts functions via e^{h \partial_x} f(x) = f(x + h) for h \in \mathbb{R}, and the heat semigroup operator e^{t \Delta} for t > 0, where \Delta is the Laplacian, generating solutions to the heat equation \partial_t u = \Delta u. These operators extend the finite-order case, where the series truncates, and are well-defined on entire functions of exponential type, preserving properties like reality of zeros under iteration. In terms of , infinite-order differential operators are continuous when acting on Gevrey classes \mathcal{G}^s, spaces of functions where higher derivatives satisfy |\partial^k f(x)| \leq C^{k+1} (k!)^s for some C > 0 and s \geq 1, with analytic functions corresponding to s=1. This continuity holds for symbols in Gevrey classes, ensuring boundedness in norms adapted to the growth of derivatives. The symbol of such an operator, obtained via , is an in the frequency variable \xi, extending the principal symbol concept from finite-order operators to \phi(i\xi), where \phi is of exponential type. Applications of infinite-order differential operators appear prominently in solving partial differential equations (PDEs) through formal power series methods, where they facilitate the construction of fundamental solutions or semigroups for evolution equations. For instance, operators like e^{t \Delta} directly yield the Green's function for the heat equation, while more general \phi(\Delta_{\theta,\omega}) solve Cauchy problems for second-order PDEs with variable coefficients, converging in spaces of entire functions to provide asymptotic behaviors or zero distributions of solutions.

Bidifferential Operators

A bidifferential operator is a B: C^\infty(M) \times C^\infty(M) \to C^\infty(M) that acts as a differential operator in each argument separately, generalizing the notion of bilinear forms to incorporate . Locally, on an in \mathbb{R}^n, a bidifferential operator of total order at most k takes the form B(f,g) = \sum_{|\alpha| + |\beta| \leq k} a_{\alpha\beta}(x) \, \partial^\alpha f(x) \, \partial^\beta g(x), where the coefficients a_{\alpha\beta} are smooth functions and \alpha, \beta are multi-indices. This expression ensures bilinearity and the property that, for fixed g, B(\cdot, g) is a differential operator of order at most k in the first variable, and analogously for fixed f. The of B is defined as the maximum of |\alpha| + |\beta| over all nonzero terms, measuring the highest combined degree of . Composition of bidifferential operators preserves this structure: if B has order k and C has order l, then C \circ B has order at most k + l, as compose additively. The operators satisfy the standard in each argument separately, arising from the for . Representative examples include the zeroth-order product operator B(f,g) = f g, which is simply multiplication, and the first-order operator B(f,g) = f \partial_x g in one dimension, combining multiplication in the first argument with differentiation in the second. In number theory, Rankin–Cohen brackets provide higher-order examples, defined for modular forms f_1, f_2 of weights \lambda', \lambda'' and order \ell by R_{\lambda',\lambda''}^{\lambda' + \lambda'' + \ell}(f_1, f_2)(z) = \sum_{j=0}^\ell \binom{\lambda' + \lambda'' + \ell - 1 - j}{\ell - j} \frac{f_1^{(\ell - j)}(z) f_2^{(j)}(z)}{(2i\pi)^{\ell}}, a bidifferential operator of order \ell that is invariant under the modular group. In quantum field theory, Wick products, such as the normal-ordered bilinear form on fields \partial \phi \cdot \partial \psi, function as bidifferential operators by subtracting vacuum expectations to ensure proper renormalization. The formal adjoint B^* of a bidifferential operator B is defined such that yields \int B(f,g) \, h = \int f \, B^*(g,h) (up to boundary terms), resulting in B^*(f,g) = \sum (-1)^{|\beta|} a_{\alpha\beta} \partial^\alpha g \, \partial^\beta f after transposing arguments and applying signs from the adjoint of each derivative (\partial^* = -\partial). This introduces sign changes depending on the order in the second argument. Bidifferential operators are closely related to tensor products of differential operators: the space of such operators of order at most k corresponds to elements in the tensor product \mathrm{Diff}^k(M) \otimes \mathrm{Diff}^k(M), where \mathrm{Diff}^k(M) is the module of differential operators of order \leq k on M, acting diagonally on pairs of functions via (P \otimes Q)(f \otimes g) = P(f) Q(g). This structure underlies their role in deformation quantization and .

Microdifferential Operators

Microdifferential operators arise in as a refinement of differential operators, enabling precise localization of their action on the T^*M of a smooth manifold M. These operators act on Lagrangian distributions, which are distributions associated to Lagrangian submanifolds of T^*(M \times M) and generalize smooth and their singularities in . Formally, a microdifferential P of order m is defined via an oscillatory integral representation: (Pu)(x) = \int e^{i\phi(x,y,\theta)} a(x,y,\theta) u(y) \, dy \, d\theta, where \phi is a non-degenerate function whose is a submanifold \Lambda \subset T^*(M \times M), and the amplitude a belongs to the symbol class S^m(T^*M \times T^*M), consisting of smooth functions satisfying estimates |\partial^\alpha_x \partial^\beta_y \partial^\gamma_\theta a| \leq C (1+|\theta|)^{m - |\gamma|} for multi-indices \alpha, \beta, \gamma. This structure allows microdifferential operators to capture propagation phenomena in , extending the classical Leibniz rule to infinite-order formal series while preserving algebraic properties like . A central concept in the is microlocal ellipticity, which occurs when the principal p_m(x,y,\theta) \in S^m(T^*M \times T^*M) is invertible on \Lambda away from the characteristic variety \mathrm{Char}(P) = \{(x,y,\theta) \in \Lambda \mid p_m(x,y,\theta) = 0\}. Elliptic microdifferential operators propagate singularities along bicharacteristic strips in the , ensuring that singularities of solutions to equations Pu = f follow the flow of the principal . This propagation of singularities is crucial for analyzing hyperbolic and elliptic partial differential equations, where the operator dictates how wavefronts evolve microlocally. For instance, in wave equations, microlocal ellipticity guarantees finite speed, with singularities confined to conical neighborhoods of the in . Pseudodifferential operators provide a key example of microdifferential operators of order zero, where the phase function is the standard bilinear form \phi(x,y,\theta) = (x-y) \cdot \theta, and the Lagrangian \Lambda is the conormal bundle to the diagonal in M \times M. In this case, the full symbol admits an asymptotic expansion a(x,\theta) \sim \sum_{j=0}^\infty a_{m-j}(x,\theta) with a_{m-j} homogeneous of degree m-j in \theta, enabling the operator to be expressed as P = \sum_{k=0}^\infty \frac{(-i)^k}{k!} \partial_\theta^k a(x,\theta) D_x^k in local coordinates. This representation highlights their role in smoothing or amplifying singularities based on the symbol's behavior. The relation to wavefront sets underscores the microlocal nature of these operators: the wavefront set \mathrm{WF}(u) \subset T^*M \setminus 0 measures the singular directions of a u, and a microdifferential P smooths singularities outside its set, meaning that if (x,\xi) \notin \mathrm{Char}(P), then u is smooth microlocally near (x,\xi) implies Pu is smooth there. More precisely, \mathrm{WF}(Pu) \cap (T^*M \setminus \mathrm{Char}(P)) \subset \mathrm{WF}(u) \cap (T^*M \setminus \mathrm{Char}(P)), ensuring that P does not introduce new singularities away from the characteristic variety. This property is foundational for proving local solvability and hypoellipticity in PDE theory. As an advanced generalization, Fourier integral operators extend microdifferential operators by allowing arbitrary clean canonical relations—Lagrangian immersions \Lambda \subset T^*M \times T^*N \setminus 0—rather than restricting to graph-like structures over the diagonal. These operators, also realized via oscillatory integrals with symbols in appropriate classes, model changes of variables and processes, preserving microlocal ellipticity and wavefront set propagation in broader geometric settings.

References

  1. [1]
    Differential Operator -- from Wolfram MathWorld
    Differential Operator ; (xD^~)^1, = xD^~ ; (xD^~)^2, = xD^~+x^2D^~^2 ; (xD^~)^3, = xD^~+3x^2D^~^2+x^3D^ ; (xD^~)^4, = xD^~+7x^2D^~^2+6x^3D^.
  2. [2]
    [PDF] Some Notes on Differential Operators - MIT OpenCourseWare
    A. Introduction. In Part 1 of our course, we introduced the symbol D to denote a func- tion which mapped functions into their derivatives. In other words,.
  3. [3]
    [PDF] Chapter 10: Linear Differential Operators and Green's Functions
    Differential operators are important for the study of differential equations and we would like to analyze them in spite of their lack of continuity. There are ...Missing: mathematics | Show results with:mathematics
  4. [4]
    Differential operator – Knowledge and References - Taylor & Francis
    These operators are used extensively in mathematical physics and engineering to describe physical phenomena such as fluid flow, electromagnetic fields, and ...
  5. [5]
    Fundamental Solutions for Differential Operators and Applications
    This book introduces fundamental solutions for differential operators and their applications to boundary value problems in mathematical physics, applied  ...
  6. [6]
    10.133 Finite order differential operators - Stacks project
    ... differential operator of order k (see proof of Lemma 10.133.3). By the ... This is a pleasant calculation using the Leibniz rule which is left to the reader. \ ...
  7. [7]
    [PDF] Elliptic Operators and Analytic Index
    Definition: Let M be a smooth compact manifold. We say that L : C∞(M) → C∞(M) is a differential operator. (on M) if L is of the form above in any chart Uα ...
  8. [8]
    A History of Mathematics/Modern Europe/Euler, Lagrange, and ...
    Apr 5, 2014 · Euler introduced in 1766 the ​name "calculus of variations," and did much to improve this science along the lines marked out by Lagrange.
  9. [9]
    Earliest Uses of Symbols of Calculus
    Dec 1, 2004 · Dx y was introduced by Louis François Antoine Arbogast (1759-1803) in "De Calcul des dérivations et ses usages dans la théorie des suites et ...
  10. [10]
    None
    Summary of each segment:
  11. [11]
    Motivation for and history of pseudo-differential operators
    Mar 21, 2011 · The language was invented to provide a context for talking about solution operators for differential equations.History of differential forms and vector calculus - MathOverflowHistorical developement of analysis and partial differential equations ...More results from mathoverflow.net
  12. [12]
    A history of Quantum Mechanics - MacTutor - University of St Andrews
    In 1901 Ricci and Levi-Civita published Absolute differential calculus. It ... Schrödinger introduced operators associated with each dynamical variable.
  13. [13]
    [PDF] Introduction to Partial Differential Operators - Nikolai Nowaczyk
    1.1.1 Definition (Differential Operator). Let U ⊂ Rn be open. A complex (partial) differential. operator on U of order k, k ∈ N, (a ”PDO”) is a C-linear map P ...
  14. [14]
    [PDF] GR lecture 5 Covariant derivatives, Christoffel connection ...
    In other words, such an operator ∇µ requires some extra geometric structure on our spacetime. When it exists, ∇µ is called a covariant derivative. The ...
  15. [15]
    [PDF] the dirac operator - UChicago Math
    Mar 30, 2012 · The Dirac operator is clearly a first-order differential operator. We can work out its symbol fairly easily. The symbol1 of the connection ...
  16. [16]
    [PDF] Chapter 6 Partial Differential Equations
    is a “backward” heat-equation operator. The corresponding “backward”. Green ... dimensional differential operator. For other than rectangular co ...
  17. [17]
    [PDF] differential operators on euclidean spaces - MIT Mathematics
    Differential operators have the following properties: Proposition 1.3. We have. (1) Diffm(Rn) ⊂ Diffm+1(Rn) for all m ≥ 0;. (2) the identity operator I lies ...
  18. [18]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    where we use the multi-index notation for partial derivatives explained in Sec- ... multi-indices, we define the multi-index (α + β) by α + β = (α1 + β1,α2 ...
  19. [19]
    [PDF] Introduction to pseudodifferential operators 1. Kohn-Nirenberg ...
    Pseudodifferential operators, in the Kohn-Nirenberg calculus, are defined by p(x, D) attached to a symbol p(x, ξ) in the simplest symbol class Sm.<|separator|>
  20. [20]
    [PDF] Linear Analysis on Manifolds: Notes for Math 7376, Spring 2016
    Jun 14, 2016 · For this reason, the somewhat unusual choice was made to use pseudodifferential operators on manifolds in order to prove key results about ...
  21. [21]
    [PDF] Pseudodifferential Operators - MIT Mathematics
    Left and right quantization are both bijections, but from this formula it is clear ... [4]. , The Weyl calculus of pseudo-differential operators, Comm. Pure ...
  22. [22]
    [PDF] 18.156 – Spring 2008 – Graduate Analysis Elliptic regularity and ...
    Feb 7, 2008 · differential operator (elliptic or not) gives continuous linear maps. (2.50) P(z, D) : Hs+m loc. (Ω) −→ Hs loc(Ω), ∀ s ∈ R, P(z, D) : C ...
  23. [23]
    Partial Differential Equations: Second Edition - AMS Bookstore
    I use Partial Differential Equations to prepare my students for their Topic exam, which is a requirement before starting working on their dissertation. The book ...
  24. [24]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    These are notes from a two-quarter class on PDEs that are heavily based on the book Partial Differential Equations by L. C. Evans, together with other sources ...
  25. [25]
    [PDF] Differential operators and Fourier methods
    May 26, 2016 · The Fourier transform and pseudo-differential operators. A very important operator is the Fourier transformation F, it is an integral operator.
  26. [26]
    [PDF] 1. Symbols of linear partial differential operators
    aα(x)ξα. It is a bijective mapping from differential operators to the polynomials of ξ with smooth coefficients. However, since differential operators do not ...<|control11|><|separator|>
  27. [27]
    [PDF] Fourier integral operators. II
    It involves a first order differential operator depending on A 1 which acts on the principal symbol of A s. The solvability of the equation A1A s = B with given ...
  28. [28]
    Some notes on Weyl quantisation - Terry Tao
    Oct 7, 2012 · I decided to write some notes on the Weyl functional calculus (also known as Weyl quantisation), and to sketch the applications of this calculus ...Missing: right | Show results with:right
  29. [29]
    [PDF] Chapter 4 Linear Differential Operators
    A linear differential operator is a formal operator, like p0(x)∂n + ... + pn(x), with a specified domain, such as L2[a,b], and boundary conditions.
  30. [30]
    [PDF] A Short Course on Duality, Adjoint Operators, Green's Functions ...
    Aug 6, 2004 · The basic technique for obtaining the formal adjoint for differential operators posed in higher space dimensions is the divergence theorem.<|control11|><|separator|>
  31. [31]
    [PDF] The duality between the gradient and divergence operators on ...
    Oct 15, 2012 · It is common knowledge that the formal adjoint of the gradient operator, grad, on an n-dimensional domain is minus the divergence operator, div.
  32. [32]
    [PDF] Class 12. Linear differential operators and the fundamental
    Oct 8, 2024 · We now extend the concept of linear differential operators to smooth manifolds. It should be clear how to define local differential operators ...
  33. [33]
    [PDF] 8 Self-adjointness
    The position operator x and the momentum operator p are self- adjoint. Both operators are essentially self-adjoint when restricted to either. C∞ c (R) or S ...
  34. [34]
    [PDF] THE WEYL ALGEBRAS - School of Mathematics & Statistics | Science
    This is the defining relation of the first Weyl algebra which can be viewed as the ring of differential operators on k[x] with polynomial coefficients. There ...
  35. [35]
    (PDF) ON DERIVATIONS OF DIFFERENTIAL ORE EXTENSIONS
    Oct 17, 2022 · k[x][t;d/dx] is called the Weyl algebra A1, and it can be presented by A1=k⟨x, t :tx−xt = 1⟩. It is clear that the set {xi ...
  36. [36]
    [PDF] Ore Localizations and Irreducible Representations of the First Weyl ...
    The first Weyl algebra, denoted by A = At = At(C), is defined as a C-algebra with two generators q and p and a relation [P, q] = pq - qp = 1, where C is the ...
  37. [37]
    [PDF] d-modules.pdf
    Jan 28, 2025 · where Q is a differential operator of order k with real-analytic coefficients, such that ... that satisfies the Leibniz rule ∇(fs) = df ...
  38. [38]
    [PDF] Heisenberg-Weyl algebra revisited: Combinatorics of words and paths
    The Heisenberg–Weyl algebra, which underlies virtually all physical representations of Quantum. Theory, is considered from the combinatorial point of view.
  39. [39]
    None
    Below is a merged summary of the sections on Differential Operators \( D^k(U) \), Manifolds, Bundles \( E, F \), \( \text{Diff}^k(E,F) \), and Filtered Structure from "Analysis on Manifolds," combining all the information from the provided summaries into a concise yet comprehensive response. To retain maximum detail, I will use a table in CSV format for clarity and density, followed by a narrative summary to tie it all together. The table will capture key definitions, properties, and relationships across the sections, while the narrative will provide context and additional insights.
  40. [40]
    [PDF] Coordinate-Independent Computations on Differential Equations
    Mar 27, 1998 · operators can be defined in a nicely coordinate-independent way on manifolds. A linear. PDE on a manifold then takes the form. Lf = g;. (3.4).
  41. [41]
    [PDF] lecture 19: differential operators on manifolds
    Differential operators on manifolds. We are aiming at extending the definition of semiclassical pseudodifferential operators from Rn to manifolds.
  42. [42]
    [PDF] Lectures on D-modules. - Yale Math
    Feb 1, 2020 · Introduction. Part 1: Differential operators. I. Algebraic background. 1. Filtered rings and modules.
  43. [43]
    [PDF] Differential operators of infinite order and the distribution of zeros of ...
    Benz [2] applied the operator 1/φ(D), φ ∈ L-P, to investigate the distribution of zeros of certain exponential polynomials. The operators φ(D), φ ∈ L-P, play a ...
  44. [44]
    [PDF] Infinite order differential operators acting on entire ... - arXiv
    Oct 9, 2020 · functions with exponential bounds play a crucial role in the continuity of infinite order differential operators acting on these two classes ...
  45. [45]
    None
    ### Summary of arXiv:math/0311213
  46. [46]
    DIFFERENTIAL OPERATORS OF INFINITE ORDER WITH ...
    Differential operators of infinite order with symbols that are infinitely differentiable functions (of Gevrey class) in some domain in are considered.
  47. [47]
    Pseudodifferential operators of infinite order and Gevrey classes
    Feb 6, 2008 · Viene definita una classe di operatori pseudodifferenziali di ordine infinito che agiscono su spazi di funzioni Gevrey e sui loro duali.
  48. [48]
    bilinear differential operators: projectively equivariant symbol and ...
    Mar 10, 2014 · A bilinear differential operator. A ∈ Pk λ1,λ2,μ is of the form. (6). A = k. =0 i+j= ai,j (x) di dxi. ⊗ dj dxj where ai,j (x) are smooth ...
  49. [49]
    None
    ### Summary of Bi-Differential Operators in Deformation Quantization from arXiv:math/0111094
  50. [50]
    [PDF] Sobolev spaces and -differential operators - arXiv
    Oct 29, 2020 · We also introduce and study the notion of a ∇-bidifferential operator (a bilinear version of differential oper- ators), obtaining results ...
  51. [51]
    [PDF] On the generalized bilinear differential equations ( ) - IOSR Journal
    In this paper we would like tofurther confirm the kind of generalized bilinear differential operator and examine when is the linear superposition principle ...
  52. [52]
    A generating operator for Rankin–Cohen brackets - ScienceDirect
    Aug 15, 2025 · ... ℓ − j f 1 ( z ) ∂ z ℓ − j ∂ j f 2 ( z ) ∂ z j , which is the Rankin–Cohen bidifferential operator R λ ′ , λ ″ λ ‴ ( f 1 , f 2 ) at ( λ ...
  53. [53]
    Tensor products of holomorphic representations and bilinear ...
    ... tensor product H ν 1 ⊗ H ν 2 modulo the subspace of functions vanishing of ... Bilinear differential operators. Representations of Lie groups.