Fact-checked by Grok 2 weeks ago

Canonical coordinates

In , canonical coordinates consist of generalized position coordinates q_i and their conjugate momenta p_i, forming the fundamental variables in the of , where the system's is governed by Hamilton's equations \dot{q}_i = \frac{\partial H}{\partial p_i} and \dot{p}_i = -\frac{\partial H}{\partial q_i}, with [H](/page/H+) denoting the . These coordinates parameterize the , a $2N-dimensional manifold for a system with N , enabling a structure that preserves the geometric properties of motion. The concept originated in the development of during the 1830s, with introducing the in 1834 and formalizing canonical transformations in 1837 as a means to simplify the while preserving the form of Hamilton's equations. Jacobi's initial theorem on these transformations, though lacking a full proof at the time, laid the groundwork for later advancements; proofs emerged in the mid-19th century from mathematicians like Adolphe Desboves and William Donkin, but provided the modern variational proof in 1899, establishing their role in and integrability. By the early 20th century, canonical coordinates became central to through Dirac's , which maps Poisson brackets to commutators, bridging classical and quantum descriptions. Key aspects of canonical coordinates include their transformation properties: a canonical transformation maps (q, p) to new variables (Q, P) via generating functions (e.g., F_1(q, Q, t) or F_2(q, P, t)), ensuring the new Hamiltonian K(Q, P, t) yields equivalent dynamics while often simplifying the problem, such as rendering coordinates cyclic. These transformations preserve Poisson brackets \{q_i, p_j\} = \delta_{ij} and the symplectic form, making them essential for analyzing conserved quantities and stability. In practice, they facilitate action-angle variables for integrable systems like the or , where action integrals J_i = \oint p_i \, dq_i yield frequencies and quantization conditions in semiclassical approximations. Applications extend to diverse fields, including accelerator physics for beam dynamics, where canonical coordinates model particle trajectories under constraints, and nonlinear dynamics for studying chaos via portraits. In modern contexts, they underpin numerical integrators for long-term simulations in and , ensuring over extended times.

Introduction and Basics

Definition

In Hamiltonian mechanics, canonical coordinates are defined as conjugate pairs (q^i, p_i), where the q^i (for i = 1, \dots, n) represent generalized position coordinates and the p_i represent their corresponding conjugate momenta, describing the state of a mechanical system with n . These coordinates parameterize the of the system, a $2n-dimensional manifold that encompasses all possible configurations and momenta, providing a complete specification of the system's dynamical state at any instant. The role of canonical coordinates is central to the Hamiltonian formulation, where the time evolution of the system is governed by Hamilton's equations—a set of $2n first-order partial differential equations derived from the Hamiltonian function H(q^i, p_i, t)—offering a symmetric and elegant description of the dynamics compared to second-order formulations. These coordinates form a canonical basis that adheres to fundamental algebraic relations underpinning the structure and invariance of , with further details addressed in later sections. The selection of such coordinates aligns with the inherent of the .

Historical Development

The concept of canonical coordinates emerged from efforts to reformulate in a more analytical framework, building on Joseph-Louis Lagrange's introduction of in his 1788 work Mécanique Analytique. Lagrange's approach emphasized variational principles and the use of arbitrary coordinates to describe mechanical systems, providing a foundation for later developments in descriptions without directly formulating momentum-conjugate pairs. William Rowan Hamilton advanced this framework in the 1830s through his reformulation of dynamics, introducing canonical coordinates as pairs of position and momentum variables in his 1834 paper "On a General Method in Dynamics" and the 1835 "Second Essay on a General Method in Dynamics," both published in the Philosophical Transactions of the Royal Society. Hamilton's innovation stemmed from analogies between optics and mechanics, leading to the characteristic and principal functions that enabled the description of systems via partial differential equations, marking a shift toward a unified treatment of conservative mechanical systems. Carl Gustav Jacob Jacobi further evolved the theory in the 1840s, extending Hamilton's methods to time-dependent and non-conservative forces in works such as his 1837 paper in Crelle's Journal and later in Vorlesungen über Dynamik (1842–1843). Jacobi's contributions included refinements to techniques and the of Hamilton's partial equations, enhancing the applicability of formulations to complex problems like the . In the early 20th century, formalized the geometric underpinnings of canonical coordinates within , notably in his 1922 Leçons sur les invariants intégraux, where he applied forms to mechanical problems, including the symplectic form \sum p_i dq_i - H dt. Cartan's exterior and development of forms from 1899 onward provided an abstract manifold-based structure, influencing the field's transition to modern geometry. The canonical framework bridged to in the 1920s through and , who adapted Poisson brackets from classical into quantum commutation relations in their respective 1925 papers: Dirac's "The Fundamental Equations of " and Heisenberg's "Über quantentheoretische Umdeutung kinematischer und mechanischer Beziehungen." This correspondence enabled the quantization of canonical variables, laying groundwork for and transformation theories of .

Classical Mechanics

Relation to Lagrangian Mechanics

In Lagrangian mechanics, the motion of a system is formulated using q^i and their velocities \dot{q}^i, with the L(q, \dot{q}) typically expressed as the difference between kinetic and potential energies. This approach derives from the principle of least , where the S = \int L \, dt is . To bridge to the Hamiltonian formulation, canonical coordinates are introduced via the conjugate momenta p_i = \frac{\partial L}{\partial \dot{q}^i}, which represent the generalized momenta associated with each coordinate. This definition arises from a , which switches the independent variables from velocities \dot{q}^i to momenta p_i by considering L as a of q and \dot{q}, and constructing the convex dual. The resulting is given by H(q, p) = \sum_i p_i \dot{q}^i - L(q, \dot{q}), where the velocities \dot{q}^i are expressed as functions of q and p by inverting the momentum relations. In practice, this yields H as the total energy in terms of coordinates and momenta, facilitating analysis in phase space. For a simple example, consider a particle of mass m in Cartesian coordinates, where the Lagrangian is L = \frac{1}{2} m \dot{x}^2 - V(x). The conjugate momentum is then p_x = m \dot{x}, the linear momentum, and the Hamiltonian becomes H = \frac{p_x^2}{2m} + V(x). The transformation is well-defined when the mapping from \dot{q}^i to p_i is invertible, which holds if L is strictly convex in the velocities—commonly the case when the kinetic energy is quadratic in \dot{q}^i, as in L = T(q, \dot{q}) - V(q) with T = \frac{1}{2} \sum_{ij} a_{ij}(q) \dot{q}^i \dot{q}^j and positive-definite metric a_{ij}. Under these conditions, the inverse exists, ensuring H(q, p) is uniquely determined and single-valued.

Poisson Bracket Formulation

In canonical coordinates (q^i, p_i) for a system with n , the of two smooth functions f and g on the is defined by \{f, g\} = \sum_{i=1}^n \left( \frac{\partial f}{\partial q^i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q^i} \right). \tag{1} This bilinear operation, introduced by in his 1809 memoir on , encodes the symplectic structure of and generates infinitesimal canonical transformations. The coordinates (q^i, p_i) are canonical if they satisfy the fundamental Poisson bracket relations \{q^i, q^j\} = 0, \quad \{p_i, p_j\} = 0, \quad \{q^i, p_j\} = \delta^i_j, \tag{2} where \delta^i_j is the . These conditions ensure that the Poisson bracket preserves the standard symplectic form under transformations, distinguishing canonical coordinates from general ones. The governs the of any function f(q, p, t) via Hamilton's equation \frac{df}{dt} = \{f, H\} + \frac{\partial f}{\partial t}, \tag{3} where H(q, p, t) is the . For time-independent f, this reduces to \dot{f} = \{f, H\}, linking the bracket directly to the system's dynamics. As an illustrative example, consider a in one dimension with H = p^2 / 2m. The canonical conditions yield \{q, p\} = 1, so \dot{q} = \{q, H\} = p/m and \dot{p} = \{p, H\} = 0, implying constant velocity and uniform motion. The Poisson bracket exhibits key algebraic properties: bilinearity, meaning \{af + bg, h\} = a\{f, h\} + b\{g, h\} and similarly for the second argument; antisymmetry, \{f, g\} = -\{g, f\}; and the Jacobi identity, \{f, \{g, h\}\} + \{g, \{h, f\}\} + \{h, \{f, g\}\} = 0. These ensure the bracket defines a Lie algebra on the space of observables, facilitating the formulation of conserved quantities and symmetries.

Geometric Formulation

Cotangent Bundles

In classical mechanics, the configuration space of a system is modeled as a smooth manifold Q, while the phase space, which encodes both positions and momenta, is naturally identified with the cotangent bundle T^*Q. This bundle consists of all covectors over Q, with the projection map \pi: T^*Q \to Q sending each covector to its base point in the configuration space. The structure of T^*Q provides the geometric foundation for canonical coordinates, enabling the formulation of Hamiltonian dynamics in a coordinate-invariant manner. Local coordinates on T^*Q are induced by choosing coordinates (q^i) on Q, yielding canonical coordinates (q^i, p_i) on T^*Q, where i = 1, \dots, n and n = \dim Q. Here, the q^i represent position coordinates, while the p_i are the components of the covector p \in T^*_q Q with respect to the dual basis \{dq^i\}. These coordinates are "canonical" in the sense that they align positions and momenta in a natural pairing, facilitating the transition from Lagrangian to Hamiltonian descriptions. The cotangent bundle T^*Q is equipped with a canonical one-form \theta, known as the tautological or Liouville form, defined intrinsically by \theta_\alpha(\xi) = \alpha(\pi_* \xi) for \alpha \in T^*Q and \xi \in T_\alpha (T^*Q). In local canonical coordinates, this takes the expression \theta = \sum_i p_i \, dq^i. This one-form satisfies the pullback property: for any smooth one-form \alpha on an open set of Q, the pullback \alpha^* \theta = \alpha. The canonical symplectic two-form on T^*Q is then obtained as the exterior derivative \omega = -d\theta. Locally, this yields \omega = \sum_i dq^i \wedge dp_i, which is closed (d\omega = 0) and non-degenerate, endowing T^*Q with its natural structure. This form is independent of the choice of coordinates on Q and serves as the primitive for the geometry of . A concrete example arises when the configuration space is Q = \mathbb{R}^n, in which case T^*Q \cong \mathbb{R}^{2n} with the standard canonical coordinates (q^1, \dots, q^n, p_1, \dots, p_n). Here, the one-form is \theta = \sum_{i=1}^n p_i \, dq^i and the two-form is \omega = \sum_{i=1}^n dq^i \wedge dp_i, reproducing the familiar of n-dimensional Cartesian mechanics. This setting underlies many standard applications, such as the or dynamics.

Symplectic Structure

A is a pair (M, \omega), where M is a even-dimensional manifold and \omega is a closed, non-degenerate 2-form on M. The closedness condition, d\omega = 0, ensures that the form satisfies the requirements for a symplectic structure, while non-degeneracy means that for every point p \in M and nonzero v \in T_p M, there exists w \in T_p M such that \omega(v, w) \neq 0. This structure underpins the geometry of in , where canonical coordinates q^i, p_i naturally arise. The Darboux theorem guarantees the existence of local canonical coordinates in which the symplectic form takes its standard appearance. Specifically, around any point on a (M, \omega), there exist coordinates (q^1, \dots, q^n, p_1, \dots, p_n) such that \omega = \sum_{i=1}^n dq^i \wedge dp_i. This canonical form highlights how canonical coordinates adapt to the , eliminating local invariants and allowing the manifold to be locally modeled on the standard \mathbb{R}^{2n} with the form \sum dq^i \wedge dp_i. From the form \omega, the Poisson bracket of two smooth functions [f, g](/page/F&G): M \to \mathbb{R} is defined as \{f, g\} = \omega(X_f, X_g), where X_f is the associated to f, satisfying df = \iota_{X_f} \omega. The interior product \iota_{X_f} contracts \omega along X_f, yielding the 1-form df, which links the of f to the structure and ensures that X_f generates the flow preserving \omega. This formulation extends the classical \{q^i, p_j\} = \delta^i_j to arbitrary functions on the manifold. Liouville's theorem states that the Hamiltonian flow preserves the volume in phase space, meaning that the Liouville measure \frac{\omega^n}{n!} is invariant under the time evolution generated by any Hamiltonian H. This follows from the fact that the Hamiltonian vector field X_H is divergence-free with respect to this measure, as \mathcal{L}_{X_H} \left( \frac{\omega^n}{n!} \right) = 0, where \mathcal{L} denotes the Lie derivative. Consequently, incompressible flow in phase space reflects the symplectic preservation under dynamics. For an example, consider the T^* \mathbb{R}^n as , equipped with the canonical symplectic form \omega = \sum dq^i \wedge dp_i. In the 2D case (n=1), this reduces to \omega = dq \wedge dp on \mathbb{R}^2, ensuring that Hamiltonian flows are area-preserving maps, as the flow \phi_t satisfies \phi_t^* \omega = \omega and thus preserves the area form \omega.

Formal Development and Applications

Canonical Transformations

In Hamiltonian mechanics, a canonical transformation is a change of coordinates in phase space from (q_i, p_i) to (Q_i, P_i) that preserves the form of Hamilton's equations of motion. Such transformations maintain the canonical structure, ensuring that the new Hamiltonian K(Q, P, t) generates dynamics identical to the original H(q, p, t) up to a total time derivative. Equivalently, a map \phi: (q, p) \to (Q, P) is canonical if it preserves the Poisson brackets, satisfying \{Q_i, Q_j\}_{q,p} = 0, \{P_i, P_j\}_{q,p} = 0, and \{Q_i, P_j\}_{q,p} = \delta_{ij}, or if it preserves the symplectic form such that \phi^*\omega = \omega, where \omega = \sum_i dq_i \wedge dp_i. Canonical transformations are often generated by a scalar function F, known as a , which relates the old and new variables through partial derivatives. There are four standard types for one degree of freedom (generalizing to multiple): F_1(q, Q, t) with p_i = \frac{\partial F_1}{\partial q_i} and P_i = -\frac{\partial F_1}{\partial Q_i}; F_2(q, P, t) with p_i = \frac{\partial F_2}{\partial q_i} and Q_i = \frac{\partial F_2}{\partial P_i}; F_3(p, Q, t) with q_i = -\frac{\partial F_3}{\partial p_i} and P_i = -\frac{\partial F_3}{\partial Q_i}; and F_4(p, P, t) with q_i = -\frac{\partial F_4}{\partial p_i} and Q_i = \frac{\partial F_4}{\partial P_i}. The new Hamiltonian relates to the old by K(Q, P, t) = H(q, p, t) + \frac{\partial F}{\partial t}, ensuring the transformation preserves the structure. A key property is that the pdq - PdQ = dF holds for the appropriate type of F, directly linking to the invariance of the symplectic 2-form. Point transformations, a special case, depend only on the coordinates Q = Q(q), with momenta transforming as P_i = \sum_j p_j \frac{\partial q_j}{\partial Q_i} to maintain canonicity; these preserve the Poisson brackets if the Jacobian determinant is nonzero. Extended canonical transformations incorporate explicit time dependence, allowing Q = Q(q, p, t) and P = P(q, p, t), which is useful for time-dependent systems while still satisfying the symplectic condition M^T J M = J, where M is the matrix and J is the \begin{pmatrix} 0 & I \\ -I & 0 \end{pmatrix}. An illustrative example is the rotation in phase space for the one-dimensional harmonic oscillator with Hamiltonian H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 q^2. The transformation to action-angle variables q = \sqrt{\frac{2I}{\ m \omega}} \sin \phi, p = \sqrt{2 I m \omega} \cos \phi (generated by F_2(q, P) = -\frac{m \omega q^2}{2 \tan P}) rotates the elliptical orbits into circles, yielding the simplified Hamiltonian K = \omega I independent of \phi, and preserves the fundamental Poisson bracket \{ \phi, I \} = 1. The provides a systematic to find canonical transformations that simplify the , often reducing it to a constant or zero in the new coordinates. For time-independent systems, it is H\left(q, \frac{\partial W}{\partial q}\right) = E, where W(q) is the generating the transformation via P_i = \frac{\partial W}{\partial Q_i}; solving this yields coordinates where the new depends only on constants of motion, facilitating of the equations. For the , W = \frac{1}{2} m \omega q^2 \cot \alpha (with E = \frac{1}{2} m \omega A^2) generates action-angle variables directly.

Hamilton's Equations and Modern Uses

In canonical coordinates (q^i, p_i), the dynamics of a Hamiltonian system are governed by Hamilton's equations, which arise from the fundamental Poisson brackets \{q^i, H\} = \frac{\partial H}{\partial p_i} and \{p_i, H\} = -\frac{\partial H}{\partial q^i}, where H is the Hamiltonian function. These yield the first-order differential equations \dot{q}^i = \frac{\partial H}{\partial p_i}, \quad \dot{p}_i = -\frac{\partial H}{\partial q^i}. This formulation encapsulates the time evolution of the system while preserving the symplectic structure inherent to the phase space. A simple illustration is the one-dimensional , with H = \frac{p^2 + q^2}{2} (in units where mass and frequency are ). Substituting into Hamilton's equations gives \dot{q} = p and \dot{p} = -q, whose solutions describe periodic motion with conserved energy. Canonical coordinates provide the foundation for quantization in , where the classical Poisson brackets \{q^i, p_j\} = \delta^i_j are promoted to commutators [ \hat{q}^i, \hat{p}_j ] = i \hbar \delta^i_j for operator-valued observables in the . This correspondence ensures that quantum dynamics recover classical limits via Ehrenfest's theorem. The canonical formalism extends to quantum field theory (QFT), where Hamilton's equations apply to infinite-dimensional phase spaces of field configurations \phi(\mathbf{x}) and conjugate momenta \pi(\mathbf{x}), enabling the quantization of relativistic fields while respecting and unitarity. In numerical simulations of classical Hamiltonian systems, integrators—such as the explicit fourth-order schemes—discretize Hamilton's equations while preserving the form, yielding superior long-term compared to non-symplectic methods; this is particularly vital in for accurate N-body orbital predictions over extended timescales. In modern , the Arnowitt-Deser-Misner (ADM) formalism recasts Einstein's equations as a constrained using canonical coordinates on the space of spatial metrics and their momenta, facilitating simulations of mergers and . Similarly, in , the Kolmogorov-Arnold-Moser (KAM) theorem demonstrates that most invariant tori in integrable systems persist under small perturbations, provided the frequency vectors satisfy a Diophantine condition, explaining the onset of global chaos in perturbed canonical systems.

References

  1. [1]
    [PDF] Chapter 4 Canonical Transformations, Hamilton-Jacobi Equations ...
    The simplest case is for a single dimension of canonical coordinates (q, p). If H(q, p) = α1 then p = p(q, α1). There are two types of periodic motion to ...
  2. [2]
    [PDF] Hamiltonian Dynamics - CERN Indico
    Nov 8, 2022 · The coordinates (q, p) are canonical variables, and the space of canonical variables is known as phase space. • The phase space may exhibit ...
  3. [3]
    The History of Canonical Transformations in Hamilton-Jacobi Theory
    The concept of a canonical transformation was introduced by Jacobi in 1837 [9], but really only became a central part of the theory more than fifty years later.
  4. [4]
    15.3: Canonical Transformations in Hamiltonian Mechanics
    Mar 21, 2021 · Canonical transformations are the foundation of Hamiltonian mechanics; they underlie Hamilton-Jacobi theory and action-angle variable theory.Generating functions · Type 1: F = F 1 ⁡ ( q , Q , t ) · Type 2: F = F 2 ⁡ ( q , P , t ) − Q...
  5. [5]
    [PDF] Hamiltonian Mechanics
    The resulting 2N Hamiltonian equations of motion for qi and pi have an elegant symmetric form that is the reason for calling them canonical equations. Although ...<|control11|><|separator|>
  6. [6]
    [PDF] Hamiltonian Mechanics and Symplectic Geometry
    Physicist's discussions of Hamiltonian mechanics often assume that one can globally choose “canonical coordinates” on phase space and identify it with (R2n,ω0) ...
  7. [7]
    [PDF] The Early History of Hamilton-Jacobi Dynamics 1834–1837
    May 2, 2023 · As we observed in 8 2.3, Hamilton's invention of canonical coordinates and equations of motion was preceded by certain innovations of ...
  8. [8]
    [PDF] Early History of Symplectic Geometry
    The coordinates are canonical coordinates for the symplectic form. Dar- boux's theorem in symplectic geometry was first stated in 1953 by Paulette. Libermann ...
  9. [9]
    [PDF] The Birth of Quantum Mechanics: A Historical Study Through ... - arXiv
    This paper explores the historical development of the theory of quantum mechanics between 1900 and 1927 by chronological exami-.
  10. [10]
    [PDF] Chapter 15 - The Hamiltonian method
    The Hamiltonian method is a third way to solve mechanics problems, similar to the Lagrangian method, and is useful in statistical and quantum mechanics.
  11. [11]
    S 0273-0979(03)00984-4 - American Mathematical Society
    Apr 9, 2003 · around “Poisson brackets”, introduced by Siméon-Denis Poisson (1781-1840) in his work on the three-body problem in celestial mechanics. In ...
  12. [12]
    Poisson Brackets - Richard Fitzpatrick
    The classical Poisson bracket of two dynamical variables, and , is defined [55] (2.3) where and are regarded as functions of the coordinates and momenta, and , ...<|control11|><|separator|>
  13. [13]
    Poisson Brackets
    ∂f. ∂pi. ∂g. ∂qi −. ∂f. ∂qi. ∂g. ∂pi . The structure functions of the canonical Poisson bracket are. {qi,qj} = {pi,pj} = 0, {pi,qj} = δ j i . By Darboux's ...
  14. [14]
    [PDF] 1 The basic set-up 2 Poisson brackets
    The phase space of the system is parametrized by canonical coordinates q = (q1,...,qn) and p = (p1,...,pn). 2. We are given a Hamiltonian function H(q, p,t). 3.
  15. [15]
    [PDF] Poisson Brackets and Constants of the Motion - Galileo and Einstein
    Time Evolution in Phase Space: Poisson Brackets and Constants of the Motion ... H f = then (. ) ,i i. f p q is a constant of the motion, also called an ...
  16. [16]
    [PDF] Jacobi Identity for Poisson Brackets: A Concise Proof - arXiv
    Aug 20, 2004 · The bilinearity, antisymmetry property and cel- ebrated Jacobi identity, etc., are some of the key properties that are respected by PBs.
  17. [17]
    [PDF] Cotangent Bundles
    canonical one-form is given by Θ = pi dqi, and (6.1.3) holds. Proof. If (q1,... ,qn,p1,... ,pn) are coordinates ...Missing: Arnold | Show results with:Arnold
  18. [18]
    [PDF] A GUIDE TO SYMPLECTIC GEOMETRY - Williams College
    May 6, 2022 · In particular, we know that cotangent coordi- nates for cotangent bundles are Darboux coordinates, so pretty much everything we have ...
  19. [19]
    [PDF] Lectures on Mechanics
    ... canonical cotangent coordinates of T∗Q. 2.3.1 Proposition. There is a unique 1-form Θ on T∗Q such that in any choice of canonical cotangent coordinates,. Θ ...
  20. [20]
    Symplectic Form -- from Wolfram MathWorld
    A symplectic form on a smooth manifold is a smooth closed 2-form on which is nondegenerate such that at every point , the alternating bilinear form on the ...Missing: two- | Show results with:two-
  21. [21]
    symplectic manifold - PlanetMath
    Mar 22, 2013 · Definition 1.​​ A symplectic manifold is a pair (M,ω) consisting of a smooth manifold. M and a closed 2-form (http://planetmath.org/ ...Missing: two- | Show results with:two-
  22. [22]
    Darboux's theorem (symplectic geometry) - PlanetMath.org
    Mar 22, 2013 · Darboux's theorem implies that there are no local invariants in symplectic geometry, unlike in Riemannian geometry, where there is curvature.
  23. [23]
    [PDF] from hamiltonian systems to poisson geometry - UChicago Math
    Let (M,ω) be a symplectic manifold and take H ∈ C∞(M). Then we define the symplectic Hamiltonian vector field of H to be the vector-field XH such that ω ...<|separator|>
  24. [24]
    4 The Hamiltonian Formalism - DAMTP
    The central idea of Liouville's theorem – that volume of phase space is constant – is somewhat reminiscent of quantum mechanics. Indeed, this is the first of ...
  25. [25]
    [PDF] Symplectic Flows and Maps and Volume Preservation
    In particular, all Hamiltonian flow maps preserve volume in phase space. This is a sort of. (qualitative) invariant property, but not a first integral. 16.2.2 ...
  26. [26]
    [PDF] Symplectic Geometry and Integrable Systems (MATH 538-003)
    May 2, 2020 · For example, (x, y) 7→ (x + 1, y) is an area-preserving map of the right-half plane to itself. (with respect to the standard area form dx ∧ dy), ...
  27. [27]
    [PDF] PHY411 Lecture notes – Canonical Transformations
    Sep 27, 2023 · The coordinate transformation is canonical as long as we define the new coordinate and momenta using equations 5. Similar choices can be made ...<|control11|><|separator|>
  28. [28]
    Mathematical Methods of Classical Mechanics | SpringerLink
    In stockIn this text, the author constructs the mathematical apparatus of classical mechanics from the beginning, examining all the basic problems in dynamics.
  29. [29]
    The Principles of Quantum Mechanics - P. A. M. Dirac
    The standard work in the fundamental principles of quantum mechanics, indispensable both to the advanced student and to the mature research worker.