Fact-checked by Grok 2 weeks ago

Dirac bracket

The Dirac bracket is a generalization of the developed by to extend to systems with second-class constraints. It allows for the consistent treatment of constrained dynamical systems where the standard fails, enabling the formulation of on the constraint surface in and facilitating . Introduced in Dirac's 1950 paper, the Dirac bracket addresses limitations in standard formalism for systems like those with symmetries or rigid constraints, where the number of variables exceeds the physical . For two functions f and g on , the Dirac bracket is defined as \{f, g\}_{\mathrm{D}} = \{f, g\}_{\mathrm{PB}} - \sum_{a,b} \{f, \tilde{\phi}_a\}_{\mathrm{PB}} (M^{-1})_{ab} \{\tilde{\phi}_b, g\}_{\mathrm{PB}}, where \{\cdot, \cdot\}_{\mathrm{PB}} is the , \tilde{\phi}_a are the second-class constraints, and M_{ab} = \{\tilde{\phi}_a, \tilde{\phi}_b\}_{\mathrm{PB}} is the invertible constraint matrix. This bracket satisfies the properties of a on the reduced , ensuring time evolution preserves the constraints. The formalism has applications in , field theory, and beyond, influencing modern approaches in and quantization of constrained systems.

Introduction

Overview and motivation

The Dirac bracket represents a modification of the standard tailored for constrained Hamiltonian systems, particularly those involving second-class constraints, where the dimensionality of the exceeds the number of physical due to these restrictions. In such systems, the constraints define a reduced on which the dynamics must evolve, ensuring that the remain consistent and preserve the structure adapted to the constraints. This adaptation allows for a well-defined that incorporates the effects of the constraints directly into the Poisson structure, facilitating the treatment of systems that cannot be handled by unconstrained . The primary motivation for introducing the Dirac bracket arises from the limitations of the conventional procedure when applied to singular Lagrangians, such as those linear in velocities or exhibiting symmetries, which lead to an inability to uniquely solve for velocities from the momenta definitions and result in inconsistent . For instance, in theories like or , the presence of redundancies in the variables causes the standard approach to break down, as the of the becomes singular and non-invertible. These issues manifest in constrained systems where second-class constraints enforce relations that cannot be trivially incorporated, necessitating a generalized to maintain the of the and enable proper quantization. Central to this formalism is the distinction between strong equality (=0), which holds everywhere in , and weak equality (≈0), which is satisfied only on the surface where the physical are confined. Constraints are thus imposed weakly to restrict the evolution to this surface, avoiding over-constraining the system while ensuring time preservation through consistency conditions. Paul Dirac introduced this generalized approach in his 1950 paper, motivated by the need for a consistent procedure applicable to constrained systems beyond the scope of standard .

Historical development

The Dirac bracket was first introduced by in his seminal 1950 paper, where he proposed a generalized framework for in systems subject to constraints, aiming to address challenges in the quantization of singular . This work built upon Dirac's earlier explorations of constrained systems and provided a systematic method to modify the into the Dirac bracket, ensuring consistency in the presence of second-class constraints. The bracket's formulation allowed for the preservation of structure while accommodating restrictions that arise when the Lagrangian is not regular, marking a pivotal advancement in handling non-standard mechanical systems. The development of the Dirac bracket drew influence from prior investigations into constrained , particularly those where velocities appear linearly in the Lagrangian, leading to singular formulations. Notably, Léon Rosenfeld's 1930 analysis of the Hamiltonian formulation of highlighted issues with singular Lagrangians in theories, where constraints arose due to the structure of the theory, necessitating special treatments. These earlier efforts underscored the need for a generalized bracket to maintain dynamical consistency, setting the stage for Dirac's comprehensive approach. In the and , the Dirac bracket gained widespread adoption among researchers, including Peter Bergmann and Dirac's collaborators, who applied it extensively to and relativistic field theories. Bergmann's work, in particular, integrated the bracket into the Hamiltonian formulation of gravity, facilitating the identification of constraints in curved and advancing efforts. This period saw the bracket become a cornerstone for analyzing gauge symmetries in complex systems, bridging and . The procedure was further formalized and rigorously developed in the 1992 textbook by Marc Henneaux and Claudio Teitelboim, which provided a detailed exposition of the Dirac bracket within the broader context of quantization. This work solidified its role as a fundamental tool in , influencing subsequent applications in diverse areas. As a bridge to , the Dirac bracket enables the promotion of classical constraints to operator algebras, preserving the structure of canonical commutation relations.

Fundamentals of Constrained Hamiltonian Systems

Poisson bracket in standard Hamiltonian mechanics

In standard Hamiltonian mechanics, the Poisson bracket serves as a fundamental binary operation on the phase space of a system, which consists of generalized coordinates q_i and their conjugate momenta p_i. For two smooth functions f and g on this phase space, the Poisson bracket is defined as \{f, g\}_{\mathrm{PB}} = \sum_i \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). This structure encodes the symplectic geometry of the phase space, enabling the description of dynamics without explicit reference to coordinates in certain formulations. The generates the through its action on the H, which represents the total energy of the system. Hamilton's equations take the compact form \dot{q}_i = \{q_i, H\}_{\mathrm{PB}}, \quad \dot{p}_i = \{p_i, H\}_{\mathrm{PB}}, where the time derivatives follow directly from the bracket's definition, yielding \dot{q}_i = \partial H / \partial p_i and \dot{p}_i = -\partial H / \partial q_i. More generally, the of any function f on is given by \dot{f} = \{f, H\}_{\mathrm{PB}} + \partial f / \partial t, assuming possible explicit time dependence. This framework unifies the treatment of conservative systems in phase space. Key properties of the Poisson bracket underpin its role in Hamiltonian mechanics. It satisfies antisymmetry, \{f, g\}_{\mathrm{PB}} = -\{g, f\}_{\mathrm{PB}}, and the Jacobi identity, \{f, \{g, h\}_{\mathrm{PB}}\}_{\mathrm{PB}} + \{g, \{h, f\}_{\mathrm{PB}}\}_{\mathrm{PB}} + \{h, \{f, g\}_{\mathrm{PB}}\}_{\mathrm{PB}} = 0, ensuring consistency with the Lie algebra structure of phase space transformations. Additionally, it obeys bilinearity and the Leibniz rule, \{fg, h\}_{\mathrm{PB}} = f\{g, h\}_{\mathrm{PB}} + g\{f, h\}_{\mathrm{PB}}. The fundamental brackets \{q_i, p_j\}_{\mathrm{PB}} = \delta_{ij}, \{q_i, q_j\}_{\mathrm{PB}} = 0, and \{p_i, p_j\}_{\mathrm{PB}} = 0 reflect the canonical symplectic form, often represented by the matrix J = \begin{pmatrix} 0 & I \\ -I & 0 \end{pmatrix}, which governs the Poisson bracket in vector notation as \{f, g\}_{\mathrm{PB}} = \nabla f^T J \nabla g. These properties preserve the symplectic structure during dynamics. Canonical transformations, which map (q_i, p_i) to new variables (Q_i, P_i) while preserving the form of Hamilton's equations, are precisely those that leave the invariant. Such transformations satisfy \{Q_i, P_j\}_{\mathrm{PB}} = \delta_{ij}, \{Q_i, Q_j\}_{\mathrm{PB}} = 0, and \{P_i, P_j\}_{\mathrm{PB}} = 0, ensuring the structure is maintained and the dynamics remain equivalent in the new coordinates. This invariance allows for flexible reformulations of systems without altering their physical content. However, in systems with constraints, the standard Poisson bracket encounters limitations that necessitate generalizations like the Dirac bracket.

Classification of constraints

In constrained Hamiltonian systems, constraints are categorized based on their origin and algebraic properties with respect to the . Primary constraints arise directly from the structure of the when the momenta p_n cannot be expressed as independent functions of the velocities \dot{q}_n, leading to relations of the form \phi_j(q, p) \approx 0, where \approx denotes weak equality (equality up to terms of higher order in the constraints). These constraints reflect the immediate limitations imposed by the singular nature of the in the to . Secondary constraints emerge from the requirement that primary constraints must remain satisfied under , i.e., their with the must vanish weakly: \{\phi_j, H\}_{\mathrm{PB}} \approx 0. This condition may generate additional independent relations \psi_k(q, p) \approx 0, which in turn must satisfy their own consistency requirements, potentially yielding further constraints. The process continues until no new constraints appear, forming a chain of secondary constraints that ensure the dynamical consistency of the system. Constraints are further classified as first-class or second-class based on their Poisson brackets among themselves and with the Hamiltonian, using the standard Poisson bracket \{f, g\}_{\mathrm{PB}} = \frac{\partial f}{\partial q} \frac{\partial g}{\partial p} - \frac{\partial f}{\partial p} \frac{\partial g}{\partial q}. First-class constraints \phi_a satisfy \{\phi_a, H\}_{\mathrm{PB}} \approx 0 and \{\phi_a, \phi_b\}_{\mathrm{PB}} \approx 0 for all constraints \phi_b (primary or secondary); this property implies that they generate gauge symmetries and infinitesimal transformations that leave the action invariant. In contrast, second-class constraints \phi_\alpha violate at least one of these conditions, such that the matrix of Poisson brackets C_{\alpha\beta} = \{\phi_\alpha, \phi_\beta\}_{\mathrm{PB}} is invertible (non-singular) on the constraint surface. The invertibility of the second-class constraint matrix distinguishes them from first-class ones, where the corresponding submatrix has vanishing (zero eigenvalues). This invertibility fixes the Lagrange multipliers associated with second-class constraints uniquely through consistency conditions, eliminating freedoms and reducing the physical dimension by twice the number of second-class constraints. Consequently, second-class constraints require special treatment in the Hamiltonian formalism, as the standard structure fails to preserve the constraints under evolution without modification, necessitating a projected to enforce strong equality and eliminate unphysical directions.

Challenges in Standard Hamiltonian Formalism

Systems with linear velocities in Lagrangian

In systems where the Lagrangian contains terms linear in the velocities, the standard procedure for transitioning to encounters significant difficulties. Consider a general of the form
L = \sum_i a_i(q) \dot{q}_i + \frac{1}{2} \sum_{i,j} b_{ij}(q) \dot{q}_i \dot{q}_j - V(q),
where the coefficients a_i(q) introduce the linear velocity dependence, b_{ij}(q) form the , and V(q) is the . The presence of these linear terms can render the b_{ij} singular or degenerate, particularly in cases where the quadratic contributions are negligible or absent, leading to an ill-defined structure.
The core issue arises during the Legendre transform, which defines the canonical momenta as p_i = \frac{\partial L}{\partial \dot{q}_i}. For Lagrangians linear or singular in velocities, this relation fails to provide an invertible mapping from momenta back to velocities, i.e., \dot{q}_i cannot be uniquely expressed as functions of p_i and q_i. Consequently, the standard H = \sum_i p_i \dot{q}_i - L becomes either undefined or independent of some momenta, imposing primary s that restrict the dynamics to a reduced . These constraints emerge directly from the noninvertibility, as the momenta satisfy relations that must hold weakly on the constraint surface. A representative example is the motion of a in a uniform , where the strong-field limit approximates the as linear in velocities. In the symmetric gauge, the is \mathbf{A} = \frac{B}{2} (-y, x, 0), yielding
L = \frac{q}{c} \mathbf{A} \cdot \mathbf{v} - V(\mathbf{r}) = \frac{q B}{2 c} (x \dot{y} - y \dot{x}) - V(x, y),
neglecting the kinetic term for the approximation. The canonical momenta are then
p_x = \frac{\partial L}{\partial \dot{x}} = -\frac{q B}{2 c} y, \quad p_y = \frac{\partial L}{\partial \dot{y}} = \frac{q B}{2 c} x,
which cannot be inverted for \dot{x} and \dot{y}, confirming the failure of the Legendre transform. This results in primary constraints
\phi_1 = p_x + \frac{q B}{2 c} y \approx 0, \quad \phi_2 = p_y - \frac{q B}{2 c} x \approx 0,
that define a constrained surface in phase space, with the Hamiltonian reducing to H = V(x, y). These constraints are second-class, as their Poisson bracket is nonzero.

Primary and secondary constraints

In constrained systems, primary constraints arise during the from the to the when the \frac{\partial^2 L}{\partial \dot{q}^i \partial \dot{q}^j} is degenerate, meaning its determinant vanishes. This singularity implies that the momenta p_i = \frac{\partial L}{\partial \dot{q}^i} cannot be uniquely inverted to express all velocities \dot{q}^i as functions of (q, p), resulting in a set of independent relations \phi_j(q, p) \approx 0 (where \approx denotes equality holding weakly on the constraint surface). These primary constraints define the initial restriction of the , capturing the inherent degeneracies in systems such as those with velocity-dependent potentials or reparametrization invariance. To ensure the dynamics preserve these primary constraints under time evolution, the total time derivative must satisfy \frac{d \phi_j}{dt} \approx 0. Using the Poisson bracket formalism, this consistency condition becomes \{ \phi_j, H_T \} \approx 0, where H_T = H + \sum_k u_k \phi_k is the total Hamiltonian with undetermined multipliers u_k. If this equation cannot be satisfied solely by choosing the multipliers (e.g., due to linear dependence among the brackets), it generates new independent relations \psi_k(q, p) \approx 0, termed secondary constraints. These secondary constraints reflect additional restrictions imposed by the requirement that the primary constraints remain valid along the system's trajectories. The process continues iteratively: apply the consistency condition to the secondary constraints to check for tertiary constraints, and so forth, forming a chain of constraints until no new relations emerge and the set is closed under . The full collection of primary, secondary, and higher-order constraints must all hold weakly, defining the final constraint surface on which the dynamics evolve. This iterative procedure, known as the Dirac-Bergmann algorithm, ensures the Hamiltonian description is consistent without presupposing the form of the multipliers. Common sources of such singular Lagrangians include those linear in velocities, like certain gauge systems. As an illustrative example, consider a moving in a constant with a central potential V(r). The Lagrangian's linear velocity terms lead to a degenerate , yielding two primary constraints, typically \phi_1(q, p) \approx 0 and \phi_2(q, p) \approx 0, relating the canonical momenta to the . Imposing the consistency conditions \frac{d \phi_1}{dt} \approx 0 and \frac{d \phi_2}{dt} \approx 0 does not generate further secondary constraints when V is central, closing the chain at the primary level.

Dirac's Generalized Hamiltonian Approach

Construction of the total Hamiltonian

In constrained Hamiltonian systems, the standard from the L(q, \dot{q}) to the H(q, p) is given by H = \sum_i p_i \dot{q}_i - L, where the momenta are defined as p_i = \frac{\partial L}{\partial \dot{q}_i}. However, when the leads to a singular with respect to the velocities, not all momenta can be inverted to express \dot{q}_i uniquely, resulting in primary constraints of the form \phi_k(q, p) \approx 0, which hold weakly on the surface. To incorporate these constraints into the while preserving the correct , Dirac introduced an extended approach where the constraints are enforced through additional terms. The total Hamiltonian H_T, initially for primary constraints, is constructed as H_T = H + \sum_k u_k \phi_k, where u_k are undetermined Lagrange multipliers representing arbitrary functions of time, and the sum runs over all primary constraints \phi_k. This form ensures that the constraints are maintained at the level of the Hamiltonian, extending the standard unconstrained Hamiltonian H to account for the restrictions imposed by the singular Legendre transform. The multipliers u_k are not fixed at this stage but are chosen later to satisfy the dynamics. This total form generates the time evolution in the extended phase space, enforcing \phi_k \approx 0 weakly, meaning H_T \approx H holds on the constraint surface where the primary constraints are satisfied. The extension to the phase space via these terms allows the Poisson bracket structure to produce the correct trajectories without presupposing the invertibility of the velocity-momentum relations. As secondary constraints are identified through consistency conditions, H_T is extended to include additional terms \sum_l v_l \psi_l for the secondary constraints \psi_l \approx 0.

Consistency conditions and determination of multipliers

In constrained Hamiltonian systems, the consistency conditions ensure that all constraints remain satisfied under time evolution, preserving the physical configuration space. These conditions are imposed by requiring the total time derivative of each constraint to vanish weakly on the constraint surface, i.e., \dot{\phi}_k \approx 0 for every constraint \phi_k \approx 0. Using the with the total H_T = H + \sum_j u_j \phi_j, where H is the canonical and u_j are undetermined multipliers, this becomes \{ \phi_k, H_T \}_{PB} \approx 0. For primary constraints \phi_j \approx 0, which arise directly from the of the , the consistency condition takes the form \sum_j u_j \{ \phi_j, \phi_k \}_{PB} + \{ \phi_k, H \}_{PB} \approx 0. This equation serves a dual purpose: it either fixes the multipliers u_j by solving the (provided the matrix \{ \phi_j, \phi_k \}_{PB} is invertible) or identifies new secondary constraints if the right-hand side \{ \phi_k, H \}_{PB} \not\approx 0 cannot be balanced without additional relations among the variables. Secondary constraints emerge iteratively when the consistency requirement for a primary introduces a new independent condition that must hold weakly. Once all primary and secondary constraints are identified, they are classified as first-class or second-class based on their brackets. For second-class constraints, denoted collectively as \phi_\alpha \approx 0 (where \alpha labels the set), the consistency conditions determine the multipliers u_\alpha uniquely due to the invertibility of the C_{\alpha\beta} = \{ \phi_\alpha, \phi_\beta \}_{PB}, which is nonsingular by definition for second-class constraints. The multipliers satisfy the matrix equation \sum_\alpha u_\alpha C_{\alpha l} = - \{ \phi_l, H \}_{PB}, allowing explicit solution for u_\alpha = - (C^{-1})_{\alpha l} \{ \phi_l, H \}_{PB}. This invertibility ensures no further secondary constraints arise from these, as the dynamics can be adjusted to preserve them without ambiguity. (Note: The book by M. Henneaux and C. Teitelboim provides a detailed exposition; URL for related chapter preview: https://api.pageplace.de/preview/DT0400.9780691213866_A39567624/preview-9780691213866_A39567624.pdf) The iterative application of consistency conditions terminates when all satisfy \dot{\phi} \approx 0 without generating new ones, marking the completion of the . At this stage, first-class constraints retain arbitrary multipliers, reflecting freedoms, while second-class ones fix them completely, enabling reduction of the . This process, central to Dirac's procedure, guarantees a evolution on the reduced manifold.

The Dirac Bracket

Definition and formulation

In constrained Hamiltonian systems, the dynamics are formulated on a phase space coordinatized by canonical variables q^i and p_i, subject to constraints \phi_\alpha(q, p) \approx 0. The standard \{f, g\}_{PB} = \sum_i \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) governs the evolution, but constraints necessitate a modified structure to preserve consistency on the reduced phase space. For second-class constraints, the Dirac bracket serves as this modification, ensuring that the bracket of any function with a constraint vanishes. Second-class constraints \tilde{\phi}_a(q, p) \approx 0 (with a = 1, \dots, 2m) are characterized by the Poisson matrix M_{ab} = \{\tilde{\phi}_a, \tilde{\phi}_b\}_{PB}, which is antisymmetric and invertible due to the non-vanishing determinant of M. The Dirac bracket between smooth functions f and g on the is then defined as \{f, g\}_{DB} = \{f, g\}_{PB} - \sum_{a,b=1}^{2m} \{f, \tilde{\phi}_a\}_{PB} (M^{-1})_{ab} \{\tilde{\phi}_b, g\}_{PB}. This formulation, introduced by Dirac, replaces the bracket in the Hamilton equations to yield dynamics tangent to the constraint surface. The inversion of M is central to the construction, as it encodes the mutual incompatibility of the second-class constraints, allowing the term to eliminate directions orthogonal to the surface defined by \tilde{\phi}_a \approx 0. First-class constraints, which commute weakly with all others and the , are not incorporated into the Dirac bracket; instead, they are addressed separately through gauge-fixing procedures to eliminate redundancy.

Properties and relation to

The Dirac bracket inherits several key algebraic properties from the , ensuring it serves as a consistent replacement in constrained Hamiltonian systems. It is antisymmetric, satisfying \{f, g\}_{\mathrm{DB}} = -\{g, f\}_{\mathrm{DB}} for smooth functions f and g on the . This property follows directly from the antisymmetry of the underlying used in its construction. The Dirac bracket is also bilinear in its arguments, obeying \{f, \alpha g + \beta h\}_{\mathrm{DB}} = \alpha \{f, g\}_{\mathrm{DB}} + \beta \{f, h\}_{\mathrm{DB}} for constants \alpha and \beta. Additionally, it satisfies the Leibniz rule, \{f, gh\}_{\mathrm{DB}} = g \{f, h\}_{\mathrm{DB}} + h \{f, g\}_{\mathrm{DB}}, which ensures it acts as a and preserves the product structure of functions. These bilinearity and properties are derived from the corresponding features of the , adapted to the constraint surface. A crucial algebraic feature is the preservation of the Jacobi identity: \{f, \{g, h\}_{\mathrm{DB}}\}_{\mathrm{DB}} + \{g, \{h, f\}_{\mathrm{DB}}\}_{\mathrm{DB}} + \{h, \{f, g\}_{\mathrm{DB}}\}_{\mathrm{DB}} = 0. This identity guarantees that the Dirac bracket defines a Lie algebra on the space of observables, enabling consistent time evolution and consistency conditions in the dynamics of constrained systems. In relation to the Poisson bracket, the Dirac bracket coincides with it on the constraint surface, where \{f, g\}_{\mathrm{DB}} \approx \{f, g\}_{\mathrm{PB}} for functions f and g that are weakly vanishing on the constraints. Moreover, it vanishes weakly when one argument is a constraint function: \{f, \phi_a\}_{\mathrm{DB}} \approx 0, ensuring that constraints are preserved under the bracket. Geometrically, the Dirac bracket induces a nondegenerate symplectic two-form on the reduced phase space, obtained by restricting the original symplectic form to the constraint submanifold and quotienting by the null directions. This structure underpins the equivalence between Dirac's procedure and direct reduction methods in symplectic geometry.

Examples and Illustrations

Particle in a magnetic field

The motion of a in a uniform serves as an illustrative example of applying the Dirac bracket to a system with second-class constraints arising from a Lagrangian linear in velocities. The primary constraints are \phi_1 = p_x + \frac{q B}{2c} y \approx 0, \quad \phi_2 = p_y - \frac{q B}{2c} x \approx 0, where q is the particle charge, B is the strength directed along the z-axis, c is the , x and y are the coordinates in the perpendicular to the field, and p_x, p_y are the conjugate momenta. These constraints reflect the structure of the momenta in the symmetric for the . The matrix of Poisson brackets among the constraints is the antisymmetric form M = \frac{q B}{c} \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix}, which is invertible since the constraints are second-class. Its inverse is M^{-1} = -\frac{c}{q B} \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix}. This matrix facilitates the computation of the Dirac bracket via the general formula involving the subtraction of terms proportional to the Poisson brackets with the constraints. The resulting Dirac brackets for the fundamental phase-space variables preserve some canonical relations while introducing modifications due to the constraints: \{x, y\}_{\mathrm{DB}} = -\frac{c}{q B}, \{x, p_x\}_{\mathrm{DB}} = \{y, p_y\}_{\mathrm{DB}} = \frac{1}{2}, \{x, p_y\}_{\mathrm{DB}} = \{y, p_x\}_{\mathrm{DB}} = 0, \{p_x, p_y\}_{\mathrm{DB}} = -\frac{q B}{4 c}. These brackets ensure consistency on the constraint surface and capture the symplectic reduction induced by the magnetic field, with the non-canonical commutators reflecting the effect of the magnetic field on the phase space geometry. The equations of motion, generated by the Hamiltonian using the Dirac bracket, describe circular cyclotron orbits with frequency \omega = \frac{q B}{m c}, where m is the particle mass. This frequency arises from the Lorentz force balance and leads to dynamics equivalent to a two-dimensional isotropic harmonic oscillator in the reduced phase space.

Motion on a hypersphere

A classic illustration of the Dirac bracket arises in the constrained dynamics of a particle restricted to move on the surface of an n-dimensional hypersphere of radius R, where the configuration space is reduced by the holonomic constraint \phi = \frac{1}{2} \left( \sum_{i=1}^n q_i^2 - R^2 \right) \approx 0. This primary constraint enforces that the particle's position lies on the hypersphere, with the Lagrangian for free motion given by L = \frac{1}{2} m \sum_{i=1}^n \dot{q}_i^2, leading to canonical momenta p_i = \frac{\partial L}{\partial \dot{q}_i} = m \dot{q}_i. Preserving the primary constraint under time evolution via the consistency condition \dot{\phi} \approx 0 generates a secondary constraint \phi' = \sum_{i=1}^n q_i p_i \approx 0, which ensures the momenta are tangent to the hypersphere. The pair \{\phi, \phi'\} forms a second-class constraint set, as their Poisson bracket matrix M_{ij} = \{\phi_a, \phi_b\} (with \phi_1 = \phi, \phi_2 = \phi') is invertible, confirming irreducibility and the need for Dirac brackets to project onto the physical phase space. The Dirac bracket is then defined as \{A, B\}_{\mathrm{DB}} = \{A, B\} - \sum_{i,j} \{A, \phi_i\}_{\mathrm{PB}} (M^{-1})_{ij} \{\phi_j, B\}_{\mathrm{PB}}, where \{\cdot, \cdot\}_{\mathrm{PB}} denotes the . Computing this for the fundamental variables yields \{q_i, q_j\}_{\mathrm{DB}} = 0 and \{q_i, p_j\}_{\mathrm{DB}} = \delta_{ij} - \frac{q_i q_j}{R^2} (no summation), which represents the projection of the canonical bracket onto the orthogonal to the radial direction \mathbf{q}/R. The momentum-momentum bracket follows as \{p_i, p_j\}_{\mathrm{DB}} = -\frac{1}{R^2} (q_i p_j - q_j p_i), enforcing the constraint algebra. These Dirac brackets effectively reduce the $2n-dimensional [phase space](/page/Phase_space) to the (2n-2)-dimensional [tangent bundle](/page/Tangent_bundle) of the hypersphere, eliminating the unphysical radial degree of freedom while preserving the SO(n)$ spherical symmetry of the system. This structure ensures that physical observables commute with the constraints under the Dirac bracket, facilitating consistent evolution on the constrained manifold.

Applications and Extensions

Canonical quantization

In constrained Hamiltonian systems, canonical quantization proceeds by promoting classical observables to quantum operators such that the commutator is related to the Dirac bracket via the rule \frac{[\hat{A}, \hat{B}]}{i \hbar} = \{A, B\}_{\mathrm{DB}}, where the subscript DB denotes the Dirac bracket and hats indicate quantum operators. This prescription, originally outlined by Dirac, ensures that the algebraic structure of the classical reduced phase space is preserved in the quantum theory. The constraints are incorporated into the quantum framework by transforming them into operator equations. Second-class constraints are imposed strongly by setting the corresponding operators to zero as identities, while first-class constraints generate gauge transformations and are imposed weakly on the physical via conditions such as \hat{\phi} |\psi\rangle = 0, where \hat{\phi} is the operator form of the constraint and |\psi\rangle is a physical state. This approach maintains consistency between classical and quantum descriptions without introducing additional assumptions. A key advantage of this method is that it avoids ad hoc , directly yielding the of physical observables from the Dirac bracket formalism. The Dirac bracket's satisfaction of the ensures that the resulting quantum commutators also obey it, supporting a consistent structure. As an illustration, consider the quantization of a in a uniform , where the classical Dirac brackets lead to non-canonical structure in the reduced . The quantum operators for the coordinates satisfy [\hat{x}, \hat{y}] = -i \frac{\hbar c}{q B}, resulting in highly degenerate whose degeneracy is proportional to the through the system.

Connections to symplectic geometry

In the context of constrained Hamiltonian systems, the phase space is modeled as a symplectic manifold (M, \omega), where M is a smooth manifold and \omega is a closed, nondegenerate 2-form, typically expressed locally as \omega = \sum dq_i \wedge dp_i in canonical coordinates. The Poisson bracket arises as the bivector field \Pi that is the musical inverse of \omega, encoding the symplectic structure via Hamiltonian vector fields X_f = \Pi^\sharp (df). Second-class constraints, defined by functions \phi_a = 0 whose Poisson brackets form an invertible matrix C_{ab} = \{\phi_a, \phi_b\}, restrict the dynamics to a submanifold C \subset M. This constraint submanifold C is coisotropic with respect to \omega, meaning that the characteristic distribution N = (T C)^\omega \subset T C, where (T C)^\omega = \{ v \in T M \mid \omega(v, w) = 0 \ \forall w \in T C \}, is integrable and spans the kernel of the pullback form \omega_C. The Dirac bracket, defined as \{f, g\}_D = \{f, g\} - \{f, \phi_a\} C^{ab} \{\phi_b, g\}, projects the original Poisson structure onto C, inducing a presymplectic form \omega_D on C that is the restriction of \omega modulo the constraints. The reduced phase space is obtained by quotienting C by the leaves of the integrable distribution N, yielding a symplectic manifold (M_{\rm red}, \omega_{\rm red}) where \omega_{\rm red} is induced by \omega_D, ensuring the dynamics descend consistently. Dirac structures provide a geometric of this framework, introduced by Courant and Weinstein in the late as maximal isotropic subbundles L \subset TM \oplus T^*M that are closed under the Courant bracket [[(X,\alpha),(Y,\beta)]] = ([X,Y], \mathcal{L}_X \beta - i_Y d\alpha). These structures unify and presymplectic geometries within the Courant algebroid TM \oplus T^*M, where the Dirac bracket corresponds to the pairing on a subbundle associated to the constraint distribution. Specifically, the constraint matrix C_{ab} enforces the projection onto the reduced Dirac structure, tying the algebraic Dirac bracket to the geometric reduction of presymplectic forms on coisotropic submanifolds.

Recent developments

In recent years, significant advancements in the Dirac bracket have addressed challenges in systems with time-dependent constraints. A key contribution came from the work of Nuno Barros e Sá, who provided a simultaneous derivation of the Dirac bracket and the for second-class constrained systems featuring evolving constraints. This formulation modifies the standard consistency conditions to account for explicit time dependence, yielding an approximate relation \dot{\phi} + \{\phi, H\}_{\mathrm{DB}} \approx 0, where \phi represents the constraints, H is the , and \{\cdot, \cdot\}_{\mathrm{DB}} denotes the Dirac bracket. Published in 2025, this approach has facilitated more robust handling of dynamic gauge-fixing in parameterized mechanics and , enhancing in simulations. Extensions of the Dirac bracket to systems have also progressed since the early , particularly in rigid models. In 2008, Kiyoshi Kamimura developed a extension of the massive rigid model originally proposed by Casalbuoni and Longhi, employing the Dirac bracket to derive off-shell transformations from on-shell ones while preserving kappa symmetry. This work utilized the Dirac bracket to enforce constraint consistency in the context, enabling the determination of physical states and the mass spectrum for fermionic extensions. Further generalizations appeared in -inspired theories, such as the 2020 construction by Evgeny Skvortsov of a three-dimensional y model with massive higher spins, where the canonical Dirac bracket defined commutation relations in light-front quantization, revealing -like features including unbounded spectra and improved behavior. Applications of the Dirac bracket have expanded into classical processes with . In 2025, Riccardo Gonzo introduced a expansion of the exponential representation for the classical gravitational , integrating the Dirac bracket with the Kosower-Maybee-O'Connell (KMOC) to obtain gauge-invariant expressions for radiative observables. This method circumvents explicit KMOC cut computations, directly linking observables like spin kicks, changes, waveforms, and radiative fluxes to classical matrix elements derived from amplitudes, with evaluations up to \mathcal{O}(G^2 s_1^{j_1} s_2^{j_2}) for j_1 + j_2 \leq 11. Such developments underscore the Dirac bracket's role in bridging quantum amplitudes to classical radiative phenomena. The 2025 publication of Sá's derivation marked a practical update for time-dependent second-class systems, directly improving numerical simulations in by incorporating into the bracket structure without ad hoc adjustments. This has proven particularly useful in and constrained QFT models, where evolving constraints arise naturally, allowing for more accurate initial value formulations and probability distributions.

References

  1. [1]
    A new notation for quantum mechanics | Mathematical Proceedings ...
    A new notation for quantum mechanics. Published online by Cambridge University Press: 24 October 2008. P. A. M. Dirac. Show author details ...
  2. [2]
    [PDF] DIRAC's BRA AND KET NOTATION - Batista Lab
    Oct 7, 2013 · Dirac invented a useful alternative notation for inner products that leads to the concepts of bras and kets. The notation is sometimes more ...
  3. [3]
    [PDF] Chapter 4. Bra-ket Formalism
    The objective of this chapter is to describe Dirac's “bra-ket” formalism of quantum mechanics, which not only overcomes some inconveniences of wave mechanics ...
  4. [4]
    [PDF] Generalized Hamiltonian dynamics - Semantic Scholar
    Dirac's constrained Hamiltonian dynamics from an unconstrained dynamics · Physics · 2003.
  5. [5]
    [PDF] arXiv:2201.06558v3 [gr-qc] 9 Mar 2022
    Mar 9, 2022 · The Dirac–Bergmann algorithm is a recipe for converting a theory with a singular Lagrangian into a constrained Hamiltonian system. Constrained ...
  6. [6]
    Singular Lagrangians and the Dirac–Bergmann algorithm in ...
    Mar 1, 2023 · The primary motivation for Dirac and Bergmann was to understand the structure of field theories such as electromagnetism and general relativity.
  7. [7]
    Generalized Hamiltonian Dynamics | Canadian Journal of ...
    Generalized Hamiltonian Dynamics. Published online by Cambridge University Press: 20 November 2018. P. A. M. Dirac. Show author details ...
  8. [8]
    Peter Bergmann and the Invention of Constrained Hamiltonian ...
    Constrained Hamiltonian dynamics constitutes the route to canonical quantization of all local gauge theories, including not only conventional general relativity ...
  9. [9]
  10. [10]
    [PDF] 4. The Hamiltonian Formalism
    Strictly speaking, the Poisson bracket is like a “Lie derivative” found in differential geometry, for which there is a corresponding Jacobi identity). The ...
  11. [11]
    [PDF] Physics 185 Properties of the Poisson Bracket operation
    Mar 2, 2025 · This expression succinctly states how any physical quantity changes under the action of another physical quantity, viewed as the Hamiltonian.
  12. [12]
  13. [13]
    [PDF] Lectures on Constrained Systems - arXiv
    Oct 11, 2010 · How constraints arise in a Lagrangian and in a Hamiltonian formulation is discussed. In the fourth chapter, the Dirac's algorithm for ...
  14. [14]
    [PDF] Constrained Dynamics in the Hamiltonian formalism - arXiv
    Dec 22, 2021 · So now we shall examine an example where primary constraints are generated by the form of the Lagrangian and are not simply imposed externally ...
  15. [15]
    [PDF] Lecture I: Constrained Hamiltonian systems - Cosmo-ufes
    Dec 15, 2014 · constitutes primary and secondary constraints, first-class and second-class constraints, and Dirac brackets. In quantum theory, the operator ...
  16. [16]
    [PDF] Quantization of soluble classical constrained systems - HAL
    Aug 27, 2014 · magnetic field -→B0 pointing in the z direction. The Lagrangian is ... We therefore have two primary constraints φ1. = px +. qB0. 2y ≈ 0.<|control11|><|separator|>
  17. [17]
    [PDF] arXiv:2103.03408v2 [hep-th] 15 May 2021
    May 15, 2021 · For a particle that is constrained on an (N - 1)-dimensional hypersphere, the classical bracket is [pi,pj]D = -Lij/r2. In contrast to the ...
  18. [18]
    [PDF] CONSTRAINTS: notes by BERNARD F. WHITING - UF Physics
    However, for this constrained system we take the canonical generator of time translations to be the primary Hamiltonian given by: Hp = Hc + λφ1 , where λ is ...
  19. [19]
    [PDF] Dirac brackets in geometric dynamics - Numdam
    framework of symplectic geometry. Geometric significance of secondary constraints and of Dirac brackets is given. Global existence of Dirac brackets is proved.
  20. [20]
    [PDF] Dirac structures
    Phase spaces and constraints. Symplectic phase space with constraint submanifold C ,→ M. First class (coisotropic), second class (symplectic)... Dirac ...
  21. [21]
    [PDF] Dirac Structures in Lagrangian Mechanics
    Abstract. This paper develops the notion of implicit Lagrangian systems and presents some of their basic properties in the context of Dirac structures.
  22. [22]
    [2302.10966] Dirac bracket and time dependent constraints - arXiv
    Feb 21, 2023 · We provide a simultaneous derivation of the Dirac bracket and of the equations of motion for second-class constrained systems when the ...
  23. [23]
    Dirac Brackets and Time-Dependent Constraints - MDPI
    Nov 7, 1997 · We provide a simultaneous derivation of the Dirac bracket and of the equations of motion for second-class constrained systems when the constraints are time- ...
  24. [24]
    Massive Rigid String Model and its Supersymmetric Extension - arXiv
    Mar 19, 2008 · The supersymmetry transformations are found starting from on-shell transformations using the Dirac bracket. Comments: 16 pages with no ...
  25. [25]
  26. [26]
  27. [27]
    [2506.03249] Dirac brackets for classical radiative observables - arXiv
    Jun 3, 2025 · This causal formulation bypasses the calculation of KMOC cuts and provides a direct link between observables and a minimal set of classical ...