Fact-checked by Grok 2 weeks ago

Matrix mechanics

Matrix mechanics is a foundational formulation of quantum mechanics developed by Werner Heisenberg in 1925 and formalized by Max Born and Pascual Jordan later that year. It represents physical observables, such as position and momentum, as infinite-dimensional Hermitian matrices, with quantum states described by column vectors, and employs non-commutative multiplication to capture the discrete nature of atomic spectra and transitions. This approach marked a departure from classical physics by focusing exclusively on observable quantities, like transition probabilities, rather than unobservable trajectories. Heisenberg's initial insight arose from attempts to explain the spectral lines of and other atoms using only measurable frequencies and intensities, leading to his seminal paper in Zeitschrift für Physik in July 1925. Born and quickly recognized the matrix algebra implicit in Heisenberg's equations, publishing "Zur Quantenmechanik" in November 1925, which introduced the term "" and established the commutation relation [q, p] = i\hbar as a core postulate. A follow-up paper by , Heisenberg, and in 1926 extended the theory to systems with multiple , proving its consistency with and Bohr's . Key mathematical features include the use of matrices for dynamical variables, where the elements q_{mn} represent transition amplitudes between stationary states m and n, and time evolution governed by equations analogous to Hamilton's but with matrix products. This was successfully applied by to derive the spectrum and selection rules in , validating its predictive power. In , Erwin Schrödinger's wave mechanics emerged as an alternative, and Schrödinger himself demonstrated their mathematical equivalence that same year through a proof showing that matrix elements correspond to Fourier coefficients of wave functions. John von Neumann later formalized this equivalence within the Hilbert space framework in 1927-1928, unifying the field under formalism. Matrix mechanics laid the groundwork for modern , influencing the development of the in 1927 and subsequent extensions like . Its operator-based approach remains particularly effective for problems involving and the , though it is less intuitive for continuous systems compared to wave mechanics. The original papers, now recognized as pivotal in the quantum revolution, continue to be studied for their innovative blend of physics and mathematics.

Historical Development

Heisenberg's Epiphany on Heligoland

In June 1925, , then 23 years old, retreated to the remote island of to alleviate severe hay fever and recover from exhaustion caused by his intensive studies in . This isolation provided him the solitude needed to grapple with the limitations of the , which he had been exploring through his recent collaboration with on the correspondence principle—a framework linking quantum transitions to classical radiation in the high-energy limit. had grown increasingly dissatisfied with visualizable models like orbits in , viewing them as misleading because they invoked unobservable paths rather than directly addressing experimental data. During his stay, Heisenberg's breakthrough emerged from a deliberate shift away from classical mechanical descriptions toward quantities that could be directly observed, such as the frequencies and intensities of lines produced by transitions. He proposed representing these observables through arrays of transition amplitudes between states, inspired by the of classical motion but stripped of any spatial assumptions. This conceptual pivot, which discarded the "pictures" of orbiting electrons in favor of abstract mathematical relations, laid the intuitive groundwork for what would become matrix mechanics. The pivotal moment came one sleepless night when Heisenberg feverishly calculated the quantum analog of the , a model central to vibrations. He constructed infinite arrays for and variables, discovering that their products did not commute, leading to a set of relations that successfully reproduced the known levels and selection rules without relying on classical . This non-commutativity hinted at the commutation relations that would formalize , marking the epiphany's core insight into a purely algebraic approach to quantum phenomena.

The Three Foundational Papers

The foundational development of matrix mechanics unfolded through three seminal papers published in Zeitschrift für Physik in 1925, marking a rapid collaborative effort among , , and . These works shifted from classical intuitions to a based on observable quantities represented by non-commuting arrays, later recognized as matrices. The sequence began with Heisenberg's solitary contribution, followed by joint refinements that formalized the theory and addressed its broader implications. Heisenberg's paper, received by the on 29 July 1925 and published in volume 33, pages 879–893, introduced the core idea of non-commuting dynamical variables treated as infinite arrays for and . Titled " quantentheoretische Umdeutung kinematischer und mechanischer Beziehungen," it proposed a grounded in spectroscopic data, such as transition frequencies and intensities, rather than unobservable trajectories like orbits. Heisenberg represented observables through with amplitudes that formed array-like structures, emphasizing that products of variables, such as and , do not commute, as "classically x(t)y(t) is always equal to y(t)x(t), [but] this does not need to be the case in ." This approach yielded approximate solutions for the and anharmonic oscillator, demonstrating consistency with known quantum rules. Building directly on Heisenberg's arrays, which Born identified as matrices from linear , the second paper by and Jordan, received on 27 1925 and published in volume 34, pages 858–888, formalized as the mathematical foundation of . Titled "Zur Quantenmechanik," it introduced the term "" and defined operations like non-commutative multiplication for arrays representing physical quantities, such as coordinates and momenta, with elements oscillating at quantum frequencies. The authors also introduced state representations akin to vectors and proved by showing the is diagonal, with its time derivative vanishing, thus establishing a key theorem: the energy levels satisfy Bohr's frequency condition. This work clarified the non-commutativity as essential, linking matrix elements to radiative transitions. The collaborative "three-man paper" (Dreimännerarbeit) by Heisenberg, , and , received on 16 November 1925 and published in volume 35, pages 557–615 (1926), provided a comprehensive under the "Zur Quantenmechanik ." Extending the prior works to multi-degree-of-freedom systems, it presented a general framework with diagonal energy matrices, quantum , and quantization rules for , including the possibility of half-integer quantum numbers, later extended to intrinsic and confirmed experimentally. The paper derived general quantum conditions from correspondence principles and solved problems like the perturbatively, solidifying matrix mechanics as a systematic . These papers, all appearing in Zeitschrift für Physik within months, received swift engagement from the physics community despite initial obscurity in Heisenberg's matrix-less terminology. Born's recognition of the array structure spurred the collaborations, and by early 1926, peers like Wolfgang Pauli and Erwin Schrödinger were critiquing and extending the ideas, signaling matrix mechanics' emergence as a viable quantum framework. The rapid publications—spanning July to November 1925—reflected the urgency to resolve atomic inconsistencies, influencing subsequent developments like Dirac's independent formulation.

Heisenberg's Reasoning and Key Insights

Heisenberg's development of matrix mechanics marked a profound philosophical shift in , rejecting the visualization of unobservable electron paths in atomic models such as those proposed by Bohr. Instead, he advocated basing exclusively on directly measurable quantities, arguing that attempts to describe hidden trajectories led to inconsistencies with experimental data. This observability principle critiqued the causal inefficacy of such orbits, emphasizing empirical observables over speculative mechanisms. Central to this approach was the emphasis on spectral lines as the fundamental data of , with transition amplitudes interpreted as representing probabilities of quantum jumps between stationary states. Heisenberg posited that could be described through frequencies and intensities derived from these transitions, rather than continuous motions, aligning the with observed spectra. This focus on probabilities over deterministic paths transformed into a predictive framework for measurable effects, such as line intensities in emission spectra. Heisenberg drew heavily on Bohr's to bridge the quantum and classical regimes, particularly for large quantum numbers where quantum predictions should asymptotically match classical results. This principle guided the formulation of transition probabilities by associating quantum amplitudes with classical components in the high-energy limit, ensuring between the two theories. By applying it, Heisenberg reinterpreted classical equations in terms of quantum observables, linking discrete spectral frequencies to classical harmonic overtones. The reasoning was influenced by the older of Bohr and Sommerfeld, which introduced quantization rules and selection principles but struggled with in non-harmonic systems. Heisenberg addressed these challenges, particularly the difficulties in treating anharmonic oscillators where classical methods failed under quantum conditions, by seeking a general algebraic framework applicable to all systems. As a testing ground, he initially applied the approach to the , where exact solutions reinforced the viability of focusing on observables.

Initial Matrix Formulation

In the initial formulation of matrix mechanics, physical observables such as q and p were represented as infinite-dimensional matrices, where the matrix elements q_{mn} and p_{mn} correspond to transition amplitudes between discrete quantum states labeled by quantum numbers m and n. This approach discretized continuous classical variables into matrix indices tied to stationary states, with the elements incorporating time-dependent factors like q_{mn} e^{2\pi i \nu_{mn} t}, where \nu_{mn} are the Bohr frequencies associated with energy differences between states. States in this framework were described by column vectors, whose components served as probability amplitudes for the system to occupy the corresponding eigenstates. These vectors facilitated the representation of superpositions and transitions, aligning with the probabilistic interpretation emerging from the theory. The product of two s, such as qp, was defined through , yielding a new matrix with elements (qp)_{kn} = \sum_m q_{km} p_{mn}, which inherently revealed the non-commutativity of quantum s since, in general, qp \neq pq. Time evolution of an matrix g followed the equation \dot{g}_{nm} = 2\pi i \nu_{nm} g_{nm}, ensuring consistency with the discrete spectrum of quantum frequencies. This matrix structure, motivated by Heisenberg's to focus on transitions rather than unobservable trajectories, provided the foundational algebraic tool for quantum calculations.

Recognition and Nobel Prize

Upon its publication in 1925, matrix mechanics encountered initial skepticism from some physicists, who viewed its abstract, non-visual approach as a radical departure from and Bohr's own model. This stemmed from the theory's reliance on unobservable mathematical arrays rather than intuitive trajectories, leading some to question its physical interpretability. Acceptance accelerated in 1926 when applied matrix mechanics to derive the energy levels and spectrum of the , reproducing Bohr's results exactly and extending them to perturbations like electric and , thus validating the theory against experimental data. Pauli's calculation demonstrated the method's power and internal consistency, dispelling much of the early reservation and establishing matrix mechanics as a viable alternative to older quantum models. Max and Pascual contributed essential mathematical rigor by recognizing Heisenberg's arrays as matrices and deriving the non-commutative multiplication rules in their 1925 collaborative paper, transforming the initial sketch into a complete . Nevertheless, the for 1932 was awarded exclusively to Werner "for the creation of ," reflecting the committee's attribution of primary credit to his originating insight despite the collaborative development. later received the 1954 for his probabilistic interpretation of quantum states, while was never honored, primarily due to his involvement with the Nazi Party, including enlisting in an SA Stormtrooper unit in 1933, which led to him being shunned in the post-war scientific community. Erwin Schrödinger's wave mechanics, introduced in 1926, was demonstrated by Schrödinger himself to be equivalent to in his 1926 paper, where he showed that Heisenberg's matrix elements correspond to Fourier developments of the atom's electric moment, interpretable through classical electrodynamics. This formulation was excluded from the 1932 prize owing to the nomination cycle's focus on Heisenberg's earlier work as the foundational breakthrough. The Nobel Committee prioritized the matrix approach's precedence in establishing non-commutative quantum dynamics, awarding Schrödinger the 1933 prize jointly with for new forms of . Matrix mechanics' introduction of non-commutative observables marked it as the first abandoning classical commutativity, exerting lasting influence on the field, notably inspiring Paul Dirac's 1926 q-number algebra and relativistic extensions that unified quantum rules with . This non-commutative foundation became central to modern and Dirac's prediction of .

Mathematical Formulation

Quantum Harmonic Oscillator Solution

In matrix mechanics, the serves as a foundational example to illustrate the formalism and verify its consistency with established quantization rules from the . The q and p are represented as infinite matrices in a basis of states, where the matrix elements q_{nm} and p_{nm} correspond to transition amplitudes between states labeled by quantum numbers n and m. These elements are constructed such that off-diagonal contributions vanish except for adjacent states, reflecting the correspondence principle for high quantum numbers where classical periodic motion emerges. The for the system is expressed in matrix form as H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 q^2, with the squares denoting . The matrix elements of H are calculated assuming the tridiagonal : diagonal elements H_{nn} represent energies, while neighboring off-diagonal elements H_{n,n\pm1} arise from the products p^2 and q^2. Specifically, the matrix elements take the form q_{n,n+1} = \sqrt{\frac{(n+1)\hbar}{2 m \omega}} (up to phases), ensuring compliance with the [q, p] = i \hbar. Diagonalization of the is performed by finding the eigenvalues through the or iterative methods, transforming the representation to a basis where H is diagonal. This yields the energy eigenvalues E_n = \hbar \omega \left( n + \frac{1}{2} \right), with n = 0, 1, 2, \dots, spaced by multiples of \hbar \omega and including the \frac{1}{2} \hbar \omega. These levels match the semiclassical predictions for the oscillator, demonstrating the theory's validity. The off-diagonal elements in the original basis, particularly those connecting states with \Delta n = \pm 1, govern quantum transitions and determine the probabilities for or of photons at frequencies \nu_{n,n+1} = \omega / 2\pi. These elements align with selection rules \Delta n = \pm 1, providing a matrix-mechanical interpretation of intensities in systems.

Canonical Commutation Relations

In matrix mechanics, introduced a novel approach to derive the for quantum observables by analogy with classical , but adapted to the non-commuting arrays representing physical quantities. To mimic the classical time in the quantum , employed a "differentiation trick," treating the indices of the matrices—corresponding to quantum numbers of states—as analogous to time variables. This allowed him to approximate using finite differences, effectively bridging the classical structure to quantum algebra. The derivation begins with the classical equation of motion \dot{q} = \frac{\partial H}{\partial p}, where H is the Hamiltonian. In the quantum formulation, observables q and p are infinite matrices, and their products do not commute. Heisenberg postulated that the quantum analog of the time evolution should satisfy \frac{dq}{dt} = \frac{1}{i\hbar} [q, H], where [q, H] = qH - Hq is the . To operationalize \frac{dq}{dt} for matrix elements q_{mn}, he considered a finite difference approximation: \frac{dq}{dt} \approx \frac{q_{m,n+1} - q_{m,n}}{\Delta t}, with the index shift \Delta n = 1 playing the role of a small time increment \Delta t, informed by the discrete spectrum of quantum numbers and the correspondence principle. This leads to the relation [q, H] = i\hbar \frac{\partial q}{\partial t}, ensuring consistency with classical limits for large quantum numbers. Max Born and Pascual Jordan formalized this insight in their subsequent analysis, recognizing that the finite difference condition imposed by Heisenberg on the matrix elements corresponds to the canonical commutation relation [q, p] = i\hbar for the fundamental position and momentum observables, where the identity is implied. For a general Hamiltonian H(q, p), the equations of motion generalize to \dot{q} = \frac{1}{i\hbar} [q, H] = \frac{\partial H}{\partial p} and \dot{p} = \frac{1}{i\hbar} [p, H] = -\frac{\partial H}{\partial q}, with the partial derivatives treated formally as in classical mechanics but evaluated using the non-commuting products. This structure preserves energy conservation (\dot{H} = 0) only if the commutation relation holds. For systems with multiple degrees of freedom, Born and Jordan extended the relation to [q_j, p_k] = i\hbar \delta_{jk}, with all other commutators vanishing ([q_j, q_k] = [p_j, p_k] = 0), ensuring the algebra respects the independence of coordinates. This multi-variable form arises naturally from applying the single-pair commutation rule to polynomial expansions of H in the q_j and p_k, maintaining the correspondence to classical Poisson brackets \{q_j, p_k\} = \delta_{jk} in the limit \hbar \to 0, where commutators scale as i\hbar times the Poisson brackets. The relation underpins the algebraic consistency of matrix mechanics, as verified in applications like the quantum harmonic oscillator.

Heisenberg Picture and Equations of Motion

In matrix mechanics, the provides a framework where the remains time-independent, while the matrices representing physical observables evolve dynamically according to the . This approach aligns naturally with the operator-based formulation introduced by Heisenberg, , and , emphasizing the evolution of non-commuting dynamical variables rather than wave functions. The time dependence is incorporated into the observables themselves, allowing direct analogy to through structures. The time evolution of an arbitrary operator A in the Heisenberg picture is defined via the unitary transformation A(t) = U^\dagger(t) A U(t), where U(t) = e^{-i H t / \hbar} is the unitary evolution operator and H is the time-independent Hamiltonian matrix. To derive the explicit equation of motion, consider the time derivative: \frac{d}{dt} A(t) = \frac{d}{dt} \left[ U^\dagger(t) A U(t) \right] = \left( \frac{d U^\dagger}{dt} \right) A U(t) + U^\dagger(t) A \left( \frac{d U}{dt} \right). From the Schrödinger equation for the evolution operator, i \hbar \frac{d U}{dt} = H U, it follows that \frac{d U}{dt} = -\frac{i}{\hbar} H U and, since H is Hermitian, \frac{d U^\dagger}{dt} = \frac{i}{\hbar} U^\dagger H. Substituting these yields: \frac{d}{dt} A(t) = \frac{i}{\hbar} U^\dagger H A U - \frac{i}{\hbar} U^\dagger A H U = \frac{i}{\hbar} U^\dagger (H A - A H) U = \frac{i}{\hbar} [H, A(t)], or equivalently, i \hbar \frac{d}{dt} A(t) = [A(t), H]. This is the Heisenberg equation of motion, first systematically derived in the context of matrix mechanics by generalizing classical Poisson brackets to quantum commutators divided by i \hbar. The equation holds for any observable A, building on the canonical commutation relations such as [q, p] = i \hbar. As illustrative examples, consider the position x and momentum p operators. For a free particle with Hamiltonian H = \frac{p^2}{2m}, the commutators yield [x, H] = \frac{i \hbar}{m} p and [p, H] = 0. Thus, the Heisenberg equations give \dot{x} = \frac{p}{m} and \dot{p} = 0, mirroring classical free-particle motion. For the quantum harmonic oscillator with H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 x^2, one finds [x, H] = \frac{i \hbar}{m} p and [p, H] = -i \hbar m \omega^2 x, leading to \dot{x} = \frac{p}{m} and \dot{p} = -m \omega^2 x, again reproducing the classical oscillator dynamics. These evolutions can be solved explicitly, yielding x(t) = x(0) \cos(\omega t) + \frac{p(0)}{m \omega} \sin(\omega t) and p(t) = p(0) \cos(\omega t) - m \omega x(0) \sin(\omega t). In the Heisenberg picture, the interpretation emphasizes that expectation values of observables, \langle A(t) \rangle = \sum_{mn} c_m^* c_n A_{mn}(t), evolve according to the same formal equation as the operators themselves, i \hbar \frac{d}{dt} \langle A \rangle = \langle [A, H] \rangle. In the classical correspondence limit—where \hbar \to 0 or for systems with large quantum numbers—the off-diagonal matrix elements contribute negligibly, and the expectation values satisfy the canonical equations of classical mechanics, bridging quantum and classical descriptions.

Conservation of Energy and Other Laws

In matrix mechanics, the conservation of energy follows directly from the structure of the theory. The Hamiltonian operator H, representing the total energy, is time-independent, and its commutator with itself vanishes trivially: [H, H] = 0. Applying the Heisenberg equation of motion, the time derivative of H is given by \frac{dH}{dt} = \frac{i}{\hbar} [H, H] = 0, implying that the energy remains constant throughout the system's evolution. This result was established in the foundational formulation, ensuring that energy eigenvalues, which are the diagonal elements of H, determine stationary states without temporal variation. More generally, matrix mechanics incorporates an analog of , linking symmetries of the to conserved quantities. If an Q corresponding to a commutes with H, i.e., [Q, H] = 0, then \frac{dQ}{dt} = 0, making Q a constant of the motion. For instance, in a translationally invariant system where H does not explicitly depend on position coordinates, the total P satisfies [P, H] = 0, leading to . This framework extends classical principles to the quantum domain through the non-commutative algebra of observables. A key application arises in systems with , such as a particle in a central potential V(r). Here, the H = \frac{p^2}{2m} + V(r) commutes with each component of the \mathbf{L}, yielding [L_i, H] = 0 for i = x, y, z. Consequently, is conserved, simplifying the solution of problems like the quantum by allowing separation into radial and angular parts. This relation preserves the vector nature of \mathbf{L} and its magnitude L^2, mirroring classical orbital conservation under central forces. Additionally, matrix mechanics ensures the conservation of probability through the unitarity of . Observables are represented by Hermitian matrices, guaranteeing real eigenvalues and orthogonal eigenvectors, while the time-evolution generated by H is unitary, preserving the normalization of state vectors. This unitarity maintains the total probability across all possible measurement outcomes, providing a foundational for the probabilistic of quantum states.

Extensions and Applications

Relation to Wave Mechanics

In 1926, Erwin Schrödinger introduced wave mechanics through a series of papers, proposing a that describes via continuous wave functions, and applied it to solve the problem, yielding energy levels identical to those obtained earlier via matrix mechanics. This parallelism demonstrated that wave mechanics could reproduce the discrete spectral lines predicted by Heisenberg, , and Jordan's matrix approach, despite their differing conceptual foundations. Schrödinger initially criticized mechanics for its abstract, non-intuitive nature, viewing it as a formal manipulation of mathematical arrays disconnected from physical visualization, and sought to replace it with his more geometrically intuitive picture. In a 1926 paper, he demonstrated a partial equivalence between the two formulations by showing how matrix elements could be expressed as integrals over functions, though he expressed reservations about the full isomorphism of their structures. The complete mathematical equivalence was rigorously established by in his 1932 formulation, where both matrix mechanics and wave mechanics were unified as operator algebras on an infinite-dimensional , with matrices representing operators acting on wave functions as basis expansions. This framework resolved earlier ambiguities and provided a common abstract setting for quantum observables and states. The historical tension subsided through the development of transformation theory by , , and in 1927, which bridged the formulations by introducing unitary transformations between matrix and wave representations, enabling calculations in either picture while confirming their physical predictions. Matrix mechanics proved particularly advantageous for handling discrete energy spectra and transition probabilities, while wave mechanics excelled in describing continuous phenomena like ; in modern quantum mechanics, these are integrated seamlessly within the paradigm.

Ehrenfest Theorem

In matrix mechanics, Ehrenfest's theorem establishes that the expectation values of position q and momentum p evolve according to equations analogous to the classical Hamilton's equations of motion. Specifically, for a Hamiltonian H(q, p), \frac{d}{dt} \langle q \rangle = \left\langle \frac{\partial H}{\partial p} \right\rangle, \quad \frac{d}{dt} \langle p \rangle = -\left\langle \frac{\partial H}{\partial q} \right\rangle, where \langle \cdot \rangle denotes the expectation value with respect to the quantum state. This result was first derived by Paul Ehrenfest in 1927, demonstrating the approximate validity of classical mechanics within quantum theory. The proof relies on the , where operators evolve according to the equation i \hbar \frac{dA}{dt} = [A, H] for a time-independent operator A, combined with the linearity of expectation values \langle [A, B] \rangle = \langle A B \rangle - \langle B A \rangle. For the standard Hamiltonian H = \frac{p^2}{2m} + V(q), the canonical commutation relation [q, p] = i \hbar yields \frac{d}{dt} \langle q \rangle = \frac{1}{i \hbar} \langle [q, H] \rangle = \frac{\langle p \rangle}{m}, since [q, H] = [q, p^2 / 2m] = i \hbar p / m. Similarly, \frac{d}{dt} \langle p \rangle = \frac{1}{i \hbar} \langle [p, H] \rangle = -\left\langle \frac{dV}{dq} \right\rangle, as [p, V(q)] = -i \hbar \frac{dV}{dq}. These follow directly from the general commutation rules [q, f(p)] = i \hbar \frac{\partial f}{\partial p} and [p, g(q)] = -i \hbar \frac{\partial g}{\partial q} for functions f and g. The theorem implies that quantum mechanics recovers classical mechanics in the limit \hbar \to 0 or for macroscopic systems with large quantum numbers, where the relative uncertainties in position and momentum become negligible, allowing expectation values to trace classical trajectories closely. For the quantum harmonic oscillator with H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 q^2, the expectation values \langle q \rangle and \langle p \rangle oscillate exactly as in the classical case with frequency \omega, independent of the initial state, due to the quadratic form of the potential.

Transformation Theory

Transformation theory in quantum mechanics, primarily developed by and in the mid-1920s, establishes a mathematical framework for unifying different representations of through unitary transformations. These transformations allow observables, represented as operators A, to be mapped to new operators A' = U A U^\dagger, where U is a satisfying U^\dagger U = I, ensuring the preservation of algebraic structures such as the canonical commutation relations [q, p] = i\hbar. This approach demonstrates the equivalence of matrix mechanics and wave mechanics by showing that both can be viewed as specific realizations within the same abstract space. In his 1926 work, Dirac introduced an abstract formulation of , treating states and observables in an infinite-dimensional rather than concrete matrices or functions. This formalism uses linear operators on the to represent physical quantities, with state vectors denoting quantum states, enabling a representation-independent treatment of . Dirac's notation emphasized the inner product between states, laying the groundwork for later symbolic tools, and highlighted how transformations between bases maintain the probabilistic interpretation of quantum amplitudes. Jordan complemented Dirac's ideas by focusing on the infinite-dimensional nature of the in his 1927 contributions, where he rigorously developed unitary transformations for continuous spectra and introduced completeness relations to ensure the resolution of the operator, \int |n\rangle\langle n| = I, over a complete . These relations guarantee that any state can be expanded in the basis, facilitating transformations between discrete matrix representations and continuous wave functions without loss of information. Jordan's emphasis on the mathematical consistency of these operations in infinite dimensions addressed potential divergences in early matrix formulations. The culmination of Dirac and Jordan's transformation theory is a general quantum that operates solely in the abstract , rendering the theory independent of any particular representation. This unification resolved apparent discrepancies between and wave mechanics, establishing as a consistent theory of linear operators on , with unitary transformations serving as the bridge between equivalent descriptions of the same physical reality.

Selection and Sum Rules

In matrix mechanics, selection rules govern the allowed transitions between quantum states, determined by the non-vanishing matrix elements of the perturbation operator, such as the dipole moment for electric dipole transitions. These rules emerge directly from the algebraic structure of the infinite matrices representing observables, where off-diagonal elements \langle m | \hat{O} | n \rangle vanish unless specific conditions on the quantum numbers m and n are met. For the quantum harmonic oscillator, a foundational example solved using matrix methods, the position operator \hat{x} has non-zero matrix elements only when the change in quantum number satisfies \Delta n = \pm 1, as derived from the recurrence relations in the oscillator's infinite matrix representation. This restriction arises because higher-order differences lead to zero contributions in the matrix products enforcing the equations of motion, ensuring that only adjacent states couple under linear perturbations like \hat{x}. Generalizing to atomic systems, these selection rules extend to dipole transitions in multi-electron atoms, where the matrix elements \langle f | \hat{\mathbf{r}} | i \rangle for the position operator vanish unless the angular momentum quantum numbers change by \Delta l = \pm 1 and \Delta m_l = 0, \pm 1, reflecting the vector nature of the dipole operator. This framework, developed in early matrix mechanics applications, predicts forbidden transitions (e.g., \Delta l = 0) by the orthogonality of spherical harmonics in the matrix elements, providing a quantum explanation for observed spectral line absences in atomic spectra. Such rules enabled precise predictions of allowed radiative transitions, aligning experimental intensities with theoretical expectations without invoking wave functions. Sum rules in matrix mechanics further constrain transition intensities through global relations derived from the commutation relations and the of the basis. The general f-sum rule states that the sum of oscillator strengths f_{mn} over all final states m from an initial n equals the number of electrons [Z](/page/Z), i.e., \sum_m f_{mn} = [Z](/page/Z), where f_{mn} = \frac{2m_e (E_m - E_n)}{\hbar^2} |\langle m | x | n \rangle|^2 for one . This follows from inserting the relation \sum_m |m\rangle\langle m| = \hat{1} into the double [\hat{x}, [\hat{H}, \hat{p}]] = i\hbar [\hat{x}, \hat{p}], leveraging the [\hat{x}, \hat{p}] = i\hbar to yield a model-independent result. A key application is the Thomas-Reiche-Kuhn (TRK) sum , a specific form of the f-sum for atomic s, stating \sum_n \frac{2m_e (E_n - E_0)}{\hbar^2} |\langle n | \mathbf{r} | 0 \rangle|^2 = N for N s in the , derived similarly from position-momentum commutation in the . Originally formulated using the correspondence principle but rigorously proven in the matrix framework, the TRK relates total transition strengths to electron count, explaining the saturation of oscillator strengths in atomic spectra and validating intensity distributions observed in lines. These sum rules provided early tests of matrix , confirming its predictive power for spectral intensities across hydrogen-like and multi-electron atoms.

References

  1. [1]
    The 1925 Born and Jordan paper “On quantum mechanics”
    Feb 1, 2009 · This paper introduced matrices to physicists. We discuss the original postulates of quantum mechanics, present the two-part discovery of the law of commutation.II. BACKGROUND TO “ON... · III. THE ORIGINAL... · VI. THE ENERGY THEOREMS
  2. [2]
    None
    ### Summary of Matrix Mechanics Section
  3. [3]
    1925: the first papers on quantum mechanics - Europhysics News
    May 6, 2025 · A review of the foundational papers on matrix mechanics of Werner Heisenberg, Max Born, Pascual Jordan and Paul Dirac, which were written 100 years ago<|control11|><|separator|>
  4. [4]
    [PDF] On Quantum Mechanics - Neo-classical physics
    The Ansätze that Heisenberg recently gave will be developed into a systematic theory of quantum mechanics (initially for systems of one degree of freedom).
  5. [5]
    [PDF] on quantum mechanics ii - m. born, w. heisenberg and p. jordan
    Abstract: The quantum mechanics developed in Part I of this paper from. Heisenberg's approach is here extended to systems having arbitrarily many degrees of ...
  6. [6]
    [PDF] Tran_sub_et_all_Am_J.. - Georgetown University
    Jul 23, 2025 · In 1925, Heisenberg, Born, and Jordan developed matrix mechanics as a strategy to solve quantum- mechanical problems. While finite-sized ...
  7. [7]
    [PDF] 1 Why were two theories (Matrix Mechanics and Wave Mechanics ...
    A recent rethinking of the early history of Quantum Mechanics deemed the late 1920s agreement on the equivalence of Matrix Mechanics and Wave Mechanics, ...
  8. [8]
    [PDF] Understanding Heisenberg's 'magical' paper of July 1925 - arXiv
    Apr 1, 2004 · Heisenberg, Physics and Beyond (London, Allen & Unwin, 1971), page 61. 54 W. Pauli, “Über das Wasserstoffspektrum vom Standpunkt der neuen ...
  9. [9]
  10. [10]
    [PDF] On quantum-theoretical reinterpretation of kinematic and ...
    “Über quantentheoretische Umdeutung kinematischer und mechanischer Beziehungen,” Zeit. Phys. 33. (1925), 879-893. On the quantum-theoretical reinterpretation.
  11. [11]
  12. [12]
  13. [13]
    3. The Development of Quantum Mechanics (1925 – 1927)
    ... matrix mechanics”, culminating in the long “three-man-paper”, by Born, Heisenberg and Jordan, submitted on 16 November 1925. Further developments followed ...<|control11|><|separator|>
  14. [14]
    June/July 1925: Werner Heisenberg pioneers quantum mechanics
    Jul 1, 2025 · Heisenberg's matrix mechanics fixed the holes in quantum theory by taking physics into the realm of pure abstraction and math.Missing: epiphany | Show results with:epiphany
  15. [15]
    [PDF] Understanding Heisenberg's “magical” paper of July 1925
    returning from Heligoland, and, although I think I under- stand quantum ... Heisenberg, Physics and Beyond 共Allen & Unwin, London, 1971兲, p. 61. 64W ...
  16. [16]
    (PDF) Heisenberg's observability principle - Academia.edu
    Heisenberg's observability principle critiques Bohr's atomic model by eliminating causally idle electron orbits. The paper argues observability should not ...
  17. [17]
    “The Heisenberg Method”: Geometry, Algebra, and Probability in ...
    The mathematical correspondence principle motivated Heisenberg's decision to retain the equations of classical mechanics, while, necessarily, introducing ...Missing: unobservable | Show results with:unobservable
  18. [18]
    Bohr's Correspondence Principle
    Oct 14, 2010 · The correspondence principle was first introduced in the context of the old quantum theory, which was developed in the period between 1900 and ...
  19. [19]
    Werner Heisenberg's Path to Matrix Mechanics
    Heisenberg did not know it, but his inspired insight on noncommutative mathematics pointed directly to a matrix formulation of quantum mechanics, which was ...Missing: epiphany | Show results with:epiphany
  20. [20]
    [PDF] Chapter 2 - Heisenberg's Matrix Mechanics - faculty.fairfield.edu
    Heisenberg's theory, as was developed immediately by Born, Jordan, Pauli and others, is capable of producing the Hydrogen spectra, the selection rules and ...
  21. [21]
    Quantum mechanics and a preliminary investigation of the hydrogen ...
    A recent paper by Heisenberg provides the clue to the solution of this question, and forms the basis of a new quantum theory.
  22. [22]
    The hydrogen atom in electric and magnetic fields: Pauli's 1926 article
    Feb 1, 2003 · The results obtained by Pauli in his 1926 article on the hydrogen atom made essential use of the dynamical so(4) symmetry of the bound states. ...
  23. [23]
    Matrix mechanics mis-prized: Max Born's belated nobelization
    Oct 18, 2023 · The “Matrix mechanics [MM]” of Born, Jordan, and Heisenberg and its application to the hydrogen atom by Pauli were in hand before Schrödinger ...
  24. [24]
    Quantum theory and the Nobel prize - Physics World
    Aug 2, 2002 · Meanwhile, the 1933 prize was shared by Erwin Schrödinger and Paul Dirac for “the discovery of new productive forms of atomic theory”. The ...
  25. [25]
    Paul Dirac: a genius in the history of physics - CERN Courier
    Aug 15, 2002 · From this thought he quickly developed a quantum theory that was based on non-commuting dynamical variables. ... Heisenberg's matrix mechanics.
  26. [26]
    Über quantentheoretische Umdeutung kinematischer und ...
    Über quantentheoretische Umdeutung kinematischer und mechanischer Beziehungen. Published: December 1925. Volume 33, pages 879–893, (1925); Cite this ...
  27. [27]
  28. [28]
  29. [29]
    Heisenberg Equation of Motion
    Equation (244) shows how the dynamical variables of the system evolve in the Heisenberg picture. It is denoted the Heisenberg equation of motion.Missing: primary sources
  30. [30]
    Zur Quantenmechanik. II. | Zeitschrift für Physik A Hadrons and nuclei
    Zur Quantenmechanik. II. ... Article PDF. Download to read the full article text. Similar content being viewed by others ...
  31. [31]
    Quantisierung als Eigenwertproblem - 1926 - Annalen der Physik
    Quantisierung als Eigenwertproblem. E. Schrödinger,. E. Schrödinger ... First published: 1926. https://doi.org/10.1002/andp.19263851302. Citations ...
  32. [32]
    Quantisierung als Eigenwertproblem - 1926 - Annalen der Physik
    Quantisierung als Eigenwertproblem. E. Schrödinger,. E. Schrödinger. Zürich ... First published: 1926. https://doi.org/10.1002/andp.19263840404. Citations ...
  33. [33]
    Quantum Theory and Mathematical Rigor
    Jul 27, 2004 · In the late 1920s, von Neumann developed the separable Hilbert space formulation of quantum mechanics, which later became the definitive one ( ...
  34. [34]
    Why were Matrix Mechanics and Wave ... - ScienceDirect.com
    The Matrix Mechanics was an algebraic approach employing the technique of manipulating matrices. The Wave Mechanics, in contrast, employed differential ...
  35. [35]
    Ehrenfest Theorem - Richard Fitzpatrick
    are functions that can be expanded as power series, are easily proved using the fundamental commutation relations, Equations (3.32). ... using the Heisenberg or ...
  36. [36]
    None
    ### Summary of Ehrenfest's Theorem from the Provided Document
  37. [37]
    The physical interpretation of the quantum dynamics - Journals
    In Heisenberg's original matrix mechanics it was assumed that the elements of the diagonal matrix that represents the energy are the energy levels of the ...<|control11|><|separator|>
  38. [38]
    Über eine neue Begründung der Quantenmechanik
    Die vier bisher entwickelten Formen der Quantenmechanik: die Matrizentheorie, die Theorie von Born und Wiener, die Wellenmechanik und die Theorie derq-Zahl.
  39. [39]
    On the theory of quantum mechanics - Journals
    The new mechanics of the atom introduced by Heisenberg may be based on the assumption that the variables that describe a dynamical system do not obey the ...
  40. [40]
    Pascual Jordan
    Physics Today article detailing Pascual Jordan's life, contributions to quantum mechanics, and his involvement with the Nazi Party, including enlisting in the SA in 1933, which contributed to his lack of Nobel recognition.
  41. [41]
    Relation between the Quantum Mechanics of Heisenberg, Born and Jordan and the Matrix Mechanics
    Erwin Schrödinger's 1926 paper demonstrating the equivalence between wave mechanics and matrix mechanics via Fourier analysis of the atom's electric moment.