Fact-checked by Grok 2 weeks ago

S -matrix

The S-matrix, also known as the scattering matrix, is a unitary operator in quantum mechanics and quantum field theory that relates the asymptotic initial states of particles at t \to -\infty to their asymptotic final states at t \to +\infty, thereby encoding the probabilities and amplitudes for all possible outcomes in scattering processes. It is defined as S = \lim_{t_\text{out}, t_\text{in} \to \mp \infty} U(t_\text{out}, t_\text{in}), where U is the time-evolution operator, ensuring that the total probability is conserved through its unitarity property S^\dagger S = 1. In practice, the S-matrix elements \langle f | S | i \rangle provide the transition amplitudes from initial state |i\rangle to final state |f\rangle, often expressed as S = 1 + iT, where T is the transition matrix related to interaction strengths. The concept of the S-matrix was first introduced by in 1937 as a tool to describe nuclear reactions using resonating group structures, marking an early attempt to systematize without full solutions. It gained prominence in the through Werner Heisenberg's development of S-matrix theory, proposed as an alternative framework to for handling high-energy particle interactions, emphasizing observable data over unobservable fields. Heisenberg's approach, detailed in papers from 1943 onward, aimed to resolve divergences in by focusing solely on the S-matrix's analytic properties derived from and unitarity. This theory influenced subsequent work, including the bootstrap models of the 1960s and , before being largely integrated into standard by the 1970s. Key properties of the S-matrix include its analyticity in the complex energy plane, which stems from (effects cannot precede causes) and locality of interactions, leading to dispersion relations that connect real and imaginary parts of . Unitarity not only conserves probability but also implies the optical theorem, $2 \Im \langle f | T | i \rangle = \sum_n \langle f | T^\dagger | n \rangle \langle n | T | i \rangle, linking the forward scattering 's imaginary part to total cross-sections. Poincaré invariance ensures the S-matrix commutes with Lorentz transformations, making it a covariant object suitable for relativistic theories. In modern physics, the S-matrix remains central to perturbative calculations in and electroweak theory, as well as approaches like the S-matrix bootstrap, where it serves as the primary for constraining effective theories without relying on underlying Lagrangians. Recent applications extend to methods in experiments and even connections to and chaos in . As of 2024, there has been a resurgence in S-matrix research, highlighted by workshops like the S-Matrix Marathon and advances in the S-matrix bootstrap program.

Historical Development

Origins in Early Quantum Mechanics

The concept of the S-matrix emerged from early efforts in to describe processes without relying on the full solution of the , building on foundational ideas about transition probabilities. In 1927, introduced key precursors through his work on the quantum mechanics of collision processes, where he developed methods to compute scattering amplitudes, laying the groundwork for descriptions of particle transitions. This approach gained practical momentum in with 's 1937 proposal of the S-matrix as a mapping initial to final states in reactions. Wheeler formulated the S-matrix within the resonating group structure method to analyze light nuclei and nuclear interactions, allowing predictions of reaction outcomes—such as cross-sections for particle collisions—directly from asymptotic wave functions, bypassing the computationally intensive task of solving the complete many-body for complex nuclear systems. Wheeler's non-relativistic framework was significantly extended by during , amid growing frustrations with the infinities plaguing perturbative calculations in (QFT). In papers published in Zeitschrift für Physik from 1943 to 1944, Heisenberg proposed the S-matrix as a fundamental tool for relativistic particle interactions, prioritizing observable scattering amplitudes over unphysical intermediate states and emphasizing principles like and unitarity to circumvent renormalization issues. Heisenberg introduced the unitary property as a core postulate, ensuring probability conservation in transitions between free-particle states, which became central to the theory's predictive power for high-energy processes.

Evolution in Quantum Field Theory

In the post-World War II era, the S-matrix formalism underwent significant refinement within relativistic (QFT), particularly through the development of dispersion relations in the 1950s. These relations, derived from and unitarity principles, established the analyticity of the S-matrix in the complex energy plane, enabling predictions for scattering amplitudes without relying on underlying field Lagrangians. Pioneering contributions came from Geoffrey Chew and , who applied dispersion relations to pion-nucleon scattering, connecting asymptotic behavior at high energies to low-energy experimental data and resolving issues in phenomenology. A hallmark of this evolution was the bootstrap hypothesis, proposed by Geoffrey Chew in the early , which posited that the S-matrix could be self-consistently determined solely from its general principles—such as unitarity, crossing symmetry, and analyticity—without invoking elementary particles or fundamental fields. This approach aimed to derive particle masses, widths, and coupling constants as bound states emerging from the S-matrix dynamics itself, offering a democratic view of s as composite excitations. The hypothesis gained traction in physics, where it integrated with , developed in the late and by Tullio Regge and others, to model high-energy scattering via Regge poles and trajectories that interpolated between resonances and particle exchanges. Regge theory successfully described forward scattering peaks and total cross-section behaviors in hadron collisions, providing a framework for extrapolating experimental data beyond perturbative regimes. The S-matrix approach demonstrated early empirical success in interpreting 1950s experiments on pion photoproduction, where dispersion relations predicted multipole amplitudes that matched observations from photon-nucleon interactions at energies up to several hundred MeV, validating the method's non-perturbative power before the dominance of techniques. By the 1970s, however, the rise of (QCD) as the standard theory of strong interactions—confirmed by and data—led to a decline in pure S-matrix methods, as QCD provided a field-theoretic basis for confinement and quark-gluon dynamics. Despite this, S-matrix principles persisted in effective field theories for low-energy phenomena and saw a revival in the 21st century through the AdS/CFT correspondence, where bulk scattering amplitudes in map to boundary correlators, bridging and S-matrix analyticity.

Conceptual Foundations

Motivation from Scattering Processes

In scattering processes, particles interact over long distances, transitioning from initial free-particle states to final states that are also asymptotically free, allowing the S-matrix to characterize these interactions without resolving the full dynamics of the collision. This approach arises from the practical difficulties in solving the time-dependent for multi-particle systems, where the complexity of entangled wavefunctions and makes exact solutions infeasible, shifting emphasis to probabilities between distant asymptotic regimes. As time approaches ±∞, interactions effectively vanish, defining incoming waves from the past and outgoing waves to the future, which the S-matrix maps directly. The elements of the S-matrix, denoted as S_{fi} = \langle f | S | i \rangle, where |i\rangle and |f\rangle are initial and final asymptotic states, yield transition probabilities |S_{fi}|^2 that quantify the likelihood of specific outcomes in experiments. These probabilities underpin the calculation of differential and total cross-sections, essential for predicting reaction rates and comparing theory with experimental data, such as in particle accelerators where cross-sections measure interaction strengths. Historically, the S-matrix concept emerged in to describe collision outcomes in light nuclei without full wavefunction computations, as introduced by in 1937 for resonating group structures. In , further motivated its development in 1943 to circumvent ultraviolet divergences plaguing perturbative calculations, by parameterizing observables like scattering amplitudes directly through an integral formulation that avoids differential equations and focuses solely on asymptotic behaviors. This strategy ensured finite, unitarity-constrained results for physical processes, preserving probability conservation amid theoretical uncertainties.

General Role and Applications

The S-matrix serves as the fundamental operator in that encodes the probabilities for transitions between free-particle asymptotic states, allowing direct computation of key observables in experiments. Specifically, the cross-section for a process is proportional to the square of the magnitude of the relevant S-matrix element, \frac{d\sigma}{d\Omega} \propto | \langle f | S | i \rangle |^2, where |i\rangle and |f\rangle denote the initial and final states, while the total cross-section is obtained by integrating this over all solid angles. These relations enable precise predictions of interaction strengths without relying on intermediate off-shell states. In perturbative , the S-matrix can be expanded using the to compute these elements order by order in the . Beyond cross-sections, the S-matrix connects to other observables, including decay rates via the optical theorem, which relates the imaginary part of the forward to the total cross-section and thus to sum rules over decay channels. Resonances appear as poles in the complex energy plane of the S-matrix, determining their positions and widths, while extracts phase shifts that characterize the energy dependence of interactions. These features are crucial for interpreting collider data and modeling unstable particles. The S-matrix finds broad applications across physics disciplines. In , it is used to analyze neutron-proton scattering, where phase shifts derived from S-matrix elements fit experimental data to constrain nucleon-nucleon potentials. In , the scattering matrix approach describes quantum electron transport in mesoscopic systems, relating transmission probabilities to conductance via Landauer-Büttiker formalism. Analogously, in , the S-matrix framework for wave propagation through multilayer structures parallels the Jones matrix, which handles polarization transformations in birefringent media. In axiomatic quantum field theory, the S-matrix is constrained by fundamental postulates: unitarity ensures probability conservation, analyticity arises from causality and locality, and crossing symmetry relates scattering in different channels, providing a non-perturbative foundation independent of Lagrangian details. Modern applications leverage these properties in effective field theories for low-energy quantum chromodynamics (QCD), where S-matrix bootstrap methods bound low-energy constants in chiral perturbation theory to describe pion scattering and nucleon interactions.088) Additionally, S-matrix unitarity imposes bounds on beyond-Standard-Model physics, guiding searches for new particles at the (LHC) by constraining deviations in high-energy scattering amplitudes.

S-matrix in Non-Relativistic Quantum Mechanics

Definition in One Dimension

In one-dimensional non-relativistic quantum mechanics, the S-matrix provides a compact description of particle scattering by a localized potential V(x), relating the asymptotic forms of the incoming and outgoing wavefunctions. The scattering problem is governed by the time-independent Schrödinger equation, -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V(x) \psi(x) = E \psi(x), where E > 0 is the energy of the incident particle, m is its mass, and \hbar is the reduced Planck's constant. For large |x|, where V(x) \to 0, the solutions reduce to free-particle plane waves with wave number k = \sqrt{2mE}/\hbar. The general asymptotic behavior of the wavefunction \psi(x) is thus \psi(x) \sim A e^{ikx} + B e^{-ikx} \quad (x \to -\infty), \psi(x) \sim C e^{ikx} + D e^{-ikx} \quad (x \to +\infty), where A and D represent incoming amplitudes from the left and right, respectively, while B and C are outgoing amplitudes to the left and right. The S-matrix is defined as the $2 \times 2 unitary matrix that maps the incoming amplitudes to the outgoing ones: \begin{pmatrix} B \\ C \end{pmatrix} = S \begin{pmatrix} A \\ D \end{pmatrix}, with S = \begin{pmatrix} r & t' \\ t & r' \end{pmatrix}. Here, t (or t') is the transmission coefficient, giving the amplitude for the particle to pass through the potential from left to right (or right to left), while r (or r') is the reflection coefficient, giving the amplitude for backscattering from left (or right). For time-reversal invariant and real potentials, t = t' and |r|^2 + |t|^2 = 1, ensuring conservation of probability current. The explicit form of S is obtained by solving the Schrödinger equation numerically or analytically, matching the wavefunction and its derivative at boundaries within the potential region to determine the coefficients.

Unitary Property

In one-dimensional , the unitarity of the S-matrix arises directly from the conservation of probability, as encoded in the derived from the . The probability density is \rho(x, t) = |\psi(x, t)|^2, and the is \frac{\partial \rho}{\partial t} + \frac{\partial j}{\partial x} = 0, where the is j(x, t) = \frac{\hbar}{2mi} \left( \psi^* \frac{\partial \psi}{\partial x} - \psi \frac{\partial \psi^*}{\partial x} \right). For stationary scattering states of the form \psi(x, t) = \psi(x) e^{-iEt/\hbar}, the density \rho is time-independent, implying \frac{\partial j}{\partial x} = 0 and thus a j throughout . Consider a scattering state with incidence from the left, where the asymptotic is \psi(x) \sim e^{ikx} + r e^{-ikx} as x \to -\infty and \psi(x) \sim t e^{ikx} as x \to +\infty. The current at x \to -\infty evaluates to j(-\infty) = \frac{\hbar k}{m} (1 - |r|^2), while at x \to +\infty it is j(+\infty) = \frac{\hbar k}{m} |t|^2. Equating these constant currents yields |t|^2 + |r|^2 = 1. A similar relation holds for incidence from the right. The full S-matrix, which linearly relates the outgoing wave amplitudes to the incoming ones as \begin{pmatrix} b_L \\ b_R \end{pmatrix} = S \begin{pmatrix} a_L \\ a_R \end{pmatrix} (with a_L, a_R incoming and b_L, b_R outgoing amplitudes from left and right), takes the form S = \begin{pmatrix} t & r' \\ r & t' \end{pmatrix}. Unitarity S^\dagger S = I follows from these flux conservation conditions, ensuring the norms of incoming and outgoing states are preserved. This unitarity has direct implications for elastic scattering: the transmission probability T = |t|^2 and reflection probability R = |r|^2 satisfy T + R = 1, conserving the total incident probability flux. In one dimension, unitarity also implies the optical , which relates the imaginary part of the forward transmission amplitude \operatorname{Im} t(k) to the total scattering cross-section (analogous to the reflection probability in 1D). Specifically, \sigma_\text{tot} = \frac{2}{k} \operatorname{Im} f^{(+)}(k, +), where f^{(+)}(k, +) is the forward related to t(k). For multi-channel scattering, such as when the particle has internal (e.g., ), the S-matrix generalizes to a larger acting on the space of open at a given ; unitarity S^\dagger S = I holds in this , or block-unitarity for degenerate channel subspaces, again ensuring probability across all accessible channels. Time-reversal further preserves this unitarity by relating the S-matrix to its in a basis of time-reversed states.

Time-Reversal Symmetry

In one-dimensional non-relativistic , time-reversal symmetry constrains the S-matrix for processes in systems without or absorption. The time-reversal operator T, an , acts on a as T \psi(x, t) = \psi^*(x, -t), where * denotes complex conjugation, effectively reversing the direction of time while complex-conjugating the . For a time-independent Hamiltonian H that commutes with T, satisfying T H T^{-1} = H, this invariance implies that if \psi(x, t) is a to the , then T \psi(x, t) is also a . Applying T to the asymptotic states maps an incident wave from the left (or right) to the time-reversed outgoing wave, leading to specific relations among the S-matrix elements for real-valued, time-independent potentials. In particular, the is the same from both sides, t = t'. For potentials even under spatial reversal, V(-x) = V(x), the coefficients satisfy r = -r', where r and r' are the amplitudes for from the left and right, respectively. These relations ensure the symmetry of the S-matrix, S(k) = S^T(-k), for real-valued potentials. The derivation proceeds by considering the time-reversed solution: under T, the roles of incoming and outgoing waves are interchanged, and for a real potential, the reversed process corresponds to incidence from the opposite direction, yielding the equality of transmission amplitudes and the relation between reflections. This symmetry holds as long as the potential is real and time-independent, preserving the invariance T H T^{-1} = H. Time-reversal is broken by mechanisms such as , introduced via an imaginary component in the potential (e.g., V(x) + i W(x)), or by magnetic fields, which couple through a odd under T, resulting in non-reciprocal where t \neq t' and the reflection coefficients become unrelated. In such cases, may differ directionally, and the S-matrix loses its . In relativistic , these extend through the CPT theorem, which guarantees that any local, Lorentz-invariant theory is invariant under the combined charge conjugation (C), (P), and time-reversal (T) transformations, imposing analogous symmetries on the S-matrix elements. Unitarity serves as a complementary , ensuring conservation independently of time reversal.

Scattering Coefficients

In one-dimensional quantum mechanics, the scattering coefficients are defined in terms of the elements of the S-matrix, which connects the amplitudes of incoming and outgoing plane waves on either side of the scattering potential. For an incident wave from the left, the asymptotic wave function is expressed as \psi(x) \to e^{ikx} + r e^{-ikx} for x \to -\infty and \psi(x) \to t e^{ikx} for x \to +\infty, where r is the amplitude and t is the . The transmission coefficient T = |t|^2 represents the probability that the incident particle is transmitted through the potential, while the reflection coefficient R = |r|^2 is the probability of . The unitary property of the S-matrix, arising from the conservation of , imposes the constraint T + R = 1 for processes. This relation ensures that the incident flux is fully accounted for by and alone, without loss. Experimentally, these coefficients are determined by measuring the ratios of probability currents in the asymptotic regions: the incident is proportional to |1|^2 v, the reflected to |r|^2 v, and the transmitted to |t|^2 v, where v = \hbar k / m is the , yielding T and R directly from measurements. The energy dependence of the coefficients is a key feature, with T(E) varying smoothly or oscillatory for energies above any potential barrier, reflecting effects in the process. Near specific energies, T(E) displays sharp resonant peaks, corresponding to quasi-bound states where the particle is temporarily trapped before or ; these resonances manifest as poles of the S-matrix in the complex energy plane just above the real axis. In one dimension, the basis states are plane waves, providing a direct to the partial-wave in higher dimensions, where the S-matrix is expanded in and the l=0 (s-wave) component mirrors the 1D plane-wave . The optical theorem in one dimension links the imaginary part of the forward to the total probability ( in 1D), providing an additional consistency check on unitarity.

Optical Theorem in One Dimension

The optical theorem in one dimension is the analog of the relation in higher dimensions that connects the total scattering cross section to the imaginary part of the forward . In the one-dimensional case, derived from S-matrix , the theorem states that the total cross section σ_tot = (2/k) Im f(0), where k is the wave number and f(0) is the forward in a flux-normalized form. This relation holds for short-range potentials and reflects the conservation of probability . The derivation follows directly from the unitarity of the S-matrix, S^\dagger S = I, which ensures the conservation of probability in scattering processes. For an initial state |i\rangle, the unitarity condition implies \sum_f |S_{fi}|^2 = 1, where the sum is over all final states f. In the elastic scattering limit, the forward amplitude corresponds to the diagonal S-matrix element S_{ii}, and the imaginary part arises from the discontinuity across the cut in the complex energy plane. Specifically, 2 Im S_{ii} = \sum_{f \neq i} |S_{fi}|^2, linking the imaginary part of the forward amplitude to the total probability of transition to other states, interpreted as the "cross section" comprising transmission and reflection probabilities. In one dimension, with scattering coefficients t for transmission and r for reflection satisfying |t|^2 + |r|^2 = 1, this yields the flux-normalized form where the imaginary part of the forward amplitude (related to t - 1 in asymptotic wave conventions) equals the total scattering probability normalized by incident flux. A similar flux-normalized expression involves the reflection coefficient as (2/k) Im r(-k) = |t(k)|^2, where r(-k) is the reflection amplitude for incident waves with momentum - \hbar k, connecting the imaginary part to the transmission probability under time-reversal symmetry. This admits an interpretation in terms of shadow or the paradox within one-dimensional waveguides, where the scatterer creates a forward "shadow" due to , leading to an equal to twice the geometric "size" at high energies—half from actual and half from the diffractive shadow. In 1D quantum or wave propagation, this manifests as the between incident and scattered waves reducing the transmitted , with the quantifying the via the forward imaginary part. The theorem applies specifically to elastic scattering processes without absorption, where the S-matrix remains unitary due to probability . In cases with absorption or inelastic channels, unitarity is relaxed (||S|| ≤ 1), and the relation generalizes to include absorption in the total , but the pure elastic form requires no loss mechanisms.

Examples in One-Dimensional Potentials

Transfer Matrix Relation

In one-dimensional quantum , the transfer matrix provides an alternative parameterization to the S-matrix by relating the coefficients of the wavefunction on the left and right sides of a scattering potential. Consider the asymptotic form of the wavefunction: to the left of the potential (x → -∞), ψ(x) = A e^{ikx} + B e^{-ikx}, where A is the of the incoming wave from the left and B is the reflected wave; to the right (x → +∞), ψ(x) = a e^{ikx} + b e^{-ikx}, with a the transmitted wave and b the incoming wave from the right. The M connects these coefficients via \begin{pmatrix} A \\ B \end{pmatrix} = M \begin{pmatrix} a \\ b \end{pmatrix}, where M is a 2×2 matrix. The unitarity of the S-matrix implies that det M = 1, ensuring conservation of probability current. The transfer matrix relates directly to the S-matrix, which typically expresses outgoing amplitudes in terms of incoming ones as \begin{pmatrix} B \ a \end{pmatrix} = S \begin{pmatrix} A \ b \end{pmatrix}. By inverting the transfer matrix relation and adjusting for the direction conventions (swapping incoming and outgoing labels), the S-matrix elements can be expressed in terms of M^{-1}, such as S_{11} = -M_{21}/M_{11} and S_{21} = 1/M_{11} for incidence from the left. This connection highlights the equivalence of the two formalisms while emphasizing different aspects of the scattering process. A key advantage of the transfer matrix approach lies in its multiplicative structure for composite systems, such as layered potentials, where the total M is the product of individual transfer matrices for each layer, making it ideal for numerical computations of scattering in complex, piecewise-defined potentials. In systems invariant under time reversal, such as those with real-valued potentials, the transfer matrix exhibits symmetry properties M_{11} = M_{22} and M_{12} = -M_{21}, reflecting the reciprocity of transmission and the relation between left and right reflections.

Finite Square Well

The finite square well potential serves as a key example for computing the S-matrix in one-dimensional non-relativistic , particularly illustrating bound states and resonant . Defined as V(x) = -V_0 for |x| < a and V(x) = 0 elsewhere, with V_0 > 0 and a > 0, this attractive potential supports a finite number of bound states for -V_0 < E < 0 and scattering states for E > 0. For scattering states (E > 0), the time-independent Schrödinger equation -\frac{\hbar^2}{2m} \frac{d^2 \psi}{dx^2} + V(x) \psi = E \psi is solved piecewise. In the exterior regions (|x| > a), the wave number is k = \sqrt{2mE}/\hbar, yielding plane wave solutions. Inside the well (|x| < a), the effective wave number is \kappa = \sqrt{2m(E + V_0)}/\hbar, also yielding oscillatory solutions. The general wave function for incidence from the left is e^{ikx} + r(k) e^{-ikx} & x < -a, \\ A \cos(\kappa x) + B \sin(\kappa x) & -a \leq x \leq a, \\ t(k) e^{ikx} & x > a, \end{cases} $$ where $ r(k) $ and $ t(k) $ are the reflection and transmission amplitudes, respectively.[](https://farside.ph.utexas.edu/teaching/qmech/qmech.pdf) Continuity of $ \psi(x) $ and $ \psi'(x) $ at $ x = \pm a $ provides four equations for the coefficients $ A $, $ B $, $ r(k) $, and $ t(k) $. Solving this system yields the transmission amplitude $$ t(k) = \left[ \cos(2 \kappa a) - i \frac{k^2 + \kappa^2}{2 k \kappa} \sin(2 \kappa a) \right]^{-1}, $$ up to an overall phase factor $ e^{-i 2 k a} $ accounting for the free propagation distance across the well, and the reflection amplitude $$ r(k) = i \frac{k^2 - \kappa^2}{2 k \kappa} \sin(2 \kappa a) \, t(k). $$ These expressions relate directly to the S-matrix elements for the symmetric potential, with the unitarity condition ensuring $ |t(k)|^2 + |r(k)|^2 = 1 $.[](https://ocw.mit.edu/courses/8-04-quantum-physics-i-spring-2016/79c490b96f77e1374c42d7119a940091_MIT8_04S16_LecNotes17.pdf)[](https://farside.ph.utexas.edu/teaching/qmech/qmech.pdf) The transmission probability $ T(k) = |t(k)|^2 $ exhibits perfect transmission ($ T = 1 $, $ R = 0 $) at discrete resonances where $ \sin(2 \kappa a) = 0 $, or $ 2 \kappa a = n \pi $ for integer $ n \geq n_{\min} $, corresponding to energies $ E_n = \frac{n^2 \pi^2 \hbar^2}{8 m a^2} - V_0 > 0 $. These resonances occur at the quasi-bound state energies analogous to those of the infinite square well, manifesting as sharp peaks in $ T(E) $ versus $ E $. For example, with well strength parameter $ z_0 = a \sqrt{2 m V_0}/\hbar \approx 13\pi/4 $, plots of $ T $ versus $ E/V_0 $ show multiple peaks reaching unity, separated by intervals reflecting the well's quantization, with broader low-energy behavior near $ E \approx 0 $.[](https://ocw.mit.edu/courses/8-04-quantum-physics-i-spring-2016/79c490b96f77e1374c42d7119a940091_MIT8_04S16_LecNotes17.pdf) In the limit $ V_0 \to \infty $, the finite well approaches the infinite square well, where [transmission](/page/transmission) vanishes for all $ E > 0 $ as the particle is confined, but for large finite $ V_0 $, $ T(E) $ remains small except at narrow resonant peaks near the infinite well energies $ E_n^{\infty} = \frac{n^2 \pi^2 \hbar^2}{8 m a^2} $, with peak heights of 1 and widths scaling as $ 1/V_0 $. This highlights the role of the finite depth in enabling resonant tunneling without violating unitarity.[](https://farside.ph.utexas.edu/teaching/qmech/qmech.pdf) ### Finite Square Barrier The finite square barrier potential is defined as $ V(x) = V_0 $ for $ |x| < a $ and $ V(x) = 0 $ otherwise, with $ V_0 > 0 $, representing a repulsive scatterer that particles of energy $ E $ may tunnel through if $ E < V_0 $. Inside the barrier region, for $ E < V_0 $, the time-independent Schrödinger equation yields evanescent wave solutions of the form $ e^{\pm \kappa x} $, where $ \kappa = \sqrt{2m(V_0 - E)} / \hbar $ is the decay constant, with $ m $ the particle mass and $ \hbar $ the reduced Planck's constant; these solutions describe the penetration and decay of the wave function without oscillatory propagation.[](https://openstax.org/books/university-physics-volume-3/pages/7-6-the-quantum-tunneling-of-particles-through-potential-barriers)[](https://phys.libretexts.org/Bookshelves/Quantum_Mechanics/Introductory_Quantum_Mechanics_(Fitzpatrick)/04%3A_One-Dimensional_Potentials/4.02%3A_Square_Potential_Barrier) The S-matrix elements for this potential, particularly the transmission amplitude $ t(k) $, are obtained by matching the wave functions and their derivatives at the boundaries $ x = \pm a $, where $ k = \sqrt{2mE} / \hbar $ is the wave number outside the barrier. For $ E < V_0 $, the exact transmission coefficient is $ T = |t(k)|^2 = \left[ 1 + \frac{V_0^2 \sinh^2(\kappa \cdot 2a)}{4E(V_0 - E)} \right]^{-1} $, reflecting the probability of tunneling through the full barrier width $ 2a $. At low energies where $ \kappa \cdot 2a \gg 1 $, this approximates to $ T \approx 16 \frac{E}{V_0} \left(1 - \frac{E}{V_0}\right) e^{-2\kappa \cdot 2a} $, demonstrating the exponential suppression of transmission due to the barrier's opacity.[](https://openstax.org/books/university-physics-volume-3/pages/7-6-the-quantum-tunneling-of-particles-through-potential-barriers)[](https://phys.libretexts.org/Bookshelves/Quantum_Mechanics/Introductory_Quantum_Mechanics_(Fitzpatrick)/04%3A_One-Dimensional_Potentials/4.02%3A_Square_Potential_Barrier) For $ E > V_0 $, the waves inside the barrier become oscillatory with wave number $ q = \sqrt{2m(E - V_0)} / \hbar $, and the exact transmission amplitude takes the form $ t(k) = \left[ \cos(q \cdot 2a) - i \frac{k^2 + q^2}{2 k q} \sin(q \cdot 2a) \right]^{-1} $, up to an overall [phase factor](/page/Phase_factor) $ e^{-i k \cdot 2a} $ accounting for the free propagation distance across the barrier. This leads to an oscillatory [transmission coefficient](/page/Transmission_coefficient) $ T = |t(k)|^2 = \left[ 1 + \frac{V_0^2 \sin^2(q \cdot 2a)}{4E(E - V_0)} \right]^{-1} $, where $ T $ varies between near-zero and near-unity values depending on the phase $ q \cdot 2a $. Consequently, the [reflection coefficient](/page/Reflection_coefficient) $ R = 1 - T $ approaches 1 at specific energies above the barrier where $ \sin(q \cdot 2a) = \pm 1 $, resulting from destructive interference of the transmitted waves, analogous to the Ramsauer-Townsend effect observed in atomic scattering.[](https://phys.libretexts.org/Bookshelves/Quantum_Mechanics/Introductory_Quantum_Mechanics_(Fitzpatrick)/04%3A_One-Dimensional_Potentials/4.02%3A_Square_Potential_Barrier)[](https://link.aps.org/doi/10.1103/PhysRevB.97.235418) The unitarity of the S-matrix is preserved in this case, as $ |r(k)|^2 + |t(k)|^2 = 1 $, which aligns with the one-dimensional optical theorem relating the reflection and transmission coefficients.[](https://phys.libretexts.org/Bookshelves/Quantum_Mechanics/Introductory_Quantum_Mechanics_(Fitzpatrick)/04%3A_One-Dimensional_Potentials/4.02%3A_Square_Potential_Barrier) ## S-matrix in Relativistic Quantum Field Theory ### Interaction Picture In quantum field theory, the interaction picture serves as a perturbative framework for analyzing scattering processes, enabling the systematic computation of S-matrix elements through time-dependent perturbation theory. This picture, formalized in the unification of earlier approaches by Dyson, separates the dynamics into free evolution and interaction effects, crucial for handling asymptotic behaviors in relativistic theories. The total Hamiltonian is split as $ H = H_0 + V(t) $, where $ H_0 $ is the free-field [Hamiltonian](/page/Hamiltonian) governing non-interacting particles, and $ V(t) $ represents the interaction potential, often local in [spacetime](/page/Spacetime) for relativistic invariance. Free fields under $ H_0 $ obey the Klein-Gordon equation $ (\square + m^2) \phi = 0 $ for scalars or the [Dirac equation](/page/Dirac_equation) $ (i \gamma^\mu \partial_\mu - m) \psi = 0 $ for spinors, ensuring [canonical quantization](/page/Canonical_quantization) and positive-energy solutions.[](https://www.damtp.cam.ac.uk/user/tong/qft/three.pdf) Operators in the interaction picture evolve solely under the free [Hamiltonian](/page/Hamiltonian): $ A_I(t) = e^{i H_0 t} A_S e^{-i H_0 t} $, where $ A_S $ denotes the time-independent Schrödinger-picture operator; this makes interaction-picture fields $ \phi_I(t) $ and $ \psi_I(t) $ satisfy their respective free [equations of motion](/page/Equations_of_motion). In contrast, states evolve according to the interaction term: $ i \frac{d}{dt} |\psi(t)\rangle_I = V_I(t) |\psi(t)\rangle_I $, with $ V_I(t) = e^{i H_0 t} V_S(t) e^{-i H_0 t} $. Field operators in the [Heisenberg picture](/page/Heisenberg_picture), $ A_H(t) = e^{i H t} A_S e^{-i H t} $, incorporate full dynamics but are related to interaction-picture operators via the full evolution operator.[](https://web2.ph.utexas.edu/~vadim/Classes/2008f/dyson.pdf)[](https://www.damtp.cam.ac.uk/user/tong/qft/three.pdf) The solution for state evolution is the time-ordered exponential $ U_I(t, t_0) = T \exp\left( -i \int_{t_0}^t dt' \, V_I(t') \right) $, where $ T $ ensures chronological ordering of non-commuting operators at different times, preventing ambiguities in the perturbative expansion. This operator underpins the [Dyson series](/page/Dyson_series) for higher-order corrections, directly leading to S-matrix elements as limits of $ U_I $ from early to late times.[](https://www.damtp.cam.ac.uk/user/tong/qft/three.pdf) The [interaction picture](/page/Interaction_picture) thus facilitates the treatment of asymptotic states, where free-particle descriptions hold far from interactions.[](https://web2.ph.utexas.edu/~vadim/Classes/2008f/dyson.pdf) ### Asymptotic In and Out States In [quantum field theory](/page/Quantum_field_theory), the asymptotic in-states describe the configuration of free particles approaching from the distant past, corresponding to the limit $ t \to -\infty $ in the [interaction picture](/page/Interaction_picture), where interactions are negligible and particles propagate as solutions to the free-field equations.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) Similarly, the out-states capture free particles receding to the distant future as $ t \to +\infty $, again behaving as free-particle states in the [interaction picture](/page/Interaction_picture).[](https://edu.itp.phys.ethz.ch/hs12/qft1/Chapter10.pdf) These states form the basis for defining [scattering](/page/Scattering) processes, as the S-matrix connects the in-states to the out-states, encoding the effects of interactions during the intermediate evolution.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) The in- and out-states are constructed from the respective vacua using creation operators of the free fields. For a single-particle in-state with [momentum](/page/Momentum) $ \mathbf{p} $, it is given by $ | \mathbf{p}, \text{in} \rangle = a^\dagger_{\text{in}}(\mathbf{p}) | 0, \text{in} \rangle $, where $ a^\dagger_{\text{in}}(\mathbf{p}) $ is the creation operator for the asymptotic in-field, normalized such that the states have unit norm in the relativistic sense, $ \langle \mathbf{p}, \text{in} | \mathbf{p}', \text{in} \rangle = (2\pi)^3 2 E_{\mathbf{p}} \delta^3(\mathbf{p} - \mathbf{p}') $.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) Multi-particle in-states are built by applying multiple creation operators to the in-vacuum, for example, for two bosons, $ | \mathbf{p}_1, \mathbf{p}_2; \text{in} \rangle = a^\dagger_{\text{in}}(\mathbf{p}_1) a^\dagger_{\text{in}}(\mathbf{p}_2) | 0, \text{in} \rangle $, symmetrized appropriately for identical particles.[](https://edu.itp.phys.ethz.ch/hs12/qft1/Chapter10.pdf) The out-states follow analogously, $ | \mathbf{p}, \text{out} \rangle = a^\dagger_{\text{out}}(\mathbf{p}) | 0, \text{out} \rangle $, ensuring the asymptotic vacua $ |0, \text{in}\rangle $ and $ |0, \text{out}\rangle $ coincide in the full theory.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) In the [Heisenberg picture](/page/Heisenberg_picture), the full interacting fields $ \phi_H(t, \mathbf{x}) $ approach the free asymptotic fields $ \phi_{\text{as}}(t, \mathbf{x}) $ as $ |t| \to \infty $, reflecting the dilution of interactions over large distances and times.[](https://edu.itp.phys.ethz.ch/hs12/qft1/Chapter10.pdf) This convergence justifies treating the in- and out-states as eigenstates of the free [Hamiltonian](/page/Hamiltonian) $ H_0 = \int \frac{d^3 p}{(2\pi)^3} E_{\mathbf{p}} a^\dagger(\mathbf{p}) a(\mathbf{p}) $, with the full [Hamiltonian](/page/Hamiltonian) $ H $ agreeing with $ H_0 $ on the subspace of asymptotic states.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) The in- and out-states are related through the S-matrix operator via $ | f, \text{out} \rangle = S | i, \text{in} \rangle $, where the unitary nature of $ S $ preserves probabilities, assuming the interactions are adiabatically switched on and off.[](https://edu.itp.phys.ethz.ch/hs12/qft1/Chapter10.pdf) This connection underpins crossing symmetry, where [analytic continuation](/page/Analytic_continuation) of S-matrix elements relates [scattering](/page/Scattering) processes across different kinematic channels.[](https://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf) ### Formal Definition of the S-matrix In relativistic [quantum field theory](/page/Quantum_field_theory), the S-matrix is formally defined within the [interaction picture](/page/Interaction_picture) as the [unitary operator](/page/Unitary_operator) that asymptotically maps incoming particle states to outgoing states as interactions switch on and off adiabatically. The explicit expression is S = \lim_{T \to +\infty, , t \to -\infty} U_I(T, t), where $ U_I(T, t) $ denotes the time-evolution operator in the [interaction picture](/page/Interaction_picture), given by the time-ordered exponential U_I(T, t) = \mathcal{T} \exp\left( -i \int_t^T V_I(t') , dt' \right). Here, $ V_I(t) = e^{i H_0 t} V e^{-i H_0 t} $ is the interaction part of the [Hamiltonian](/page/Hamiltonian) in the [interaction picture](/page/Interaction_picture), with $ H_0 $ the free [Hamiltonian](/page/Hamiltonian) and $ V $ the interaction potential.[](https://pages.uoregon.edu/soper/QFT/interactions.pdf) This limit assumes that interactions vanish sufficiently rapidly at early and late times to allow free-particle asymptotics.[](http://sites.science.oregonstate.edu/~stetza/COURSES/ph654/ShortBook.pdf) The S-matrix acts on asymptotic in-states as $ |f, \text{out} \rangle = S |i, \text{in} \rangle $, connecting the [Hilbert space](/page/Hilbert_space) of free incoming particles to that of free outgoing particles.[](http://sites.science.oregonstate.edu/~stetza/COURSES/ph654/ShortBook.pdf) Its unitarity follows directly from the unitarity of the full time-evolution operator, yielding $ S^\dagger S = I $, which preserves probabilities and implies [conservation](/page/Conservation) of particle number in the absence of [absorption](/page/Absorption).[](https://pages.uoregon.edu/soper/QFT/interactions.pdf) Locality of interactions further imposes the cluster [decomposition](/page/Decomposition) property on S, ensuring that distant systems evolve independently.[](https://arxiv.org/pdf/math-ph/0509047.pdf) The physical content of the S-matrix resides in its matrix elements $ \langle f, \text{out} | S | i, \text{in} \rangle $, which encode the scattering amplitudes for transitions from initial state $ |i, \text{in} \rangle $ to final state $ |f, \text{out} \rangle $.[](https://pages.uoregon.edu/soper/QFT/interactions.pdf) These amplitudes include a forward-scattering delta function plus interaction contributions. The S-matrix relates to the transition operator (T-matrix) via $ S = I + i T $, where the T-matrix isolates the non-trivial scattering effects; its elements connect to momentum-space amplitudes through the [LSZ reduction formula](/page/LSZ_reduction_formula), which extracts on-shell matrix elements from correlation functions without deriving the full procedure here.[](https://pages.uoregon.edu/soper/QFT/interactions.pdf) Causality in [quantum field theory](/page/Quantum_field_theory) implies analytic properties for the S-matrix: it is analytic in the upper half of the [complex](/page/Complex) momentum-squared plane (away from bound-state poles), arising from the time-ordered nature of propagators and the iε prescription that enforces retarded response. ## Perturbative Methods ### Time Evolution Operator In the [interaction picture](/page/Interaction_picture) of [quantum field theory](/page/Quantum_field_theory), the [time evolution](/page/Time_evolution) operator $ U(t_2, t_1) $ describes the dynamics of states under the influence of an [interaction Hamiltonian](/page/Hamiltonian) $ V_I(t) $, which is the interaction term transformed to evolve with the free [Hamiltonian](/page/Hamiltonian) $ H_0 $. This operator is formally defined as the time-ordered exponential U(t_2, t_1) = \mathcal{T} \exp\left( -i \int_{t_1}^{t_2} V_I(t) , dt \right), where $ \mathcal{T} $ denotes the time-ordering operation, ensuring that operators at later times are placed to the left of those at earlier times in the expansion. This non-perturbative expression, introduced by [Freeman Dyson](/page/Freeman_Dyson), captures the full [time evolution](/page/Time_evolution) in the [interaction picture](/page/Interaction_picture) without assuming weak coupling. The S-matrix, which encodes scattering amplitudes between asymptotic states, is obtained as the limit $ S = U(\infty, -\infty) $, assuming the interaction vanishes at early and late times. Key properties of $ U(t_2, t_1) $ include unitarity, satisfying $ U^\dagger(t_2, t_1) U(t_2, t_1) = 1 $, which follows from the Hermitian nature of the total Hamiltonian and ensures probability conservation in [scattering](/page/Scattering) processes, and the initial condition $ U(t, t) = 1 $. Dyson's formulation of time-ordering provides a rigorous way to handle the non-commutativity of operators at different times, enabling consistent perturbative calculations while defining the operator exactly. The [interaction picture](/page/Interaction_picture) simplifies the treatment of perturbations by separating the free evolution $ e^{-i H_0 t} $ from the [interaction](/page/Interaction) effects, in contrast to the full [time evolution](/page/Time_evolution) operator $ U_\text{full}(t) = e^{-i H t} $, where $ H = H_0 + V $. To justify the asymptotic limits in the S-matrix, the [adiabatic theorem](/page/Adiabatic_theorem) is invoked, ensuring that states evolve freely under $ H_0 $ as $ t \to \pm \infty $ when the [interaction](/page/Interaction) is [switched on](/page/Switched_On) and off sufficiently slowly, thus connecting initial and final free-particle states to the interacting dynamics.[](https://doi.org/10.1007/BF01035766) This framework underpins the identification of in- and out-states as those evolved by $ U $ from the distant past and future, respectively. ### Dyson Series Expansion The Dyson series provides a perturbative expansion for the S-matrix, which is defined as the time-evolution [operator](/page/Operator) $ S = U(\infty, -\infty) $ in the [interaction picture](/page/Interaction_picture) of relativistic [quantum field theory](/page/Quantum_field_theory), where the full [Hamiltonian](/page/Hamiltonian) is split into free and interaction parts. This expansion arises from iteratively solving the time-dependent [Schrödinger equation](/page/Schrödinger_equation) in the [interaction picture](/page/Interaction_picture), yielding a [power series](/page/Power_series) in the [coupling constant](/page/Coupling_constant) of the interaction potential $ V_I(t) $. The series begins with the identity [operator](/page/Operator) and incorporates higher-order corrections through multiple time integrals of the interaction, ensuring proper chronological ordering to handle non-commuting operators at different times.[](https://doi.org/10.1103/PhysRev.75.1736) The iterative form of the expansion for the time-evolution operator $ U(t, t_0) $ is given by U(t, t_0) = 1 + \sum_{n=1}^\infty U_n(t, t_0), where the first few terms are $ U_1(t, t_0) = -i \int_{t_0}^t dt_1 \, V_I(t_1) $ and $ U_2(t, t_0) = (-i)^2 \int_{t_0}^t dt_1 \int_{t_0}^{t_1} dt_2 \, T \left[ V_I(t_1) V_I(t_2) \right] $, with $ T $ denoting the time-ordering operator that arranges products of operators in decreasing order of their time arguments. In general, the $ n $-th order term is \begin{equation} U_n(t, t_0) = \frac{(-i)^n}{n!} \int_{t_0}^t dt_1 \cdots \int_{t_0}^t dt_n \, T \left[ V_I(t_1) \cdots V_I(t_n) \right], \end{equation} where the factor of $ 1/n! $ accounts for the indistinguishability of the integration variables under time-ordering, and the integrals extend over all times with the $ T $-product enforcing the correct operator sequence. This form was derived by [Dyson](/page/Dyson) to systematically compute S-matrix elements order by order in the [fine-structure constant](/page/Fine-structure_constant).[](https://doi.org/10.1103/PhysRev.75.1736) To evaluate matrix elements of these time-ordered products between free-[field](/page/Field) states, [Wick's theorem](/page/Wick's_theorem) is applied, which decomposes the $ T $-product into a sum of normal-ordered terms plus all possible full contractions (pairings of [field](/page/Field) operators) using free-[field](/page/Field) propagators. In the free-[field](/page/Field) limit, only fully contracted terms contribute to connected diagrams, facilitating the computation of [scattering](/page/Scattering) amplitudes. This theorem, essential for higher-order calculations, reduces the complexity by replacing operator products with c-number contractions. The convergence of the Dyson series is asymptotic rather than absolute, meaning it provides accurate approximations for weak couplings but diverges for strong interactions; nonetheless, it serves as the foundation for renormalization procedures in [quantum electrodynamics](/page/Quantum_electrodynamics). Diagrammatically, each term in the series corresponds to Feynman graphs, where vertices represent interactions and lines denote propagators, offering a visual and computational tool for summing infinite sets of diagrams at fixed order.[](https://doi.org/10.1103/PhysRev.75.1736) The equivalence between the nested integral form and the time-ordered [integral](/page/Integral) form is proven using Heaviside step functions (theta functions) to express the time-ordering operator explicitly. For the second-order term, for instance, T \left[ V_I(t_1) V_I(t_2) \right] = \theta(t_1 - t_2) V_I(t_1) V_I(t_2) + \theta(t_2 - t_1) V_I(t_2) V_I(t_1), and integrating over all $ t_1, t_2 \in [t_0, t] $ yields twice the nested [integral](/page/Integral) $ \int_{t_0}^t dt_1 \int_{t_0}^{t_1} dt_2 V_I(t_1) V_I(t_2) $ (up to the symmetric part if operators commute), with the $ 1/2! $ factor restoring equality; this generalizes to arbitrary $ n $ by summing over all $ n! $ permutations weighted by theta functions.[](http://scipp.ucsc.edu/~haber/ph215/TimeOrderedExp.pdf) ### Relation to Transition Operators In perturbative [quantum field theory](/page/Quantum_field_theory), the S-matrix relates asymptotic in- and out-states and is decomposed as $ S = 1 + i T $, where $ T $ is the [transition](/page/Transition) operator (or T-matrix operator) that encodes the non-trivial [scattering](/page/Scattering) interactions.[](https://web.physics.ucsb.edu/~mark/ms-qft-DRAFT.pdf) The [leading-order term](/page/Leading-order_term) in the perturbative expansion of $ T $ is $ T^{(1)} = -i \int_{-\infty}^{\infty} dt \, V_I(t) $, with $ V_I(t) $ the interaction Hamiltonian in the [interaction picture](/page/Interaction_picture), and higher-order contributions arise from connected diagrams that capture the irreducible interactions between particles, excluding disconnected [vacuum](/page/Vacuum) fluctuations.[](https://www.damtp.cam.ac.uk/user/tong/qft/qfthtml/S3.html) This decomposition ensures that the identity part of $ S $ accounts for no-scattering forward propagation, while $ iT $ isolates the transition amplitudes. The matrix elements of the T-operator are given by $ \langle f | T | i \rangle = i (2\pi)^4 \delta^4(P_f - P_i) M_{fi} $, where $ |i\rangle $ and $ |f\rangle $ are initial and final multi-particle states with total four-momenta $ P_i $ and $ P_f $, respectively, and $ M_{fi} $ is the Lorentz-invariant [scattering amplitude](/page/Scattering_amplitude).[](https://web.physics.ucsb.edu/~mark/ms-qft-DRAFT.pdf) This form arises from the extraction of connected Feynman diagrams in the perturbative expansion, which automatically excludes [vacuum](/page/Vacuum) bubbles—disconnected loops that do not contribute to physical [scattering](/page/Scattering) processes but appear in the full time-evolution [operator](/page/Operator). The connected [structure](/page/Structure) of $ T $ is enforced by normalizing the S-matrix elements against the [vacuum](/page/Vacuum) persistence amplitude, ensuring unitarity and [causality](/page/Causality) in the theory.[](https://www.damtp.cam.ac.uk/user/tong/qft/qfthtml/S3.html) Historically, the T-matrix formalism bridges old-fashioned [perturbation theory](/page/Perturbation_theory), which relies on non-covariant time-ordered expansions and [energy conservation](/page/Energy_conservation) at each [vertex](/page/Vertex), with modern covariant methods using Feynman diagrams and manifest Lorentz-invariant [propagators](/page/Propagator). In old-fashioned approaches, intermediate states are summed over positive and negative energies without immediate covariance, leading to equivalent results for S-matrix elements but complicating higher-order calculations; covariant [perturbation theory](/page/Perturbation_theory), via the [Dyson series](/page/Dyson_series), generates the T-operator orders more systematically while preserving [relativity](/page/Relativity). In contemporary applications, the T-matrix serves as the irreducible two-particle kernel in the Bethe-Salpeter equation, describing bound-state wave functions as $ \Psi(P, q) = G(q, P) T(P, q) \Psi(P, q) $, where $ G $ is the two-particle [propagator](/page/Propagator) and $ T $ sums irreducible interactions beyond ladder approximations.[](https://theses.hal.science/tel-04544120v1/file/2023TOU30275.pdf)

References

  1. [1]
    [PDF] Quantum Field Theory I Chapter 10 10 Scattering Matrix
    Let us therefore discuss the asymptotic particle states and their scattering matrix in more detail. 10.1 Asymptotic States. First we need to understand ...
  2. [2]
    S-matrices for simple quantum systems | American Journal of Physics
    Nov 1, 2023 · The rows of an S-matrix contain all possible final states, while the columns contain all possible initial states. Each matrix element equals the ...INTRODUCTION · II. QUANTUM SYSTEMS AND... · VII. SYMMETRY HELPS...
  3. [3]
    [PDF] On the Mathematical Description of Light Nuclei by the Method of ...
    1,834 (1933). 26See John A. Wheeler, “Wave Functions for Large Arguments by the Amplitude Phase. Method,” Phys. Rev. 52, 1123 (1937). 27Cf. Mott and Massey ...
  4. [4]
    The S-Matrix Is the Oracle Physicists Turn To in Times of Crisis
    May 23, 2024 · In this way, physicists have fully defined theories of gluons and other particles purely with S-matrices, avoiding Lagrangians.
  5. [5]
    From S-matrix theory to strings: Scattering data and the commitment ...
    The S -matrix approach to elementary particle theory originated with Werner Heisenberg, who proposed an S -matrix theory for quantum electrodynamics in 1943.
  6. [6]
    Physics of the analytic S-matrix - ScienceDirect.com
    Jan 10, 2024 · We will learn how different analytic features of the S-matrix are really consequences of causality, locality of interactions, unitary propagation, and so on.
  7. [7]
    [2211.15594] Explorations in the Space of S-Matrices - arXiv
    Nov 28, 2022 · S-matrix is one of the fundamental observables of the quantum theory of relativistic particles. There have been attempts to understand the ...<|control11|><|separator|>
  8. [8]
    The quantum theory of the emission and absorption of radiation
    P. A. M. Dirac, Proceedings A, 1927. On the theory of quantum mechanics. Paul Adrien Maurice Dirac, Proceedings A, 1926. The fundamental equations of quantum ...
  9. [9]
    On the Mathematical Description of Light Nuclei by the Method of ...
    John A. Wheeler, "Wave Functions for Large Arguments by the Amplitude Phase Method," Phys. Rev. 52, 1123 (1937); Mott and Massey, [14], p. 275; M. Born and P ...Missing: matrix | Show results with:matrix
  10. [10]
    Die „beobachtbaren Größen“ in der Theorie der Elementarteilchen
    Die Arbeit gliedert sich in folgende Abschnitte: I. Die beobachtbaren Größen und ihre mathematische Darstellung. a) Formulierung der Grundannahmen. b) Die ...Missing: S- | Show results with:S-
  11. [11]
    [PDF] From S-matrix theory to strings: scattering data and the commitment ...
    Mar 11, 2024 · Heisenberg's original S-matrix theory and the strong interaction S-matrix program are linked by work on dispersion relations in the 1950s.
  12. [12]
    [PDF] Causality and dispersion relations and the role of the S-matrix in the ...
    Jul 16, 2012 · The S-matrix boostrap project, which was vigorously proposed (notably by Chew), was based on the conjecture that a unique S-matrix (primarily of.Missing: 1950s | Show results with:1950s
  13. [13]
    "Bootstrap": A Scientific Idea? - Science
    "Bootstrap": A Scientific Idea?: The place of the bootstrap idea in science ... Chew, G. F., The Analytic S Matrix (1966). Google Scholar. Chew, G. F. ...
  14. [14]
    THE S-MATRIX THEORY OF STRONG INTERACTIONS
    THE S-MATRIX THEORY OF STRONG INTERACTIONS ... Many UC-authored scholarly publications are freely available on this site because of the UC's open access policies.Missing: paper | Show results with:paper
  15. [15]
    $$S$-Matrix Theory of Strong Interactions without Elementary Particles
    G. F. Chew, The S -Matrix Theory of Strong Interactions (W. A. Benjamin and Company, Inc., New York, 1961); M. Gell-Mann, in Proceedings of the Sixth Annual ...Missing: seminal | Show results with:seminal
  16. [16]
    [PDF] A retrospective look at Regge poles - arXiv
    In the second and third sections, we recall the role played by the Theory of the S-Matrix in obtaining dispersion relations in strong- interaction processes. In ...
  17. [17]
    Theory of Pion Photoproduction | Phys. Rev.
    The photoproduction of pions from nucleons is calculated in the Tamm-Dancoff approximation to the relativistic pseudoscalar theory.
  18. [18]
    [PDF] arXiv:2112.08150v1 [nucl-th] 15 Dec 2021
    Dec 15, 2021 · A theory of pion photoproduction was constructed in the 1950's. Kroll and Ruder- man [1] were the first to derive model-independent ...Missing: 1950s | Show results with:1950s
  19. [19]
    The boundary S-matrix and the AdS to CFT dictionary - hep-th - arXiv
    Mar 4, 1999 · An S-matrix analog is defined for anti-de Sitter space by constructing in and out states that asymptote to the timelike boundary.Missing: revival | Show results with:revival
  20. [20]
    [PDF] 10. Scattering Theory - DAMTP
    It turns out that the information about these bound states can be extracted from the S-matrix, which we constructed purely from knowledge of the scattering ...
  21. [21]
    A multiple scattering theory approach to solving the time-dependent ...
    Jun 29, 2015 · The solution describes these quantum phenomena as a function of time and is related to the fundamental issues (such as measuring time) of ...Missing: challenges | Show results with:challenges
  22. [22]
    S-matrix and cross sections (simplified)
    In scattering experiments in particle physics, usually two particles (target and projectile) interact, producing many particles.
  23. [23]
    [PDF] 3.3 cross-sections - Purdue Physics
    Mar 3, 2010 · §3.3 CROSS-SECTIONS. So far we have derived a procedure for calculating the transition probability am- plitudes Sfi for any process in QED.
  24. [24]
    [PDF] Dreams of a Physics Without Lagrangians: The Long S-Matrix ...
    Mar 29, 2023 · He believed that the divergence difficulties facing quantum field theory (QFT) stemmed from the use of a differential time evolution equation.
  25. [25]
    Scattering, the Scattering Matrix and Cross-Sections (Chapter 12)
    Second, we relate the S-matrix to quantities that can actually be measured in our laboratory, the so-called cross-sections.
  26. [26]
    [PDF] 50. Resonances - Particle Data Group
    Dec 1, 2023 · The analyticity principle of the S-matrix germane to quantum scattering theories, dictates that only poles and branch points can exist on the ...
  27. [27]
    The formation of a rational S matrix using phase shifts from elastic ...
    Three linear methods for constructing a rational S matrix from phase shifts are presented and compared. Applications to 1S0 neutron-proton elastic scattering ...
  28. [28]
    Scattering matrix approach to the description of quantum electron ...
    Aug 8, 2014 · We consider the scattering matrix approach to quantum electron transport in meso- and nano-conductors. This approach is an alternative to the more conventional ...
  29. [29]
    S Matrix Theory - an overview | ScienceDirect Topics
    In the BSA convention, the scattering matrix is known as the Sinclair matrix, while in the FSA convention, the scattering matrix is recognized as the Jones ...
  30. [30]
    The Scattering matrix : Properties and one-dimensional examples
    Jun 1, 2022 · The scattering matrix is used to relate particle states before and after undergoing a scattering process. It is used in quantum mechanics, scattering theory ...
  31. [31]
    None
    ### Summary of Optical Theorem in One-Dimensional Quantum Scattering
  32. [32]
    Multichannel framework for singular quantum mechanics
    The channels are linked by a fully unitary S-matrix, which governs all possible scenarios, including cases with an apparent nonunitary behavior as viewed from ...
  33. [33]
    [PDF] Time Reversal - University of California, Berkeley
    We can see already from this example that time reversal in quantum mechanics is represented by an antilinear operator, since a linear operator is unable to map ...
  34. [34]
  35. [35]
  36. [36]
    [PDF] Notes on the CPT theorem 1. Introduction In relativistic quantum field ...
    Mar 2, 2021 · In relativistic quantum field theory, the CPT theorem is a collection of results showing that the dynamics are invariant under a transformation ...
  37. [37]
    Illustrations of loosely bound and resonant states in atomic nuclei
    Feb 1, 2022 · In quantum mechanics, resonances are often defined as peaks of the transmission coefficient. They are also defined as isolated poles of the S- ...
  38. [38]
    [PDF] one-dimensional scattering in quantum mechanics - Sites do IFGW
    The division into an unperturbed wave and a scattered wave is useful for the treatment in perturbation theory where one begins with the unperturbed wave as the ...
  39. [39]
    Optical theorem in dimensions | Phys. Rev. A
    Dec 5, 2005 · The quantum-mechanical optical theorem in $N$-dimensional space is derived in a simple way from S-matrix theory.
  40. [40]
    Multiple scattering model of the quantum random Lorentz gas
    Apr 7, 2022 · ... extinction paradox ... In one dimension, it is customary to define the transmission and reflection coefficients, that we denote A T and A ...
  41. [41]
  42. [42]
    [PDF] Transfer matrix in scattering theory - arXiv
    Sep 22, 2020 · Here we discuss the transfer matrix, reflection and transmission amplitudes, Jost solutions, spectral singularities and their time-reversal,.
  43. [43]
    [PDF] 1 Transfer Matrix - Princeton University
    Time-reversal invariance and conservation of the current density impose strong conditions on the form of the transfer matrix M, regardless of the specific form ...
  44. [44]
    [PDF] Chapter 17: Resonant transmission and Ramsauer–Townsend
    Apr 26, 2016 · What one has at play is quantum-mechanical resonant scattering. The cross section going to zero means the reflection coefficient going to zero, ...
  45. [45]
    [PDF] Quantum Mechanics - Richard Fitzpatrick
    These lecture notes outline a single semester course on non-relativistic quantum mechanics which is primarily intended for upper-division undergraduate ...
  46. [46]
    [PDF] arXiv:2103.07096v3 [quant-ph] 14 Jul 2021
    Jul 14, 2021 · The transmission amplitude is given as t(k) = ¯h2k/(¯h2k − imV0 ... Ka tan(Ka/2) = κa and − Ka cot(Ka/2) = κa, (5). K = p2m(V0 + E)/¯h ...
  47. [47]
    7.6 The Quantum Tunneling of Particles through Potential Barriers
    Sep 29, 2016 · We substitute Equation 7.79 into Equation 7.71 and obtain the exact expression for the transmission coefficient for the barrier,. T ( L , E ) ...
  48. [48]
  49. [49]
    Scattering on a rectangular potential barrier in nodal-line Weyl ...
    Jun 12, 2018 · These angles depend on the barrier width. Such a behavior is related to the Ramsauer-Townsend effect for a nonrelativistic quantum particle.
  50. [50]
    [PDF] 3. Interacting Fields
    The interaction picture is a hybrid of the two. We split the Hamiltonian up ... where the unitary operator S is known as the S-matrix. (S is for ...
  51. [51]
    [PDF] Interaction Picture and Dyson Series - UT Physics
    The interaction picture separates transitions from wave-packet spreading. The Dyson series is the evolution operator for the interaction picture.Missing: paper | Show results with:paper
  52. [52]
    [PDF] Quantum Field Theory - DAMTP
    Heisenberg picture, and φ1I to denote the field in the interaction picture. ... we developed for S-matrix elements; namely Dyson's formula and Wick's theorem or,.<|control11|><|separator|>
  53. [53]
    [PDF] 1 Interactions and Green functions - University of Oregon
    We will follow the argument in Peskin and Schroeder quite closely. Before we try to find the scattering matrix, we will investigate something a little simpler, ...
  54. [54]
    [PDF] A Very Short Introduction to Quantum Field Theory
    Nov 21, 2007 · THE INTERACTION PICTURE AND THE S-MATRIX do the quantum fields ϕ(x) evolve in time, and given this information, how can we calculate the ...<|control11|><|separator|>
  55. [55]
  56. [56]
  57. [57]
  58. [58]
    [PDF] The time evolution operator as a time-ordered exponential
    It is instructive to see how this is accomplished for the n = 2 term of the series (with the general proof for the nth term in the series left to the reader).Missing: functions | Show results with:functions
  59. [59]
    [PDF] Quantum Field Theory - UCSB Physics
    Quantum field theory is the basic mathematical language that is used to describe and analyze the physics of elementary particles. The goal of this.
  60. [60]
    3 Interacting Fields‣ Quantum Field Theory by David Tong - DAMTP
    3.1 The Interaction Picture. There's a useful viewpoint in quantum mechanics to describe situations where we have small perturbations to a well-understood ...
  61. [61]
    Why do we need an old perturbation theory?
    May 6, 2014 · When time-dependent perturbation theory is used within a relativistic covariant framework, it normally bears the name of S-matrix theory. In ...Virtual particles and S-matrix - Physics Stack ExchangeOld-fashioned perturbation theory and contribution from resonancesMore results from physics.stackexchange.comMissing: methods | Show results with:methods
  62. [62]
    [PDF] Equivalence of Light-Front and Covariant Field Theory - arXiv
    Abstract. In this paper we discuss the relation between the standard covariant quantum field theory and light-front field theory. We define covariant theory ...
  63. [63]
    [PDF] Exploring new exchange-correlation kernels in the Bethe-Salpeter ...
    Apr 12, 2024 · to derive the T-matrix approximation from the conventional form of Hedin's equations. 5.2 An alternative form of Hedin's equations. Following ...<|control11|><|separator|>