Fact-checked by Grok 2 weeks ago

Quantum jump

A quantum jump, also known as a , is the discontinuous transition of an (or other quantum particle) from one energy level to another in an atom or molecule, traditionally viewed as instantaneous and without passing through intermediate states, typically accompanied by the absorption or emission of a single whose precisely matches the difference between the levels. This phenomenon, first conceptualized by in his 1913 model of the , fundamentally explains the nature of atomic spectra and marks a departure from , where such changes would occur gradually. In Bohr's atomic model, electrons orbit the in stable, quantized "stationary" states defined by the principal n (an integer starting from 1), with the of each level for given by the formula E_n = -\frac{13.6 \, \text{eV}}{n^2}, where the negative sign indicates bound states and n = 1 represents the . An in a lower state (n_i) can absorb a to "jump" to a higher state (n_f > n_i), requiring \Delta E = E_f - E_i > 0; conversely, a jump downward emits a with \Delta E = h\nu, where h is Planck's constant and \nu is the frequency of the emitted light, producing the sharp spectral lines observed in elements like . These quantized transitions resolved key puzzles, such as the stability of atoms against classical electromagnetic radiation loss, and laid the groundwork for modern quantum theory. The concept of quantum jumps extends beyond the into full and , where it describes abrupt changes in the of systems interacting with their , including during measurement. Recent studies, such as a 2019 experiment on superconducting qubits, have shown that quantum jumps can exhibit precursors and gradual dynamics in certain systems, revealing a more nuanced picture beyond pure instantaneity. In , quantum jumps manifest as random, unpredictable events in open systems, such as the intermittent fluorescence of a single or , where the system toggles between a "bright" (emitting photons) and a "dark" metastable . A landmark experimental confirmation came in 1986 with the observation of these jumps in a single trapped mercury , revealing non-classical and validating the nature of quantum transitions without predictable timing. Today, quantum jumps inform applications in , precision , and tests of , underscoring the inherently probabilistic evolution of microscopic systems.

Fundamental Concepts

Definition and Historical Context

A quantum jump refers to an abrupt, discontinuous transition of a quantum system, such as an or , between energy states in a bound system, in stark contrast to the gradual, continuous changes predicted by . This phenomenon was first conceptualized by in his 1913 atomic model, where he introduced the term "" to describe instantaneous shifts of electrons between stationary orbits, enabling the explanation of spectral lines observed in atomic emission spectra. The historical roots of quantum jumps trace back to early 20th-century efforts to resolve inconsistencies in , particularly in explaining and the . Max Planck's introduction of in 1900 provided the foundational idea that is exchanged in discrete packets, laying the groundwork for quantized transitions. extended this in 1905 by applying the quantum hypothesis to light, proposing that photons are absorbed or emitted in whole units during interactions with matter, which directly influenced Bohr's model of atomic transitions. In Bohr's 1913 model, electrons occupy fixed "stationary states" without radiating energy, and quantum jumps occur with the or of a whose \nu satisfies the condition \nu = (E_i - E_f)/h, where E_i is the initial , E_f is the final , and h is Planck's constant. This semi-classical framework marked a pivotal shift from continuous to discrete quantum behavior, though it did not yet incorporate probabilistic elements. The concept evolved further into through Werner Heisenberg's in 1925 and Erwin Schrödinger's wave mechanics in 1926, which provided a more complete theoretical description of such transitions.

Theoretical Basis in Quantum Mechanics

In , quantum jumps represent abrupt, non-adiabatic transitions between discrete eigenstates of a system's , contrasting with the continuous evolution typical of classical systems. These jumps occur when a disrupts the evolution, causing the to shift from one energy eigenstate to another. The fundamental governing equation is the time-dependent , i \hbar \frac{\partial}{\partial t} |\psi(t)\rangle = \hat{H}(t) |\psi(t)\rangle, where \hbar is the reduced Planck's constant, |\psi(t)\rangle is the in , and \hat{H}(t) is the operator, which may include both the unperturbed and external . For a time-independent unperturbed \hat{H}_0, the eigenstates |n\rangle satisfy \hat{H}_0 |n\rangle = E_n |n\rangle, forming stationary states where the probability density |\psi_n(\mathbf{r}, t)|^2 = |\phi_n(\mathbf{r})|^2 remains constant over time, with the wave function evolving only through a e^{-i E_n t / \hbar}. Such states exhibit no intrinsic transitions without , highlighting the spectrum's role in prohibiting classical-like gradual changes. Transitions arise primarily through time-dependent perturbation theory, where a small interaction \hat{H}'(t) couples the initial state |i\rangle to final states |f\rangle. Radiative jumps, involving photon emission or absorption, are dominated by the electric dipole term in the interaction Hamiltonian, \hat{H}' \propto -\mathbf{d} \cdot \mathbf{E}, with \mathbf{d} the dipole moment and \mathbf{E} the electric field. Non-radiative mechanisms include higher-order multipolar interactions or collisional de-excitation via environmental coupling, which do not involve light quanta but still induce state changes. The transition rate \Gamma_{i \to f} from initial to final state is quantified by Fermi's golden rule, \Gamma_{i \to f} = \frac{2\pi}{\hbar} \left| \langle f | \hat{H}' | i \rangle \right|^2 \rho(E_f), where \langle f | \hat{H}' | i \rangle is the perturbation matrix element and \rho(E_f) is the density of states at the final energy E_f, assuming a weak, time-harmonic perturbation and Markovian approximation. This first-order result applies when the perturbation is slow compared to the energy difference but leads to irreversible decay for continuum final states. The probabilistic nature of jumps manifests in exponential decay laws for excited states, with lifetime \tau = 1/\Gamma, yielding survival probability P(t) = e^{- \Gamma t} derived from integrating the transition rate over time. Between jumps, the state evolves unitarily under the , potentially maintaining superpositions of eigenstates; however, a is triggered by or decoherence, collapsing the superposition to a single eigenstate via non-unitary projection, consistent with the measurement postulate. This collapse ensures discrete outcomes, with no classical trajectory analog due to the quantized energy levels, though early models like Bohr's quantized orbits served as a phenomenological precursor.

Transitions in Bound Systems

Atomic Electron Transitions

In atomic systems, quantum jumps manifest as discrete transitions of valence or inner-shell electrons between bound orbitals, releasing or absorbing energy in the form of photons or through non-radiative processes. These jumps occur in multi-electron atoms where electrons occupy orbitals characterized by quantum numbers n (principal), l (azimuthal), m_l (magnetic), and m_s (spin), with interactions between electrons complicating the simple hydrogenic picture. The energy levels for hydrogen-like atoms (single-electron systems with nuclear charge Z) are given by the formula E_n = -\frac{13.6 \, Z^2}{n^2} \, \text{eV}, where n is the principal quantum number ranging from 1 to ∞, establishing the quantized framework for possible transition energies. For multi-electron atoms, this hydrogenic model is generalized using the , which approximates the many-body wavefunction as a of single-particle orbitals and self-consistently solves for densities to account for repulsion and effects, yielding more accurate energies that deviate from the simple 1/n² scaling due to screening. Radiative quantum jumps in atoms primarily involve electric dipole (E1) transitions, where an changes its orbital by Δl = ±1 and by Δm = 0, ±1, adhering to strict selection rules that determine allowed versus forbidden pathways based on and conservation. Forbidden transitions (e.g., those violating Δl = ±1) occur at much lower rates via higher-order multipoles like (M1) or electric quadrupole (E2). In , prominent examples include the in the (transitions from n > 1 to n = 1, with wavelengths 91–122 nm) and the in the visible (n > 2 to n = 2, wavelengths 364–656 nm), each corresponding to specific jumps that produce characteristic spectral lines observed in emission or absorption. The spectral lines arising from these atomic transitions exhibit natural linewidth due to the finite lifetime τ of the , governed by the , with the Lorentzian given by Δν = 1/(2πτ). This broadening reflects the of the population, and the occurrence of individual quantum jumps follows statistics, indicating random, probabilistic timing with no intermediate states during the transition. Transition rates for allowed radiative jumps in atomic systems typically span nanoseconds to microseconds, depending on the and coupling to the , as calculated from time-dependent in . Non-radiative processes, such as decay in inner-shell vacancies, compete with ; here, an outer-shell fills the core hole while simultaneously ejecting another , dissipating energy without emission and dominating for high-Z atoms or deep shells where rates exceed 10^{15} s^{-1}.

Molecular Electronic and Vibrational Transitions

In molecules, quantum jumps involve electronic transitions between different surfaces (PES), where the electronic state changes while the nuclei initially remain fixed due to the rapid timescale of the process compared to nuclear motion. These transitions are described by the Born-Oppenheimer approximation, separating electronic and nuclear , with PES representing the effective potential for nuclear motion in each electronic state. The probability and intensity of such jumps depend on the geometry of the PES, particularly the displacement between equilibrium nuclear configurations in the and excited states. The Franck-Condon principle explains the predominantly vertical nature of these electronic transitions, as the electron jump occurs much faster than rearrangement, leading to an initial vibrational state in the upper PES determined by the overlap of vibrational wavefunctions from the lower and upper surfaces. The transition intensity is proportional to the square of the Franck-Condon overlap integral between these wavefunctions, which favors transitions to vibrational levels where the probability distributions align closely. This principle, originally formulated for diatomic molecules, applies broadly to polyatomics and accounts for the distribution of vibrational excitation following the electronic jump. Vibronic transitions couple these electronic jumps with changes in vibrational quantum numbers, resulting in spectra that reveal both electronic and vibrational structure. In the harmonic approximation, vibrational energy levels on a PES are quantized as E_v = \hbar \omega \left( v + \frac{1}{2} \right), where v = 0, 1, 2, \dots is the vibrational and \omega is the vibrational derived from the force constant and . For accurate description at higher energies, anharmonic corrections—such as those from the —are included to account for level convergence and limits, modifying the energy spacing and enabling transitions. Rotational motion further complicates these jumps, introducing fine structure in the spectra due to the coupling of electronic-vibrational changes with rotational levels. Under the rigid rotor approximation, rotational energies are given by E_J = \frac{J(J+1) \hbar^2}{2I}, where J = 0, 1, 2, \dots is the rotational quantum number and I is the moment of inertia. Electric dipole selection rules for rotational transitions in electronic spectra typically allow \Delta J = \pm 1, producing P-branch lines (\Delta J = -1, lower frequency) and R-branch lines (\Delta J = +1, higher frequency) flanking the band origin, with Q-branch (\Delta J = 0) forbidden in many cases due to symmetry. Compared to atomic electron transitions, molecular quantum jumps exhibit longer excited-state lifetimes, typically ranging from picoseconds to milliseconds, owing to the denser density of vibronic states that facilitates intramolecular relaxation pathways. In unbound or crossing PES scenarios, these jumps can lead to , where the molecule breaks apart directly, or predissociation, involving nonradiative to a repulsive state that broadens spectral lines and shortens lifetimes. A representative example is in organic dyes like squaraines, where vibronic between electronic excited states and low-frequency vibrational modes produces structured emission spectra, enabling applications in probing .

Experimental Realizations

Early Spectroscopic Observations

The observation of discrete spectral lines in the emission and absorption spectra of atoms provided the earliest indirect evidence for quantum jumps, as these lines implied abrupt transitions between discrete energy levels rather than continuous changes predicted by . In 1814, systematically mapped hundreds of dark absorption lines in the solar spectrum using a prism spectroscope, revealing fixed positions that could not be explained by continuous atmospheric or instrumental effects; these "" were later understood as resulting from atomic absorption processes involving quantized energy shifts in stellar atmospheres. Although Fraunhofer did not interpret them in terms of atomic structure, his precise measurements established as a tool for probing atomic phenomena. A pivotal example of discovery came during the total solar eclipse of August 18, 1868, when French astronomer and English spectroscopist independently observed a bright yellow emission line at 587.6 nm in the Sun's , distinct from known or other terrestrial elements; Lockyer termed the hypothetical element responsible "" based on this unidentified line. This line, later confirmed as neutral helium's principal series, highlighted how solar spectra could reveal atomic transitions invisible in laboratory conditions, reinforcing the discreteness of spectral features. was not isolated on until 1895, underscoring the extraterrestrial origin of such observations. By the late 19th century, empirical patterns emerged in laboratory spectra of . In 1885, derived a formula fitting the wavelengths of visible emission lines observed in gas discharges, expressed as \lambda = 364.56 \times \frac{n^2}{n^2 - 4} nm for integer n > 2, accurately predicting lines like H\alpha at 656.3 nm and H\beta at 486.1 nm without theoretical justification. This empirical relation suggested underlying discrete states, though Balmer viewed it as a mathematical curiosity. In 1888, Swedish physicist generalized Balmer's formula to all series, proposing \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R \approx 1.097 \times 10^7 m^{-1} is the and n_1 < n_2 are integers; this unified visible, ultraviolet, and later infrared lines, implying a common mechanism of discrete level transitions. Further series confirmed this pattern. In 1908, German physicist identified the infrared series of hydrogen using a Rowland grating spectrometer on electrically excited gas, measuring lines from n_2 \geq 4 to n_1 = 3, such as the principal line at 1875.1 nm, which fit Rydberg's formula and extended the evidence for quantized structure beyond the visible range. These observations relied on ensemble averages from many atoms, inferring jumps from line positions and intensities rather than single events. In the 1890s, Albert A. Michelson employed his interferometer to resolve the fine widths of spectral lines, such as sodium D-lines at ~10^{-5} cm^{-1}, demonstrating their intrinsic narrowness consistent with sharp energy level differences, though Doppler and pressure broadening complicated direct quantum interpretations. Theoretical synthesis arrived in 1913 with Niels Bohr's model of the hydrogen atom, which posited stationary electron orbits with quantized angular momentum L = n \hbar (where \hbar = h/2\pi) and radiative transitions via instantaneous "jumps" between levels, emitting photons of frequency \nu = (E_{n_2} - E_{n_1})/h; this quantitatively reproduced Balmer and Rydberg formulas, with energy levels E_n = -13.6 eV/n^2, marking the first explicit link between discrete spectra and quantum transitions. By the 1930s, refinements incorporated relativity. In 1928, Paul Dirac's relativistic quantum equation for the electron yielded exact fine structure splittings in hydrogen, such as the 0.365 cm^{-1} separation in the Balmer \alpha line, arising from spin-orbit coupling and Darwin term corrections, providing a more precise framework for spectral details while preserving the jump paradigm. These early spectroscopic insights, though indirect, laid the groundwork for understanding quantum jumps as fundamental to atomic stability and radiation.

Modern Quantum Optics and Ion Trap Experiments

In modern quantum optics, cavity quantum electrodynamics (cQED) experiments with single atoms confined in high-Q optical cavities have enabled precise observation and manipulation of quantum jumps through strong coherent coupling between the atom and the quantized cavity field. These setups typically involve neutral atoms, such as rubidium, injected into a high-finesse Fabry-Pérot cavity where the atom-cavity interaction exceeds decay rates, realizing the strong coupling regime. Vacuum Rabi oscillations occur as the atom and cavity exchange a single photon back and forth at the vacuum Rabi frequency g, manifesting as oscillatory energy transfer observable in the cavity transmission spectrum. The Jaynes-Cummings model theoretically describes this dynamics, predicting atom-photon entanglement and the emergence of dressed states that lead to vacuum Rabi splitting of $2g in the spectrum for a single atom. Seminal experiments in 2004 demonstrated this splitting for one trapped cesium atom in an optical cavity, with g/2\pi \approx 34 MHz, directly visualizing the quantum jump between atomic and photonic excitations. Ion trap experiments complement cQED by providing long-lived, isolated systems for detecting individual quantum jumps via fluorescence monitoring, particularly in alkaline-earth ions like ^{40}\mathrm{Ca}^+ and ^{88}\mathrm{Sr}^+. These ions are confined in linear Paul traps and laser-cooled to millikelvin temperatures (around 0.1 mK) using Doppler and sideband cooling techniques, achieving motional ground states essential for high-fidelity control. Quantum jumps are observed through optical shelving: the ion cycles between a bright ground state and an excited state under resonant laser excitation (e.g., 397 nm for \mathrm{Ca}^+), but occasionally transitions to a long-lived metastable "dark" state (e.g., D_{5/2} at 729 nm for \mathrm{Ca}^+), halting fluorescence for durations up to seconds and producing characteristic dark periods. Detection of these jumps relies on photon counting at the cycling transition wavelength, with modern setups achieving >99% fidelity in distinguishing bright and dark states. The pioneering 1986 experiment with a single \mathrm{Ba}^+ ion in an rf trap first resolved these jumps as abrupt interruptions in resonance fluorescence at 493 nm, confirming the metastable $5D_{5/2} state involvement. The statistics of these quantum jumps follow a Poisson process, with exponentially distributed waiting times between transitions governed by the metastable state's lifetime \tau, yielding a probability P(t) = e^{-t/\tau} for no jump in time t. In \mathrm{Sr}^+ ions, shelving to the $4D_{5/2} state has been probed at 674 nm detection, revealing dark periods consistent with \tau \approx 390 ms and Poissonian jump rates. Post-2010 advancements in linear Paul traps have integrated quantum jumps for heralded state preparation, enabling protocols like ion-photon entanglement for quantum teleportation with >90% fidelity, where jump detection non-destructively projects the ion into a known state. Multiple-ion experiments further refine lifetime measurements, as in 2018 studies with up to three ^{138}\mathrm{Ba}^+ ions, where correlated jumps allow sub-Poissonian statistics analysis and precision spectroscopy of the D_{5/2} state.

Applications and Extensions

In Spectroscopy and Laser Technology

In , quantum jumps between discrete s of atoms and molecules produce characteristic and spectra that map the underlying quantum structure. occurs when a excites an from a lower to a higher energy state, corresponding to specific wavelengths that reveal spacings, while follows the reverse jump, releasing s at those same wavelengths. These spectral lines enable precise identification of atomic and molecular in and astrophysical environments. Raman spectroscopy exploits virtual quantum jumps, where incident light scatters inelastically via intermediate virtual states without populating real excited levels, providing insights into vibrational and rotational transitions. Time-resolved spectroscopy further probes the ultrafast dynamics of these jumps, capturing transient states in molecular systems with precision to study relaxation and . In astrophysical spectroscopy, such transitions allow remote identification of elements in stars and nebulae by analyzing lines from quantum jumps in ionized gases. Lasers rely on quantum jumps for amplification, achieving by optically pumping atoms or molecules to excited states, where —triggered by incoming —induces synchronized downward jumps, producing coherent light. Four-level laser systems, such as Nd:YAG, enhance efficiency by separating the pumping, upper lasing, lower lasing, and ground states, minimizing reabsorption losses during jumps. The rates of these processes are quantified by : A_{21} for (probability per unit time) and B_{21} = B_{12} for and (proportional to photon density). Linewidth broadening from quantum jumps and environmental interactions affects laser coherence, limiting spectral purity; narrower linewidths, below 1 MHz in stabilized systems, are essential for high-resolution applications. In (NMR) and electron spin resonance (ESR) spectroscopy, quantum jumps between spin states under enable probing of molecular environments, with single-spin detection revealing jump statistics over minute timescales. Tunable lasers like and Ti:sapphire systems exploit molecular quantum jumps across broad spectral ranges (e.g., 650–1100 nm for Ti:sapphire), enabling selection for precise spectroscopic interrogation. These technologies underpin advancements in precision measurement and , from analysis to stellar composition studies.

In Quantum Computing and Information Science

In superconducting qubits, quantum jumps manifest as abrupt transitions from the to the , often modeled as bit-flip or errors due to relaxation processes. These jumps are and governed by the qubit's energy relaxation time T1, which typically ranges from microseconds to over 1 millisecond in state-of-the-art devices as of 2025, enabling coherent operations on timescales shorter than T1. For instance, in (cQED) architectures, such jumps have been directly observed in artificial atoms, where the lifetime determines the jump rate. This relaxation is a primary source of decoherence in processing, limiting fidelities and necessitating robust mitigation strategies. In cQED systems, quantum jumps are monitored through dispersive readout, where the state shifts the resonance frequency of a coupled , allowing non-demolition of the without directly exciting it. This technique uses quantum trajectory theory to describe how populations in the act as an effective heat bath, inducing jumps and that reduce measurement signal-to-noise ratios. Recent advancements in low-noise amplifiers have enabled detection of these jumps with , paving the way for in scalable quantum processors. Building on ion trap precursors, such monitoring has been adapted to solid-state platforms for continuous weak . Quantum jump detection plays a central role in error correction protocols, where syndrome measurements identify errors without collapsing the logical qubit state. In hardware-efficient schemes, such as those using concatenated bosonic codes with ancillas, jumps are detected via controlled-X gates that map phase-flip errors to measurable s, achieving logical error rates below 2% per cycle in distance-5 codes. The further suppresses unwanted jumps by applying frequent, weak observations that "freeze" the qubit evolution, increasing survival probabilities in noisy intermediate-scale quantum (NISQ) devices and extending for fault-tolerant operations. These methods leverage jump statistics to implement fault-tolerant gates, as demonstrated in experiments where repeated extractions correct errors with exponential suppression of bit-flips. Extensions of quantum jumps appear in quantum repeaters, where atomic spontaneous emissions—manifesting as jumps—facilitate entanglement distribution over long distances. In cavity-based nodes, atoms are entangled with emitted photons, which are sent to remote users; successful Bell-state measurements on the atoms confirm shared entanglement, doubling the effective channel length and enabling secure key rates beyond classical limits. The underscores limitations in these protocols, as it prohibits perfect copying of pre-jump quantum states, ensuring that jump-induced entanglement cannot be duplicated without introducing errors and preserving quantum security.

Interpretations Beyond Physics

The metaphorical use of "quantum jump," often interchangeably termed "quantum leap," emerged prominently in the 1970s and 1980s amid a surge of public interest in quantum physics, fueled by media portrayals of scientific breakthroughs and the broader "quantum hype" in popular science literature. This figurative adoption shifted the term from its precise scientific meaning—a discrete transition between atomic energy levels—to denote abrupt, transformative changes in various domains. The 1989–1993 NBC television series Quantum Leap, which depicted a physicist time-traveling by "leaping" into different historical figures to alter events, played a pivotal role in embedding the phrase in mainstream culture, amassing a dedicated audience and syndication that reinforced its association with dramatic, non-literal shifts. The series was revived in 2022 by NBC, running for two seasons until its cancellation in 2024, with episodes beginning to stream on Netflix in August 2025, further perpetuating the metaphorical usage.) In business and marketing contexts, the term became jargon for "," exemplified by ' 1992 speech at MIT's Sloan School of Management, where he described technological convergences as creating "something that's a forward." Tech advertisements and corporate strategies frequently invoke it to promise revolutionary advancements, such as in software or , emphasizing scale and speed over incremental progress. Self-help literature has similarly co-opted the for personal growth, portraying "quantum leaps" as sudden mindset shifts or success accelerations, as in books like YOU²: A High Velocity Formula for Multiplying Your Personal Effectiveness in Quantum Leaps by Price Pritchett, which advises readers to bypass gradual change through bold, intuitive actions. Physicists in the , including through outreach efforts like those highlighted in Alan Sokal's of scientific misuse in cultural discourse, warned that such applications dilute the term's rigor, fostering pseudoscientific interpretations that conflate atomic phenomena with everyday transformations. The cultural impact of this is evident in its permeation of and media, with phrases like "quantum leap forward" appearing in memes, motivational speeches, and casual discourse to signify major progress, often detached from physics. Fan-generated content around , including memes riffing on the show's time-jumping premise, has sustained its visibility online, contributing to a broader pop . Educational on misconceptions has documented public confusion about quantum concepts, often portraying them as mystical rather than probabilistic events. Post-2000s, media efforts have aimed to clarify the distinction, distinguishing the scientific —an abrupt but observable transition—from its exaggerated metaphorical counterpart. Outlets like have explained how popular usage, amplified by the TV series and trends, promotes unverified ideas like "quantum jumping" into alternate realities via , lacking empirical support. This evolution reflects growing scientific communication initiatives to combat dilution, emphasizing the term's origins in while acknowledging its linguistic entrenchment.

Philosophical and Interpretational Debates

The in centers on the apparent discontinuity of quantum jumps, particularly in the context of wavefunction collapse during measurement. In the , developed primarily by and around 1927, quantum jumps are understood as instantaneous collapses of the wavefunction from a superposition of states to a single definite outcome upon , bridging the quantum and classical realms but introducing a non-unitary process that lacks a clear physical mechanism. This view posits that the act of measurement induces the jump, rendering the process inherently probabilistic and observer-dependent, though it does not specify the precise boundary between quantum and classical systems. In contrast, the decoherence program, advanced by Wojciech Zurek in the 1980s and 1990s, explains the appearance of quantum jumps without invoking a true of the wavefunction. Decoherence arises from the inevitable interaction of a quantum system with its environment, which rapidly suppresses interference between superposition branches, leading to an effective classical-like behavior and the illusion of definite outcomes. Zurek's framework of einselection emphasizes how environmentally induced decoherence selects preferred states, resolving the by showing that jumps are emergent phenomena rather than fundamental discontinuities, thus preserving the unitary evolution of the full . Alternative interpretations further challenge the reality of quantum jumps. In the proposed by Hugh Everett in 1957, there is no collapse; instead, quantum jumps manifest as branching of the universal wavefunction into parallel worlds, each realizing a different outcome of the measurement without violating unitarity. Similarly, , formulated by in 1952, describes particles following continuous, deterministic trajectories guided by the wavefunction, making quantum jumps appear discontinuous only from the perspective of the evolving , while the underlying motion remains smooth and non-local. These interpretations fuel ongoing debates about the ontological status of discontinuity in quantum jumps. , in his 1927 critiques during the , objected to the Copenhagen view's acceptance of fundamental discontinuities, arguing that they violated the continuity of physical processes and introduced an unacceptable element of randomness into nature's laws. More recent discussions highlight potential time-symmetry violations in quantum jumps, as the irreversible nature of measurement outcomes contrasts with the time-reversible unitary evolution of isolated systems, suggesting a preferred direction for quantum processes akin to thermodynamic arrows of time. In the 2020s, (RQM), originally proposed by in 1996 but extended in recent works, reframes jumps as relative to specific observers, eliminating absolute collapse by treating quantum states as interactions between systems, which has implications for resolving foundational puzzles like the preferred basis problem. The philosophical implications of quantum jumps extend to challenges against classical . By introducing irreducible probabilities or branching realities, jumps undermine the Laplacian of a fully predictable , suggesting that may hold only at the level of the full wavefunction in some interpretations, while appearing indeterministic in observed outcomes. This tension also connects to broader issues in , such as the , where Hawking radiation's apparent loss of during resembles a non-unitary jump, prompting proposals that unitary evolution and decoherence-like mechanisms preserve across horizons.

References

  1. [1]
    Quantum Theory - FSU Chemistry & Biochemistry
    The Bohr model consists of four principles: 1) Electrons assume only certain orbits around the nucleus. These orbits are stable and called "stationary" orbits.
  2. [2]
    Nats 101, S01-16
    The electron cannot exist anywhere between the two radii. Such a process, that of moving from one energy level to another is called a quantum leap, or quantum ...
  3. [3]
    Quantum Physics
    When an atom undergoes a discontinuous transition, or quantum jump, its energy changes abruptly by a sharply defined amount, and a photon of that energy is ...
  4. [4]
    [PDF] Observation of Quantum Jumps in a Single Atom
    electron moves to the metastable state, the fluores- cence from the strongly ... Simplified optical energy-level diagram for HgII. Work of the U. S. ...
  5. [5]
  6. [6]
    I. On the constitution of atoms and molecules - Taylor & Francis Online
    On the constitution of atoms and molecules. N. Bohr Dr. phil. Copenhagen. Pages 1-25 | Published online: 08 Apr 2009.<|control11|><|separator|>
  7. [7]
    [PDF] Einstein's Proposal of the Photon Concept-a Translation
    Of the trio of famous papers that Albert Einstein sent to the Annalen der Physik in 1905 only the paper proposing the photon concept has been unavailable in ...
  8. [8]
    [PDF] Philosophical Magazine Series 6 I. On the constitution of atoms and ...
    It will further be shown that from this theory we are led to a theory of the constitution of molecules. In the present first part of the paper the mechanism of".
  9. [9]
    [PDF] Quantum-theoretical re-interpretation of kinematic and mechanical ...
    HEISENBERG. The present paper seeks to establish a basis for theoretical quantum mechanics founded exclusively upon relationships between quantities which in ...Missing: matrix | Show results with:matrix
  10. [10]
    [PDF] 1926-Schrodinger.pdf
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The analogy is to ...
  11. [11]
    [PDF] V. Time Dependence A. Energy Eigenstates Are Stationary States
    Note that in quantum mechanics time, t , plays a fundamentally different role than the position, qˆ , or momentum, pˆ . The latter two are represented by ...<|control11|><|separator|>
  12. [12]
    Fermi's golden rule: its derivation and breakdown by an ideal model
    Apr 23, 2016 · Abstract page for arXiv paper 1604.06916: Fermi's golden rule: its derivation and breakdown by an ideal model.
  13. [13]
    Elucidating Fermi's golden rule via bound-to-bound transitions in a ...
    Mar 1, 2013 · In this paper, we use two methods to resolve this mathematical dilemma, and in doing so provide a concrete demonstration of the consistency ...Missing: original | Show results with:original
  14. [14]
    Aspects of Time-Dependent Perturbation Theory | Rev. Mod. Phys.
    Jul 1, 1972 · The Dirac variation-of-constants method has long provided a basis for perturbative solution of the time-dependent Schrödinger equation.
  15. [15]
    [PDF] Elements of Atomic Structure in Multi-Electron Atoms
    When we obtain the Hartree-Fock estimate for the energy, the result will depend only on the quantum numbers (nℓ), and not on the magnetic quantum numbers (mℓ,ms) ...
  16. [16]
    The hydrogen atom
    The energy levels scale with Z2, i.e. En = -Z2*13.6 eV/n2. It takes more energy to remove an electron from the nucleus, because the attractive force that must ...Missing: Z²/ n²
  17. [17]
    [PDF] Hartree-Fock Method - Rutgers Physics
    ... Fock method we get for the ground state energy of He atom -2.85516 Hartree while the exact ground state energy is -2.86166 Hartree. The equlibrium distance ...
  18. [18]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  19. [19]
    Spectroscopy - The Hydrogen Atom - Harper College
    The Balmer series releases light in the visible region of the electromagnetic spectrum. The Lyman series, with longer arrows, requires the higher energy of the ...
  20. [20]
    Transitions - Hydrogen Energy Levels - NAAP - UNL Astronomy
    Lyman, Balmer ... Balmer were named the Balmer series. It was later understood that the Balmer lines are created by energy transitions in the Hydrogen atom.
  21. [21]
    Primary Atomic Frequency Standards at NIST - PMC - PubMed Central
    Ignoring noise for the moment, the linewidth Δνa of the atomic ... 5), which has a natural linewidth of 1/(2πτD) = 1.7 Hz (where τD = 90 ms is ...
  22. [22]
    Electron Spin Resonance at the Single Atom Level | NIST
    Aug 20, 2025 · However, this process is typically very fast, in the order of nanoseconds to microseconds.
  23. [23]
    Auger processes in the 21st century - PMC - PubMed Central
    Auger processes are non-radiative processes where an atom emits Auger electrons after an inner shell electron is removed, often with low energies.
  24. [24]
  25. [25]
    [PDF] in 1814, Joseph von Fraunhofer discovered dark lines in the solar ...
    In 1814, Fraunhofer used a glass prism to split sunlight into its rainbow colors. To his amazement, the resulting color fan contained about six hundred dark ...
  26. [26]
    The High-Flying, Death-Defying Discovery of Helium
    Mar 23, 2021 · The English astronomer Norman Lockyer spotted that same yellow line independently in October 1868. This was months after the eclipse, but just ...
  27. [27]
  28. [28]
    Observation of the Vacuum Rabi Spectrum for One Trapped Atom
    Dec 3, 2004 · The vacuum Rabi splitting for single intracavity atoms has been observed with atomic beams in both the optical [11–13] and microwave regimes [14]Abstract · Article Text
  29. [29]
    Shelved optical electron amplifier: Observation of quantum jumps
    Jun 30, 1986 · We demonstrate here the direct observation of quantum jumps between the 6 2 S 1 / 2 state and the 5 2 D 5 / 2 state of an individual laser-cooled B a + ions.
  30. [30]
    Lifetime of the level of from quantum jumps with single and multiple ...
    Mar 14, 2018 · For the experiments, we trap Ba + 138 ions in a Paul trap. The electrode surfaces facing the ions are hyperbolically shaped and made from oxygen ...Missing: post- | Show results with:post-
  31. [31]
    Spectroscopy 101 – How Absorption and Emission Spectra Work
    Each of the absorption lines corresponds to a specific electron jump. The shortest wavelength/highest energy light (violet 410 nm) causes the electron to jump ...
  32. [32]
    Absorption, Emission, and Spectroscopy
    Any transition from a higher to a lower energy level will result in the emission of a photon. But since there are only certain, well-defined, quantized energy ...
  33. [33]
    [PDF] Atomic Astrophysics and Spectroscopy
    Spectroscopy is the science of light–matter interaction. It is one of the most powerful scientific tools for study- ing nature. Spectroscopy is dependent on ...
  34. [34]
    Time Resolved Raman Scattering of Molecules - ACS Publications
    Oct 24, 2022 · The scattered photon emitted in the Raman process causes the decay from a virtual state with energy determined by the pulse frequency to a ...Introduction · Methods · Results and Discussion · Conclusion
  35. [35]
    Capturing Ultrafast Quantum Dynamics with Femtosecond and ...
    In this Perspective, we provide an overview of the progress in applying the recently developed technique of femtosecond to attosecond time-resolved soft X-ray ...
  36. [36]
    Theoretical Atomic Astrophysics - Uppsala University
    Aug 9, 2024 · Atomic astrophysics can be viewed as the bridge between macroscopic astronomy and the microscopic world at the atomic and molecular level. To ...
  37. [37]
    15.2: The Dynamics of Transitions can be Modeled by Rate Equations
    Oct 25, 2025 · This page discusses Einstein's three processes of atomic spectral line formation: spontaneous emission, stimulated emission, and absorption.
  38. [38]
    Four-level and Three-level Laser Gain Media - RP Photonics
    All lasers based on neodymium-doped laser gain media, except those operated on the ground-state transition around 0.9–0.95 μm, are four-level lasers. Neodymium ...What are Four-level and Three... · Three-Level Systems · Four-Level Systems
  39. [39]
    The quantum theory of the emission and absorption of radiation
    The paper presents a non-relativistic quantum theory of radiation emission and absorption, using energy and phase of radiation components as dynamical ...<|separator|>
  40. [40]
    Linewidth – bandwidth, laser, spectral, line width, measurement ...
    The fundamental quantum noise limit for the laser linewidth is essentially resulting from noise of the amplification process and vacuum noise entering the laser ...
  41. [41]
    Coherent Control of a Single Nuclear Spin Qubit | Phys. Rev. Lett.
    Dec 9, 2014 · Quantum jumps of the nuclear spin occur on minute-long time scales ... ESR pulse precedes the NMR pulse, as sketched in Fig. 2(a) . The ...
  42. [42]
    Titanium–sapphire Lasers - RP Photonics
    titanium–sapphire lasers are lasers based on a Ti:sapphire gain medium. They can be widely wavelength-tunable and/or generate femtosecond light pulses.Missing: jumps | Show results with:jumps
  43. [43]
    Observation of quantum jumps in a superconducting artificial atom
    Sep 15, 2010 · Quantum jumps are abrupt transitions from an excited to ground state, observed in a superconducting artificial atom (qubit) coupled to a ...
  44. [44]
    [PDF] Fluctuations of Energy-Relaxation Times in Superconducting Qubits
    In this report, we spectrally and temporally resolve T 1 of frequency-tunable Xmon transmon qubits [1-3,17,30]. The spectral data is used to identify defects ...
  45. [45]
    Dispersive regime of circuit QED: photon-dependent qubit ... - arXiv
    Oct 8, 2008 · Using quantum trajectory theory, we show that this heat bath can induce quantum jumps in the qubit state and reduce the achievable signal-to- ...
  46. [46]
    To catch and reverse a quantum jump mid-flight - Nature
    Jun 3, 2019 · Experimentally, quantum jumps were first observed in an atomic ion driven by a weak deterministic force while under strong continuous energy ...Missing: computing | Show results with:computing
  47. [47]
    Hardware-efficient quantum error correction via ... - Nature
    Feb 26, 2025 · The repetition code, using ancilla transmons for syndrome measurement, corrects cat qubit phase flips. We study the performance and scaling of ...
  48. [48]
    Observation of the Quantum Zeno Effect on a NISQ Device
    ### Summary of Quantum Zeno Effect on a NISQ Device
  49. [49]
    Fault-tolerant detection of a quantum error - Science
    We demonstrate a fault-tolerant error-detection scheme that suppresses spreading of ancilla errors by a factor of 5, while maintaining the assignment fidelity.
  50. [50]
    A Quantum Repeater Node Demonstrating Unconditionally Secure Key Distribution
    ### Summary of Quantum Repeater Using Atomic Entanglement with Photon Emission
  51. [51]
    The difference between a quantum leap and a quantum jump - SYFY
    Sep 16, 2022 · In scientific circles, a quantum jump is just another name for a quantum leap. Two terms for describing the same weird behavior. In the public ...Missing: influence | Show results with:influence
  52. [52]
    Sokal Hoax - 'Transgressing the Boundaries: Towards a ...
    In 1996, physicist Alan Sokal performed an experiment after becoming frustrated with what he saw as the abuses of science and mathematics by writers and ...
  53. [53]
    Transcript: Steve Jobs Speaks at MIT Sloan Distinguished Speaker ...
    Jun 21, 2019 · And what I mean by that is, enough technology, usually from fairly diverse places, comes together, and makes something that's a quantum leap ...Missing: 1980s | Show results with:1980s
  54. [54]
    How To Take A Quantum Leap In Your Business - Forbes
    Aug 5, 2022 · To take a quantum leap in your business, you must abandon the process on which you currently depend. Moderate growth can become a safety net.
  55. [55]
    Wanting to Quantum Leap?. These books will expand you, FAST.
    Jul 11, 2023 · My top 3 favorites: YOU2: A High Velocity Formula for Multiplying Your Personal Effectiveness in Quantum Leaps by Price Pritchett
  56. [56]
    10 Hilarious Quantum Leap Memes - Screen Rant
    Mar 31, 2020 · The series Quantum Leap was a cultural phenomenon long before memes or even the internet existed. Each week, families gathered around the ...
  57. [57]
    [PDF] Quantum Physics Education Research over the Last Two Decades
    Nov 1, 2021 · Abstract: Quantum physics is an essential field of science education research, which reflects the high relevance of research on quantum ...
  58. [58]
    Four common misconceptions about quantum physics
    Nov 8, 2022 · This view is widely held in public. Quantum physics is supposedly impossible to understand, including by physicists. But from a 21st-century ...Missing: 2010s | Show results with:2010s
  59. [59]
    Copenhagen Interpretation of Quantum Mechanics
    May 3, 2002 · The Copenhagen interpretation was the first general attempt to understand the world of atoms as this is represented by quantum mechanics.
  60. [60]
    Models of wave-function collapse, underlying theories, and ...
    Apr 2, 2013 · The Copenhagen interpretation does not solve the quantum measurement problem, nor does it explain the absence of macroscopic superpositions. The ...
  61. [61]
    Many-Worlds Interpretation of Quantum Mechanics
    Mar 24, 2002 · The Many-Worlds Interpretation (MWI) of quantum mechanics holds that there are many worlds which exist in parallel at the same space and time as our own.
  62. [62]
    Bohmian Mechanics - Stanford Encyclopedia of Philosophy
    Oct 26, 2001 · Bohmian mechanics is a quantum theory where particles' positions are determined by a wave function and a 'guiding equation', completing the ...
  63. [63]
    [PDF] The Bohr and Einstein debate: Copenhagen Interpretation challenged
    Jul 28, 2018 · Einstein's criticism of Bohr's view of quantum theory began at the fifth Solvay. Conference in Brussels in 1927. Einstein would invent thought ...
  64. [64]
    Noninvasiveness and time symmetry of weak measurements
    Feb 27, 2013 · Even in the ground state, quantum fluctuations allow jumps between the wells, which turn out to be non-time symmetric under weak measurement. , ...
  65. [65]
    Relational Quantum Mechanics - Stanford Encyclopedia of Philosophy
    Oct 8, 2019 · Relational Quantum Mechanics (RQM) is the most recent among the interpretations of quantum mechanics which are most discussed today.Missing: 2020s | Show results with:2020s
  66. [66]
    Relative Facts of Relational Quantum Mechanics are Incompatible ...
    May 23, 2023 · We derive a GHZ-like contradiction showing that relative facts described by these statements are incompatible with quantum theory.Missing: 2020s jumps<|separator|>
  67. [67]
    Quantum theory and determinism
    Jul 3, 2014 · Quantum theory and determinism usually do not go together. A natural combination is quantum theory and randomness. Indeed, when in the end of ...<|separator|>
  68. [68]
    Does quantum theory imply the entire Universe is preordained?
    Dec 19, 2023 · The popular idea that quantum physics implies everything is random and nothing is certain might be as far from the truth as it could possibly be.