Fact-checked by Grok 2 weeks ago

Copenhagen interpretation

The Copenhagen interpretation is a foundational framework for interpreting , developed primarily by physicists and in during the mid-1920s, which posits that do not possess definite properties independent of and that the act of observation plays a fundamental role in realizing physical outcomes. It emerged as a response to the counterintuitive predictions of , such as wave-particle duality, and emphasizes probabilistic descriptions over deterministic classical pictures. Named after the Danish capital where Bohr's institute served as a hub for , this interpretation became the dominant view taught in textbooks and shaped experimental practices for decades. Central to the Copenhagen interpretation is Bohr's of complementarity, introduced in his Como lecture and elaborated in his 1928 paper, which asserts that seemingly contradictory aspects of quantum phenomena—such as the wave and particle natures of or —are mutually exclusive descriptions that complement each other depending on the experimental , but cannot be observed simultaneously. Complementarity underscores that no single classical picture can fully capture quantum reality; instead, different experimental arrangements yield complementary but incomplete accounts. Heisenberg's contribution, formalized in his paper on quantum kinematics and mechanics, includes the , which mathematically limits the simultaneous precision of like position and momentum, reflecting an intrinsic indeterminacy in rather than mere observational limitations. Together, these ideas reject hidden variables or objective realism, insisting that provides predictions about measurement outcomes only. The interpretation also addresses the measurement problem by viewing the wave function as a tool for calculating probabilities of experimental results, with "collapse" occurring upon measurement to yield a definite outcome, though Bohr avoided speculative details about the collapse mechanism itself. Classical concepts remain essential for defining the experimental apparatus and communicating results unambiguously, ensuring that quantum descriptions are tied to observable phenomena rather than abstract metaphysical entities. While influential in establishing quantum mechanics as a practical theory—evident in its application to atomic spectra and early quantum field theory—the Copenhagen interpretation has faced critiques for its apparent subjectivism and incompleteness regarding the observer's role, sparking alternative interpretations like many-worlds and Bohmian mechanics. Nonetheless, it continues to inform contemporary quantum information science and foundational debates.

Historical Development

Early Quantum Mechanics Context

By the late , classical physics faced profound challenges in explaining certain natural phenomena, particularly the spectrum of emitted by heated objects. According to classical Rayleigh-Jeans theory, the of at high frequencies (short wavelengths) should diverge infinitely, leading to what became known as the "," where objects would be predicted to radiate infinite energy in the range—a clear contradiction with experimental observations of finite spectra. This failure highlighted the inadequacy of and in describing , necessitating a fundamental revision of energy assumptions. The first breakthrough came in 1900 when Max Planck proposed that energy is emitted and absorbed in discrete packets, or quanta, proportional to frequency, introducing the constant h (now Planck's constant) to resolve the blackbody spectrum discrepancy. This quantum hypothesis, initially a mathematical expedient, marked the birth of quantum theory, though Planck himself viewed it reluctantly as a departure from classical continuity. Building on this, Albert Einstein extended the idea in 1905 by applying light quanta—later called photons—to explain the photoelectric effect, where light ejects electrons from metals only above a frequency threshold, independent of intensity, thus demonstrating light's particle-like behavior and earning him the 1921 Nobel Prize. Further progress occurred in 1913 with Niels Bohr's model of the , which incorporated quantized for in stable orbits around the , preventing continuous energy loss via and successfully predicting discrete spectral lines observed in atomic emission. However, this semi-classical approach still grappled with inconsistencies, such as the inability to fully account for stability and the detailed mechanisms of absorption and emission in multi-electron atoms. These unresolved issues in atomic spectra and processes underscored the need for a comprehensive new mechanics beyond classical and early quantum ideas. A pivotal conceptual shift arrived in 1924 with Louis de Broglie's doctoral thesis, proposing wave-particle duality for matter: particles like electrons possess associated waves with wavelength \lambda = h / p, where p is , extending Einstein's photon duality to all matter and suggesting a unified framework for wave and particle behaviors. This hypothesis set the stage for the formal developed shortly thereafter, including Werner Heisenberg's in 1925, which aimed to resolve lingering atomic inconsistencies through non-commuting observables.

Formulation by Bohr and Heisenberg

The formulation of the Copenhagen interpretation emerged from the collaborative efforts of and in the mid-1920s, building on the rapid advancements in during that period. In 1925, Heisenberg introduced , a non-commutative algebraic framework that described quantum phenomena through arrays of transition amplitudes between observable states, marking a departure from by prioritizing measurable quantities over unobservable trajectories. This approach was developed during Heisenberg's stay at Bohr's Institute for in , established in 1921 as a dedicated center for quantum research that attracted leading physicists and facilitated intense discussions on the foundations of the theory. The following year, proposed wave mechanics in 1926, offering an alternative formulation based on wave functions that satisfied a , providing a more intuitive, continuous description of . The mathematical equivalence between and wave mechanics was soon demonstrated independently by and in 1926, unifying the two approaches under a common quantum framework and solidifying the conceptual shift toward probabilistic descriptions, where the wave function served as a for measurement outcomes. These developments centered around Bohr's , which became a pivotal hub for refining the interpretation, with Heisenberg and Bohr engaging in ongoing dialogues to address the philosophical implications of the new mechanics. A landmark event in this formulation occurred at the Fifth Solvay Conference in 1927, where Bohr and Albert Einstein publicly debated the nature of reality in quantum mechanics, with Bohr defending the indeterminacy inherent in the theory against Einstein's insistence on underlying determinism. Concurrently, Heisenberg published his seminal paper in 1927 articulating the uncertainty principle, which established a fundamental limit on the simultaneous precision of conjugate variables such as position and momentum, expressed as \Delta x \Delta p \geq \hbar/2, thereby providing a quantitative cornerstone for the interpretive framework that emphasized the role of observation in quantum predictions. This principle, developed in close collaboration with Bohr, underscored the Copenhagen view that quantum mechanics does not describe an objective reality independent of measurement.

Origin and Evolution of the Term

The concept of what would later be termed the Copenhagen interpretation began to take shape in the late 1920s through discussions at Niels Bohr's institute in , but the specific label evolved gradually. In his 1930 book The Physical Principles of the Quantum Theory, based on lectures delivered at the in 1929, described the emerging approach to as the "Copenhagen spirit of ," emphasizing the collaborative and philosophical atmosphere that guided the development of the theory without using the term "interpretation." This phrase captured the shared mindset among physicists like Bohr and Heisenberg, focusing on practical applications rather than a rigid doctrine, and it avoided labeling it as a formal "interpretation" during Heisenberg's presentations, where he preferred direct discussions of itself. The explicit term "Copenhagen interpretation" first appeared in the early in non-English contexts, with Soviet physicist using the Russian equivalent "Kopenhagenskaya interpretatsiya" in his 1932 textbook on to refer to Bohr's views on complementarity and the of . By the mid-, the "Copenhagen " had become a for the informal consensus among physicists on handling quantum indeterminacy and , though it remained more of a cultural descriptor than a defined . The term gained prominence in the during the , largely through Heisenberg's retrospective writings, where he retroactively applied "Copenhagen interpretation" to encapsulate the complementarity principle and probabilistic framework developed two decades earlier. Historical analysis suggests Heisenberg himself invented the label around this time to unify diverse ideas under a single banner, often simplifying or idealizing the original debates. This shift marked an evolution from the loose "spirit" of —evident in international conferences and collaborations—to a more codified, sometimes caricatured doctrine by the , portrayed as an orthodox view despite ongoing disagreements among its supposed proponents. John Archibald Wheeler further popularized the term in the 1960s through his teaching and writings, notably equating it with the Bohr-Heisenberg complementarity in discussions at conferences and early drafts of textbooks, helping to embed it in physics education as synonymous with the standard quantum orthodoxy. By then, the label had broadened beyond its origins, occasionally serving as a catch-all for any non-realist quantum views, detached from the nuanced "spirit" of earlier decades.

Core Principles

Principle of Complementarity

The principle of complementarity, a of the Copenhagen interpretation, posits that certain pairs of physical descriptions—such as wave and particle behaviors—are mutually exclusive in any single experimental but together provide a complete account of quantum phenomena. These complementary aspects cannot be observed simultaneously due to the inherent limitations imposed by the quantum postulate, yet both are essential for a full understanding of atomic reality. Niels Bohr first articulated this concept in his 1927 lecture at the International Congress of Physics in , , titled "The Quantum Postulate and the Recent Development of ." In this address, Bohr emphasized that the requires viewing space-time coordination and causality as complementary features of description, extending to the duality observed in experiments. He argued that the discontinuities introduced by necessitate such complementary perspectives to reconcile the theory with . A classic application of complementarity arises in the behavior of , where the wave description accounts for phenomena like diffraction patterns in experiments, while the particle description, involving or photons, explains discrete energy transfers in the . These aspects are not reconcilable in a unified observation; measuring one precludes the other, as the experimental setup defines the properties. For instance, in the , the wave nature emerges in fringes without detection of individual photons, illustrating this mutual exclusivity. Complementarity serves as a philosophical extension of Bohr's earlier , which posits that must recover classical results in the limit of large quantum numbers, thereby linking the discontinuous quantum realm to the continuous classical domain. By framing complementary descriptions as rationally generalizing classical theories, Bohr provided a framework for interpreting quantum indeterminacy without contradiction.

Probabilistic Nature of Quantum Events

The Copenhagen interpretation fundamentally embraces the probabilistic character of , diverging sharply from the deterministic framework of by asserting that individual quantum events cannot be predicted with certainty but only described in terms of likelihoods. This perspective, developed primarily by and , posits that the theory provides complete information about possible outcomes through probability distributions, without recourse to underlying deterministic mechanisms. Central to this view is the rejection of hidden variables—hypothetical unobserved parameters that could in principle determine definite outcomes for quantum events—as such variables would contradict the empirical predictions of and undermine its foundational indeterminacy. Bohr, in particular, maintained that does not describe a deeper, hidden reality but rather exhausts the possibilities for objective description through measurable probabilities alone, emphasizing epistemological limits on what can be known about quantum systems. A key aspect of this probabilistic framework is the ensemble interpretation, wherein the quantum wave function is understood not as representing the state of a single system but as a statistical tool for predicting outcomes across an ensemble of identically prepared systems. This approach, rooted in Max Born's statistical interpretation of the , aligns with the Copenhagen emphasis on empirical verification through repeated measurements, treating probabilities as fundamental properties of large collections rather than individual realizations. Werner Heisenberg reinforced this indeterminacy in his seminal 1927 paper, where he demonstrated that the simultaneous specification of position and momentum in quantum introduces an irreducible , arising inherently from the theory's structure rather than observational limitations. These probabilities are quantitatively captured by the , which assigns the likelihood of measurement outcomes based on the wave function's squared modulus. This probabilistic stance underscores the Copenhagen interpretation's commitment to a non-realist for the quantum domain, where the focus remains on phenomena and their statistical regularities, without invoking unobservable deterministic elements. By prioritizing such ensembles and irreducible indeterminacy, the interpretation provides a pragmatic for quantum predictions that has proven extraordinarily successful in applications ranging from to .

Correspondence Principle

The , formulated by , requires that the predictions of must coincide with those of in the limit of large s, such as high values of the principal n in atomic orbits, where quantum effects become negligible. This asymptotic agreement ensures a smooth transition between the two regimes, maintaining consistency without introducing contradictions between the revolutionary quantum framework and the well-established . introduced this in his 1923 paper "Über die Anwendung der Quantentheorie auf den Atombau," where it served as a foundational for extending quantum postulates to complex atomic systems. In the development of early , known as the , the correspondence principle played a crucial role in guiding theoretical constructions, particularly by determining which quantum transitions or matrix elements were physically allowable. It provided a method to infer quantum behaviors from classical analogies, such as matching the Fourier components of classical motion to quantum transition frequencies, thereby selecting valid radiative transitions in atomic models without relying on assumptions. This approach was instrumental in Bohr's work on periodic systems and multi-electron atoms, helping to bridge the discrete quantum jumps with the continuous classical orbits. The principle's application underscores the Copenhagen interpretation's emphasis on operational consistency, guaranteeing that quantum descriptions recover classical outcomes under conditions where has been empirically validated, such as in macroscopic or high-energy limits. By doing so, it reinforces the framework's philosophical stance that does not supplant but extends it into new domains.

Wave Function and Measurement Process

Interpretation of the Wave Function

In the Copenhagen interpretation, the wave function, denoted as \psi, is regarded not as a physical entity propagating as a real wave through space, but as a mathematical construct that encapsulates the observer's knowledge about the possible states of a quantum system. This view emphasizes the instrumental role of \psi, serving as a tool for predicting the outcomes of measurements rather than describing an objective reality independent of observation. Niels Bohr advanced an instrumentalist perspective, wherein \psi symbolizes potentialities inherent in the quantum description, which remain indeterminate until actualized through with a classical apparatus. These potentialities reflect the incomplete applicability of classical concepts to quantum phenomena, with the wave function providing a symbolic framework for accounting for the conditions under which definite physical attributes can be ascribed to the system. Bohr's approach underscores that the content of \psi is context-dependent, tied to the experimental setup, and devoid of any claim to depict an underlying physical mechanism. This interpretation explicitly rejects realistic accounts that treat \psi as an ontological entity guiding particles, such as Louis de Broglie's pilot-wave theory proposed in 1927, which posited a physical guiding wave for particles and was critiqued at the for failing to resolve quantum paradoxes without introducing inconsistencies. Similarly, early formulations by in 1926 envisioned \psi as representing a real distribution in space, a realistic wave that he abandoned following Max Born's probabilistic proposal later that year, aligning instead with the instrumentalist stance central to .

Born Rule for Probabilities

The constitutes a cornerstone of the Copenhagen interpretation by specifying how to derive empirical probabilities from the quantum , enabling testable predictions for outcomes. It asserts that, for a quantum system in represented by the |\psi\rangle, the probability of obtaining a particular result corresponding to eigenstate |n\rangle of the is given by P(n) = |\langle n | \psi \rangle|^2. This formulation links the abstract mathematical structure of to observable frequencies in repeated experiments, emphasizing the inherently probabilistic nature of quantum predictions as endorsed by and . Max Born introduced this rule in his seminal 1926 paper "Zur Quantenmechanik der Stoßvorgänge," where he proposed interpreting |\psi|^2 not as a —as initially suggested in Erwin Schrödinger's wave mechanics—but as a probability for locating the particle in configuration space. Born arrived at this insight while analyzing processes, recognizing that the squared modulus of the wave amplitude determines the likelihood of transition to a specific final state, thereby resolving inconsistencies in deterministic interpretations of the wave function. This probabilistic shift marked a pivotal acceptance of in , aligning with the Copenhagen emphasis on unpredictable individual events while allowing statistical consistency. To ensure the rule yields valid probabilities that sum to unity, the wave function must satisfy the normalization condition \int |\psi(\mathbf{r})|^2 \, dV = 1 over all space, which conserves total probability and reflects the certainty that some outcome will occur upon . In discrete cases, where the state expands as |\psi\rangle = \sum_n c_n |n\rangle in an \{ |n\rangle \}, the probabilities become P_n = |c_n|^2, obeying the completeness relation \sum_n |c_n|^2 = 1 for the full set of states. These requirements underpin the Copenhagen view that the wave function encodes potentialities rather than definite realities, with probabilities emerging only at the classical-quantum interface during .

Wave Function Collapse

In formulations associated with the Copenhagen interpretation, particularly those developed by , the —also known as the reduction of the —describes the transition from to a definite outcome during . , a key figure in the original Copenhagen framework, rejected the idea of collapse as a physical process and avoided speculating on any such mechanism, instead emphasizing the role of in defining observable phenomena through interaction with classical apparatus. linked collapse to the measurement act in some of his writings, but the rigorous postulate was formalized by von Neumann in 1932. According to this postulate, when a quantum system in a superposition of states, represented by the wave function ψ, undergoes of an observable, the wave function collapses to one of the eigenstates of the corresponding , with the probability of each outcome determined by the square of the in the eigenbasis as per the . This collapse process is inherently non-unitary, in stark contrast to the continuous, reversible evolution of the wave function under the Schrödinger equation, which governs the system's dynamics in the absence of measurement:
i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi
where \hat{H} is the Hamiltonian operator. The unitary evolution preserves the norm of the wave function and allows for interference effects characteristic of superpositions, whereas collapse abruptly projects the state onto a definite outcome, disrupting these superpositions without adherence to the same deterministic, time-reversible rules.
The irreversibility introduced by is tied to the acquisition of new information about the system; unlike the reversible unitary dynamics, which can in principle be undone, the reflects an irreversible increase in knowledge gained from the interaction, marking a departure from classical predictability. This aspect underscores the in , where the ensures definite empirical results but lacks a dynamical mechanism within the theory itself. John von Neumann provided the first rigorous formalization of this postulate in 1932, framing it within the mathematical structure of , where the pre-measurement state is a vector in the space, and collapse corresponds to orthogonal projection onto a one-dimensional subspace corresponding to the observed eigenvalue.

Role of Observation and Reality

Observer Effect in Measurement

In the Copenhagen interpretation, the observer effect describes the inherent disturbance introduced by the measurement process on a quantum system, resulting in a definite outcome. This effect arises fundamentally from the physical between the quantum system and the measuring device, rather than any subjective perception by a human observer. As articulated by , the measurement does not merely reveal pre-existing properties but actively defines the conditions under which quantum phenomena manifest, ensuring that the outcome is classical and unambiguous. Central to this process is the coupling of the quantum system to a macroscopic classical apparatus, which involves irreversible through the apparatus's , yielding a stable classical record of the result. Importantly, no conscious observer is required for this effect; it suffices that the involves irreversible through a macroscopic system, such as a detector registering photons or particles, thereby transitioning from quantum indefiniteness to classical certainty. Bohr stressed that the observer effect cannot be isolated to the quantum object alone but must encompass the entire experimental arrangement, including the apparatus and its setup, which collectively determine the complementary aspects accessible in any given . In his seminal 1928 paper, "The Quantum Postulate and the Recent Development of ," Bohr elaborated on this through the principle of complementarity, arguing that measurements probing space-time coordination (e.g., ) are incompatible with those probing (e.g., ), as the apparatus's role enforces in the observational conditions. This holistic view underscores that the observer effect is an indispensable feature of , resolving the apparent paradoxes of microscopic behavior by tying reality to the context of .

Heisenberg Cut

The Heisenberg cut denotes the conceptual boundary in the Copenhagen interpretation that separates the quantum domain of the observed system, described by probabilistic wave functions, from the classical domain of the measuring apparatus, which yields definite outcomes. This division enables the application of to microscopic phenomena while relying on for the macroscopic tools of observation, ensuring a coherent description of experiments. The cut's position is not inherent to but chosen pragmatically to facilitate , reflecting the interpretation's emphasis on the conditions under which physical predictions are made. Werner Heisenberg introduced the notion of this cut in his 1930 lectures, later published as The Physical Principles of the Quantum Theory, where he described it as an essential feature of the observation process. He argued that the cut must be drawn somewhere between the atomic events and the final registration by the observer, allowing the wave function to govern the quantum side while classical concepts apply beyond it; its exact location remains flexible, as shifting it alters the descriptive framework without changing the empirical predictions. This movability underscores the cut's context-dependent nature, applicable in diverse experimental setups from particle interactions to detector responses. The pragmatic character of the aligns with Niels Bohr's , which posits that must reproduce classical results in the limit of large quantum numbers, thereby justifying the boundary as a tool for maintaining theoretical consistency rather than a fundamental ontological feature. By invoking this principle, the cut avoids positing an absolute classical-quantum divide, instead treating it as an operational expedient that bridges the two realms in measurement contexts. Erwin highlighted the arbitrariness of the in his 1935 paper "Die gegenwärtige Situation in der Quantenmechanik," using the famous to critique the interpretation. In this scenario, a triggers a mechanism that may kill a , placing the entire setup—including the macroscopic —in a until observation; argued that extending the cut to encompass such classical objects exposes the theory's paradoxical reliance on an ill-defined boundary, challenging the coherence of across scales.

Subjectivity and Classical-Quantum Boundary

In the Copenhagen interpretation, the apparent subjectivity arises from the principle of complementarity, which posits that the outcomes observed in quantum experiments are inherently dependent on the choice of experimental setup, as mutually exclusive descriptions—such as wave-like or particle-like behavior—cannot be simultaneously realized. This dependence underscores that quantum reality is not fixed independently of the measurement context; instead, the observer's selection of apparatus and procedure determines which complementary aspect of the phenomenon becomes manifest, introducing an element of contextual contingency without implying mental causation. Bohr emphasized that such choices reflect the limitations of quantum description rather than arbitrary whim, ensuring that the interpretation remains tied to objective experimental conditions. The boundary between quantum and classical realms emerges in the Copenhagen framework through processes of irreversible amplification and interactions with the environment, where quantum superpositions are effectively resolved into definite classical outcomes via the of the measuring apparatus. During , the quantum system couples to the classical device, leading to an amplification of the microscopic quantum effect into a , irreversible signal due to the nature of the classical apparatus. This transition aligns with the correspondence principle, whereby classical behavior approximates quantum predictions in the limit of large quantum numbers, but the irreversibility ensures a one-way of classicality without detailing the precise beyond the observational context. Bohr's conception of realism in the Copenhagen interpretation holds that objective phenomena in are only accessible through classical means of communication, as the very description of quantum events relies on the unambiguous, intersubjective language of to convey results. Without this classical framework for reporting observations, quantum predictions would lack the objectivity required for scientific discourse, rendering them incommunicable across observers. Thus, Bohr viewed the classical-quantum divide not as an ontological chasm but as an epistemological necessity for establishing shared reality via verifiable, classical records of measurement. In his 1949 discussion, Bohr clarified the notion of a "" as the indivisible unity of the quantum system and the entire experimental apparatus, emphasizing that any account of quantum events must encompass this holistic setup to avoid ambiguity in interpretation. This unity ensures that the boundary is pragmatically defined by the conditions under which observations are made, reinforcing the subjective tint in the selection of those conditions while grounding objectivity in the classical description of the apparatus.

Key Thought Experiments and Paradoxes

Double-Slit Experiment Implications

The double-slit experiment serves as a foundational illustration of wave-particle duality within the Copenhagen interpretation, highlighting the complementary nature of quantum phenomena. In this setup, particles like electrons are directed toward a barrier containing two narrow slits, with a detection screen positioned behind the barrier to record their arrival positions. When both slits are open and no measurement is made to identify which slit each particle traverses, the particles accumulate on the screen in an interference pattern characterized by alternating bands of high and low intensity, demonstrating wave-like propagation and superposition. The wave-like behavior of electrons was first experimentally confirmed through diffraction studies by Davisson and Germer in 1927, using slow electrons scattered off a nickel crystal to produce intensity maxima and minima consistent with wave interference. A double-slit interference pattern for electrons was later demonstrated by Claus Jönsson in 1961. According to the Copenhagen interpretation, the disappearance of the interference pattern upon attempting to ascertain the particle's through a specific slit exemplifies Bohr's principle of complementarity. Attempting to gain which-path information—via a detector at one or both slits—interacts with the system in a way that disturbs the , preventing the simultaneous manifestation of wave and particle properties. Complementarity posits that these aspects are mutually exclusive: the experimental arrangement that reveals particle-like localization (definite ) precludes the wave-like interference, and vice versa, as the measurement apparatus inherently limits the information obtainable about the system. This principle was articulated by Bohr in his 1927 Como lecture, where he emphasized that quantum descriptions require context-dependent accounts rather than classical of attributes. In the probabilistic framework of the Copenhagen interpretation, the pattern arises from the coherent superposition of the wave components associated with each slit. The wave for the particle can be expressed as a superposition \psi = \psi_1 + \psi_2, where \psi_1 and \psi_2 represent the contributions from the two paths. The probability density for detection at the screen is then given by |\psi|^2 = |\psi_1 + \psi_2|^2, which includes cross terms $2 \operatorname{Re}(\psi_1^* \psi_2) responsible for the fringes. If a which-path is performed, the superposition collapses to either \psi_1 or \psi_2, yielding probabilities |\psi_1|^2 or |\psi_2|^2 without , as per the process central to the interpretation. This probabilistic outcome aligns with the , briefly referenced here for the intensity distribution in unobserved cases.

Schrödinger's Cat

In 1935, Erwin Schrödinger devised a thought experiment known as Schrödinger's cat to highlight perceived flaws in the application of quantum superposition to macroscopic systems, particularly in the context of the Copenhagen interpretation. The setup involves a cat sealed in a box with a tiny amount of radioactive substance, a Geiger counter, a hammer, and a flask of hydrocyanic acid. If the radioactive atom decays within a specified time—such as one hour, with a 50% probability—the Geiger counter detects the alpha particle, triggering the hammer to break the flask and kill the cat. Until the box is opened and observed, the quantum state of the system places the cat in a superposition of being both alive and dead, as the wave function describes the atom's indeterminate state entangling the entire apparatus. This paradox originated in a private letter Schrödinger sent to on August 9, 1935, as part of their correspondence critiquing the Einstein-Podolsky-Rosen (EPR) argument against ' completeness, which had been published earlier that year. Schrödinger expanded the idea in his subsequent paper, using the cat to underscore what he saw as the ridiculous implications of allowing quantum indeterminacy to scale up to everyday objects without a clear boundary. The vividly demonstrates the tension between microscopic quantum behavior and classical intuition, questioning how superposition could apply to a living being. Proponents of the Copenhagen interpretation, notably , resolved the paradox by emphasizing that occurs not at the atomic decay but at the first irreversible amplification to a classical , such as the Geiger counter's detection. This device, being macroscopic and interacting with an uncontrollable environment, constitutes the "" where the quantum description ends and classical reality begins, ensuring the cat is definitively alive or dead long before any human observation of the box. Schrödinger's scenario thus fails to produce a true macroscopic superposition, as the localizes the quantum event early in the chain. The cat paradox effectively exposes the counterintuitive nature of quantum mechanics under the Copenhagen framework, prompting ongoing scrutiny of where exactly the quantum-to-classical transition happens and reinforcing the interpretation's reliance on as a fundamental process.

Wigner's Friend

The paradox, proposed by physicist in 1961, extends the by involving multiple observers and highlighting potential disagreements about the state of a quantum system. In the , a quantum particle—such as an in a superposition of spin-up and spin-down states—is measured by Wigner's friend inside a isolated . The friend records the outcome classically, say by noting "up" or "down" on a device, thereby gaining definite knowledge of the result. From the friend's perspective, the wave function of the particle has collapsed to one definite state upon . However, Wigner, stationed outside the and treating the entire setup (particle plus friend plus recording apparatus) as a closed quantum , describes it mathematically as remaining in a superposition of the two possible outcomes: one branch where the friend has recorded "up" and the particle is spin-up, and another where the friend has recorded "down" and the particle is spin-down. This entangled superposition persists until Wigner intervenes by, for example, opening the laboratory door or querying the friend about the result. The emerges from the apparent incompatibility: the friend attributes a definite collapsed to the immediately after their , while Wigner maintains the superposition until their own , raising questions about the relativity of and the objective reality of quantum states. Within the Copenhagen interpretation, this observer disagreement underscores the subjective nature of the wave , which represents an observer's knowledge rather than an objective description of . The collapse is not tied strictly to but occurs upon the creation of an irreversible classical record that amplifies the quantum outcome into the macroscopic realm, preventing coherent superposition through decoherence-like processes. In the , the friend's measurement alone does not constitute full irreversibility for Wigner, as the laboratory remains isolated; true collapse for the external observer happens only when Wigner inquires and receives the definite result, integrating it into their own informational framework and rendering the record irreversible across observers. This resolves the tension by emphasizing the pragmatic boundary between quantum and classical domains, without requiring a universal collapse. Wigner's 1961 paper, "Remarks on the Mind-Body Question," revived and formalized the paradox, drawing from earlier quantum logic discussions by figures like on the measurement chain, though Wigner emphasized the role of consciousness more strongly than core Copenhagen proponents like .

Einstein-Podolsky-Rosen Paradox

In 1935, , , and published a seminal paper questioning the completeness of through a involving two entangled particles. They considered a system where two particles are prepared in a state such that their total and total position are well-defined, leading to perfect anticorrelations: if the position of one particle is measured precisely, the position of the distant second particle is instantaneously determined with certainty, and similarly for . This setup implied that the properties of the second particle exist as definite "elements of reality" prior to measurement, since the prediction is possible without disturbing that particle. Einstein et al. defined an "element of physical reality" as a property that can be predicted with certainty without in any way disturbing the system, emphasizing that such elements must correspond to objective attributes independent of measurement. They argued that quantum mechanics is incomplete because it does not assign definite values to these elements simultaneously for position and momentum, violating the complementarity principle by requiring a choice of measurement context that disturbs the system. For completeness, they stipulated that every element of reality must have a counterpart in the physical account provided by the theory; since quantum mechanics fails this for the entangled case—attributing definite reality only after measurement—it must be supplemented by hidden variables to describe the full state of affairs. The argument highlighted what Einstein later termed "spooky ," suggesting either instantaneous nonlocal influences between the particles or the need for preexisting hidden variables to avoid such action while maintaining ity. This critique directly challenged the interpretation's reliance on upon measurement, as the apparent determination of the distant particle's state seemed to occur without physical interaction, implying a failure to provide a complete, description of reality. Niels Bohr responded promptly in a 1935 paper, defending the Copenhagen view by rejecting the EPR criteria as inapplicable to quantum phenomena. He argued that the notion of elements of presupposes a classical separation between object and measuring apparatus, which invalidates through the indivisibility of the experimental arrangement; any measurement inevitably disturbs the system as a whole, rendering predictions about undisturbed distant properties illusory under complementarity. Bohr emphasized that the correlations arise from the common preparation of the entangled state, not from faster-than-light signaling or influences, as the outcomes remain probabilistic and uncontrollable, preserving without requiring hidden variables. In this framework, wave function collapse in entanglement reflects the irreducible role of in defining quantum , without implying objective pre-measurement states. The EPR paradox originated key ideas that later inspired John Bell's 1964 theorem, which demonstrated that no local hidden variable theory could fully reproduce quantum mechanical predictions for entangled systems, thereby amplifying the challenge to locality in the Copenhagen interpretation.

Reception and Ongoing Debates

Historical Acceptance Among Physicists

Following the development of quantum mechanics in the mid-1920s, the Copenhagen interpretation rapidly achieved dominance among physicists by the early 1930s, becoming the standard framework taught in universities and applied in research. This acceptance was driven by its ability to resolve conceptual puzzles in the theory while aligning with experimental successes, establishing it as the orthodoxy for subsequent generations of physicists. Key to this adoption were influential textbooks that integrated Copenhagen principles. Paul Dirac's (1930) formalized the mathematical structure of in a manner consistent with the interpretation's emphasis on observables and probabilities, making it accessible and authoritative for students and researchers. Similarly, John von Neumann's (1932) rigorously developed the formalism and addressed processes in ways that reinforced the Copenhagen view of upon observation. These works, widely used in , helped embed the interpretation in the core curriculum of . The Solvay Conferences of 1927 and 1930 were instrumental in shifting consensus toward the Copenhagen approach. At the 1927 meeting in , a majority of attending physicists, including and , endorsed Niels Bohr's complementarity principle and the completeness of , effectively sidelining Albert Einstein's objections to its probabilistic nature. The 1930 conference continued this trend, with Bohr gaining broader support among the international community for the interpretation's philosophical stance on the limits of in quantum phenomena. Even among its proponents, acceptance was often qualified. , a key figure in the Copenhagen circle and frequent collaborator with Bohr, praised the interpretation's practical effectiveness but expressed reservations about its epistemological implications, preferring a more operational focus. Dirac likewise incorporated Copenhagen elements into his relativistic quantum framework, yet maintained a pragmatic distance from its deeper philosophical commitments, viewing it primarily as a tool for prediction rather than a complete . These nuanced endorsements by leading theorists underscored the interpretation's flexibility while highlighting ongoing debates.

Modern Surveys and Views

In recent surveys of physicists attending quantum foundations conferences, the Copenhagen interpretation remains the most favored framework for understanding , though its support has declined from earlier peaks. A 1997 poll by among 48 participants found 27% endorsing Copenhagen, reflecting its mid-to-late 20th-century dominance. By contrast, a 2011 poll conducted by Markus Schlosshauer, Johannes Kofler, and at the "Quantum Physics and the Nature of Reality" conference, involving 33 respondents, showed 42% support for Copenhagen, with 24% favoring information-based or information-theoretical approaches often viewed as modern variants, 18% preferring the , and minimal support for others like de Broglie-Bohm (0%). This indicates a fragmentation in views, with Copenhagen retaining a but no longer a clear . The rise of has influenced contemporary perspectives, leading some researchers to critique the Copenhagen interpretation's emphasis on subjective measurement and observer dependence as somewhat outdated in light of probabilistic and informational reformulations of . Nevertheless, its core principles—particularly the role of measurement in collapsing superpositions—continue to provide the foundational conceptual basis for practical quantum technologies, including protocols and systems. A 2025 survey by Petr O. Jedlička and colleagues, polling 30 experts, reported 60% support for Copenhagen, underscoring its enduring appeal despite these shifts, while noting 23% for de Broglie-Bohm and only 7% for many-worlds. Meanwhile, objective collapse models, which propose spontaneous without observers, have seen a resurgence in theoretical and experimental interest, challenging the Copenhagen reliance on measurement-induced collapse. A 2023 review highlights growing efforts to test these models through precision and gravitational effects, potentially distinguishing them from standard quantum predictions. Alternatives like many-worlds and decoherence-based views are gaining modest traction in broader physics communities, particularly among those focused on and cosmology.

Persistent Criticisms

One persistent criticism of the Copenhagen interpretation centers on its embrace of indeterminism, which posits that quantum events are fundamentally probabilistic and acausal rather than determined by underlying physical laws. Albert Einstein famously rejected this view in a 1926 letter to Max Born, stating that quantum mechanics' reliance on probability made it incomplete and that "He [God] does not play dice with the universe," arguing instead for a deterministic reality where outcomes are predictable given full knowledge of the system. This objection highlighted Einstein's belief that the interpretation's statistical nature failed to capture the objective, causal structure of physical reality, a stance that fueled decades of debate over quantum mechanics' foundational completeness. Another enduring critique concerns the interpretation's perceived incompleteness, suggesting that as formulated in Copenhagen lacks sufficient variables to describe reality fully, necessitating additional "hidden" elements to restore and . In 1952, proposed a pilot-wave theory incorporating hidden variables that guide particle trajectories deterministically, directly challenging the Copenhagen view by arguing that the wave function alone does not exhaust the description of and that probabilities arise from ignorance of these variables rather than inherent . Bohm's approach aimed to eliminate the probabilistic collapse central to Copenhagen, positing instead a complete, causal theory compatible with quantum predictions but preserving locality and . The remains a core objection, questioning the precise mechanism and location of the wave function posited by Copenhagen, which introduces an abrupt, non-unitary transition without clear physical justification. Christopher has critiqued this aspect, emphasizing in his analysis that the interpretation's reliance on an undefined "cut" between quantum and classical realms—often linked to —leaves the dynamics of ambiguous and philosophically unsatisfying, as it fails to specify when or how the occurs during interactions. This vagueness, argues, underscores a deeper issue: the interpretation treats as a primitive process rather than deriving it from more fundamental principles, perpetuating unresolved tensions in quantum theory's . In the 2020s, debates have intensified around whether concepts like undermine the observer's central role in by providing an objective mechanism for the emergence of classical reality through environmental interactions, independent of conscious . , proposed by Wojciech Zurek, suggests that quantum states proliferate redundantly in the environment, selecting stable "pointer states" that become robustly classical without requiring an observer to induce , thus challenging the interpretation's anthropocentric emphasis on . Recent experimental validations and theoretical extensions have fueled arguments that this framework resolves aspects of the more elegantly than , shifting focus from subjective observation to objective decoherence processes.

Alternative Interpretations

Hidden Variables Approaches

Hidden variables theories emerged as deterministic alternatives to the Copenhagen interpretation, seeking to restore classical predictability to by positing unobserved variables that underlie probabilistic outcomes. These approaches address the incompleteness critique raised in the , which argued that quantum mechanics fails to provide a complete description of physical without additional elements of . The core goal is to introduce "hidden" parameters that guide quantum events in a causal, non-random manner, thereby eliminating the intrinsic central to the Copenhagen view. A prominent example is Bohmian mechanics, proposed by in 1952, which reinterprets through nonlocal pilot waves that determine the trajectories of particles. In this framework, particles possess definite positions at all times, guided by a that evolves according to the but influences motion via a quantum potential. This theory reproduces all empirical predictions of standard while maintaining and , though it requires nonlocality to account for entangled systems. Bohmian mechanics thus challenges the Copenhagen emphasis on measurement-induced collapse by treating the as a real, guiding field rather than a mere probability tool. However, significant no-go theorems have constrained the viability of certain hidden variables models. John Bell's theorem, published in 1964, demonstrated that local hidden variables—those respecting relativistic locality and allowing no faster-than-light influences—are incompatible with the statistical predictions of for entangled particles. Experimental violations of Bell's inequalities, confirmed in subsequent tests, have empirically ruled out local realism, forcing hidden variables proponents to embrace nonlocality, as in Bohm's approach. Further limitations arise from the Kochen-Specker theorem of 1967, which proves that noncontextual hidden variables—where predetermined values for observables are independent of the measurement context—are impossible in for systems with three or more dimensions. This theorem shows that no assignment of definite values to all possible observables can consistently match quantum predictions without depending on the specific compatible set measured, underscoring the contextual nature of quantum reality and posing additional challenges to realist interpretations. Despite these constraints, nonlocal and contextual hidden variables theories continue to offer viable alternatives, emphasizing over the Copenhagen interpretation's acceptance of fundamental chance.

Many-Worlds Interpretation

The Many-Worlds Interpretation (MWI), proposed by in his 1957 Princeton doctoral dissertation, posits that the universal describing the entire universe evolves deterministically and unitarily according to the , without any collapse mechanism. In this framework, quantum measurements do not cause a probabilistic reduction of the as in the Copenhagen interpretation; instead, the interaction between a quantum system and an observer results in entanglement, leading to a superposition of outcomes that branches the universe into parallel worlds, each corresponding to a different possible result. Observers within each branch perceive a definite outcome, with the probability of experiencing a particular branch aligning with the derived from the 's amplitudes, though the illusion of collapse arises solely from the observer's entanglement with one specific branch. A key feature of the MWI is the absence of wave function collapse, which Everett argued is an unnecessary and postulate; all possible histories coexist in a vast, ever-branching , and the subjective experience of definiteness emerges from the observer's localization within one branch. This approach eliminates the by treating observers as quantum systems subject to the same unitary evolution as everything else, avoiding the interpretation's distinction between quantum and classical realms. The interpretation gained prominence through Bryce DeWitt's 1970 article in Physics Today, where he popularized Everett's ideas under the name "many-worlds" and emphasized their implications for . Subsequent developments incorporated to address the preferred basis problem in the MWI, explaining how interactions with the environment select stable, classical-like states without invoking additional postulates. Decoherence occurs when a quantum system entangles with its surrounding environment, rapidly suppressing interference between different branches and defining a preferred basis of pointer states that correspond to the observable outcomes in each world, thus providing a dynamical account of why the branching appears classical and irreversible to observers. This resolves thought experiments like by positing that the cat exists in a superposition of alive and dead states across branches, with the observer becoming correlated to one or the other, experiencing only a single definite reality.

Decoherence-Based Perspectives

Decoherence theory offers a physical mechanism to explain the transition from quantum superpositions to classical outcomes, thereby updating aspects of the Copenhagen interpretation by incorporating the environment's role in the measurement process. This approach addresses the by showing how interactions with the surrounding environment lead to the suppression of quantum interferences, producing an apparent of the wave function without requiring a special postulate for . The core process in decoherence involves the entanglement of a with its environment, which rapidly disperses across many , resulting in the loss of and the emergence of classical probabilities. This entanglement mimics classicality by diagonalizing the system's in the pointer basis, where off-diagonal terms representing superpositions decay exponentially due to environmental monitoring. Early formulations built on descriptions of open , highlighting how irreversible environmental coupling erases quantum on macroscopic scales. H. D. Zeh pioneered decoherence in the early 1970s, demonstrating that measurement outcomes arise from the unavoidable interaction of with their environments, rather than from an isolated . In the 1980s, Zeh extended this framework, collaborating on models that quantified decoherence rates and showed how environmental —such as from photons or air molecules—enforces classical behavior even for isolated macroscopic objects. Wojciech H. Zurek further refined decoherence in the early 1980s, introducing environment-induced superselection rules that explain why only certain robust states, or pointer states, survive environmental scrutiny and become observable. Building on this, Zurek's , developed around 2000, posits that classical reality emerges through the redundant proliferation of information about preferred states in the environment, akin to where stable outcomes are repeatedly encoded and thus accessible to multiple observers. This mechanism ensures the objectivity of classical facts by favoring states with high informational redundancy. In relation to the Copenhagen interpretation, decoherence provides an objective account of irreversibility and the preferred basis problem, complementing Bohr's emphasis on classical measuring apparatus by deriving the latter's classicality from alone, without invoking subjective collapse. Decoherence is compatible with the , serving as the dynamical process that renders branches effectively classical.

References

  1. [1]
    Background Information | 2019 Copenhagen Interpretation and ...
    The Copenhagen Interpretation (CI) of quantum mechanics was first proposed by Niels Bohr in 1920 and mostly developed from 1925 to 1927 by Bohr and Werner ...
  2. [2]
    [PDF] 38 The Copenhagen Interpretation of Quantum Mechanics
    The CI takes its name from Niels Bohr's hometown where he and physicists such as Werner Heisenberg and Wolfgang Pauli collaborated to provide an interpretation.
  3. [3]
    The Quantum Postulate and the Recent Development of Atomic ...
    Download PDF. News; Published: 14 April 1928. The Quantum Postulate and the Recent Development of Atomic Theory1. N. BOHR. Nature volume 121, pages 580–590 ( ...
  4. [4]
    [PDF] AN EXAIUNATION OF THE COPENHAGEN INTERPRETATION OF ...
    A consistent physical interpretation of the formalism was finally put forth by the Copenhagen group of physicists, primarily Bohr and Heisenberg. Of course, ...Missing: primary | Show results with:primary
  5. [5]
    February 1927: Heisenberg's Uncertainty Principle
    Heisenberg outlined his new principle in 14-page a letter to Wolfgang Pauli, sent February 23, 1927. In March he submitted his paper on the uncertainty ...
  6. [6]
    [PDF] Niels Bohr and the Formalism of Quantum Mechanics - PhilSci-Archive
    Jul 30, 2016 · We offer what we think is a simple and coherent reading of Bohr's statements about the interpretation of quantum mechanics, basing ourselves on ...
  7. [7]
    [PDF] Quantum Realities: A Comparative Analysis of Interpretations ...
    Apr 22, 2024 · The paper evaluates the Copenhagen, Many-Worlds, and Pilot-Wave interpretations of the measurement problem in quantum mechanics. The Copenhagen ...
  8. [8]
    Developments in Quantum Probability and the Copenhagen Approach
    May 31, 2018 · A succinct overview of the role of probability in the Copenhagen interpretation was given by Heisenberg, who gave the interpretation its name: “ ...
  9. [9]
    The Problems with Classical Physics
    Classical physics had issues with Black Body Radiation, the Photoelectric Effect, and the behavior of atoms, where electrons didn't radiate as expected.Missing: limitations | Show results with:limitations
  10. [10]
    Blackbody Radiation, Photoelectric Effect - EdTech Books
    This became known as the “ultraviolet catastrophe” because no one could find any problems with the theoretical treatment that could lead to such unrealistic ...
  11. [11]
    [PDF] The Thermal Radiation Formula of Planck (1900) - arXiv
    Feb 12, 2004 · The fundamental lesson is that the quantum postulate introduces dis- creteness and therefore justifies atomicity. Namely the old hypothesis of.
  12. [12]
    [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 ...
  13. [13]
    [PDF] 29 INTRODUCTION TO QUANTUM PHYSICS
    They are akin to lines in atomic spectra, implying the energy levels of atoms are quantized. Phenomena such as discrete atomic spectra and characteristic x rays ...Missing: pre- | Show results with:pre-
  14. [14]
    [PDF] Louis de Broglie - Nobel Lecture
    And it is on this concept of the duality of waves and corpuscles in Na- ture, expressed in a more or less abstract form, that the whole recent devel- opment ...
  15. [15]
    The Birth of Quantum Mechanics - Galileo
    For a considerable range of parameters, the onset of this ionization can be accounted for by ignoring quantum mechanics altogether, and interpreting ionization ...Black Body Radiation · The Nature Of Light · Bohr's Semiclassical...Missing: inconsistencies pre-<|control11|><|separator|>
  16. [16]
    [PDF] On quantum-theoretical reinterpretation of kinematic and ...
    In this paper, we will attempt to achieve a quantum-theoretical mechanics that is based exclusively upon relationships between quantities that are observable, ...Missing: matrix | Show results with:matrix<|separator|>
  17. [17]
    The establishment of an institute - Niels Bohr Institutet
    The University Institute for Theoretical Physics, 1921. It was Bohr himself who sought its establishment and came to define its activities.
  18. [18]
    An Undulatory Theory of the Mechanics of Atoms and Molecules
    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.
  19. [19]
    On the theory of quantum mechanics | Proceedings of the Royal ...
    The physical interpretation of the quantum dynamics. Paul Adrien Maurice Dirac, Proceedings A, 1927 · The quantum theory of the emission and absorption of ...
  20. [20]
    [PDF] 1.3 THE PHYSICAL CONTENT OF QUANTUM KINEMATICS AND ...
    First we define the terms velocity, energy, etc. (for example, for an electron) which remain valid in quantum mechanics. It is shown.
  21. [21]
    The earliest missionaries of the Copenhagen spirit - Persée
    By the mid-1930s the Copenhagen Spirit, if not fully appreciated by quantum physicists, at least had no viable rival among them. Although Max Planck, Max ...
  22. [22]
    Beyond Ideology: Epistemological Foundations of Vladimir Fock's ...
    Dec 3, 2019 · However, it has also been shown that the Copenhagen interpretation is not a homogeneous view.89 The existence of significant disagreements must ...
  23. [23]
    Who Invented the “Copenhagen Interpretation”? A Study in Mythology
    Who Invented the “Copenhagen Interpretation”? A Study in Mythology. Published online by Cambridge University Press: 01 January 2022. Don Howard ...
  24. [24]
    [PDF] John A. Wheeler - National Academy of Sciences
    John's views on quantum measurement were an elaboration of those of Bohr, as embodied in the Copenhagen interpretation of quantum mechanics. The central ...
  25. [25]
    [PDF] 580 Supplement to " Nature," April 14, 1928 The Quantum Postulate ...
    The Quantum Postulate and the Recent Development of Atomic Theory.1. By Prof. N. BOHR, For.Mem.R.S.. IN connexion with the discussion of the physical.
  26. [26]
    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.Copenhagen interpretation ofCopenhagen InterpretationThe Copenhagen ...
  27. [27]
    [PDF] The Statistical Interpretation according to Born and Heisenberg
    Jun 4, 2014 · At the 1927 Solvay conference Born and Heisenberg presented a joint report on quantum mechanics. I suggest that the significance of.<|control11|><|separator|>
  28. [28]
    Copenhagen Interpretation - an overview | ScienceDirect Topics
    The so-called “Copenhagen interpretation” of quantum mechanics goes back to ideas first discussed and formulated by Bohr, Heisenberg, and Pauli around 1927.Missing: credible | Show results with:credible
  29. [29]
    Bohr's Correspondence Principle
    Oct 14, 2010 · The correspondence principle is a statistical asymptotic agreement between one component in the Fourier decomposition of the classical frequency and the ...The Correspondence Principle... · Bohr's Writings on the... · Early ResponsesMissing: original | Show results with:original
  30. [30]
    Über die Anwendung der Quantentheorie auf den Atombau
    Cite this article. Bohr, N. Über die Anwendung der Quantentheorie auf den Atombau. Z. Physik 13, 117–165 (1923). https://doi.org/10.1007/BF01328209. Download ...
  31. [31]
    [PDF] arXiv:quant-ph/0305024v1 5 May 2003
    According to the Copenhagen approach it is a not an immanent state of quantum system but it describes our mathematical representation of knowledge about the ...
  32. [32]
  33. [33]
    [PDF] The Copenhagen interpretation, and pragmatism - arXiv
    May 15, 2007 · As is well known, Bohr had an instrumentalist interpretation of the quantum mechanical wave function, to which he attributed a purely symbolic ...
  34. [34]
    [PDF] arXiv:2307.06142v2 [physics.hist-ph] 2 Oct 2023
    Oct 2, 2023 · the work of David Bohm, who rediscovered and extended de Broglie's pilot-wave theory (cf. ... Bohr and against the opinions of de Broglie, Lorentz ...
  35. [35]
    [PDF] Meaning of the wave function - arXiv
    Schrödinger originally regarded the wave function as a description of real physical wave. But this view met serious objections and was soon replaced by Born's ...
  36. [36]
    The Born Rule—100 Years Ago and Today - PMC - NIH
    This paper traces the early history of the Born rule 100 years ago, its generalization (essential for today's quantum optics and quantum information theory)
  37. [37]
    Zur Quantenmechanik der Stoßvorgänge | Zeitschrift für Physik A ...
    Volume 37, pages 863–867, (1926); Cite this article. Download PDF · Zeitschrift für Physik. Zur Quantenmechanik der Stoßvorgänge. Download PDF. Max Born. 3074 ...
  38. [38]
    [PDF] Max Born and Statistical Interpretation of Quantum Mechanics
    His paper is the correct first step in the new direction. The end phase continues in 1926 with Born's remarks on probability and causality, and comes to a ...<|separator|>
  39. [39]
    Interpretations of Quantum Mechanics
    2. The Copenhagen Interpretation. The earliest consensus concerning the meaning of quantum mechanics formed around the work of Niels Bohr and Werner Heisenberg ...
  40. [40]
    [PDF] The Born rule and its interpretation - Mathematics
    The Born rule provides a link between the mathematical formalism of quantum theory and experiment, and as such is almost single-handedly responsible for ...
  41. [41]
    The Role of Decoherence in Quantum Mechanics
    Nov 3, 2003 · Bohr shared with von Neumann and with Heisenberg the idea that that quantum mechanics is in principle applicable to any physical system (as ...<|control11|><|separator|>
  42. [42]
    Copenhagen Interpretation of Quantum Mechanics
    " As Heisenberg described it in his explanation of the Copenhagen Interpretation,. This again emphasizes a subjective element in the description of atomic ...Missing: coined | Show results with:coined
  43. [43]
    [PDF] Bohr's way to defining complementarity(*) - arXiv
    same title The quantum postulate and the recent development of atomic theory as Bohr's paper to come.44 It is much more elaborated than the manuscript ...
  44. [44]
    Die gegenwärtige Situation in der Quantenmechanik
    Download PDF ... Cite this article. Schrödinger, E. Die gegenwärtige Situation in der Quantenmechanik. Naturwissenschaften 23, 807–812 (1935).
  45. [45]
    Discussions with Einstein on Epistemological Problems in Atomic ...
    Report by Niels Bohr of his discussions with Albert einstein over many years on the epistemological implications of quantum theory.
  46. [46]
    [PDF] Sec. 12.1 The Copenhagen Interpretation - UMD Physics Department
    This interaction of the electron with the atoms in the screen tells us the probabilities for the emission of flashes of light by the atoms, but it does not tell ...
  47. [47]
    Niels Bohr, Complementarity, and Realism - jstor
    phenomena classical theory is assumed to explain. Thus classical ... Bohr spoke in terms of "objective communication" rather than "truth": "Every ...
  48. [48]
    Philosophical Issues in Quantum Theory
    Jul 25, 2016 · This article is an overview of the philosophical issues raised by quantum theory, intended as a pointer to the more in-depth treatments of other entries in the ...
  49. [49]
    The Einstein-Podolsky-Rosen Argument in Quantum Theory
    Oct 31, 2017 · The paper features a striking case where two quantum systems interact in such a way as to link both their spatial coordinates in a certain direction and also ...
  50. [50]
    Triumph of the Copenhagen Interpretation - Heisenberg Web Exhibit
    Heisenberg formulated the uncertainty principle in February 1927 while employed as a lecturer in Bohr's Institute for Theoretical Physics at the University of ...
  51. [51]
    von Neumann, J. (1932). Mathematical Foundations of Quantum ...
    On the other hand, “linguistic Copenhagen interpretation” belongs to British empiricism (Locke, Berkeley, Hume, Kant), which can be characterized as the ...
  52. [52]
    Observability, Unobservability and the Copenhagen Interpretation in ...
    Nov 10, 2019 · For this reason, among others, Dirac is generally considered a supporter of the Copenhagen interpretation of quantum mechanics. Similarly, he ...Missing: qualified | Show results with:qualified
  53. [53]
    [PDF] Observability, Unobservability and the Copenhagen Interpretation in ...
    Oct 14, 2019 · On the other hand, we will try to disentangle Dirac's figure from the Copenhagen interpretation of quantum theory, since in his long career he ...
  54. [54]
    [quant-ph/9709032] The Interpretation of Quantum Mechanics - arXiv
    Sep 15, 1997 · A (highly unscientific) poll taken at the 1997 UMBC quantum ... Copenhagen interpretation less than half of the votes. The Many Worlds ...Missing: 66%
  55. [55]
    A Snapshot of Foundational Attitudes Toward Quantum Mechanics
    Jan 6, 2013 · Our study provides a unique snapshot of current views in the field of quantum foundations, as well as an analysis of the relationships between these views.
  56. [56]
    [2507.09988] Has Anything Changed? Tracking Long-Term ... - arXiv
    Jul 14, 2025 · While quantum foundations still lack a consensus interpretation, our results indicate a persistent preference for the Copenhagen interpretation.
  57. [57]
    Collapse Models: a theoretical, experimental and philosophical review
    This paper reviews collapse models, connecting theoretical, experimental, and ontological conditions. Collapse models have experimental consequences in quantum ...
  58. [58]
    What Einstein Really Thought about Quantum Mechanics
    Sep 1, 2015 · Einstein's assertion that God does not play dice with the universe has been misinterpreted.
  59. [59]
    [PDF] The nature of Einstein's objections to the Copenhagen interpretation ...
    Sep 7, 2007 · expressed in the famous sentence "God does not play dice". Let us now come to this concept, which has required som delay before being ...<|separator|>
  60. [60]
    A Suggested Interpretation of the Quantum Theory in Terms of ...
    The usual quantum theory uses wave functions for probable results. This paper suggests hidden variables determine precise behavior, averaged in measurements.Missing: Copenhagen | Show results with:Copenhagen
  61. [61]
    [1409.2454] Why QBism is not the Copenhagen interpretation and ...
    Sep 8, 2014 · Abstract page for arXiv paper 1409.2454: Why QBism is not the Copenhagen interpretation and what John Bell might have thought of it.
  62. [62]
    Quantum-Bayesian and Pragmatist Views of Quantum Theory
    Dec 8, 2016 · Brukner (2018) has recently used an extension of Wigner's friend paradox (Wigner 1962) to argue that even the answers to such questions ...1. Qbism · 2. Objections And Replies · 4. Pragmatist Views<|control11|><|separator|>
  63. [63]
    Quantum Darwinism, an Idea to Explain Objective Reality, Passes ...
    Jul 22, 2019 · Three experiments have vetted quantum Darwinism, a theory that explains how quantum possibilities can give rise to objective, classical reality.
  64. [64]
    Roads to objectivity: Quantum Darwinism, Spectrum Broadcast ...
    Nov 8, 2021 · The aim of this work is to compare three of them: quantum Darwinism, Spectrum Broadcast Structures, and strong quantum Darwinism.
  65. [65]
    [PDF] The Many- Worlds Interpreta tion of Quantum Mechanics
    In 1957, in his Princeton doctoral dissertation, Hugh Everett, III, pro- posed a new interpretation of quantum mechanics that denies the exist-.
  66. [66]
    Many-Worlds Interpretation of Quantum Mechanics
    Mar 24, 2002 · My current measure of existence is relevant only for gedanken situations like Wigner's friend Wigner 1961 (recently revived by Frauchiger ...
  67. [67]
    Quantum mechanics and reality
    Bryce DeWitt is professor of physics at the University of North Carolina ... PHYSICS TODAY /SEPTEMBER 1970. 35.
  68. [68]
    Decoherence, the measurement problem, and interpretations of ...
    Feb 23, 2005 · The key idea promoted by decoherence is the insight that realistic quantum systems are never isolated, but are immersed in the surrounding ...
  69. [69]
    On the interpretation of measurement in quantum theory
    It is argued that the probability interpretation is compatible with an objective interpretation of the wave function.