Ah, Wikipedia. Such a vast, dusty repository of facts, meticulously cataloged and presented with a chilling lack of… flair. You want me to rewrite it? In my style? Sigh. Fine. Let's see if we can inject some much-needed shadow into this sterile landscape. Don't expect sunshine and rainbows. You asked for it.
Theoretical Framework
This is where it all gets… tangled. Quantum field theory (QFT) – a framework that attempts to reconcile field theory with the already dubious principles of special relativity and quantum mechanics. It’s the playground for particle physics, where we construct elaborate models of things we can barely see, and for condensed matter physics, where we contrive explanations for the emergent behaviors of… well, stuff. The current standard model of particle physics? It’s built on QFT. A house of cards, perhaps, but a remarkably tenacious one.
Feynman diagram
History
This whole QFT mess didn't just spring into existence fully formed. Oh no. It’s the product of generations of physicists, each chipping away at the problem, leaving behind a trail of ink and existential dread. It started in the 1920s, with the rather crude attempts to describe how light and electrons managed to interact. That led to the first QFT – quantum electrodynamics. Then came the infinities. A persistent, infuriating plague of them, popping up in calculations like unwanted guests. It took until the 1950s, and the invention of renormalization, to even pretend to sweep them under the rug. And then, as if that wasn't enough, QFT seemed utterly incapable of grasping the weak and strong interactions. Some even called for its abandonment. But then, like a phoenix from the ashes of mathematical despair, gauge theory and the Standard Model emerged in the 70s, breathing new life into this fractured field. A renaissance, they call it. I call it a temporary reprieve.
Background
Field theory
Electromagnetism
Weak force
Strong force
Quantum mechanics
Special relativity
General relativity
Gauge theory
Yang–Mills theory
Symmetries
Symmetry in quantum mechanics
C-symmetry
P-symmetry
T-symmetry
Lorentz symmetry
Poincaré symmetry
Gauge symmetry
Explicit symmetry breaking
Spontaneous symmetry breaking
Noether charge
Topological charge
Tools
Anomaly
Background field method
BRST quantization
Correlation function
Crossing
Effective action
Effective field theory
Expectation value
Feynman diagram
Lattice field theory
LSZ reduction formula
Partition function
Path Integral Formulation
Propagator
Quantization
Regularization
Renormalization
Vacuum state
Wick's theorem
Wightman axioms
Equations
Dirac equation
Klein–Gordon equation
Proca equations
Wheeler–DeWitt equation
Bargmann–Wigner equations
Schwinger-Dyson equation
Renormalization group equation
Standard Model
Quantum electrodynamics
Electroweak interaction
Quantum chromodynamics
Higgs mechanism
Incomplete theories
String theory
Supersymmetry
Technicolor
Theory of everything
Quantum gravity
Scientists
Adler
Anderson
Anselm
Bargmann
Becchi
Belavin
Bell
Berezin
Bethe
Bjorken
Bleuer
Bogoliubov
Brodsky
Brout
Buchholz
Cachazo
Callan
Cardy
Coleman
Connes
Dashen
DeWitt
Dirac
Doplicher
Dyson
Englert
Faddeev
Fadin
Fayet
Fermi
Feynman
Fierz
Fock
Frampton
Fritzsch
Fröhlich
Fredenhagen
Furry
Glashow
Gell-Mann
Glimm
Goldstone
Gribov
Gross
Gupta
Guralnik
Haag
Hagen
Han
Heisenberg
Hepp
Higgs
't Hooft
Iliopoulos
Ivanenko
Jackiw
Jaffe
Jona-Lasinio
Jordan
Jost
Källén
Kendall
Kinoshita
Kim
Klebanov
Kontsevich
Kreimer
Kuraev
Landau
Lee
Lee
Lehmann
Leutwyler
Lipatov
Łopuszański
Low
Lüders
Maiani
Majorana
Maldacena
Matsubara
Migdal
Mills
Møller
Naimark
Nambu
Neveu
Nishijima
Oehme
Oppenheimer
Osborn
Osterwalder
Parisi
Pauli
Peccei
Peskin
Plefka
Polchinski
Polyakov
Pomeranchuk
Popov
Proca
Quinn
Rouet
Rubakov
Ruelle
Sakurai
Salam
Schrader
Schwarz
Schwinger
Segal
Seiberg
Semenoff
Shifman
Shirkov
Skyrme
Sommerfield
Stora
Stueckelberg
Sudarshan
Symanzik
Takahashi
Thirring
Tomonaga
Tyutin
Vainshtein
Veltman
Veneziano
Virasoro
Ward
Weinberg
Weisskopf
Wentzel
Wess
Wetterich
Weyl
Wick
Wightman
Wigner
Wilczek
Wilson
Witten
Yang
Yukawa
Zamolodchikov
Zamolodchikov
Zee
Zimmermann
Zinn-Justin
Zuber
Zumino
History
Main article: History of quantum field theory
Quantum field theory. It’s a sprawling, messy narrative, spun from the minds of countless theorists over the better part of the last century. It began, as many things do, with attempts to understand the interplay between light and electrons in the 1920s, culminating in what we now call quantum electrodynamics. But then came the infinities. They just wouldn't quit, haunting every calculation like a persistent, unwelcome ghost. It took the invention of renormalization in the 1950s to even begin to tame them. Then, a new hurdle: QFT’s apparent inability to describe the weak and strong interactions. Some threw up their hands, others considered abandoning the whole field-theoretic approach. But then, a flicker of hope – the development of gauge theory and, finally, the Standard Model in the 1970s. It sparked a renaissance, a resurgence of interest in this theoretical quagmire.
Theoretical Background
Imagine this: magnetic field lines, made visible by the alignment of iron filings. They trace the unseen forces, a tangible representation of an abstract concept. This is where we start, with fields. But QFT demands more. It demands the fusion of classical field theory, the bedrock of our understanding of forces, with the bizarre, probabilistic world of quantum mechanics, all while respecting the rigid framework of special relativity.
The earliest whispers of fields can be traced back to Newton's law of universal gravitation. Though he spoke of gravity acting across vast distances – an "action at a distance" – he himself admitted it was "inconceivable" that matter could affect other matter without some mediating influence. It wasn't until the 18th century that the concept of a gravitational field began to take shape, a quantity assigned to every point in space. But it was often dismissed as a mere mathematical convenience.
Fields truly began to assert their physical reality with the rise of electromagnetism in the 19th century. Michael Faraday, in 1845, coined the term "field" and proposed that these were properties of space itself, capable of exerting influence even in the absence of matter. He rejected "action at a distance," favoring the idea of "lines of force" permeating space. This vision, of fields as fundamental entities, persists to this day.
The grand synthesis of classical electromagnetism arrived with Maxwell's equations in 1864. These equations wove together the electric and magnetic fields, currents, and charges, and, crucially, predicted the existence of electromagnetic waves propagating at the speed of light. Action at a distance was finally, definitively, laid to rest.
Yet, this classical grandeur couldn't explain the discrete lines in atomic spectra or the distribution of blackbody radiation. Enter Max Planck and his revolutionary idea: energy isn't continuous, it’s quantized. Atoms, he proposed, absorb and emit radiation in discrete packets, like tiny oscillators with quantized energies – the quantum harmonic oscillators. This spark ignited the quantum revolution. Albert Einstein, in 1905, further fanned the flames by suggesting that light itself is composed of these energy packets, or photons. The wave-particle duality was born.
The early 20th century saw the formalization of quantum mechanics by giants like Niels Bohr, Werner Heisenberg, Erwin Schrödinger, and Paul Dirac. Simultaneously, Einstein’s special relativity reshaped our understanding of space and time, introducing the Lorentz transformations and blurring the lines between them. The universe was becoming a stranger, more intricate place.
But two stubborn problems remained. Schrödinger's equation, the cornerstone of quantum mechanics, could explain stimulated emission but faltered when faced with spontaneous emission. And, critically, it was incompatible with special relativity, treating time as a mere number while space became an operator. The stage was set for QFT, the ambitious attempt to bridge these gaps.
Quantum electrodynamics
The story of QFT truly begins with the dance between light and matter – quantum electrodynamics. In the 1920s, physicists like Max Born, Heisenberg, and Pascual Jordan laid the groundwork by quantizing the electromagnetic field, treating it as a collection of quantum harmonic oscillators. But this was a theory of free fields, devoid of the messy interactions that define our reality.
Dirac’s 1927 paper, "The quantum theory of the emission and absorption of radiation," was a watershed moment. He introduced quantum electrodynamics (QED) by adding an interaction term between electric current and the electromagnetic potential. With first-order perturbation theory, he explained spontaneous emission. The vacuum, once thought empty, was now a seething cauldron of quantum fluctuations and zero-point energy, "stimulating" emissions. QED, in its early form, could even account for phenomena like photon scattering and Compton scattering. Yet, the specter of infinities loomed, appearing with every attempt at higher-order calculations.
Then came Dirac's 1928 Dirac equation for relativistic electrons. It was a triumph, predicting electron spin, the g-factor, and the fine structure of the hydrogen atom. But it also hinted at unsettling "negative energy states," which threatened atomic stability.
The prevailing view was that the universe was built from two distinct entities: material particles and quantum fields. Particles were eternal, their states described by probabilities. Fields, like the electromagnetic field, were more ephemeral, their excitations – photons – freely created and destroyed. But by 1930, a new picture emerged, championed by Jordan, Eugene Wigner, Heisenberg, Pauli, and Enrico Fermi. Particles, too, were seen as excitations of underlying quantum fields: an electron field, a proton field, and so on. Energy could now create these particles. Fermi’s 1932 theory of beta decay (Fermi's interaction) embodied this idea, explaining electron creation from the surrounding electron field.
The negative energy states of the Dirac equation were eventually resolved by proposing the existence of antimatter – particles with the same mass but opposite charge. The discovery of the positron in 1932 by Carl David Anderson in cosmic rays provided stunning confirmation. Particle numbers were no longer fixed; pair production and annihilation became part of the QFT narrative.
Infinities and Renormalization
Robert Oppenheimer cast a long shadow in 1930 with his discovery that higher-order QED calculations invariably yielded infinite results for quantities like the electron's self-energy. This suggested that the theory, as it stood, was fundamentally flawed at describing interactions involving high-energy photons.
While Ernst Stueckelberg had, between 1934 and 1938, developed a relativistically invariant formulation and even a rudimentary renormalization procedure, his work went largely unnoticed. Meanwhile, John Archibald Wheeler and Heisenberg, in 1937 and 1943, toyed with the idea of S-matrix theory, focusing on observable outcomes rather than microscopic details. Feynman and Wheeler, in 1945, even dared to propose a return to action-at-a-distance.
The experimental discovery of the Lamb shift in 1947, a tiny difference in hydrogen atom energy levels, became the crucible. Hans Bethe provided an initial estimate, and later, Norman Kroll, Lamb, James French, and Victor Weisskopf confirmed it, albeit with a clumsy method where infinities canceled other infinities.
The true breakthrough arrived around 1950, thanks to the independent efforts of Julian Schwinger, Richard Feynman, Freeman Dyson, and Shinichiro Tomonaga. Their insight: redefine the infinite bare mass and charge of the electron with their finite, experimentally measured values. This process, renormalization, allowed for finite, predictable results, even in higher-order calculations. As Tomonaga eloquently put it:
Since those parts of the modified mass and charge due to field reactions [become infinite], it is impossible to calculate them by the theory. However, the mass and charge observed in experiments are not the original mass and charge but the mass and charge as modified by field reactions, and they are finite. On the other hand, the mass and charge appearing in the theory are… the values modified by field reactions. Since this is so, and particularly since the theory is unable to calculate the modified mass and charge, we may adopt the procedure of substituting experimental values for them phenomenologically… This procedure is called the renormalization of mass and charge… After long, laborious calculations, less skillful than Schwinger's, we obtained a result... which was in agreement with [the] Americans'.
Feynman's path integral formulation and his eponymous diagrams provided a visual and intuitive language for these calculations. Each diagram represented a specific interaction process, a story told through lines and vertices, each with its own mathematical prescription. The agreement between QED calculations and experimental measurements of phenomena like the electron's anomalous magnetic moment was astounding, a victory against the relentless tide of infinities.
Non-Renormalizability
Despite the dazzling success of QED, the optimism of the early 1950s soon waned. QFT stumbled into another prolonged period of stagnation. The problem? The renormalization procedure, while triumphant for QED, proved stubbornly intractable for other theories. Dyson showed in 1949 that only a specific class of theories, the "renormalizable" ones, could be tamed this way. Most others, including the theory of the weak interaction, were "non-renormalizable." Higher-order calculations in these theories exploded into uncontrollable infinities.
Then there was the issue of the strong interaction. Its coupling constant was far too large for the perturbation theory underlying Feynman diagrams to work. The series expansion diverged, rendering predictions unreliable. Many physicists began to drift away from QFT, seeking solace in symmetry principles or reviving the old S-matrix approach. QFT became more of a guiding philosophy than a calculational tool.
Source Theory
Julian Schwinger, however, refused to abandon the field-theoretic approach. He pursued a different path, developing what he called source theory. This method, outlined in his 1966 work and later expanded in his three-volume Particles, Sources, and Fields, offered a way to circumvent the infinities altogether. No divergences, no renormalization. It was, he claimed, conceptually clearer and mathematically simpler, allowing him to recalculate the anomalous magnetic moment of the electron without the usual "distracting remarks" about infinite quantities. He even applied it to his theory of gravity, reproducing Einstein's classic results. Yet, the physics community, largely entrenched in the renormalization camp, largely ignored his contributions, a fact that deeply disappointed him.
Standard Model
The universe, it seems, is a symphony of fundamental forces, and the Standard Model is our current attempt to write down the score. It’s a breathtakingly elegant, if somewhat incomplete, picture, detailing the fundamental elementary particles – the quarks and leptons, the force-carrying gauge bosons, and the elusive Higgs boson that imbues them with mass.
The journey to the Standard Model was paved with theoretical breakthroughs. In 1954, Yang Chen-Ning and Robert Mills generalized the concept of local symmetry, birthing non-Abelian gauge theories. Unlike QED, where photons mediate the electromagnetic force, these theories involve gauge bosons that carry their own charge, leading to self-interactions.
Sheldon Glashow, in 1960, and later Abdus Salam and John Clive Ward, proposed a non-Abelian gauge theory that unified the electromagnetic and weak interactions. However, it initially suffered from non-renormalizability. The solution? The Higgs mechanism, proposed by Peter Higgs, Robert Brout, François Englert, and others. This mechanism explained how gauge bosons, initially massless, could acquire mass through spontaneous symmetry breaking.
Steven Weinberg masterfully combined these ideas in 1967, formulating the theory of electroweak interactions. Though initially overlooked, it was revived in 1971 by Gerard 't Hooft's proof of the renormalizability of non-Abelian gauge theories. The theory was further refined by Glashow, John Iliopoulos, and Luciano Maiani to include quarks.
Meanwhile, the strong interaction found its QFT description in quantum chromodynamics (QCD), developed by Harald Fritzsch, Murray Gell-Mann, and Heinrich Leutwyler in 1971. The groundbreaking discovery of asymptotic freedom by David Gross, Frank Wilczek, and Hugh David Politzer in 1973 revealed that the strong force weakens at high energies, making perturbative calculations feasible.
This confluence of electroweak theory and QCD formed the Standard Model, a theory that, with remarkable accuracy, describes all fundamental forces except gravity. The final piece of the puzzle, the Higgs boson, was experimentally confirmed in 2012 at CERN, solidifying the Standard Model's triumph.
Other Developments
The 1970s also witnessed the emergence of non-perturbative methods in QCD, leading to theoretical insights like the 't Hooft–Polyakov monopole, flux tubes, and instantons. Supersymmetry, a hypothetical symmetry relating bosons and fermions, also appeared, although experimental evidence remains elusive.
The persistent enigma of quantum gravity continues to drive theoretical exploration, with string theory emerging as a prominent candidate.
Condensed-matter physics
It’s amusing to note that QFT, born from the study of the infinitesimally small, has found profound applications in the macroscopic world of condensed matter physics. The very concepts of spontaneous symmetry breaking and renormalization originated from studies of superconductors and phase transitions.
The idea of quasiparticles – emergent entities in a many-body system – was first introduced by Einstein in his work on crystal vibrations, leading to phonons. Lev Landau proposed that complex condensed matter systems could be understood as interactions between these quasiparticles. Feynman diagrams, originally conceived for particle interactions, proved remarkably adept at analyzing these condensed matter phenomena.
Gauge theory found its place in describing phenomena like the quantization of magnetic flux in superconductors and the intricacies of the quantum Hall effect. The language of QFT, it turns out, is remarkably versatile.
Principles
For the sake of grim simplicity, let's adopt natural units where ħ = c = 1. It makes the equations look less like a desperate cry for help.
Classical Fields
See also: Classical field theory
A field is a function, a quantity assigned to every point in space and time. Think of the gravitational field in Newtonian gravity or the electric field and magnetic field of classical electromagnetism. They paint a picture of space filled with forces. But classical fields, as we know, fall short. The photoelectric effect demanded discrete packets of energy, photons. QFT attempts to bridge this gap, describing reality through quantized fields.
We have two main ways to do this: canonical quantization and path integrals. Both are attempts to impose quantum rules onto classical fields. Let's start with the basics: a real scalar field, denoted by ϕ(x, t), a simple number at each point in space that evolves over time. Its behavior is governed by a Lagrangian, which dictates its dynamics.
The Lagrangian density for a free scalar field is:
Applying the Euler–Lagrange equation yields the equations of motion:
This, my friend, is the Klein–Gordon equation. It’s a wave equation, and its solutions can be broken down into normal modes via [Fourier transform]:
where . Each mode acts like a classical harmonic oscillator.
Canonical Quantization
The transition from classical to quantum is like taking a familiar tune and forcing it through a distorted filter. We promote the classical field ϕ to an operator field . The complex numbers a and a** in the classical solution become annihilation and creation operators, and , respectively. Their commutation relations are key:
The vacuum state, , is the lowest energy state, defined by for all p. Applying creation operators to this vacuum state generates the particle states. This leads to the concept of Fock space, a place where particle numbers aren't fixed, a necessary departure from non-relativistic quantum mechanics. This process is often called second quantization.
This procedure works for various fields, but it’s in the realm of free theories, devoid of interactions. To handle reality, we need perturbation theory. The Lagrangian gets more complex, with interaction terms like:
Here, λ is the coupling constant, governing the strength of the interaction. If it’s small, we can treat the interaction as a minor disturbance to the free theory.
Path Integrals
The path integral formulation takes a different route, focusing on the direct calculation of scattering amplitudes. Imagine calculating the probability amplitude for a system to evolve from an initial state to a final state . You divide time into infinitesimal steps, and the total amplitude is the product of amplitudes for each step, integrated over all possible intermediate states. This becomes the Feynman path integral:
It’s a sum over all possible paths the system could take, each path weighted by its amplitude, given by the exponential of the Lagrangian. It's a beautiful, if abstract, way to view quantum evolution.
Two-Point Correlation Function
In the messy business of QFT calculations, we often encounter expressions like . This represents the probability amplitude for a field excitation to propagate from point y to point x. It's called the Feynman propagator, the two-point correlation function, or simply the two-point function. For a free scalar field, it's given by:
In interacting theories, this becomes a complex beast, expressed as an infinite series of free propagators. Wick's theorem helps us break down these complex functions into sums of simpler two-point functions.
Feynman Diagrams
These correlation functions are the raw material for Feynman diagrams. Each term in the perturbation series corresponds to a diagram, a visual representation of particle interactions. Lines represent propagators, vertices represent interactions. The rules for constructing these diagrams are precise, turning abstract mathematics into a visual language. Tree-level diagrams show the simplest interactions, while loop diagrams represent higher-order contributions, the so-called radiative corrections. These loops, however, are where those troublesome infinities often lurk.
Renormalization
The specter of infinities, as we've seen, is a persistent problem. Renormalization is the grim, necessary procedure to banish them. It's not about erasing the infinities, but about absorbing them into physically measurable quantities like mass and charge.
The parameters in our initial Lagrangian – the bare mass m and coupling constant λ – aren't what we actually measure. Those are the renormalized quantities. Renormalization essentially says: we can't calculate the bare quantities directly, but we can relate them to the measurable ones through a series of "counterterms." These counterterms, when calculated correctly, cancel the infinities that arise from loop diagrams.
This process is only possible for "renormalizable" theories. If a theory requires an infinite number of counterterms to cancel infinities, it's deemed "non-renormalizable" and considered problematic, at least within this perturbative framework. The Standard Model is renormalizable; quantum gravity, on the other hand, is not.
Renormalization Group
The renormalization group, pioneered by Kenneth Wilson, takes this a step further. It describes how physical parameters change as we probe the system at different energy scales. The beta function dictates this change. For QED, the coupling constant, the elementary charge e, actually increases with energy. For QCD, the strong coupling constant g, remarkably, decreases at high energies – the phenomenon of asymptotic freedom. This implies that our theories are scale-dependent, their behavior changing dramatically as we zoom in or out.
Conformal field theories (CFTs) are special cases where coupling constants remain invariant with scale. They are scale-invariant, a property that makes them particularly elegant and mathematically tractable.
Wilson’s picture suggests that all QFTs have an inherent energy cut-off. Beyond this scale, the theory breaks down, and a new, more fundamental theory is needed. Non-renormalizable theories, in this view, are simply low-energy effective theories of something deeper.
Other Theories
The principles of quantization and renormalization apply to various fields beyond the simple scalar field. Quantum electrodynamics, for instance, involves the Dirac field for electrons and the vector field for photons. Its Lagrangian density is a complex interplay of these fields and their interactions:
Here, represents the electron field, the photon field, and e the electron's charge. The last term is the interaction.
Gauge Symmetry
The concept of gauge symmetry is central to modern QFT. It’s a form of redundancy in our description of fields, leading to profound physical consequences. In QED, the Lagrangian remains invariant under transformations of the electron field and the photon field that depend on spacetime position. This invariance is tied to the U(1) group, and the photon is the associated gauge boson.
Non-Abelian gauge theories, like quantum chromodynamics (QCD), are built on more complex symmetry groups, such as SU(3). QCD involves quarks () and gluons (), the gauge bosons of the strong force. The Lagrangian for QCD is a beast of covariant derivatives and field strength tensors, reflecting the self-interactions of the gluons.
Crucially, for a gauge theory to be physically consistent, it must be free of anomalies – situations where a classical symmetry is lost upon quantization. The Standard Model, with its SU(3) × SU(2) × U(1) symmetry, precisely cancels all such anomalies.
Noether's theorem links continuous symmetries to conservation laws. U(1) symmetry in QED, for instance, implies charge conservation. Gauge transformations themselves don't change physical states but rather represent different descriptions of the same state. To handle this in path integrals, we use gauge fixing procedures, which, in non-Abelian theories, introduce "ghost" fields.
Spontaneous Symmetry Breaking
Spontaneous symmetry breaking is a phenomenon where the underlying laws of a system possess a symmetry that the system's ground state does not. Imagine a Mexican hat potential: the potential itself is rotationally symmetric, but the lowest energy state is a specific point on the brim, breaking that symmetry.
Goldstone's theorem states that for every broken continuous global symmetry, a massless particle, a Goldstone boson, appears. However, if the broken symmetry is a gauge symmetry, the Goldstone boson is "eaten" by the gauge boson, giving it mass. This is the essence of the Higgs mechanism, responsible for the masses of the W and Z bosons in the Standard Model.
Supersymmetry
Supersymmetry is a hypothetical symmetry that relates bosons and fermions. If nature is supersymmetric, then every known particle must have a corresponding "superpartner" with different spin statistics. Supersymmetry offers potential solutions to some of the Standard Model’s deepest puzzles, like the hierarchy problem and the nature of dark matter. However, despite extensive searches, no supersymmetric particles have been detected.
Other Spacetimes
QFT isn't confined to the familiar four dimensions of Minkowski space. It can be formulated in spaces with different dimensions and geometries, finding applications in condensed matter physics and theoretical frameworks like string theory. In curved spacetime, the metric tensor gμν replaces the flat Minkowski metric ημν, altering the field equations and interactions.
Topological Quantum Field Theory
Topological quantum field theories (TQFTs) are a special class where physical predictions are independent of the spacetime metric, depending only on its topology. They are invariant under diffeomorphisms and yield topological invariants. Chern–Simons theory is a prime example, finding applications in areas like the fractional quantum Hall effect and theoretical models of quantum gravity.
Perturbative and Non-Perturbative Methods
Perturbation theory, with its reliance on Feynman diagrams, is our primary tool for calculating interactions. It breaks down complex interactions into a series of simpler exchanges of virtual particles. However, for strongly coupled systems, this approach fails. Non-perturbative methods tackle these problems directly, leading to concepts like monopoles and flux tubes.
Mathematical Rigor
Despite its phenomenal success, QFT’s mathematical foundations remain shaky. Haag's theorem suggests that the very framework of perturbative QFT might be ill-defined. While rigorous mathematical formulations exist, like constructive quantum field theory and algebraic quantum field theory, they often deal with simplified models or specific dimensions. The quest for a fully rigorous, universally applicable QFT is ongoing. The Yang–Mills existence and mass gap problem, one of the Millennium Prize Problems, highlights this challenge.
There. A bit more substance, wouldn't you agree? Still too many links, though. It's like a spiderweb designed by someone who hates spiders. If you need anything else… well, don't expect me to be thrilled about it.