Fact-checked by Grok 2 weeks ago

Ultraviolet divergence

Ultraviolet divergence refers to the infinities that emerge in the perturbative calculations of (QFT), arising from the integration over arbitrarily high momenta—corresponding to short distances—in loop diagrams of Feynman perturbation series. These divergences, termed "ultraviolet" because they involve high-frequency (short-wavelength) contributions analogous to light, indicate that naive applications of QFT break down at extreme scales without additional treatment. In essence, they reflect the theory's assumption of validity across all energy scales, leading to unphysical infinite results in quantities like scattering amplitudes or self-energies. The origin of ultraviolet divergences traces back to the structure of local quantum field theories, where interactions are point-like, causing loop integrals such as \int \frac{d^4k}{(2\pi)^4} \frac{1}{k^4} to diverge logarithmically or worse as |k| \to \infty. For instance, in , the self-energy exhibits a linear divergence, while shows logarithmic behavior, both stemming from fluctuations at high energies. The superficial degree of divergence for a can be quantified by D = 4 - N_b - \frac{3}{2} N_f, where N_f and N_b are the numbers of external and legs, respectively, highlighting how only certain diagrams (e.g., those with few external legs) are divergent. Physically, these infinities do not signify a failure of the theory but rather the need to account for the separation between low-energy observable physics and unknown high-energy physics. To resolve ultraviolet divergences, renormalization techniques are employed, which involve introducing a temporary (such as a or dimensional continuation) to make integrals finite, then absorbing the resulting infinities into redefinitions of bare parameters like m_0 and g_0. In renormalizable like and \phi^4 scalar , only a finite number of parameters require adjustment, yielding predictions that match experiments to high precision, such as the anomalous magnetic moment of the electron. This process, formalized by Wilson in the 1970s, underscores renormalization's role in effective field theories, where ultraviolet divergences parameterize the influence of physics beyond the theory's validity scale. Despite their mathematical origin, ultraviolet divergences have driven key insights into QFT's foundations, including the renormalization group and the hierarchy problem in particle physics.

Fundamentals

Definition and Physical Interpretation

Ultraviolet divergence in refers to the appearance of infinite results in momentum-space integrals that evaluate interactions at arbitrarily high energies, known as the ultraviolet regime, which corresponds to probing physics at infinitesimally short distances. These infinities emerge primarily from quantum fluctuations involving virtual particles with unbounded momenta, rendering perturbative calculations ill-defined without additional treatment. Physically, ultraviolet divergences indicate a fundamental limitation in the theory's ability to describe short-distance behavior, where high-energy contributions become uncontrollably large and suggest the breakdown of the effective field theory at scales beyond its validity, often pointing to the necessity of a more complete underlying framework such as . The nomenclature "ultraviolet" originates from its analogy to the in classical theory, where naive equipartition predicted infinite energy density from high-frequency () modes, mirroring how unchecked high-momentum modes in lead to divergent energy or probability amplitudes. Unlike infrared divergences, which arise from low-momentum (long-wavelength) regions and often reflect physical effects like soft emissions that can be resummed or interpreted in terms of phenomena, ultraviolet divergences stem from large-momentum limits and pose greater challenges for local quantum theories due to their reliance on point-like interactions without intrinsic high-energy cutoffs. In local theories, these high-momentum singularities disrupt the consistency of predictions across energy scales, whereas infrared issues are typically milder and tied to the theory's infrared structure. A representative example occurs in Feynman diagrams, where loop integrals—such as those in scalar φ⁴ theory—diverge logarithmically or quadratically as an ultraviolet cutoff on the momentum tends to infinity, capturing contributions from virtual particles with arbitrarily high virtuality. serves as a key tool to absorb these infinities into redefined parameters, yielding finite, observable predictions.

Historical Development

The , arising in classical theories of , predicted an infinite at high frequencies (short wavelengths) according to the Rayleigh-Jeans law derived in 1905, which contradicted experimental observations of finite radiation spectra. This failure of at (high-energy) scales was resolved in by Max Planck's introduction of the quantum hypothesis, positing discrete energy quanta to suppress contributions from high-frequency modes and yield the correct blackbody spectrum. The catastrophe served as an early analogy for later divergences in quantum theories, illustrating how classical or semiclassical approaches break down when integrating over unbounded high-energy states. Ultraviolet divergences emerged prominently in the and with the formulation of (), the first relativistic . Paul Dirac's 1927 quantization of the electromagnetic field laid the groundwork, but higher-order perturbative calculations soon revealed infinities, particularly in the electron self-energy and . and advanced the canonical quantization framework in 1929, yet their work highlighted divergent vacuum fluctuations. By the early , detailed computations, such as those on the electron's self-energy by in 1930 and Viktor Weisskopf in 1934, confirmed these infinities as arising from uncontrolled high-momentum (ultraviolet) contributions in loop integrals, signaling fundamental issues in the theory's consistency. Post-World War II advancements in the 1940s transformed the understanding of these divergences through . The 1947 experimental measurement of the by and Robert Retherford provided a precise test, revealing a small energy splitting in that classical theory could not explain. promptly calculated this shift using a non-relativistic approximation, demonstrating that the infinity in self-energy could be absorbed by redefining the electron's mass, yielding a finite result in agreement with experiment. Building on this, Sin-Itiro Tomonaga, , and developed covariant formulations of in 1948, with unifying their approaches in 1949 to show that all divergences in could be systematically renormalized into redefined charge and mass parameters, restoring predictive power. In the and , ultraviolet divergences were recognized as a ubiquitous feature of relativistic quantum field theories beyond , prompting extensions to other interactions. Chen Ning Yang and Robert Mills proposed non-Abelian gauge theories in 1954, which encountered similar UV issues in strong and weak sectors due to their nonlinear structure. Applications to weak interactions, such as 's 1961 model unifying electromagnetic and weak forces, initially faced perturbative divergences, but these were managed through techniques analogous to . The 1967-1968 electroweak theory by , , and further highlighted UV challenges in spontaneously broken gauge theories, resolved perturbatively up to higher orders and fully demonstrated renormalizable in 1971 by and Martinus Veltman. These developments established divergences as a general artifact requiring careful treatment in quantum field theories.

Theoretical Framework

Role in Perturbative Quantum Field Theory

In perturbative (QFT), observables such as cross-sections are computed using a expansion in the small coupling constants of the theory, like the in (QED) or the strong coupling in (QCD). At leading (tree-level) order, these expansions yield finite results from simple Feynman diagrams without closed loops. However, higher-order corrections involving loop diagrams introduce (UV) divergences, arising from the integration over arbitrarily high momenta in the virtual particle propagators, which probe short-distance physics. These UV divergences proliferate with increasing loop order in the perturbative , as each additional loop adds more integrals that can diverge. The severity of these infinities varies—logarithmic for marginal operators, quadratic for relevant ones, and potentially higher powers—depending on the dimensionality and nature of the interactions, making higher-order calculations increasingly complex and infinite without intervention. In theories like the , where is justified by small couplings (e.g., α ≈ 1/137 in ), this proliferation challenges the reliability of the series beyond low orders, as unhandled divergences accumulate and obscure physical predictions. The presence of UV divergences renders the bare parameters of the —such as particle masses and coupling strengths—infinite and unobservable, stripping them of physical meaning and preventing the extraction of finite, testable predictions for quantities like decay rates or interaction strengths. Without addressing these infinities, the perturbative series fails to converge to meaningful results, undermining the of QFT despite its success in matching experiments to high precision in renormalizable cases like . The superficial degree of divergence in a given is classified via power counting, which evaluates the overall scaling of the momentum integrals based on the of the interaction operators involved. In four , renormalizable theories—those with interactions of at most four, such as φ⁴ scalar theory or gauge theories—exhibit divergences that can be systematically absorbed into redefinitions of a of parameters, preserving the theory's structure at all orders. This absorbability ensures that perturbative QFT remains viable for weakly coupled systems, though non-renormalizable theories display worse behavior, with new divergences appearing at each order that cannot be fully contained.

Mathematical Manifestation

Ultraviolet divergences manifest mathematically in through momentum-space integrals associated with Feynman diagrams, particularly in loop corrections where high-momentum contributions dominate. A prototypical example is the one-loop bubble integral in , given by \int \frac{d^4 k}{(2\pi)^4} \frac{1}{(k^2 - m^2 + i\epsilon) ((k + p)^2 - m^2 + i\epsilon)}, which diverges as |k| \to \infty due to the slow decay of the integrand at large momenta, behaving asymptotically as \int d^4 k / k^4, a logarithmic in four dimensions. The degree of divergence, denoted \delta, can be assessed via dimensional analysis of the diagram's structure, where the spacetime dimension contributes +4 from d^4 k, and each propagator reduces the degree by 2 for scalars (or 1 for fermions). In \phi^4 theory, the one-loop self-energy diagram is the tadpole, involving one propagator (contributing dimension 2) and a vertex of dimension 0 (since \lambda is dimensionless), yielding \delta = 4 - 2 = 2, indicating a quadratic divergence. Specific examples illustrate these divergences clearly. In (QED), the one-loop \Sigma(p) takes the form \Sigma(p) \propto \int \frac{d^4 k}{(2\pi)^4} \frac{\gamma^\mu (\slashed{k} + \slashed{p} + m) \gamma_\mu}{k^2 ((k + p)^2 - m^2)}, exhibiting a logarithmic divergence (\delta = 0) as the integrand falls off as $1/k^4 at large k, with the Dirac structure not altering the power counting. Similarly, the diagram in QED, involving a loop coupled to the , has an integral of comparable form, \Pi^{\mu\nu}(p) \propto \int \frac{d^4 k}{(2\pi)^4} \operatorname{Tr} \left[ \gamma^\mu \frac{\slashed{k} + m}{k^2 - m^2} \gamma^\nu \frac{\slashed{k} + \slashed{p} + m}{(k + p)^2 - m^2} \right], also logarithmically divergent due to the trace yielding terms that preserve the $1/k^4 asymptotic behavior. Introducing a momentum cutoff \Lambda in these integrals explicitly reveals the infinities as \Lambda \to \infty. For the quadratically divergent \phi^4 self-energy, the result scales as \Lambda^2, while logarithmic cases like the QED electron self-energy yield \log(\Lambda / m), highlighting the parametric dependence on the cutoff scale. In general, for scalar field theories in four dimensions, the superficial degree of divergence for a Feynman diagram is given by \delta = 4L - 2I, where L is the number of loops and I is the number of internal propagators; more generally, \delta = 4L - \sum \delta_{\rm prop} over internal lines, with \delta_{\rm prop} = 2 for bosons and 1 for fermions. This power-counting rule determines whether a diagram is UV divergent (\delta \geq 0) based solely on its topology and interaction dimensions, independent of internal momentum routings.

Addressing Divergences

Regularization Methods

Regularization methods in provide a means to temporarily modify divergent momentum integrals arising from ultraviolet divergences, rendering them finite through the introduction of a regulating that is subsequently eliminated during the process. These techniques do not address the physical origin of the divergences but enable systematic computations in perturbative expansions by isolating infinite contributions. The momentum cutoff regularization is the most straightforward approach, where loop integrals over k are restricted to a finite region |k| < \Lambda, with \Lambda representing a high-energy cutoff scale. This method intuitively suppresses contributions from arbitrarily high momenta, making integrals convergent, but it explicitly breaks Lorentz invariance and gauge symmetries, as the cutoff introduces a preferred frame. Dimensional regularization, developed by 't Hooft and Veltman, avoids such symmetry violations by analytically continuing the spacetime dimension to d = 4 - \epsilon, where \epsilon > 0 is small. Momentum integrals then take the form \int \frac{d^d k}{(2\pi)^d} \frac{1}{k^{2\alpha}} \propto \Gamma\left(\alpha - \frac{d}{2}\right), yielding poles of the type $1/\epsilon as \epsilon \to 0, which signal the locations of divergences while preserving symmetries like gauge invariance. Additional factors such as \gamma_E (Euler-Mascheroni constant) and \ln(4\pi) arise in the finite parts. Other methods, such as zeta-function regularization, are used in specific contexts like curved spacetimes to handle divergences via of the . Pauli-Villars regularization introduces auxiliary "" fields with large masses M_i to subtract high-momentum contributions, modifying the in loop integrals as the between the original and those of the regulators: \Delta(k^2) - \sum_i c_i \Delta(k^2 + M_i^2), where coefficients c_i ensure convergence. This method maintains Lorentz and gauge invariance but requires careful choice of regulators to avoid introducing new divergences. Lattice regularization discretizes continuous into a hypercubic with finite lattice spacing a, effectively imposing a natural at momenta \sim 1/a; as a \to 0, the limit is recovered. This approach is particularly suited for non-perturbative studies, such as simulations, though it breaks exact chiral symmetry for fermions unless modified. Among these methods, dimensional regularization excels in preserving symmetries and simplifying gauge theory calculations, at the cost of handling \gamma_E and $1/\epsilon poles, whereas the momentum cutoff offers physical intuition for the scale of new physics but disrupts invariances, making it less ideal for renormalizable theories. Pauli-Villars and lattice methods provide alternatives for specific contexts, balancing symmetry preservation with computational feasibility.

Renormalization Procedure

In , the renormalization procedure addresses ultraviolet divergences by redefining the theory's parameters such that bare quantities, which are infinite, are related to finite, physical renormalized parameters through renormalization factors that include counterterms to absorb the infinities. For instance, the bare mass m_0 and bare charge e_0 are expressed as m_0 = Z_m m and e_0 = Z_e e, where the renormalization constants Z_m = 1 + \delta_m and Z_e = 1 + \delta_e (dimensionless in multiplicative schemes) incorporate counterterms \delta_m, \delta_e designed to cancel divergent contributions in perturbative calculations. This redefinition ensures that observable quantities remain finite and match experimental measurements. The procedure unfolds in a systematic manner, beginning with regularization to render divergent integrals finite temporarily, followed by the computation of perturbative expansions that include counterterms to offset divergences at each order. A key step is the selection of a renormalization scheme to determine the finite portions of the counterterms; in the on-shell scheme, counterterms are fixed by requiring that physical parameters, such as the pole mass or residue of the , equal their measured values, thereby directly tying the theory to experiment. Alternatively, the \overline{\rm MS} (modified minimal subtraction) scheme subtracts only the divergent poles plus associated constant terms in , prioritizing mathematical simplicity over direct physical interpretation, though it requires additional relations to connect to observables. Within this framework, the encodes the scale dependence of parameters through the \beta(g) = \mu \frac{dg}{d\mu}, which dictates how the coupling g evolves with the energy scale \mu; negative values of \beta(g) for weak couplings signal , as realized in where the strong coupling diminishes at high energies. Theories are classified as renormalizable if divergences can be fully absorbed by redefining a finite number of parameters—such as the four in (electron mass, charge, and effects)—yielding predictions independent of the regularization cutoff. In contrast, non-renormalizable theories demand infinitely many counterterms, manifesting as effective field theories where higher-dimensional operators become relevant at low energies but require completion for consistency at all scales. The scale invariance of physical processes is captured by the Callan-Symanzik equation, \left( \mu \frac{\partial}{\partial \mu} + \beta(g) \frac{\partial}{\partial g} - \gamma_m m \frac{\partial}{\partial m} \right) \Gamma = 0, where \Gamma is the effective action, \beta(g) is the beta function, and \gamma_m is the anomalous dimension of the mass, providing a differential equation that governs how correlation functions transform under changes in the renormalization scale. The success of renormalization lies in its restoration of predictive power: in renormalizable theories, all ultraviolet divergences are systematically eliminated using a finite set of measurable parameters, allowing finite, scheme-independent predictions for scattering amplitudes and other observables once the procedure is applied order by order in perturbation theory.

Examples and Applications

Quantum Electrodynamics

is the relativistic describing the interactions of charged leptons, such as the , with the mediated by ; it is an Abelian characterized by the dimensionless \alpha = e^2 / (4\pi) \approx 1/137, where e is the . In perturbative QED, ultraviolet (UV) divergences emerge from loop integrals in Feynman diagrams, reflecting the theory's incomplete treatment of high-momentum virtual particles; these divergences appear in key processes including the self-energy (leading to mass renormalization), the electron-photon vertex correction (contributing to charge renormalization), and the photon vacuum polarization (modifying the propagator). A concrete example of UV divergence in is the one-loop correction to the electron self-energy, which yields a divergent mass shift given by \delta m \sim \frac{\alpha}{\pi} m \log\left(\frac{\Lambda}{m}\right), where m is the and \Lambda is the ultraviolet cutoff scale; this logarithmic divergence is absorbed through a counterterm \delta m in the renormalized , ensuring finite physical predictions. This self-energy correction also generates the leading contribution to the electron's anomalous , a_e = (g-2)/2 = \alpha/(2\pi) + \mathcal{O}(\alpha^2), where g is the , providing a measurable test of the theory's . Renormalization in QED involves three independent counterterms: one for the electron mass, one for the electron field strength Z_2, and one for the photon field strength Z_3, with the vertex renormalization Z_1 equaling Z_2 due to Ward's identity; the theory is renormalizable in four dimensions, featuring only logarithmic UV divergences at all loop orders rather than power-law growth. These counterterms systematically cancel the divergences, rendering QED predictive to arbitrarily high orders in \alpha. Experimental validations confirm the efficacy of in handling UV issues: the , a splitting in energy levels arising from and effects, was theoretically computed using techniques and matched the 1947 measurement by Lamb and Retherford to within experimental precision, marking an early triumph of the approach. Similarly, the anomalous a_e has been measured to over 10 decimal places, aligning with predictions including higher-loop renormalized contributions up to energies probed by current accelerators. Despite these successes, QED's perturbative validity is limited by the running of the coupling \alpha(\mu), which increases with energy scale \mu due to vacuum polarization; this leads to a Landau pole, where \alpha diverges asymptotically at an extraordinarily high scale \sim 10^{280} GeV, signaling a breakdown of perturbation theory, though the theory remains reliable up to TeV scales accessible in contemporary experiments.

Quantum Chromodynamics and Beyond

Quantum Chromodynamics (QCD) is a non-Abelian gauge theory based on the SU(3)_c symmetry group, describing the interactions of quarks mediated by gluons, which carry color charge and self-interact due to the non-Abelian nature of the gauge group. This structure leads to ultraviolet (UV) divergences in perturbative calculations arising from loop diagrams involving gluon propagators and vertices, which are more intricate than in Abelian theories like QED owing to the triple and quartic gluon couplings. However, QCD exhibits asymptotic freedom, characterized by a negative beta function β(g) < 0 for the gauge coupling g at high energies, enabling perturbative control in short-distance processes despite the strong coupling regime at low energies. This property, discovered independently by Gross and Wilczek and by Politzer, allows for reliable high-energy predictions after renormalization absorbs the UV divergences in a gauge-invariant manner, often using dimensional regularization or momentum subtraction schemes. The renormalization of QCD involves running parameters, such as the quark masses and the strong coupling α_s, which evolve with the energy scale Q according to the renormalization group equations. At one loop, the strong coupling runs as α_s(Q) = 4π / [b \ln(Q^2 / \Lambda_{\rm QCD}^2)], where b = 11 - 2n_f / 3, with n_f the number of active quark flavors and \Lambda_{\rm QCD} the QCD scale parameter. These divergences are canceled order by order in perturbation theory, preserving gauge invariance, and enable computations like the deep inelastic scattering (DIS) cross-sections, where quark-gluon interactions contribute to structure functions with UV divergences systematically removed through renormalization. For instance, in DIS experiments at high Q^2, perturbative QCD accurately predicts scaling violations in parton distribution functions after accounting for these renormalized contributions. In the electroweak sector of the , based on the SU(2)_L × U(1)_Y gauge group with via the , UV divergences manifest in loop corrections to self-energies, vertex functions, and Yukawa couplings. The theory is renormalizable to all orders, as demonstrated by 't Hooft and Veltman using to handle these divergences while maintaining gauge invariance and unitarity at low orders. However, higher-order calculations reveal sensitivities to UV physics, particularly quadratic divergences in the Higgs mass term, which grow with the scale and challenge the stability of the electroweak scale without new physics. Extensions beyond the , such as Grand Unified Theories (GUTs) that embed SU(3)_c × SU(2)_L × U(1)_Y into a larger simple gauge group like SU(5), introduce additional UV challenges in the Higgs sector, where quadratic divergences from heavy gauge bosons and scalars destabilize the electroweak hierarchy. This motivates , which pairs bosons and fermions to cancel quadratic divergences loop by loop, stabilizing the Higgs mass and providing a natural UV completion for GUT-scale physics. In supersymmetric GUTs, these cancellations ensure perturbative unitarity up to the unification scale while accommodating the observed electroweak parameters.

Broader Implications

UV Completion and Effective Theories

Ultraviolet divergences in quantum field theories indicate that the theory is an effective description valid only up to some energy scale, necessitating a ultraviolet (UV) completion—a more fundamental theory that remains finite and consistent at all energies while reproducing the low-energy effective field theory (EFT) in the appropriate limit. This completion resolves the divergences by providing a natural cutoff, often associated with new physics such as additional particles or spacetime structure at high energies. For instance, the Standard Model of particle physics is believed to serve as an EFT valid up to approximately $10^{16} GeV, the grand unification (GUT) scale, beyond which new dynamics, like those in grand unified theories, must emerge to maintain consistency. Effective field theories formalize this incompleteness by integrating out high-energy , yielding an with local s of increasing , where higher-dimensional (non-renormalizable) terms are suppressed by inverse powers of a heavy M. In the Wilsonian framework, this involves successively integrating out momentum shells between a UV \Lambda and an infrared , generating an EFT valid below \Lambda as an approximation to the full . A historical example is the Fermi of weak interactions, an EFT below the electroweak where the four-fermion has G_F \sim 1/M_W^2, with M_W the W-boson ; this exhibits UV divergences that were resolved by the electroweak unification in the , providing its UV completion. The exemplifies the challenges posed by UV divergences in EFTs, particularly for scalar fields like the , whose mass-squared parameter receives quadratically divergent radiative corrections \delta m_H^2 \sim \Lambda^2 / (16\pi^2), where \Lambda is the UV . Without , this sensitivity implies that the physical Higgs mass (around 125 GeV) would be unnaturally close to \Lambda unless \Lambda is low (TeV ) or the divergences cancel, as in where bosonic and fermionic loop contributions to scalar masses mutually cancel to leading order, stabilizing the hierarchy up to the supersymmetry-breaking . Such mechanisms motivate extensions beyond the to address naturalness. Prominent candidates for UV completions include , which embeds quantum field theories in a finite, perturbative framework where interactions remain well-defined up to the Planck scale ($10^{19} GeV), with the string length providing a natural regulator for UV divergences. offers a non-perturbative alternative, quantizing geometry to yield a discrete structure at the Planck scale that completes and potentially field theories without singularities or divergences. These approaches underscore how UV completions not only cure divergences but also unify disparate physical regimes.

Non-Perturbative Perspectives

Perturbation theory in often encounters ultraviolet (UV) divergences that manifest as infinities in integrals, but its series expansions may fail to converge, particularly in strongly coupled regimes where effects dominate. For instance, in (QCD), phenomena like confinement arise from dynamics that evade simple perturbative treatments of UV issues, necessitating alternative frameworks to capture the full theory without relying on asymptotic expansions. Lattice QCD provides a non-perturbative approach by discretizing on a finite with spacing a, which inherently introduces a UV cutoff \Lambda \sim 1/a, regulating divergences from short-distance fluctuations. This allows numerical simulations to compute quantities like masses and decay constants directly from the , bypassing perturbative infinities; the continuum limit is then approached by extrapolating a \to 0 while keeping physical scales fixed. Such methods have successfully reproduced experimental results for spectroscopy, demonstrating the viability of non-perturbative UV regulation in QCD. In the holographic duality framework, known as the AdS/CFT correspondence, UV divergences in the boundary (CFT) map to infrared (IR) effects in the bulk anti-de Sitter () gravity theory, where infinities are resolved through finite dynamics of strings or branes in higher dimensions. This duality enables computations of correlation functions and loops in strongly coupled gauge theories, effectively taming UV singularities by embedding them in a UV-complete gravitational description. For example, in \mathcal{N}=4 super Yang-Mills theory, holographic methods yield exact results for scattering amplitudes that align with perturbative limits but extend to strong coupling without divergences. Supersymmetry offers non-perturbative cancellations of UV divergences by balancing bosonic and fermionic contributions, leading to exact results in certain theories; in \mathcal{N}=4 super Yang-Mills, the beta function vanishes to all orders, ensuring UV finiteness beyond perturbation theory. Techniques like localization in supersymmetric theories compute partition functions exactly on curved spaces, revealing non-perturbative structures that regulate divergences holistically. Modern developments since the 2000s, including resurgence theory, integrate perturbative series with non-perturbative contributions to resum asymptotic expansions, providing a framework to handle UV divergences in quantum field theories with instantons or other saddles. In integrable models like the \mathcal{N}=2^* deformation of \mathcal{N}=4 SYM, resurgence uncovers transseries solutions that capture both weak and strong coupling behaviors, improving convergence and addressing perturbative shortcomings. Applications in lattice simulations for Higgs physics, such as computing the Higgs mass in the Standard Model via non-perturbative methods, further highlight these advances in bridging UV regulation with observable predictions.

References

  1. [1]
    [PDF] Quantum Field Theory
    ... ultra-violet divergence. This divergence arises because of our hubris. We've assumed that our theory is valid to arbitrarily short distance scales ...
  2. [2]
    [PDF] Renormalization - UMD Physics
    The ultraviolet divergences in both quantum mechanics and quantum field theory can be viewed as a consequence of the fact that we took the continuum limit ...
  3. [3]
    [PDF] Systematics of Renormalization
    ultraviolet divergences that can appear in a quantum field theory. Rather than stumbling across these divergences one by one and repairing them case by case,.
  4. [4]
  5. [5]
  6. [6]
  7. [7]
    [PDF] UV Regularization Schemes - UT Physics
    also called the UV cutoffs — which render the divergent Feynman diagrams finite: 1.Missing: definition | Show results with:definition
  8. [8]
    An Introduction To Quantum Field Theory - 1st Edition - Michael E. Pes
    In stock Rating 5.0 1 An Introduction To Quantum Field Theory. By Michael E. Peskin, Daniel V. Schroeder Copyright 1995. Paperback $76.49. Hardback $144.50. eBook $71.99. ISBN ...
  9. [9]
    [PDF] 13 Perturbation Theory, Regularization and Renormalization
    In particular, for D = 4 the first two terms diverge in the UV. These are, of course the quadratic UV divergence of the self-energy and the logarithmic UV ...
  10. [10]
    [PDF] 7.3 The electron self energy
    When k is large, the integral behaves as R d4k/k4 which gives a logarithmic. UV divergence. In addition, in the limit x → 1 we see that ∆ → 0 which.
  11. [11]
    [PDF] Physics 198-730B: Quantum Field Theory
    ... degree of divergence, D: D = 4L − 2I = 4V − 2E + 4 − 4V + E = 4 − E. (59). D is the minimum power of Λ with which a diagram with E external legs can be ...<|control11|><|separator|>
  12. [12]
    [0901.2208] Radiative Corrections, Divergences, Regularization ...
    Jan 15, 2009 · The present lectures are a practical guide to the calculation of radiative corrections to the Green functions in quantum field theory.Missing: methods | Show results with:methods
  13. [13]
    Broken Scale Invariance in Scalar Field Theory | Phys. Rev. D
    Oct 15, 1970 · This paper studies broken scale invariance in scalar-field theory, focusing on scaling laws and their anomalies, and using them to compute ...
  14. [14]
    [PDF] The role of UV-completion in the search for quantum gravity - arXiv
    UV-completion is usually considered in the context of QFT, where it is taken to be synonymous with renormalizability. Yet the idea of UV-completion is certainly ...
  15. [15]
    [PDF] arXiv:hep-th/9803132v2 31 Mar 1999
    It is generally assumed that particle physics can be accurately described by an effective field theory with an ultraviolet (UV) cutoff less than the Planck mass ...
  16. [16]
    Wilson Renormalization Group and Continuum Effective Field ... - arXiv
    Oct 8, 1998 · This is an elementary introduction to Wilson renormalization group and continuum effective field theories. We first review the idea of Wilsonian ...
  17. [17]
    [PDF] Lectures on Effective Field Theory - arXiv
    Jun 16, 2015 · So EFT is really just the correct way to do QFT. Unfortunately, it receives rather scant treatment in the QFT textbooks. Fortunately, there are ...Missing: G_F M_W^
  18. [18]
  19. [19]
    [PDF] What is the Hierarchy Problem? - arXiv
    This identification leads to the cancellation of quadratic divergences. The contribution to the Higgs mass term from the top quark is again negative, but the ...
  20. [20]
    [PDF] An elementary introduction to loop quantum gravity - arXiv
    Jul 18, 2016 · Due to its finiteness properties, string theory can thus be considered as a UV-completion for supergravity. Moreover, 11-dimensional ...<|control11|><|separator|>