Fact-checked by Grok 2 weeks ago

Dimensional regularization

Dimensional regularization is a in used to handle ultraviolet divergences in perturbative calculations by analytically continuing the dimension from the physical value of four to a general non-integer d = 4 - 2\epsilon, where \epsilon > 0 is small, and then taking the limit \epsilon \to 0 after isolating divergent poles in the results. This method redefines loop integrals over momenta in d dimensions, employing analytic extensions of functions like the to evaluate them, which manifest divergences as simple poles in \epsilon rather than more complex forms. The poles are then subtracted through counterterms, yielding finite physical predictions. Introduced independently in 1972 by C. G. Bollini and J. J. Giambiagi for general renormalization purposes, as well as by G. 't Hooft and M. Veltman specifically for gauge theories, dimensional regularization quickly became a standard tool due to its mathematical elegance. Independent contributions around the same time came from F. Cicuta and E. Montaldi, and J. F. Ashmore, further solidifying its foundations. Unlike cutoff regularization, which introduces an arbitrary high-energy scale that can violate symmetries, dimensional regularization preserves gauge invariance, Lorentz invariance, and other symmetries of the theory without adding unphysical parameters, making it particularly suitable for non-Abelian gauge theories like . It also automatically handles overlapping divergences and eliminates quadratic divergences in scalar theories, simplifying computations. The technique is widely applied in the for higher-order corrections, effective field theories, and beyond-Standard-Model physics, often in conjunction with schemes like modified minimal subtraction () to define the renormalization procedure. Its use extends to gravitational theories and contexts, where maintaining dimensional consistency is crucial.

Introduction

Definition and

In perturbative (QFT), ultraviolet (UV) divergences emerge from loop integrals evaluated in four-dimensional , where high-momentum contributions cause integrals to diverge, complicating the computation of physical quantities such as scattering amplitudes. These divergences arise particularly in Feynman integrals representing quantum corrections, rendering naive calculations infinite and necessitating regularization techniques to extract finite results. Dimensional regularization addresses these issues by analytically continuing the dimension from four to a general value d = 4 - 2\epsilon, where \epsilon is a small positive that approaches zero at the end of the calculation. This continuation transforms the divergent four-dimensional integrals into finite expressions in d for appropriate \epsilon, with the original UV divergences manifesting as simple poles in \epsilon upon expansion around \epsilon = 0. (IR) divergences can similarly appear as poles, allowing systematic treatment of both types. The primary purpose of dimensional regularization is to preserve key symmetries of QFT, including invariance and Lorentz invariance, more effectively than cutoff-based methods, as the continuation maintains identities and the structure of the theory irrespective of the dimension. It simplifies the process by isolating divergences into explicit poles that can be subtracted without introducing spurious terms or breaking symmetries, facilitating the absorption of infinities into redefined parameters and fields. A common companion to dimensional regularization is the minimal subtraction () scheme, which removes only the divergent poles in \epsilon from the renormalization constants, leaving finite parts untouched to yield scheme-independent physical predictions after further processing. This approach ensures consistency order by order in while avoiding unnecessary subtractions that could complicate higher-order calculations.

Historical Context

Dimensional regularization was introduced independently in 1972 by C. G. Bollini and J. J. Giambiagi for general purposes, as well as by G. 't Hooft and M. Veltman specifically for theories, with further contributions from F. and E. Montaldi, and J. F. Ashmore. This approach addressed the challenges of ultraviolet divergences in perturbative calculations while preserving invariance, enabling the proof of renormalizability for theories like (QCD) and the electroweak sector. During the , dimensional regularization gained widespread adoption in calculations central to the , especially following 't Hooft's demonstration of the renormalizability of non-Abelian gauge theories, which facilitated precise higher-order computations in electroweak unification efforts by researchers such as , , and . Its elegance in handling both and divergences made it indispensable for verifying the consistency of the unified electroweak theory. Key milestones in the technique's evolution include its use in perturbative matching for simulations from the 1980s onward, where it complements discrete methods by providing continuum-limit results to extract physical parameters like masses and coupling constants. This synergy supported advancements in QCD simulations and precision phenomenology.

Mathematical Foundations

Spacetime Dimension Continuation

In dimensional regularization, the dimension is analytically continued from the physical value of 4 to a complex parameter d = 4 - 2\epsilon, where \epsilon is a small positive parameter that controls the (UV) behavior of the theory. This continuation transforms divergent integrals in four dimensions into finite expressions in d dimensions, with divergences manifesting as simple poles of the form $1/\epsilon upon taking the limit \epsilon \to 0^+. The parameter \epsilon is chosen such that $2\epsilon accounts for the even deviation from 4 dimensions, ensuring that the analytic structure aligns with the expected Laurent expansion around d=4. The primary application of this continuation occurs in the momentum-space integrals of Feynman diagrams. A four-dimensional loop \int \frac{d^4 k}{(2\pi)^4} is replaced by \mu^{4-d} \int \frac{d^d k}{(2\pi)^d}, where the integration measure d^d k spans d-dimensional Euclidean or after , and \mu is an arbitrary mass-dimensional scale factor with dimensions of . The factor \mu^{4-d} ensures dimensional consistency, as the measure d^d k has mass dimension d, while physical quantities like constants retain their canonical dimensions only when this adjustment is included; without it, dimensionless quantities in four dimensions would acquire spurious powers of \mu. This prescription preserves the formal structure of the perturbative expansion while rendering UV divergences explicit as poles in \epsilon. For theories involving fermions, such as or the , the Dirac matrices and spinors must also be defined in d dimensions. The gamma matrices \gamma^\mu (\mu = 0, 1, \dots, d-1) satisfy the Clifford algebra anticommutation relation \{\gamma^\mu, \gamma^\nu\} = 2 g^{\mu\nu}, where g^{\mu\nu} is the Minkowski in d dimensions with signature (+,-,-,\dots,-), having one time-like and d-1 space-like components. Spinor fields are represented in a whose dimension is analytically continued from 4 (for Dirac spinors in four dimensions) to a f(d) such that f(4) = 4, ensuring traces like \mathrm{Tr}(\mathbb{1}) = f(d) recover the physical value at d=4. This continuation maintains Lorentz invariance in d dimensions and allows consistent evaluation of fermion propagators and vertices, with IR divergences manifesting as poles in $1/\epsilon when \epsilon \to 0 from negative values (i.e., d > 4). The propagators in the are adjusted accordingly under this dimensional shift. For a , the takes the form i / (k^2 - m^2 + i\epsilon') in , where the denominator remains formally unchanged, but the overall becomes \mu^{4-d} \int \frac{d^d k}{(2\pi)^d} \frac{1}{k^2 - m^2 + i\epsilon'}, leading to non-integer powers in the evaluation (though explicit computation is deferred). For gauge fields, the polarization sum is contracted with the d-dimensional , -g^{\mu\nu} + \frac{k^\mu k^\nu}{k^2} in Feynman gauge, preserving transversality. propagators i (\slashed{k} + m) / (k^2 - m^2 + i\epsilon') involve the slashed notation \slashed{k} = k_\mu \gamma^\mu, which contracts over d indices. These adjustments ensure that the remains gauge-invariant and in d dimensions, with the scale \mu later absorbed into the procedure.

Role of the Gamma Function

In dimensional regularization, the Gamma function arises naturally when evaluating momentum integrals in non-integer spacetime dimensions, providing an analytic continuation that isolates ultraviolet (UV) divergences as poles. Many Feynman integrals, after Wick rotation to Euclidean space and use of Feynman parametrization, reduce to forms whose results involve ratios of Gamma functions, such as \Gamma(\epsilon) or \Gamma(2 - d/2), where d = 4 - 2\epsilon is the continued dimension and \epsilon \to 0^+. These poles emerge because \Gamma(z) \sim 1/z as z \to 0, corresponding to divergences at integer dimensions like d=4. A prototypical example is the one-loop integral \int \frac{d^d k}{(2\pi)^d} \frac{1}{(k^2)^a}, which evaluates to (4\pi)^{-d/2} \frac{\Gamma(a - d/2)}{\Gamma(a)} \mu^{2(a - d/2)}, where \mu is an arbitrary mass scale introduced to track dimensions. This formula, derived via hyperspherical coordinates and the representation of the radial integral in terms of Gamma functions, holds for \mathrm{Re}(a) > d/2 and is analytically continued thereafter. The Gamma function in the numerator captures the UV behavior, producing a pole when a - d/2 \to 0, while the denominator ensures normalization. The Laurent expansion of the Gamma function around small arguments further separates divergent and finite contributions: \Gamma(\epsilon) = \frac{1}{\epsilon} - \gamma_E + O(\epsilon), where \gamma_E \approx 0.577 is the Euler-Mascheroni constant. This expansion allows the divergent $1/\epsilon term to be isolated and subtracted systematically during renormalization, while the finite part, including -\gamma_E, contributes to the renormalized quantities without introducing unphysical parameters. The role of the is crucial for maintaining gauge invariance and renormalizability in quantum theories, as it enables subtractions that preserve symmetries without arbitrary cutoffs or regulators that might break them. This analytic structure facilitates the analysis and higher-order calculations by providing a uniform treatment of divergences.

Computational Procedure

Evaluating Feynman Integrals

In dimensional regularization, the evaluation of Feynman integrals begins with a Wick rotation to transform the Minkowski space integral into a convergent Euclidean form. This step replaces the integration measure \int d^d k (in Minkowski signature) with i \int d^d k_E, where k_E denotes the Euclidean momentum, ensuring the integral converges for dimensions d < 2\alpha in scalar propagators of the form $1/(k^2 + \Delta)^\alpha. The rotation avoids oscillatory behavior in the time component and facilitates analytical continuation, a procedure standard in perturbative quantum field theory calculations. To handle multiple propagators in loop integrals, Feynman parameters are introduced to combine denominators into a single expression. For two propagators, the identity \frac{1}{A B} = \int_0^1 dx \frac{1}{[x A + (1-x) B]^2} is applied, generalizing to N propagators via the multinomial form \frac{1}{\prod_{i=1}^N A_i^{\alpha_i}} = \frac{\Gamma(\sum \alpha_i)}{\prod \Gamma(\alpha_i)} \int d^N x \delta(1 - \sum x_i) \frac{1}{[\sum x_i A_i]^{\sum \alpha_i}}, where x_i are parameters satisfying \sum x_i = 1. After parameter integration, the momentum integral simplifies to a Gaussian form amenable to evaluation. This technique, essential for reducing complexity in one- and multi-loop diagrams, preserves the dimensional continuation. The resulting scalar integral after Wick rotation and parameterization takes the standard form for a one-loop tadpole or bubble: I = \int \frac{d^d k_E}{(2\pi)^d} \frac{1}{(k_E^2 + \Delta)^\alpha} = \frac{1}{(4\pi)^{d/2}} \frac{\Gamma\left(\alpha - \frac{d}{2}\right)}{\Gamma(\alpha) \Delta^{\alpha - d/2}}, valid for \operatorname{Re}(d) < 2\alpha and extended analytically. For the bubble topology with two propagators, the parameter integral yields a similar structure with \Delta depending on external momenta and masses. The arises naturally from the volume of the d-dimensional in the angular integration. Divergences manifest as poles in the Laurent expansion around d = 4, parameterized as d = 4 - 2\epsilon with \epsilon \to 0. The integral expands as I = \frac{1}{\epsilon} c_{-1} + c_0 + \epsilon c_1 + \cdots, where coefficients c_i are computed via series expansion of the Gamma functions and other terms, isolating ultraviolet divergences as simple poles while finite parts provide physical contributions. This pole structure enables systematic renormalization. The role of the Gamma function in generating these poles is central, as detailed in the mathematical foundations.

Handling UV and IR Divergences

In dimensional regularization, ultraviolet (UV) divergences arise in Feynman integrals as simple poles of the form $1/\epsilon in the Laurent expansion around \epsilon = 0, where the spacetime dimension is continued to d = 4 - 2\epsilon with \epsilon > 0. These poles originate from terms involving the , such as \Gamma(\epsilon) \approx 1/\epsilon - \gamma_E + O(\epsilon), where \gamma_E is the Euler-Mascheroni constant, during the evaluation of loop integrals. The subtraction of these UV poles is achieved through counterterms in schemes like or modified (\overline{\text{MS}}). In the MS scheme, only the pure $1/\epsilon poles are subtracted, while the \overline{\text{MS}} scheme additionally removes universal constants \ln(4\pi) - \gamma_E to improve perturbative convergence and scheme independence in higher orders. Infrared (IR) divergences, common in massless theories, are regulated separately by analytic continuation to d > 4 (effectively \epsilon < 0), manifesting as poles $1/\epsilon_{\text{IR}} with residues often opposite in sign to UV poles. This unified treatment with the same regularization parameter \epsilon (now distinguishing \epsilon_{\text{UV}} and \epsilon_{\text{IR}} via the sign) allows both types of singularities to appear as Laurent poles, facilitating their identification in the expansion. Unlike UV poles, which are absorbed into local counterterms, IR divergences typically cancel between virtual and real emission contributions in inclusive observables, such as decay rates or cross-sections. The renormalization procedure absorbs these poles into renormalization factors Z, ensuring finite physical predictions. For example, in , the one-loop wave function renormalization for the electron field is Z_\psi = 1 - \frac{\alpha}{4\pi \epsilon} in Feynman gauge (\alpha the ), where the pole term subtracts the UV divergence from the self-energy diagram. Scheme dependence arises in the finite parts: yields slower convergence due to retaining \ln(4\pi) - \gamma_E, whereas \overline{\text{MS}} subtracts these for more rapid series convergence in applications like precision electroweak calculations. This approach, introduced by 't Hooft and Veltman, preserves gauge invariance and symmetries without introducing unphysical parameters.

Examples

Electrostatic Potential of an Infinite Line Charge

The electrostatic potential due to an infinite line charge with uniform linear charge density λ along the z-axis can be expressed in momentum space as the Fourier transform φ(r) = ∫ \frac{d^3 k}{(2\pi)^3} \frac{e^{i \mathbf{k} \cdot \mathbf{r}}}{k^2}, where r is the perpendicular distance from the line. This integral diverges logarithmically in three spatial dimensions due to the ultraviolet behavior at large k, reflecting the infinite self-energy of the charge distribution. To apply dimensional regularization, the spatial dimension is continued to d dimensions, effectively treating the transverse space as d-1 dimensions for the line charge geometry. The integral becomes ∫ \frac{d^{d-1} k}{(2\pi)^{d-1}} \frac{e^{i \mathbf{k} \cdot \mathbf{r}}}{k^2}, which evaluates to a form proportional to the Gamma function term Γ(1 - (d-1)/2) through standard techniques involving the representation of the denominator and angular integration in general dimensions. Setting d = 3 - ε with ε → 0^+, the regularized potential takes the form φ(r) ∝ 1/r^{1 - ε/2}, where the ε dependence introduces a that isolates the . In the limit ε → 0, this recovers the physical logarithmic potential φ(r) ∼ - (λ / (2π ε_0)) \ln(r / r_0), with r_0 a reference , as the divergent 1/ε term combines with the ε \ln r contribution to yield the logarithm. This regularization demonstrates that the logarithmic is intrinsic to the two-dimensional of the , analogous to the behavior of point charges in , and the method preserves translational invariance without introducing unphysical artifacts.

One-Loop Propagator Correction

In , the one-loop correction to the is given by the function \Pi(p^2), computed from the corresponding to the integral \Pi(p^2) = \int \frac{d^d k}{(k^2 - m^2)((p - k)^2 - m^2)}, where the dimension is continued to d = 4 - 2\epsilon to regulate divergences using dimensional regularization. To evaluate this integral, is applied to combine the denominators: \frac{1}{(k^2 - m^2)((p - k)^2 - m^2)} = \int_0^1 dx \frac{1}{[x(k^2 - m^2) + (1 - x)((p - k)^2 - m^2)]^2}. after shifting the loop momentum k \to l + x p yields \Pi(p^2) = \int_0^1 dx \int \frac{d^d l}{[l^2 + \Delta]^2}, with \Delta = m^2 - x(1 - x) p^2. The inner momentum integral is then performed using the standard formula in d dimensions, involving the \Gamma(2 - d/2). Expanding around \epsilon \to 0, the gives \Pi(p^2) = \frac{1}{16\pi^2} \int_0^1 dx \left[ \frac{2}{\epsilon} - \ln\left(\frac{m^2 - x(1-x) p^2}{\mu^2}\right) + \text{finite terms} \right], where \mu is the renormalization scale introduced to maintain the coupling's dimensionality. This expression reveals the as a simple $1/\epsilon, which in the limit \epsilon \to 0 corresponds to the logarithmic expected in four dimensions for this . Renormalization proceeds by subtracting the divergent $2/\epsilon pole via counterterms in the bare , resulting in a finite correction. This finite remainder modifies the while preserving the unitarity of the underlying , as the imaginary parts from on-shell intermediate states are unaffected by the regularization scheme.

Comparisons and Alternatives

Versus Regularization

Cutoff regularization addresses ultraviolet divergences in quantum field theory by imposing a hard momentum cutoff \Lambda on loop integrals, effectively replacing the unbounded integration \int \frac{d^4 k}{(2\pi)^4} with a restricted form such as \int_{|k| < \Lambda} \frac{d^4 k}{(2\pi)^4}, where \Lambda \to \infty after renormalization. This approach leads to explicit power-law divergences; for instance, the tadpole integral in scalar field theory yields a quadratic divergence of the form \frac{\Lambda^2}{16\pi^2}. These power divergences reflect the sensitivity to high-energy scales but introduce artifacts that must be subtracted order by order. Dimensional regularization offers key advantages over cutoff methods by avoiding explicit symmetry breaking. Unlike cutoff regularization, which can violate Lorentz invariance and gauge symmetry due to the non-covariant boundary in momentum space, dimensional regularization maintains these symmetries, including chiral invariance in theories with massless fermions, as the analytic continuation to d = 4 - 2\epsilon dimensions preserves the structure of the without introducing unphysical mass terms. Moreover, divergences appear solely as simple poles in \epsilon, such as \frac{1}{\epsilon}, facilitating systematic subtraction via schemes like minimal subtraction (\overline{\rm MS}), without the quadratic or higher power terms inherent in cutoff procedures. Despite these benefits, dimensional regularization has drawbacks compared to cutoff methods. The reliance on non-integer dimensions introduces \epsilon-dependent finite terms that require scheme-specific handling, such as absorbing them into the renormalization scale \mu in the \overline{\rm MS} scheme, which can complicate comparisons across schemes. Additionally, it is primarily suited for perturbative calculations and proves challenging for non-perturbative applications, like lattice simulations, where a physical cutoff is more natural. A notable difference arises in quantum electrodynamics (QED). Cutoff regularization breaks gauge invariance, necessitating additional counterterms to restore Ward identities, as the momentum cutoff introduces terms that do not respect the transversality of the photon propagator. In contrast, dimensional regularization inherently preserves gauge invariance without such extras, allowing direct evaluation of Feynman diagrams while maintaining the theory's symmetries.

Versus Pauli-Villars Method

The Pauli-Villars (PV) regularization method, introduced in 1949, addresses ultraviolet divergences in quantum field theory by modifying the propagators in Feynman integrals through the introduction of auxiliary regulator fields with large masses. Specifically, a scalar propagator \frac{1}{k^2 - m^2 + i\epsilon} is replaced by a weighted sum \sum_{i=0}^{N} (-1)^i \frac{1}{k^2 - M_i^2 + i\epsilon}, where M_0 = m is the original mass, the M_i for i \geq 1 are much larger regulator masses taken to infinity in the limit, and N is chosen sufficiently large (often N=1 or $2) to ensure convergence by subtracting divergent contributions quadratically through the alternating signs. In contrast to dimensional regularization, which analytically continues spacetime dimensions to d = 4 - \epsilon and isolates divergences as simple poles in \epsilon, the PV approach explicitly adds fictitious heavy fields that alter the structure of the . These regulator fields generally violate the original symmetries of the , such as chiral or invariance, unless their couplings are finely tuned to restore them, which adds significant complexity especially in theories where maintaining identities is crucial.90239-6) Dimensional regularization sidesteps this issue by preserving all symmetries manifestly through its dimensional continuation, yielding cleaner expressions for counterterms without introducing unphysical particles. While PV regularization effectively suppresses divergences by the large-mass subtraction, it often leads to more cumbersome calculations in perturbative expansions, particularly beyond one loop, due to the need to track multiple regulator contributions and their interactions. This contrasts with dimensional regularization's use of identities for efficient evaluation of integrals, making the latter computationally preferable in modern applications. Historically, predates dimensional regularization by over two decades but has become less favored in contemporary perturbative owing to these practical complications in symmetry-sensitive theories like and non-Abelian gauge theories.

Applications in Quantum Field Theory

In Quantum Electrodynamics

Dimensional regularization plays a crucial role in (QED) by enabling the computation of radiative corrections while maintaining the theory's gauge invariance. It is applied to evaluate loop diagrams that would otherwise diverge, such as those contributing to the and vertex function, allowing for systematic in schemes like on-shell or minimal subtraction (MS). A primary example is the one-loop , which modifies the and requires of the and . In dimensional regularization, the ultraviolet divergent part takes the form \Sigma(p) = -\frac{e^2}{16\pi^2 \epsilon} (\slashed{p} - 4m) + \text{finite terms}, where \epsilon = (4 - d)/2 and d is the spacetime dimension. This pole structure isolates the divergence, and in the on-shell renormalization scheme, counterterms are chosen to satisfy \Sigma(\slashed{p} = m) = 0 and \frac{\partial \Sigma}{\partial \slashed{p}} \big|_{\slashed{p} = m} = 0, ensuring the physical electron mass and normalization are finite and gauge-invariant. The one-loop vertex correction, involving the electron-photon interaction, also features a UV divergence that dimensional regularization manifests as a $1/\epsilon pole. In cutoff regularization, this appears as a logarithmic term (\alpha / 2\pi) \ln(\Lambda / m), but the pole subtraction in dimensional regularization yields a finite result identical to the regulated logarithm in the appropriate limit. This finite contribution provides the leading-order anomalous of the , a_e = (g-2)/2 = \alpha / (2\pi). Dimensional regularization facilitates the computation of the β-function, describing the scale dependence of the \alpha. In the MS scheme, the pole in the constant from the propagator leads to the one-loop β-function \beta(\alpha) = \frac{2\alpha^2}{3\pi}.) A key advantage of dimensional regularization in is its automatic preservation of identities, which equate the vertex to the self-energy correction (Z_1 = Z_2) and ensure gauge invariance without introducing spurious terms, in contrast to methods that can violate these relations unless carefully adjusted.

In Non-Abelian Gauge Theories

Dimensional regularization plays a crucial role in non-Abelian gauge theories, such as (QCD), where the self-interactions of s introduce complex (UV) and (IR) divergences in perturbative calculations. In SU(N) gauge theories, the method preserves the non-Abelian structure, allowing for the systematic of Feynman diagrams involving colored particles. The regularization extends to d = 4 - 2ε dimensions, handling both UV poles (ε → 0^+) and potential IR singularities through . This framework is essential for computing higher-order corrections in processes involving gluon exchanges and quark-gluon interactions. A key example is the one-loop gluon self-energy, which receives contributions from quark loops, triple-gluon vertices, four-gluon vertices, and ghost loops due to the non-Abelian nature of the theory. The ghost loops are particularly important, as they ensure gauge invariance and cancel unphysical degrees of freedom. In dimensional regularization, the transverse part of the self-energy tensor takes the form \Pi_{\mu\nu}^{ab}(p) = \delta^{ab} \left( g_{\mu\nu} p^2 - p_\mu p_\nu \right) \frac{g^2 C_A}{16\pi^2} \frac{5/3}{\varepsilon}, where C_A = N for SU(N), and this represents the pure gluonic (including ghost) contribution to the UV pole; the full expression also includes a fermion loop term proportional to the number of flavors n_f. This result arises from evaluating the loop integrals in d dimensions and extracting the 1/ε divergence, demonstrating how dimensional regularization isolates the non-Abelian effects without introducing gauge artifacts. Dimensional regularization maintains the Slavnov-Taylor identities, which generalize identities to non-Abelian theories and ensure the consistency of at the quantum level. These identities relate Green's functions and are preserved because the regularization scheme treats all dimensions equally, avoiding the introduction of gauge-dependent parameters that could break the . This compatibility enables the use of in d dimensions, where the BRST operator remains nilpotent, facilitating the proof of renormalizability and the construction of physical elements free from unphysical polarizations. One of the landmark applications is the demonstration of in QCD, where the running coupling decreases at high energies. Using dimensional regularization in the modified minimal subtraction () scheme, the one-loop is computed from the poles in the constants of the gauge coupling, yielding \beta(g) = -\frac{11 C_A - 2 n_f}{3} \frac{g^3}{16\pi^2}, with the negative sign indicating that the theory becomes weakly coupled in the . This result, derived from the UV divergences in and loops, underpins the perturbative validity of QCD at short distances and explains phenomena like the scaling violations in . Despite these successes, dimensional regularization faces challenges with IR divergences in non-perturbative regimes, such as hadronization processes where soft emissions lead to collinear and soft singularities that signal confinement. In perturbative QCD, these IR issues manifest as poles in ε from massless propagators, requiring careful resummation or input; often, dimensional regularization is combined with simulations to match perturbative results to hadronic matrix elements, providing a hybrid approach to compute observables like fragmentation functions.

References

  1. [1]
    [PDF] Institute for Theoretical Physics *, University of Utrecht
    Abstract: A new regularization and renormalization procedure is presented. It is particularly well suited for the treatment of gauge theories.
  2. [2]
    [PDF] Dimensional regularization, Wilsonian RG, and the Naturalness ...
    Apr 29, 2022 · We begin with a thorough analysis of dimensional regularization, usually considered as a useful calculation technique deprived of any direct ...
  3. [3]
  4. [4]
    [PDF] Physics 570 — Handout on Dimensional Regularization
    Consider the following integral in n dimensions with a Euclidian metric: I1 ≡ Z dnk. 1. (k2 + a2)r . We may evaluate this making in terms of the Γ function:.Missing: d^ dk /
  5. [5]
    Regularization, Renormalization, and Dimensional Analysis - arXiv
    Dec 18, 2008 · We then introduce a Minimal Subtraction (MS) and a Modified Minimal Subtraction (MS-Bar) scheme to renormalize the result.
  6. [6]
  7. [7]
  8. [8]
    [PDF] arXiv:hep-th/9206038v1 9 Jun 1992
    We now show that this same strategy may be implemented using real- space dimensional regularization and renormalization. ... Then the one-loop self-energy ...
  9. [9]
    [PDF] 7.3 Dimensional regularization - Jefferson Lab Indico
    When we introduced cutoff regularization, we denoted a cutoff-regularized field φ by φΛ, to distinguish it from a classical field. To be consistent, here we.
  10. [10]
    [PDF] arXiv:2009.00728v2 [hep-th] 16 Jun 2021
    Jun 16, 2021 · Using a cutoff adds a mass dimension to the theory which was not there originally. If we were investigating physical observables in this theory, ...
  11. [11]
    [PDF] MITOCW | 6. Chiral Lagrangians
    I like to dimensional regularization, and so do you. Dimensional regularization here preserves chiral symmetry, which is something that we want to preserve.
  12. [12]
    [PDF] Lecture 23 Regularization II: Dimensional Regularization
    An Introduction to Quantum Field Theory, M. Peskin and D. Schroeder, Section. 10.2. • Quantum Field Theory, by C. Itzykson and J. Zuber, Section 8.1.2.
  13. [13]
    [PDF] QFT Lecture 6 - Canberra International Physics Summer School
    Pauli-Villars regularization: The essence of Paul-Villars regularization (Pauli and Villars, 1949) is to subtract from any loop integral the same loop ...<|separator|>