8
$\begingroup$

I'm currently working through Coleman's lectures on Quantum field theory and I cannot quite get my head around the whole concept of renormalization.

What I am aware of, is that you have some Lagrangian and you want to impose some normalization conditions on the fields (e.g. $< 0|\phi(0)|0>=0$). Since this cannot be satisfied as is, the Lagrangian has to be split up into a physical part which contains all terms with the physical variables (e.g. the physical masses, the physical coupling constant), and a part which contains counterterms. These can then be chosen such as to satisfy the needed normalization conditions.

Here is now my question: What is the interpretation of the variables, like masses and coupling constants, before the splitting of the Lagrangian? I don't quite get why they should not be already the physical quantities, since if for example I have a classical Lagrangian for a harmonic oscillator, the mass and frequency appearing there are simply the physical mass and the physical frequency. Why is it different in QFT? And what is the precise meaning of the variables before renormalization?

$\endgroup$
2
  • 3
    $\begingroup$ hopefully useful: physics.stackexchange.com/q/693507/84967 $\endgroup$ Commented Mar 30 at 21:51
  • $\begingroup$ Wow thanks a lot! This rwally changes how I view the whole subject of renormalization. $\endgroup$ Commented Mar 31 at 18:56

1 Answer 1

5
$\begingroup$

The Lagrangian of a relativistic (quantum) field theory for a certain set of (scalar, vector, spin $1/2$) fields takes the general form (see also https://physics.stackexchange.com/q/846005) $$ \mathcal{L} = \sum\limits_{i=1}^n c_i O_i, \tag{1} \label{1} $$
where \eqref{1} contains all (linear independent) local (hermitean) operators $O_i$ (constructed in terms of products of the basic fields and their derivatives) respecting the necessary space-time symmetries (the action integral $\int\! d^4 x \, \mathcal{L}$ must be invariant under $\mathcal{P}_+^\uparrow$) and possibly also discrete symmetries (like parity $ P$ and $CP$, etc.) and/or internal symmetries (like local or global gauge symmetries). Demanding also renormalizability, the $O_i$ in \eqref{1} are further restricted to the subset of operators with dimension $\le 4$.

In general, the coefficients $c_i$ associated with the operators $O_i$ in \eqref{1} do not correspond to (directly) observable parameters of the theory. However, they can be related to observable quantities by the following procedure: Choose $n$ suitable (independent) experimentally determined (physical) values $a_{i, \rm ph}$ (like masses, cross sections, decay widths, etc.) and compute the corresponding theoretical expressions $a_i(c_1, \ldots, c_n)$ using the (quantum) field theory defined by the Lagrangian \eqref{1}, arriving at the relations $$ a_{i, \rm ph} =a_i(c_1, \ldots, c_n), \quad i=1, \ldots,n. \tag{2} \label{2} $$ In principle, the set of equations in \eqref{2} can be inverted, expressing the original parameters $c_i$ in terms of the measured ones, $$ c_i=c_i(a_{1, \rm ph}, \ldots, a_{n, \rm ph}), \quad i=1,\ldots, n. \tag{3} \label{3} $$ Note that this approach does not necessarily require a quantized field theory but can equally well be applied to a classical model.

In the case of a quantum field theory, one encounters an additional technical problem. The occurrence of ultraviolet divergences generated by large momenta in loop integrals requires the regularization of the theory by introducing a momentum cut-off parameter $\Lambda$. As a consequence, the theoretical expression for the observable $a_i$ is modified to $$ a_i(c_1, \ldots, c_n, \Lambda). \tag{4} \label{4} $$ As \eqref{4} is supposed to reproduce the finite measured values $a_{i, \rm ph}$ in the limit $\Lambda \to \infty$, the coefficients $c_i$ have to be tuned accordingly, $$ c_i \to c_i(\Lambda), \tag{5} \label{5} $$ such that $$ a_{i, \rm ph} = \lim\limits_{\Lambda \to \infty} a_i(c_1(\Lambda), \ldots, c_n(\Lambda), \Lambda) \tag{6} \label{6} $$ holds. In contrast to the physical values $a_{i, \rm ph}$, the coefficients $c_i(\Lambda)$ (being, in general, even divergent in the limit $\Lambda \to \infty$) cannot be regarded as observable parameters of the theory.

As a simple illustration of these somewhat abstract considerations, let us consider the renormalizable relativistic quantum field of a single real scalar field $\phi$ obeying the additional discrete symmetry $\phi \to -\phi$. The associated Lagrangian reads $$ \mathcal{L}= c_0 (\partial_\mu \phi) (\partial^\mu \phi)+ c_1 \phi^2+c_2 \phi^4, \tag{7} \label{7} $$ where the (redundant) parameter $c_0$ can be eliminated by the field rescaling $\phi \to \phi/{\sqrt{2 c_0}}$. With the simultaneous renamings $c_1= -c_0 m^2$ and $c_2= - c_0^2 \lambda/3!$ we arrive at the Lagrangian $$ \mathcal{L}= \frac{1}{2} (\partial_\mu \phi) (\partial^\mu \phi) - \frac{m^2}{2} \phi^2 -\frac{\lambda}{4!} \phi^4 \tag{8} \label{8} $$ of the so-called $\phi^4$ model.

Choosing the physical mass $m_{\rm ph}$ of the associated scalar one-particle state $|p\rangle$ (with $p^2=m_{\rm ph}^2$) and the $\phi(p_1) \phi(p_2) \to \phi(p_3) \phi(p_4)$ scattering amplitude $\lambda_{\rm ph}:= |\mathcal{M}(s_0,t_0)|$ taken at some specific kinematical point characterized by the Mandelstam variables $s_0, t_0$ as our experimental input parameters, it is now our task to determine the corresponding theoretical expressions of these two observables.

Instead of taming the ultraviolet divergences by a "hard" momentum cut-off $\Lambda$ in the loop integrals, it is slightly more convenient to use dimensional regularization by formulating the theory in $d$ space-time dimensions, taking the limit $d\to 4$ at the end.

The squared physical mass $m_{\rm ph}^2$ can be extracted from the two point function \begin{align} \langle 0 |T\phi(x_1) \phi(x_2)|0\rangle &= \frac{1}{i}\int \! \frac{d^dk}{(2\pi)^d} \frac{e^{-ik(x_1-x_2)}}{m^2+\Sigma(k^2)-k^2-i \epsilon} \tag{9} \label{9} \end{align} as the pole of the momentum space propagator (see e.g. https://physics.stackexchange.com/q/835974 for further details) determined by the equation $$ m^2+\Sigma(m_{\rm ph}^2)- m_{\rm ph}^2=0. \tag{10} \label{10} $$ Expanding the self-energy function $\Sigma(k^2)$ around $k^2=m_{\rm ph}^2$, $$ \Sigma(k^2)= \Sigma(m_{\rm ph}^2) + \Sigma^\prime(m_{\rm ph}^2) (k^2- m_{\rm ph}^2)+ \ldots \tag{11} \label{11}$$ and defining the field renormalization constant by $$Z= \frac{1}{1-\Sigma^\prime(m_{\rm ph}^2)}, \tag{12} \label{12}$$ the momentum-space propagator takes the form $$ \frac{1}{i} \frac{1}{m^2+\Sigma(k^2)-k^2-i \epsilon} = \frac{1}{i} \frac{Z}{m_{\rm ph}^2-k^2} + \ldots \tag{13} \label{13} $$ in the vicinity of the pole.

Computing $\Sigma(k^2)$ up to one-loop order, one finds $$ \Sigma(k^2)=- \lambda \Delta(m^2) /2+ \mathcal{O}(\lambda^2), \tag{14} \label{14} $$ with the loop integral $$ \Delta(m^2)= \int \, \frac{d^d k}{(2\pi)^d} \frac{1}{m^2-k^2-i \epsilon}=\frac{i m^{d-2}}{(4\pi)^{d/2}} \Gamma(1-d/2) \tag{15} \label{15} $$ computed in dimensional regularization. Obviously, the Gamma function $\Gamma(1-d/2)$ exhibits a pole for $d \to4$ reflecting the ultraviolet divergence of the tadpole integral. Using eqs. \eqref{10}, \eqref{14} and \eqref{15}, one finds $$ m_{\rm ph}^2= m^2 +\frac{\lambda}{2} \frac{m^{d-2}}{(4\pi)^{d/2}} \Gamma(1-d/2)+ \mathcal{O}(\lambda^2). \tag{16} \label{16} $$ We observe that keeping $m$ fixed would lead to the nonsensical result of a divergent physical mass in the limit $d \to 4$. Instead, we have to take $m=m(d)$ such that $m_{\rm ph}$ remains finite in this limit. Inverting \eqref{16}, we find $$ m^2= m_{\rm ph}^2- \frac{\lambda}{2} \frac{m_{\rm ph}^{d-2}}{(4\pi)^{d/2}} \Gamma(1-d/2)+ \mathcal{O}(\lambda^2), \tag{17} \label{17} $$ where $(m_{\rm ph}^2-m^2) \lambda= \mathcal{O}(\lambda^2)$ was used.

The $S$-matrix element $$ \langle p_3, p_4\, {\rm out} | p_1, p_2 \,{\rm in}\rangle = (2\pi)^4 \delta^{(4)}(p_1+p_2-p_3-p_4) \mathcal{M}(p_1, p_2 \to p_3, p_4) \tag{18} \label{18}$$ of the process $\phi(p_1) \phi(p_2) \to \phi(p_3) \phi(p_4)$ is obtained from the (connected) four-point function by using the Lehman-Symanzik-Zimmerman (LSZ) reduction formula \begin{align} \langle p_3,p_4 \, {\rm out}| p_1,p_2 \, {\rm in} \rangle &=\left(\!\frac{i}{\sqrt{Z}}\!\right)^4 \int \! d^4x_1 \, d^4x_2 \, d^4x_3 \, d^4x_4 \, e^{-ip_1x_1} e^{-ip_2 x_2} e^{ip_3x_3} e^{ip_4 x_4}\\[5pt]& \; \; \;\times \left(\prod\limits_{i=1}^4 (\square_i+m_{\rm ph}^2)\right) \, \langle 0 |T\phi(x_1) \phi(x_2) \phi(x_3) \phi(x_4)|0\rangle_c . \tag{19} \label{19}\end{align} To one-loop order, one has $Z=1+ \mathcal{O}(\lambda^2)$ from eqs. \eqref{12} and \eqref{14} and the expression for the invariant scattering amplitude is found as $$ \mathcal{M}(p_1, p_2 \to p_2, p_4)= - \lambda +\frac{\lambda^2}{2}\left(B(s, m_{\rm ph}^2)+B(t,m_{\rm ph}^2)+B(u,m_{\rm ph}^2) \right) \tag{20} \label{20} $$ with the three Mandelstam variables $s=(p_1+p_2)^2$, $t=(p_1-p_3)^2$ and $u=(p_1-p_4)^2$. In arriving at \eqref{20}, we have used $\lambda^2 B(s, m^2)= \lambda^2 B(s, m_{\rm ph}^2)+ \mathcal{O}(\lambda^3)$ valid at one-loop order. The function $B(p^2, m_{\rm ph}^2)$ is given by \begin{align} B(s,m_{\rm ph}^2)&= \frac{\Gamma(2-d/2)}{(4\pi)^{d/2}} \int\limits_0^1\! d \alpha \, \left[m_{\rm ph}^2-\alpha (1-\alpha) p^2\right]^{\frac{d-4}{2}} \\[5pt] &\to -\frac{1}{8 \pi^2 (d-4)}+ \; \text{finite terms}. \tag{21} \label{21} \end{align} Note that the divergent term in \eqref{21} is independent of $p^2$.

In contrast to the case of the mass, where the physical value is uniquely determined by the energy-momentum relation $p^2=m^2$, the definition of the physical coupling constant is convention dependent. Because of the relation $s+t+u=4 m_{\rm ph}^2$, the invariant amplitude in \eqref{20} depends in fact only on the two independent variables $s, t$. The observable differential cross section in the center-of-mass system is related to $\mathcal{M}(s,t)$ by the formula $$ \frac{d \sigma}{d\Omega}= \frac{1}{64 \pi^2 s} |\mathcal{M}(s,t)|^2 \tag{22} \label{22} $$ A possible definition of the physical coupling constant is thus provided by $$ \lambda_{\rm ph} := |\mathcal{M}(s_0,t_0)|, \tag{23} \label{23} $$ where $(s_0,t_0)$ is a conveniently chosen (but otherwise arbitrary) point in the physically allowed $s$-$t$ region. Using this definition, \eqref{20} leads to the relation $$ \lambda_{\rm ph}= \lambda \left\{ 1 - \frac{\lambda}{2} {\rm Re}\left[B(s_0,m_{\rm ph}^2)+B(t_0, m_{\rm ph}^2)+B(u_0,m_{\rm ph}^2)\right]+ \mathcal{O}(\lambda^2) \right\}, \tag{24} \label{24}$$ where $u_0=4 m_{\rm ph}^2-s_0-t_0$. The coefficient $\lambda=\lambda(d)$ must be chosen in such a way that the divergences in the loop functions (c.f. \eqref{21}) are compensated in the limit $d \to 4$. Inverting \eqref{24} finally leads to $$ \lambda=\lambda_{\rm ph} \left\{ 1 + \frac{\lambda_{\rm ph}}{2} {\rm Re} \left[B(s_0,m_{\rm ph}^2)+B(t_0, m_{\rm ph}^2)+B(u_0,m_{\rm ph}^2)\right]+\mathcal{O}(\lambda_{\rm ph}^2) \right\}. \tag{25} \label{25} $$

Edit: The following addendum addresses two questions raised in the comments below in some detail.

Using \eqref{25}, the scattering amplitude \eqref{20} can now be expressed in terms of $\lambda_{\rm ph}$, \begin{align} \mathcal{M}(s,t) &=-\lambda_{\rm ph}\bigg\{ 1-\frac{\lambda_{\rm ph}}{2}\bigg[\underbrace{B(s, m_{\rm ph}^2)- {\rm Re} \,B(s_0, m_{\rm ph}^2)}_{\text{finite}} + \underbrace{B(t,m_{\rm ph}^2)-B(t_0, m_{\rm ph}^2)}_{\text{finite}} \\[5pt] & \qquad \qquad \qquad \qquad+\underbrace{B(u,m_{\rm ph}^2)-B(u_0, m_{\rm ph}^2)}_{\text{finite}} \bigg] +\mathcal{O} \lambda_{\rm ph}^2) \bigg\}, \tag{26} \label{26} \end{align} showing that the scattering amplitude at all kinematical points $(s,t)$ is uniquely determined (to the given loop order) once it is known (measured) at one arbitrary reference point $(s_0,t_0)$ in the Dalitz plot (remember the definition of $\lambda_{\rm ph}$ in \eqref{23}).

The determination of the relations $$ m=m(m_{\rm ph}, \lambda_{\rm ph}, d), \quad \lambda= \lambda(m_{\rm ph}, \lambda_{\rm ph}, d) \tag{27} \label{27} $$ requires the computation of the two- and four-point function to the desired order in perturbation theory. This approach, although the most transparent one from a conceptual point of view, leads to rather cumbersome expressions (as we have seen, already at the on-loop level). On top of that, it depends also the convention used for the definition of the physical coupling constant $\lambda_{\rm ph}$.

In practical calculations, it is therefore more convenient to take advantage of a different approach, pioneered by Gerard 't Hooft in the early 1970s. It is based on the observation that in dimensional regularization, divergences appear as poles in the complex $d$-plane and the finite part of any divergent loop integral may be defined by removing these singularities by a certain prescription.

Taking $B(0, m^2)$ (defined in \eqref{21}) as an example, we have $$ B(0,m^2)= \frac{\Gamma(2-d/2) m^{d-4}}{(4\pi)^{d/2}}=-\frac{\Gamma(3-d/2)m^{d-4}}{(d-4) (4\pi)^{d/2}}, \tag{28} \label{28} $$ exhibiting a pole at $d=4$ with residue $-2/(4\pi)^2$. In the minimal subtraction scheme (MS) one defines therefore $$ B(0,m^2)_{r} \Big|_{\rm MS}:= B(0,m^2)+\frac{2 \mu^{d-4}}{(4\pi)^2} \frac{1}{d-4}, \tag{29} \label{29} $$ being finite for $d \to 4$. The renormalization scale or running (mass) scale $\mu$ in the second term is introduced in accordance with the dimension $[B(0,m^2)]=d-4$ of the first term. Expanding \eqref{29} around $d=4$, one obtains $$ B(0,m^2)_r \Big|_{\rm MS} =-\frac{2}{(4\pi)^2} \left[\ln \frac{m}{\mu}- \frac{1}{2} \Gamma^\prime(1) - \frac{1}{2} \ln(4\pi) \right]. \tag{30} \label{30} $$ The appearance of the ugly terms $\Gamma^\prime(1)$ and $\ln(4\pi)$ can be avoided in the modified minimal subtraction scheme ($\overline{\rm MS}$), where $$ B(0,m^2)_r \Big|_{\overline{\rm MS}}:= B(0,m^2) +\frac{2 \mu^{d-4}}{(4\pi)^2}\frac{\Gamma(3-d/2)}{(4\pi)^{\frac{d-4}{2}} (d-4)}= - \frac{2}{(4\pi)^2} \ln \frac{m}{\mu}+\mathcal{O}(d-4). \tag{31} \label{31}$$ Using $\overline{\rm MS}$, we can rewrite eq. \eqref{24} as \begin{align} \lambda_{\rm ph}&= \lambda \left(1+ \frac{3 \lambda \mu^{d-4}}{(4\pi)^2}\frac{\Gamma(3-d/2)}{(4\pi)^{\frac{d-4}{2}} (d-4)} \right) \\[5pt] & \quad -\frac{\lambda^2}{2} {\rm Re} \, \left[ B(s_0,m_{\rm ph}^2)_r +B(t_0, m_{\rm ph}^2)_r +B(u_0,m_{\rm ph}^2)_r\right] + \mathcal{O}(\lambda^3), \tag{32} \label{32} \end{align} suggesting the definition of the running coupling $$ \lambda_r(\mu):= \lambda \mu^{d-4} \left(1+\frac{3 \lambda \mu^{d-4}}{(4 \pi)^2} \frac{\Gamma(3-d/2)}{(4\pi)^{\frac{d-4}{2}}(d-4)}+ \mathcal{O}(\lambda^2) \right), \tag{33} \label{33} $$ being independent of any specific reference point on the Dalitz plot. Note also that the definition \eqref{33} ensures that $\lambda_r(\mu)$ is dimensionless even for $d\ne 4$. Expressing the running coupling in terms of $\lambda_{\rm ph}$ and taking the limit $d \to 4$ yields \begin{align} \lambda_r(\mu) &= \lambda_{\rm ph} \bigg\{ 1+ \frac{3 \lambda_{\rm ph}}{(4\pi)^2} \ln \frac{\mu}{m_{\rm ph}} \\[5pt] & \qquad \qquad +\frac{\lambda_{\rm ph}}{2} {\rm Re} \left[ \overline{B}(s_0, m_{\rm ph}^2) +\overline{B}(t_0,m_{\rm ph}^2)+\overline{B}(u_0, m_{\rm ph}^2) \right] + \mathcal{O}\lambda_{\rm ph}^2) \bigg\}, \tag{34} \label{34} \end{align} where $\overline{B}(p^2,m^2):=B(p^2,m^2)-B(0,m^2)$. Defining a mass scale $\mu_0$ by $\lambda_r(\mu_0)= \lambda_{\rm ph}$, the relation \eqref{34} between the running coupling $\lambda_r(\mu)$ in the $\overline{\rm MS}$-scheme and the physical coupling $\lambda_{\rm ph}$ defined in \eqref{23} can thus be written in the compact form $$ \lambda_r(\mu)= \lambda_{\rm ph} \left(1 + \frac{3 \lambda_{\rm ph}}{(4\pi)^2}\ln \frac{\mu}{\mu_0} + \mathcal{O}(\lambda_{\rm ph}^2) \right). \tag{35} \label{35} $$

Note that further aspects of this important topic can be found in https://physics.stackexchange.com/q/693546 and https://physics.stackexchange/q/408762.

$\endgroup$
7
  • $\begingroup$ I'm aware and understand all mathematical details. My question was more on the physical meaning of the whole procedure, which you greatly answered in the first lines. Thank's a lot! $\endgroup$ Commented Apr 1 at 19:10
  • $\begingroup$ Hi Hyperon, nice answer (and I've upvoted some time ago already, but I now noticed that I have a, presumably trivial, question). In your last line, you arrived at an expression $\lambda=\lambda(s_0,t_0)$ (?) s.t. it reproduces the correct differential cross section for these parameters. But is this value sufficient (together with the mass) to predict cross sections for different momentum transfers without the input of the physical parameter $\lambda_{\mathrm{ph}}$ for the new scale? In other words, how can we make sure eq. 25 yields the correct values for eq. 20 for all $t,s$? $\endgroup$ Commented May 7 at 19:37
  • 1
    $\begingroup$ @TobiasFünke This is an important point. Indeed, once you have measured the physical mass $m_{\rm ph}$ and the scattering amplitude $\mathcal{M}(s_0,t_0)$ at one kinematical point $(s_0,t_0)$, the scattering amplitude $\mathcal{M}(s,t)$ is uniquely determined (at least to the given loop order) for all other points $(s,t)$. In this way, it is, of course, always possible to express $\lambda_{\rm ph}$ defined by eq. (23) at the point $(s_0,t_0)$ to some $\lambda^\prime_{\rm ph}$ defined by the same eq. but at another point $(s_0^\prime, t_0^\prime)$ and vice versa. $\endgroup$ Commented May 7 at 20:25
  • $\begingroup$ Thank you for your answer. I think it makes sense now to me. One more thing: What you describe in your last sentence: Is this then the "running of the coupling constant", similar to the answer here? $\endgroup$ Commented May 7 at 21:53
  • 1
    $\begingroup$ @TobiasFünke I'll explain explain the relationship/difference between the "physical" coupling $\lambda_{\rm ph}$ (as defined in eq. (23)) and the "running" coupling $\lambda_r(\mu)$ (defined e.g. in the MS-bar scheme) in an addendum to my answer. In short, the former approach gives a more transparent picture from a conceptual point of view, whereas the latter is more convenient in practical calculations. The physical content of the theory remains, of course, unaffected thereof. $\endgroup$ Commented May 8 at 4:32

Start asking to get answers

Find the answer to your question by asking.

Ask question

Explore related questions

See similar questions with these tags.