Problem #\(1\): What is the Landau-Ginzburg free energy functional \(F[m]\) for the Ising model?
Solution #\(1\): It is defined implicitly through:
\[e^{-\beta F[m]}=\sum_{\{\sigma_i\}\to_{\text{c.g.}}m(\textbf x)}e^{-\beta E_{\{\sigma_i\}}}\]
where \(E_{\sigma_i}=-E_{\text{ext}}\sum_{i=1}^N\sigma_i-E_{\text{int}}\sum_{\langle i,j\rangle}\sigma_i\sigma_j\) is the energy of a given spin microstate \(\{\sigma_i\}\) and “\(\text{c.g.}\)” is short for coarse graining the underlying \(d\)-dimensional lattice \(\Lambda\to\textbf R^d\).
Problem #\(2\): Describe how a saddle-point approximation can be used to evaluate the canonical partition function \(Z\) of a Landau-Ginzburg theory.
Solution #\(2\): In the canonical ensemble, the probability density functional \(p[\phi]\) of finding the system in a given configuration \(\phi=\phi(\textbf x)\) is given by the Boltzmann distribution:
\[p[\phi]=\frac{e^{-\beta F[\phi]}}{Z}\]
where, to ensure normalization \(\int\mathcal D\phi p[\phi]=1\) over the space of all local order parameter configurations \(\phi\), the canonical partition function \(Z\) is given by the path integral:
\[Z=\int\mathcal D\phi e^{-\beta F[\phi]}\]
In general, functional integrals (so-called because the integrand \(e^{-\beta F[\phi]}\) is a functional) are difficult to evaluate (partly because they are hard to even rigorously define!). However, whether one is doing integrals over \(\textbf R,\textbf C\) or function spaces, as long as one’s integrand looks like \(e^{-\text{something}}\), it’s always worth trying a saddle-point approximation, which in this case looks like:
\[Z\approx e^{-\beta F[\phi_*]}\]
where \(\phi_*\) is the order parameter configuration minimizing the free energy \(F=F[\phi]\). In other words, \(\phi_*\) is a stationary point of \(F[\phi]\) so that the functional derivative \(\frac{\delta F}{\delta\phi^*}=0\) vanishes.
Landau mean field theory is the special case of this saddle-point approximation in which all fluctuations \(\phi_*=\phi_*(\textbf x)\) are completely ignored, yielding a homogeneous mean field order parameter.
Problem #\(3\): Explain why, as with many other functionals in physics (e.g. the action \(S[\textbf x]\)), the Landau-Ginzburg free energy functional \(F[\phi]\) must take the form:
\[F[\phi]=\int d^d\textbf x \mathcal F\left(\phi(\textbf x),\frac{\partial\phi}{\partial\textbf x},…\right)\]
Combining this with the saddle-point approximation in Solution #\(2\), what can one conclude?
Solution #\(3\): The presence of the integral \(\int d^d\textbf x\) simply reflects the extensive nature of the free energy \(F\), while the dependence of the integrand \(\mathcal F\) on \(\phi\) and its derivatives only reflects locality.
In the special case that the free energy density \(\mathcal F\) depends only on the field \(\phi(\textbf x)\) and its gradient \(\frac{\partial\phi}{\partial\textbf x}\) (but no higher derivatives), and it obeys suitable boundary conditions, one then has the usual Euler-Lagrange equations:
\[\frac{\partial}{\partial\textbf x}\cdot\frac{\partial \mathcal F}{\partial (\partial\phi/\partial\textbf x)}=\frac{\partial\mathcal F}{\partial\textbf x}\]
or equivalently, because of the lack of explicit \(\textbf x\)-dependence in \(\mathcal F\), one has the equivalent Beltrami identity:
\[\frac{\partial \mathcal F}{\partial (\partial\phi/\partial\textbf x)}\cdot\frac{\partial\phi}{\partial\textbf x}-\mathcal F=-\mathcal F_0\]
for some constant \(\mathcal F_0\in\textbf R\).
Problem #\(3\): Suppose that free energy density \(\mathcal F(\phi)\) of a particular system (e.g. the Ising ferromagnet) is taken (on locality, analyticity and suitable symmetry grounds) to be of the form:
\[\mathcal F\left(\phi,\frac{\partial\phi}{\partial\textbf x}\right)=\frac{\alpha_2}{2}\phi^2+\frac{\alpha_4}{4}\phi^4+\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2\]
where the phenomenological coupling constants \(\alpha_2,\alpha_4,\gamma\) each may carry some \(T\)-dependence, though as far as the study of second-order phase transitions at critical points is concerned, only the \(T\)-dependence \(\alpha_2(T)\sim T-T_c\) on the quadratic \(\phi^2\) term matters, and all that needs to be assumed about the other coupling constants is their sign for all temperatures \(T\), in this case \(\alpha_4,\gamma>0\). Show that in the subcritical regime \(T<T_c\), the \(\textbf Z_2\) symmetry \(F[-\phi]=F[\phi]\) of the theory is spontaneously broken via a bifurcation into \(2\) degenerate ground states (also called vacua in analogy with QFT) representing mean-field homogeneous/ordered phases/configurations \(\phi_*(\textbf x)=\phi_0\). Show that there is also a more interesting domain wall soliton:
\[\phi_*^{\text{DW}}(x)=\phi_0\tanh\left(\sqrt{-\frac{\alpha_2}{2\gamma}}x\right)\]
that emerges upon imposing Dirichlet boundary conditions \(\lim_{x\to\pm\infty}\phi_*(\textbf x)=\pm\phi_0\) implementing a smooth transition interpolating between the two ground state phases \(\pm\phi_0\).
Solution #\(3\): The Euler-Lagrange equations for this particular free energy density \(\mathcal F\) yield a Poisson/Helmholtz-like (but nonlinear!) PDE for the on-shell field \(\phi_*(\textbf x)\):
\[\gamma\biggr|\frac{\partial}{\partial\textbf x}\biggr|^2\phi_*=\alpha_2\phi_*+\alpha_4\phi_*^3\]
Looking for a homogeneous ansatz \(\phi_*(\textbf x)=\phi_0\) yields the \(2\) degenerate ground states \(\phi_0=\sqrt{-\alpha_2/\alpha_4}\) with free energy density \(\mathcal F_0:=\mathcal F(\pm\phi_0)=-\alpha_2^2/4\alpha_4\) and corresponding Landau-Ginzburg free energy \(F_0:=F[\pm\phi_0]=L^d\mathcal F_0\) (putting the system in a box \([-L/2,L/2]^d\) to regularize the obvious IR divergence that would otherwise arise).
By contrast, reverting now to the Beltrami identity, assuming that \(\phi_*^{\text{DW}}(\textbf x)=\phi_*^{\text{DW}}(x)\) varies only along the \(x\)-direction, the PDE reduces to the nonlinear separable first-order ODE:
\[\frac{\gamma}{2}\left(\frac{d\phi_*^{\text{DW}}}{dx}\right)^2-\frac{\alpha_2}{2}(\phi_*^{\text{DW}})^2-\frac{\alpha_4}{4}(\phi_*^{\text{DW}})^4=-\mathcal F_0\]
In particular, placing the domain wall at the origin \(x=0\) so that \(\phi(0)=0\), one obtains the soliton described (for a domain wall at some other location \(x_0\in\textbf R\), just translate \(x\mapsto x-x_0\) in the \(\tanh\) function). In particular; the width of the domain wall is \(\Delta x=\sqrt{-2\gamma/\alpha_2}\) which is pretty intuitive (c.f. the formula \(\omega_0=\sqrt{k/m}\) for a mass \(m\) on a spring \(k\)). Unlike the homogeneous ground states \(\pm\phi_0\), this domain wall soliton is a non-MF stationary point of the Landau-Ginzburg free energy functional \(F\).
Problem #\(4\): By definition, the ground state(s) of any system are global minima of its energy. In particular, it is clear that the domain wall soliton \(\phi_*^{\text{DW}}(x)\) is not a ground state, having free energy \(F[\phi_*^{\text{DW}}]>F_0\); precisely how much free energy \(\Delta F_{\text{DW}}:=F[\phi_*^{\text{DW}}]-F_0\) does it cost to create such a domain wall from a homogeneous ground state phase?
Solution #\(4\): Differential equations can (and should!) often be thought of as a dance/tension between conflicting characters. Even without doing any of the math in Solution #\(3\), it should be clear that the domain wall transition cannot happen instantaneously or the free energy cost \(\int d^d\textbf x\frac{\gamma}{2}\left(\frac{d\phi}{dx}\right)^2\) from the “kinetic” term would be too great, but neither can it proceed too slowly otherwise the free energy cost \(\int d^d\textbf x\left(\frac{\alpha_2}{2}\phi^2+\frac{\alpha_4}{4}\phi^4\right)\) from the “potential” term would be too great; the domain wall \(\phi_*^{\text{DW}}(x)\) must therefore strike a balance between these two free energy costs while satisfying the boundary conditions \(\lim_{x\to\pm\infty}\phi_*^{\text{DW}}(x)=\pm\phi_0\) (cf. the virial theorem in classical mechanics). In other words, \(0<\Delta x<\infty\).
(easy to forget, but remember one is working in the subcritical \(T<T_c\) regime where \(\alpha_2<0,\alpha_4>0\) so the potential term \(\frac{\alpha_2}{2}\phi^2+\frac{\alpha_4}{4}\phi^4\) is not positive semi-definite, in particular its minimum does not lie at \(\phi_0=0\) but rather has degenerate minima at \(\phi_0=\pm\sqrt{-\alpha_2/\alpha_4}\)! This means that anywhere \(\textbf x\in\textbf R^d\) that \(\phi_*^{\text{DW}}(\textbf x)\) strays too far away from the bottom of the potential wells at \(\pm\phi_0\) costs free energy, so for this reason the domain wall cannot take too long to climb over the hump between the two minima).
Indeed, the Beltrami identity quantifies this free energy balance and one can exploit it to quickly calculate the free energy of the domain wall soliton:
\[F[\phi_*^{\text{DW}}]=\int d^d\textbf x\left(\frac{\alpha_2}{2}(\phi_*^{\text{DW}})^2+\frac{\alpha_4}{4}(\phi_*^{\text{DW}})^4+\frac{\gamma}{2}\left(\frac{d\phi_*^{\text{DW}}}{dx}\right)^2\right)=\int d^d\textbf x\left(\gamma\left(\frac{d\phi_*^{\text{DW}}}{dx}\right)^2+f_0\right)\]
the latter term is recognized as just the free energy \(F_0=\int d^d\textbf x f_0=L^df_0\) of the ground state(s) so the excess free energy cost \(\Delta F_{\text{DW}}\) of creating the domain wall is (for \(L\gg\Delta x\)):
\[\Delta F_{\text{DW}}=\gamma\int d^d\textbf x\left(\frac{d\phi_*^{\text{DW}}}{dx}\right)^2=\gamma L^{d-1}\left(\frac{\phi_0}{\Delta x}\right)^2\int_{-L/2}^{L/2}dx\space\text{sech}^4\left(\frac{x}{\Delta x}\right)\]
\[\approx\frac{\gamma L^{d-1}\phi_0^2}{\Delta x}\int_{-\infty}^{\infty}d\varphi\space\text{sech}^4\varphi=\frac{4}{3\sqrt{2}}\frac{\sqrt{-\gamma\alpha_2^3}}{\alpha_4}L^{d-1}\]
but the key point is that \(\Delta F_{\text{DW}}\sim L^{d-1}\) scales with the (hyper)area of the domain wall. Well actually, another important scaling to note is that \(\Delta F_{\text{DW}}\sim(-\alpha_2)^{3/2}\) which suggests that near criticality where \(\alpha_2\to 0\), the free energy cost \(\Delta F_{\text{DW}}\to 0\) of creating a domain wall also vanishes, while its width \(\Delta x\sim(-\alpha_2)^{-1/2}\) diverges. This suggests that domain walls become important near critical points.
Problem #\(5\): Working with the same Landau-Ginzburg system as above, estimate in \(d=1\) dimension the probability \(p_N\) that thermal fluctuations will spontaneously break the \(\textbf Z_2\) symmetry of the free energy \(F\) via the creation of \(N\ll L/\Delta x\) domain walls anywhere, and comment on the implication of this for the lower critical dimension \(d_{\ell}\) of this system.
Solution #\(5\): Because the ODE arising from the Beltrami identity was nonlinear, the superposition of a bunch of domain walls at different locations is not strictly speaking a stationary point of the free energy \(F\), but nonetheless one can sweep this under the rug and assume it is still an approximate solution provided the \(N\) domain walls are well-separated. In particular, this means their total free energy \(\approx N\Delta F_{\text{DW}}+F_0\) is also approximately additive. If one imagines discretizing the \(d=1\) line \([-L/2,L/2]\) into \(L/\Delta x\) bins each of width \(\Delta x\), then there are \(L/\Delta x\choose{N}\) choices for where to put the domain walls in (ignoring the fact that in some cases, they may not be so well-separated), so the probability of having any configuration of \(N\) domain walls is approximately:
\[p_N\approx {{L/\Delta x}\choose{N}}\frac{e^{-N\beta\Delta F_{\text{DW}}+F_0}}{Z}\approx\frac{(L/\Delta x)^N}{N!}\frac{e^{-N\beta\Delta F_{\text{DW}}+F_0}}{Z}\]
where the sparse approximation has been used \(N\ll L/\Delta x\). Alternatively, normalizing with respect to the \(N=0\) “vacuum” probability \(p_0=e^{-\beta F_0}/Z\):
\[\frac{p_N}{p_0}=\frac{(L/\Delta x)^N}{N!}e^{-N\beta\Delta F_{\text{DW}}}\]
Importantly, in \(d=1\) domain walls are free since \(\Delta F_{\text{DW}}\sim L^{1-1}=L^0\). So all the \(L\)-dependence in the expression above for \(p_N/p_0\) is contained in the entropic factor \((L/\Delta x)^N/N!\) and shows that as one takes the infinite system limit \(L\to\infty\), the probability \(p_N/p_0\) receives no exponential suppression from the \(e^{-N\beta\Delta F_{\text{DW}}}\) and instead grows unbounded. The probability that thermal fluctuations produce an even number \(N\in 2\textbf N\) of domain walls is:
\[\sum_{N=2,4,6,…}p_N=\cosh\left(\frac{Le^{-\beta\Delta F_{\text{DW}}}}{\Delta x}\right)p_0\]
and similarly for odd \(N\in 2\textbf N+1\):
\[\sum_{N=1,3,5,…}p_N=\sinh\left(\frac{Le^{-\beta\Delta F_{\text{DW}}}}{\Delta x}\right)p_0\]
and since \(\lim_{\varphi\to\infty}\tanh\varphi=1\), these two probabilities both approach the same \(e^{Le^{-\beta\Delta F_{\text{DW}}}/\Delta x}p_0/2\) as \(L\to\infty\).
More generally, any Landau-Ginzburg theory with a discrete symmetry (like the \(\textbf Z_2\) symmetry of this particular \(F\)) will possess a bunch of disconnected, degenerate ground states/vacua (like the \(\phi(\textbf x)=\pm\phi_0\) in this case) that spontaneously break that discrete symmetry, and all have \(d_{\ell}=1\) because there is very high probability that thermal fluctuations will proliferate domain walls that toggle between the degenerate ground states, preventing ordered phases from forming. Hence for instance the Ising model has no phase transitions in \(d=1\).
Problem #\(6\):

Solution #\(6\):









Problem #\(7\):

Solution #\(7\):



Note that if one does not work in natural units, then the heat capacity is instead \(C=k_B\beta^2\frac{\partial^2\ln Z}{\partial\beta^2}\) so all specific heat capacities should come with an additional factor of \(k_B\).
Problem #\(8\):

Solution #\(8\):

Problem #\(9\):

Solution #\(9\):



(NEED TO COME BACK TO THIS QUESTION!)
Problem #\(10\): Describe the quadratic approximation to the Landau-Ginzburg free energy density \(\mathcal F\) governing a \(\textbf Z_2\)-symmetric system/theory (e.g. Ising ferromagnet) described by a single, real scalar order parameter \(\phi\) in the absence of any external coupling \(E_{\text{ext}}=0\).
Solution #\(10\): On LAS (locality, analyticity, symmetry) grounds, the exact free energy density \(\mathcal F\) that knows about the all detailed microscopic physics must take the phenomenological form:
\[\mathcal F=\frac{\alpha_2}{2}\phi^2+\frac{\alpha_4}{4}\phi^4+\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2+…\]
where in principle there are also coupling constants for \(\phi^6,\phi^5\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2\biggr|\frac{\partial}{\partial\textbf x}\biggr|^2\phi\) etc. though not for couplings like \(\phi^3\) or \(1/\phi^2\). In addition, one also has to import from mean-field theory the assumption that \(\alpha_2\sim T-T_c\) (i.e. \(\alpha_2\to 0\) as \(T\to T_c\) in a linear/critical exponent \(1\) manner) and that \(\gamma,\alpha_4>0\) for all \(T\) in a neighbourhood of \(T_c\).
The quadratic approximation to \(\mathcal F\) looks slightly different depending on whether one is working in the supercritical \(T>T_c\) regime (where \(\alpha_2(T)>0\)) or the subcritical \(T<T_c\) regime (where \(\alpha_2(T)<0\)).
In the supercritical \(T>T_c\) regime, the quadratic approximation does what it sounds like, in fact drop not only all quartic, sextic, octic, decic, etc. couplings like \(\phi^4,\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^{22}\) but even quadratic couplings containing Laplacians, third derivatives, biharmonics, and all higher-derivative couplings. So at \(T>T_c\), this amounts to keeping only \(2\) couplings, a “kinetic” coupling \(\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2\) and a “Hookean quadratic potential” coupling \(\frac{\alpha_2}{2}\phi^2\):
\[\mathcal F\approx\frac{\alpha_2}{2}\phi^2+\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2\]
Notice in this case one can also replace \(\phi\mapsto\delta\phi\) everywhere:
\[=\frac{\alpha_2}{2}\delta\phi^2+\frac{\gamma}{2}\biggr|\frac{\partial\delta\phi}{\partial\textbf x}\biggr|^2\]
where the fluctuation \(\delta\phi(\textbf x):=\phi(\textbf x)-\langle\phi(\textbf x)\rangle\). This is simply because, at \(T>T_c\), \(\langle\phi(\textbf x)\rangle=0\) vanishes homogeneously (from mean-field theory).
By contrast, in the subcritical \(T<T_c\) regime, the quadratic approximation consists of \(2\) separate approximation steps:
Step #\(1\): Keep not only the zeroth and first-order quadratic couplings that were retained in the supercritical \(T>T_c\) case, but also keep the quartic \(\phi^4\) coupling:
\[\mathcal F\approx\frac{\alpha_2}{2}\phi^2+\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2\]
The reason for this is that for \(T<T_c\), \(\alpha_2(T)<0\) so the free energy density \(\mathcal F(\phi)=-\frac{|\alpha_2|}{2}\phi^2+…\) would be unbounded from below, yielding an unstable theory. By including the quartic coupling, one instead has \(2\) degenerate \(\textbf Z_2\) symmetry breaking homogeneous ordered phases \(\phi(\textbf x)=\pm\phi_0\) with \(\phi_0=\sqrt{-\alpha_2/\alpha_4}\).
However, beyond the presence of \(\alpha_4\) in \(\phi_0\), one would otherwise like to remove all other vestiges of it in \(\mathcal F\) in order to get a more quadratic-looking free energy density like in the \(T>T_c\) case. Thus:
Step #\(2\): Insert the “Reynolds decomposition” \(\phi=\phi_0+\delta\phi\) into \(\mathcal F\) and notice that the term linear in \(\delta\phi\) vanishes because \(\phi_0\) is on-shell:
\[\mathcal F\approx\mathcal F[\phi_0]-\alpha_2\delta\phi^2+\frac{\gamma}{2}\biggr|\frac{\partial\delta\phi}{\partial\textbf x}\biggr|^2+O(\delta\phi^3)\]
where the cubic and quartic couplings \(O(\delta\phi^3)=2\alpha_4\phi_0\delta\phi^3+\frac{\alpha_4}{2}\delta\phi^4\) at the level of the fluctuations \(\delta\phi\) from the mean field \(\phi_0\) are assumed negligible in this second step of the quadratic approximation for \(T<T_c\). Finally, note that the constant term \(\mathcal F[\phi_0]\) would drop out when differentiating \(\ln Z\) to compute Boltzmannian cumulants, so can safely be omitted.
In this way, the name “quadratic approximation” is justified because both the \(T>T_c\) and \(T<T_c\) cases can be unified by writing the free energy as:
\[\mathcal F\approx\frac{1}{2}\left(\mu^2\delta\phi^2+\gamma\biggr|\frac{\partial\delta\phi}{\partial\textbf x}\biggr|^2\right)\]
where the “mass coupling” \(\mu^2\geq 0\) is defined piecewise to capture both the subcritical and supercritical regimes:
\[\mu^2(T) =\begin{cases}
\alpha_{2}(T), & T \geq T_c\\
-2\,\alpha_{2}(T), & T \leq T_c
\end{cases}\]
cf. the Klein-Gordon Lagrangian density when written in natural units:
\[\mathcal L=\frac{1}{2}\partial^{\mu}\partial_{\mu}\phi-\frac{1}{2}m^2\phi^2\]

Problem #\(11\): Working within the quadratic approximation to the free energy density \(\mathcal F\) outlined in Solution #\(11\), compute the partition function \(Z\).
Solution #\(11\): This is pretty much the only case where the path integral underlying \(Z\) can be computed analytically thanks to the fact that Gaussian integrals are straightforward to do. Here is a sketch of the computation:
Step #\(1\): The free energy \(F=\int d^d\textbf x\mathcal F(\delta\phi)\) is certainly extensive so one must add up \(\int\) the chunks of free energy \(d^d\textbf x\mathcal F(\delta\phi)\) due to fluctuations \(\delta\phi\) of the order parameter \(\phi\) at each point \(\textbf x\in\textbf R^d\) in space (or \(\textbf x\in V\subset\textbf R^d\) in some suitably large volume). But conceptually, some of these fluctuations might be more short-range, rapid oscillations across \(\textbf x\), while others may be more long-range, slow envelopes. Nonetheless, it should be intuitively clear that, rather than taking the local approach of stepping through each \(\textbf x\in\textbf R^d\) and adding up the energies of all fluctuations at that point \(\textbf x\), one can take a more global approach of adding up the energies of all fluctuations across the entire space \(\textbf R^d\) that have a given wavelength \(\lambda\) (or equivalently wavenumber \(k=2\pi/\lambda\)), and stepping through all possible wavelengths, from the long (“IR”) wavelengths all the way down to the short (“UV”) wavelengths. This intuitive picture can of course be formalized by explicitly writing \(\delta\phi(\textbf x)\) as a superposition of plane wave excitations:
\[\delta\phi(\textbf x)=\int\frac{d^d\textbf k}{(2\pi)^d}\delta\phi_{\textbf k}e^{i\textbf k\cdot\textbf x}\]
whereupon, one obtains what is essentially just Plancherel’s theorem:
\[F=\frac{1}{2}\int\frac{d^d\textbf k}{(2\pi)^d}|\delta\phi_{\textbf k}|^2(\mu^2+\gamma|\textbf k|^2)\]
which uses the fact that for \(\delta\phi(\textbf x)\in\textbf R\), the Fourier transform satisfies \(\delta\phi_{-\textbf k}=\delta\phi^{\dagger}_{\textbf k}\). By viewing the system as having some large but finite volume \(V\), one can replace on dimensional grounds:
\[\int\frac{d^d\textbf k}{(2\pi)^d}\Leftrightarrow\frac{1}{V}\sum_{\textbf k}\]
where \(\sum_{\textbf k}\) means over all \(\textbf k=\frac{2\pi}{L}\textbf n\) for \(\textbf n\in\textbf Z^d\) since these are the only wavevectors compatible with periodic boundary conditions on \(\delta\phi(\textbf x)\) in a box \([-L/2,L/2]^d\) of volume \(V=L^d\). Put another way, it reduces Plancherel’s theorem for the Fourier transform to Parseval’s theorem for Fourier series (since \(\delta\phi(\textbf x)\) is now \(L\)-periodic in all \(d\) dimensions).
Finally, the existence of the plane wave basis also allows one to “rigorously” define the measure \(\mathcal D\phi=\mathcal D\delta\phi\) in the path integral for \(Z=\int\mathcal D\delta\phi e^{-\beta F[\delta\phi]}\). Intuitively, the path integral \(\int\mathcal D\delta\phi\) wants to sum over all possible fluctuations \(\delta\phi\) of the field about the mean field. But the Fourier transform allows one to explicitly parameterize this abstract space! Simply integrate over all possible choices of the “Fourier knobs/coefficients” \(\phi_{\textbf k}\) which can span any fluctuation \(\delta\phi\):
\[\int\mathcal D\delta\phi\sim\int\prod_{\textbf k}d\delta\phi_{\textbf k}\sim\int\prod_{\textbf k}d\Re \delta\phi_{\textbf k}d\Im \delta\phi_{\textbf k}\]
where the product \(\prod_{\textbf k}\) is over the same countably infinite lattice of \(\textbf k\)-wavevectors as the earlier sum \(\sum_{\textbf k}\) (actually, strictly speaking, one should only take the product over half of the entire \(\textbf k\)-space, for instance imposing \(k_x>0\). This is because if \(\delta\phi_{\textbf k}\in\textbf C\) is already known for some \(\textbf k\), then \(\delta\phi_{-\textbf k}\) is also automatically known by virtue of the reality criterion \(\delta\phi_{-\textbf k}=\delta\phi^{\dagger}_{\textbf k}\), cf. \(\sin(kx)=\frac{1}{2i}e^{ikx}+?e^{-ikx}\) where, knowing that \(\sin(kx)\in\textbf R\), one can immediately conclude that \(?=(\frac{1}{2i})^{\dagger}=-\frac{1}{2i}\). So \(\delta\phi_{\textbf k}\) and \(\delta\phi_{-\textbf k}\) are not independent knobs (imagine a “complex conjugation gear” between them), but the path integral \(\int\mathcal D\delta\phi\) only wants to integrate over independent degrees of freedom since double-counting the same fluctuation is obviously not desired. That being said, this only leads to a factor of \(2\) discrepancy which doesn’t affect any physical quantities, and the measure \(\mathcal D\delta\phi\) is only defined up to some “normalization” anyways).
At this point one writes \(|\delta\phi_{\textbf k}|^2=\Re^2\delta\phi_{\textbf k}+\Im^2\delta\phi_{\textbf k}\) and decouples all the Gaussian integrals to obtain the final result for \(Z\):
\[Z\sim\prod_{\textbf k}\sqrt{\frac{\pi V}{\beta(\mu^2+\gamma|\textbf k|^2)}}\]
Problem #\(12\): Using this \(Z\), and making the specific assumption that \(\alpha_2(T)=k_B(T-T_c)\) and that \(\gamma\) is independent of \(T\), compute the specific heat capacity \(c\) of this Landau-Ginzburg system in the supercritical \(T>T_c\) regime.
Solution #\(12\): Simply use the formula:
\[c=\frac{k_B}{V}\beta^2\frac{\partial^2\ln Z}{\partial\beta^2}\]
with cumulant generating function (ignoring the temperature-independent parts):
\[\ln Z\sim-\frac{1}{2}\sum_{\textbf k}\ln\beta(\mu^2+\gamma|\textbf k|^2)\]
So, noting that \(\partial\mu^2/\partial\beta=-1/\beta^2\):
\[c=-\frac{k_B}{2}\beta^2\frac{\partial^2}{\partial\beta^2}\int\frac{d^d\textbf k}{(2\pi)^d}\ln\beta(\mu^2+\gamma|\textbf k|^2)\]
\[=-\frac{k_B}{2}\int\frac{d^d\textbf k}{(2\pi)^d}\beta^2\frac{\partial^2}{\partial\beta^2}\left(\ln\beta+\ln(\mu^2+\gamma|\textbf k|^2)\right)\]
\[=\frac{k_B}{2}\int\frac{d^d\textbf k}{(2\pi)^d}\left(1-\frac{2}{\beta(\mu^2+\gamma |\textbf k|^2)}+\frac{1}{\beta^2(\mu^2+\gamma |\textbf k|^2)^2}\right)\]
where the \(\frac{k_B}{2}\times 1\) part reflects equipartition of quadratic degrees of freedom and is simply due to the \(\beta\) temperature dependence in \(e^{-\beta F}\) (fluctuation-dissipation theorem?). By contrast, the other terms are additional contributions to the heat capacity arising from the \(T\)-dependence in \(F\) instead, specifically in \(\mu^2=k_B(T-T_c)\).
Problem #\(13\): Clearly, the specific heat capacity \(c\) as it’s currently written suffers from a \(|\textbf k|\to\infty\) UV divergence, since it is that part of the integrand that causes the integral to diverge to \(\infty\). What should one make of this?
Solution #\(13\): To regularize this UV divergence, one has to impose a UV cutoff \(k^*\) with the property that the Fourier transform \(\delta\phi_{\textbf k}\) is only supported for \(|\textbf k|\leq k^*\) in the \(k^*\)-ball. This UV cutoff should be chosen so that \(k^*\sim 1/\Delta x\), with \(\Delta x\) being the lattice spacing of some underlying microscopic structure which has been coarse-grained away. By implementing a UV cutoff, the specific heat capacity \(c\) is made finite:
\[c=\frac{k_B}{2(2\pi)^d}|S^{d-1}|\left(\frac{(k^*)^d}{d}-\frac{2}{\beta}\int_0^{k^*}dk\frac{k^{d-1}}{\mu^2+\gamma k^2}+\frac{1}{\beta^2}\int_0^{k^*}dk\frac{k^{d-1}}{(\mu^2+\gamma k^2)^2}\right)\]
where the (hyper)surface area of the unit \(d-1\)-sphere is \(|S^{d-1}|=2\pi^{d/2}/\Gamma(d/2)\).
(aside: to prove this, notice that the \(d\)-dimensional isotropic Gaussian integral \(\int_{\textbf x\in\textbf R^d} d^d\textbf x e^{-|\textbf x|^2}\) evaluates to \(\pi^{d/2}\) when separated in Cartesian coordinates, or equivalently \(|S^{d-1}|\int_0^{\infty}d|\textbf x||\textbf x|^{d-1}e^{-|\textbf x|^2}\) in spherical coordinates. The latter integral can then be massaged into the form of a gamma function \(\Gamma(d/2)/2\) via the substitution \(z:=|\textbf x|^2\)).
For some small dimensions \(d=1,2,3,4,5\), the \(2\) integrals can be evaluated analytically, with results compiled in the table below:

or in hindsight many of these can be written more compactly via the (yet unmotivated) correlation length \(\xi:=\sqrt{\gamma}/\mu\).
Problem #\(14\): Using the results of Problem #\(14\), comment on how \(c\) behaves in various dimensions \(d\) as one approaches the critical point \(T\to T_c^+\) from above (since the results above were all computed within the supercritical \(T>T_c\) regime, though the analysis is the same in the subcritical \(T<T_c\) regime).
Solution #\(14\): As \(T\to T_c^+\), the quadratic coupling constant \(\mu^2\to 0\) vanishes. For \(d\geq 5\), both of the integrals above converge to a finite value determined by the choice of UV cutoff wavenumber \(k^*\), so \(c\) is also finite, and more precisely \(c\sim(k^*)^{d-2}\) from the first integral. For \(d=4,3\), only the first integral diverges while the second one converges, whereas for \(d=2,1\), both integrals diverge. For \(d\leq 3\), this divergence goes like \(c\sim\mu^{d-4}\) whereas for \(d=4\) it is a logarithmic divergence \(c\sim\ln\mu\) (both of these being determined in this case by the second integral).
In the case \(d\leq 3\), since \(\mu^2\sim T-T_c\) as \(T\to T_c^+\), this implies that \(c\sim (T-T_c)^{(d-4)/2}\) but in general the critical exponent \(\alpha\) is defined by the property that \(c\sim |T-T_c|^{-\alpha}\). Although this analysis was only for \(T>T_c\), one can check that for \(T<T_c\) one would have mirror behavior, so this analysis shows that, at least for \(d\leq 3\), the critical exponent is \(\alpha=2-d/2\). Thus, the contribution of fluctuations causes the critical exponent to differ from the mean-field prediction \(\alpha=0\).
Problem #\(15\): At each \(\textbf x\in\textbf R^d\), one can associate a continuous random variable \(\phi(\textbf x)\) which draws an order parameter configuration \(\phi\) from a thermal Boltzmann distribution \(p[\phi]=e^{-\beta F[\phi]}/Z\) and returns the evaluation of \(\phi\) at \(\textbf x\). Given two arbitrary positions \(\textbf x,\textbf x’\in\textbf R^d\), each giving rise to its own random variable \(\phi(\textbf x),\phi(\textbf x’)\), define the connected \(2\)-point correlation propagator \(\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle\) between \(\textbf x,\textbf x’\).
Solution #\(15\): The connected \(2\)-point correlation propagator is simply the cross-covariance of the random variables \(\phi(\textbf x),\phi(\textbf x’)\). It is thus related to the cross-correlation \(\langle\phi(\textbf x)\phi(\textbf x’)\rangle\) between \(\textbf x,\textbf x’\) and their individual expected values \(\langle\phi(\textbf x)\rangle,\langle\phi(\textbf x’)\rangle\) by the usual parallel axis theorem (sort of…):
\[\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle=\langle\phi(\textbf x)\phi(\textbf x’)\rangle-\langle\phi(\textbf x)\rangle\langle\phi(\textbf x’)\rangle\]
where, just to flesh it out explicitly, these thermal Boltzmann canonical ensemble averages look like configuration space functional integrals:
\[\langle\phi(\textbf x)\rangle=\int\mathcal D\phi\phi(\textbf x)e^{-\beta F[\phi]}\]
(remember that evaluation at \(\textbf x\) is a functional \(\phi(\textbf x)=\text{eval}_{\textbf x}[\phi]\)).
Problem #\(16\): Define the functional derivative of a functional. Evaluate the following functional derivatives:
\[\frac{\delta}{\delta f(\textbf x)}\int d^d\textbf x’\cos f(\textbf x’)\]
\[\frac{\delta}{\delta f(\textbf x)}\int d^d\textbf x’ d^d\textbf x^{\prime\prime}\frac{f(\textbf x’)f(\textbf x^{\prime\prime})}{|\textbf x’-\textbf {x}^{\prime\prime}|}\]
\[\frac{\delta}{\delta f(\textbf x)}\exp{\int d^d\textbf x’\left(\frac{1}{2}f(\textbf x’)^2+\frac{1}{2}\biggr|\frac{\partial f}{\partial\textbf x’}\biggr|^2\right)}\]
(comment: the \(1\)st and \(3\)rd functionals are local whereas the \(2\)nd functional one is nonlocal).
Solution #\(16\): Functions \(f(\textbf x)\) are like \(\infty\)-dimensional vectors whose components are indexed not by a discrete label \(i\) but rather by a continuous label \(\textbf x\in\textbf R^d\); one may as well write \(f_{\textbf x}\) or \(\langle\textbf x|f\rangle\) rather than \(f(\textbf x)\) to stress this point. The functional derivative is then conceptually no different from a partial derivative with respect to one of these infinitely many components \(f(\textbf x)\) of \(f\) and just ends up returning a vanilla function of \(\textbf x\).
Just as for a function \(F(f_1,f_2,…)\) of some real variables \(f_1,f_2,…\in\textbf R\) one has the total differential:
\[dF=\sum_i\frac{\partial F}{\partial f_i}df_i\]
So simply replacing the sum by an integral \(\sum_i\mapsto\int d^d\textbf x\), the functional derivative of a functional \(F[f]\) with respect to the variable \(f(\textbf x)\) is defined by requiring:
\[\delta F=\int d^d\textbf x\frac{\delta F}{\delta f(\textbf x)}\delta f(\textbf x)\]
where typically one takes \(\delta\phi(\textbf x)=0\) for \(\textbf x\) on the boundary of the domain of integration in \(\textbf R^d\).



Notice that functional derivatives are much easier to compute than functional integrals, reflecting the general trend from single-variable calculus that differentiation is easier than integration. In particular, for functionals which are either directly (as in the \(1\)st and \(2\)nd examples) or indirectly (as in the \(3\)rd example) related to some integral of the function, functional differentiation just boils down to partial differentiation of the integrand. Keeping in mind that \(\delta f(\textbf x)=0\) for \(\textbf x\in\partial\), the key identity in this regard, generalizing the Euler-Lagrange equations, is:
\[\frac{\delta}{\delta f(\textbf x)}\int d^d\textbf x\mathcal F\left(f,\frac{\partial f}{\partial\textbf x},\frac{\partial^2 f}{\partial\textbf x^2},…\right)=\frac{\partial\mathcal F}{\partial f}-\frac{\partial}{\partial\textbf x}\cdot\frac{\partial\mathcal F}{\partial(\partial f/\partial \textbf x)}+\frac{\partial^2}{\partial\textbf x^2}\cdot\frac{\partial^2\mathcal F}{\partial(\partial^2 f/\partial \textbf x^2)}-+…\]
Problem #\(17\): Another functional differentiation problem for fun. For any function \(f(\textbf x)\), the “orthonormality condition”:
\[\frac{\delta f(\textbf x’)}{\delta f(\textbf x)}=\delta^d(\textbf x-\textbf x’)\]
seems intuitively clear (cf. \(\langle\textbf x’|\textbf x\rangle=\delta^3(\textbf x-\textbf x’)\) in nonrelativistic quantum mechanics), but how does one go about “rigorously” interpreting and proving it?
Solution #\(17\): The idea, mentioned before, is to view \(f(\textbf x’)=\text{eval}_{\textbf x’}[f]\) as an evaluation functional so that it really can be interpreted as a functional derivative. Then, express the evaluation functional in the integral form that one is most comfortable with by convolving with a delta:
\[f(\textbf x’)=\int d^d\textbf x f(\textbf x)\delta^d(\textbf x-\textbf x’)\]
So the result follows from the considerations in Solution #\(16\).
Problem #\(18\): In general, what are the dimensions of the functional derivative \(\frac{\delta F}{\delta f}\)? Look back at both Problems #\(16,17\) and check this.
Solution #\(18\): Although one is used to intuitively reading off dimensions from regular derivatives like \([\partial f/\partial x]=[f]/[x]\), for functional derivatives the “densitized” nature of the continuum \(d^d\textbf x\) means that actually:
\[\biggr[\frac{\delta F}{\delta f(\textbf x)}\biggr]=\frac{[F]}{[f][\textbf x]^d}\]
which in particular is not equal to the naive \([F]/[f]\) (unless \(d=0\) which is boring). This is consistent with the functional derivatives in both Solutions #\(16,17\).
Problem #\(19\): Show that by “applying an external magnetic field” \(E_{\text{ext}}(\textbf x)\) to the system, one has for the original exact Landau-Ginzburg free energy bifunctional \(F[\phi,E_{\text{ext}}]\) (i.e. not the quadratic approximation \(F=F[\delta\phi]\)) the functional derivatives:
\[\langle\phi(\textbf x)\rangle=\frac{\delta\ln Z}{\delta\beta E_{\text{ext}}(\textbf x)}\]
\[\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle=\frac{\delta^2\ln Z}{\delta\beta E_{\text{ext}}(\textbf x’)\delta\beta E_{\text{ext}}(\textbf x)}\]
Solution #\(19\): The Zeeman coupling is linear, so mathematically it is a chemist’s Legendre transform of the free energy density \(\mathcal F\) from \(\phi\to E_{\text{ext}}\):
\[\mathcal F=-E_{\text{ext}}\phi+\frac{\mu^2}{2}\phi^2+\frac{\gamma}{2}\biggr|\frac{\partial\phi}{\partial\textbf x}\biggr|^2+…\]\]
Proving the \(1\)st identity concerning \(\langle\phi(\textbf x)\rangle\) is easy, and it can be directly substituted into the \(2\)nd identity to give:

where at the end one can also turn off \(E_{\text{ext}}(\textbf x)=0\) to get the ensemble average and \(2\)-point correlator in the original \(\textbf Z_2\) theory.
Problem #\(20\): Using the results of Solution #\(17\), show that, reverting back to the quadratic approximation of \(\mathcal F\):
\[\langle\phi(\textbf x)\rangle=(E_{\text{ext}}*G)(\textbf x)=\int d^d\textbf x’E_{\text{ext}}(\textbf x’)G(\textbf x-\textbf x’)\]
\[\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle=\beta^{-1}G(\textbf x-\textbf x’)\]
where \(G(\textbf x)\) is almost a stationary point of the quadratic free energy \(F\) through being the fundamental Green’s function of the Helmholtz operator \(\mu^2-\gamma\biggr|\frac{\partial}{\partial\textbf x}\biggr|^2\) on \(\textbf R^d\):
\[G(\textbf x)=\int\frac{d^d\textbf k}{(2\pi)^d}\frac{e^{-i\textbf k\cdot\textbf x}}{k^2+1/\xi^2}\]
Solution #\(20\): Essentially the same as Solution #\(11\) except that need to first complete the square in the free energy (which uses the reality of the magnetic field \(E_{\text{ext}}(\textbf x)\in\textbf R\Leftrightarrow E_{\text{ext}}^{-\textbf k}=\left(E_{\text{ext}}^{\textbf k}\right)^{\dagger}\))
\[F=\int\frac{d^d\textbf k}{(2\pi)^d}\left(\frac{\mu^2+\gamma |\textbf k|^2}{2}\biggr|\phi_{\textbf k}-\frac{E_{\text{ext}}^{\textbf k}}{\mu^2+\gamma |\textbf k|^2}\biggr|^2-\frac{|E_{\text{ext}}^{\textbf k}|^2}{2(\mu^2+\gamma |\textbf k|^2)}\right)\]
and then do the Gaussian path integral to get the original partition function (when \(E_{\text{ext}}=0\)) with an additional Plancherelian contribution from \(E_{\text{ext}}\neq 0\):
\[\ln Z=\frac{1}{2}\sum_{\textbf k}\ln\frac{\pi V}{\beta(\mu^2+\gamma |\textbf k|^2)}+\frac{\beta}{2}\int\frac{d^d\textbf k}{(2\pi)^d}\frac{|E_{\text{ext}}^{\textbf k}|^2}{\mu^2+\gamma |\textbf k|^2}\]
Or, substituting \(E_{\text{ext}}^{\textbf k}=\int d^d\textbf x E_{\text{ext}}(\textbf x)e^{-i\textbf k\cdot\textbf x}\) to revert from \(\textbf k\mapsto\textbf x\) (in anticipation that one would like to take functional derivatives with respect to \(E_{\text{ext}}(\textbf x)\)):
\[\ln Z=\frac{1}{2}\sum_{\textbf k}\ln\frac{\pi V}{\beta(\mu^2+\gamma |\textbf k|^2)}+\frac{\beta}{2}\int d^d\textbf x d^d\textbf x’ E_{\text{ext}}(\textbf x)E_{\text{ext}}(\textbf x’)G(\textbf x-\textbf x’)\]
From here, the functional derivatives are straightforward to take (and only the \(2\)nd term above matters):
(aside: it seems that \(E_{\text{ext}}(\textbf x)\) can be interpreted as some kind of fluctuation distribution which, upon convolving with the \(2\)-point correlator, solves the inhomogeneous Helmholtz equation:
\[\left(\mu^2-\gamma\biggr|\frac{\partial}{\partial\textbf x}\biggr|^2\right)\langle\phi(\textbf x)\rangle=E_{\text{ext}}(\textbf x)\]
is there an intuitive interpretation of this?)
Problem #\(21\): Check explicitly that:
\[\left(\mu^2-\gamma\biggr|\frac{\partial}{\partial\textbf x}\biggr|^2\right)G(\textbf x)=\delta^d(\textbf x)\]
Solution #\(21\):




Problem #\(22\): Show that the isotropic \(2\)-point correlator has the Ornstein-Zernicke asymptotics:
\[\beta\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle\sim\begin{cases}2^{-3d/2}\pi^{(d-1)/2}e^{-d/8}d^{(d-4)/2}\gamma^{-1}\frac{1}{r^{d-2}}, & r\ll\xi\\ 2^{-(d+3)/2}\pi^{(1-d)/2}\gamma^{-1}\frac{e^{-r/\xi}}{\xi^{(d-3)/2}r^{(d-1)/2}}, & r\gg\xi
\end{cases}\]
where \(r:=|\textbf x-\textbf x’|\) the identity \(\frac{d-3}{2}+\frac{d-1}{2}=d-2\) ensures dimensional consistency and is a good way to remember it (and indeed the \(d-2\) follows on dimensional analysis grounds).
Solution #\(22\): The derivation involves a clever saddle-point approximation:








Problem #\(23\): What are the critical exponents \(\nu,\eta\) in the mean-field, quadratic approximation to the full Landau-Ginzburg theory?
Solution #\(23\): The critical point is defined to be where the quadratic coupling \(\mu^2=0\) vanishes. So get any kind of critical exponent, this always just means: write down a formula for the quantity of interest in terms of \(\mu\sim |T-T_c|^{1/2}\). For the correlation length, this is trivial:
\[\xi=\frac{\sqrt{\gamma}}{\mu}\sim\mu^{-1}=|T-T_c|^{-1/2}:=|T-T_c|^{-\nu}\]
so \(\nu=1/2\). In particular, as \(T\to T_c\), the correlation length diverges \(\xi\to\infty\). This also means that at the critical point, the only relevant regime in the Ornstein-Zernicke \(2\)-point correlator is \(r\ll\xi\to\infty\), so:
\[\langle\delta\phi(\textbf x)\delta\phi(\textbf x’)\rangle\sim\frac{1}{r^{d-2}}:=\frac{1}{r^{d-2+\eta}}\]
So \(\eta=0\).
Problem #\(24\): Using the Ginzburg criterion, rationalize why the upper critical dimension \(d_c=4\) for the Ising model.
Solution #\(24\): Conceptually, the Ginzburg criterion is a common sense idea. It says that mean-field theory has sensible things to say about the critical point \(T\to T_c\) iff:
\[\text{fluctuations about mean field}\ll\text{mean field itself}\]
or, mathematically (integrating only within a \(\xi\)-ball due to the exponentially-decaying Ornstein-Zernicke correlation for \(|\textbf x|>\xi\)):
\[\int_{|\textbf x|\leq\xi}d^d\textbf x\langle\delta\phi(\textbf x)\delta\phi(\textbf 0)\rangle\ll\int_{|\textbf x|\leq\xi}d^d\textbf x\phi_0^2\]
\[\frac{\xi^2}{\beta\gamma}\ll\xi^d\phi_0^2\]
\[\frac{\xi^{2-d}}{\phi_0^2}\ll\beta\gamma\]
\[\frac{|T-T_c|^{\nu(d-2)}}{|T-T_c|^{2\beta}}\ll\beta\gamma\]
But at the critical point \(\beta\gamma\to\beta_c\gamma_c\) which is just finite number…so in order for the left side of the equality to also remain bounded as \(T\to T_c\), require:
\[\nu(d-2)\geq 2\beta\]
Using the mean-field exponents \(\beta=\nu=1/2\), this amounts to constraining \(d\geq 4=d_c\). This may seem a bit sketchy given that the mean-field exponents for \(\beta,\nu\) were used in the Ginzburg criterion to show that mean-field theory is consistent. Rather, it just means that MFT is self-consistent. On the other hand, for \(d<4\), MFT literally predicts its own demise!