Fluctuation effects at the onset of density wave order
with two pairs of hot spots in two-dimensional metals
Abstract
We analyze quantum fluctuation effects at the onset of charge or spin density wave order in two-dimensional metals with an incommensurate nesting () wave vector connecting two pairs of hot spots on the Fermi surface. We first compute the momentum and frequency dependence of the fermion self-energy near the hot spots to leading order in a perturbation expansion (one loop). Non-Fermi liquid behavior with a linear (in energy) quasi-particle decay rate and a logarithmically vanishing quasi-particle weight is obtained. The momentum dependence of the self-energy entails only finite renormalizations of the Fermi velocity and the Fermi surface curvature at the hot spots. The perturbative one-loop result is not self-consistent and casts doubt on the stability of the quantum critical point. We construct a self-consistent solution of the one-loop equations with self-energy feedback, where the quantum critical point is stabilized rather than being destroyed by fluctuations, while the non-Fermi liquid behavior as found in the perturbative one-loop calculation is confirmed.
I Introduction
Quantum critical fluctuations at the onset of charge or spin density wave order in two-dimensional metals destroy Fermi liquid behavior and lead to unconventional dependences on temperature and other control parameters. loehneysen07 The theory of such systems is difficult due to a complex interplay of the critical order parameter fluctuations and the gapless excitations of the electrons. A purely bosonic order parameter theory as developed by Hertz hertz76 and Millis millis93 may describe some features, but it is not generally applicable since the electronic excitations lead to singular interactions of the order parameter fluctuations. abanov04 Secondary instabilities generated by the fluctuations, especially pairing instabilities, are almost unavoidable and further complicate the analysis.
The most intensively studied case is the quantum critical point at the onset of Néel-type antiferromagnetic order. abanov03 Non-Fermi liquid behavior is obtained already at leading order in perturbation theory, but new features appear at higher orders. metlitski10_af1 Impressive analytical lee18 and numerical berg12 efforts have led to substantial progress, but a complete theory of this important universality class has not yet been constructed.
A special and particularly intricate situation arises when the density wave vector is a nesting vector of the Fermi surface, that is, when it connects Fermi points with antiparallel Fermi velocities. footnote_perfnest Charge and spin correlation functions exhibit a singularity at such wave vectors due to an enhanced phase space for low-energy particle-hole excitations. In inversion symmetric crystals with a valence band dispersion , nesting vectors are determined by the condition , where is zero or a reciprocal lattice vector, and is the Fermi energy. Since nesting vectors in isotropic continuum systems are determined by the Fermi surface radius via the simple relation , we refer to nesting vectors also as “-vectors”.
Nesting singularities are particularly pronounced in metals with reduced spatial dimensionality. Charge and spin susceptibilities in low-dimensional metals exhibit peaks at nesting vectors, so that these vectors are natural charge or spin density wave vectors. For example, the ground state of the two-dimensional Hubbard model undergoes a continuous quantum phase transition into a spin-density wave state with a wave vector satisfying the nesting condition, at least within mean-field theory. schulz90 ; igoshev10 Also -wave bond charge order generated by antiferromagnetic fluctuations in spin-fermion models for cuprates metlitski10_af2 ; sachdev13 occurs naturally with nesting wave vectors. holder12 ; punk15
Quantum criticality in a two-dimensional metal at the onset of density wave order with a nesting wave vector was first analyzed by Altshuler et al. altshuler95 They obtained non-Fermi liquid power laws for the electron self-energy and the susceptibility for the special case where the wave vector is not only a nesting vector, but also half a reciprocal lattice vector. Later it was shown that due to additional umklapp processes Landau quasi-particles actually survive, albeit with an enhanced decay rate, for the particular case where is the antiferromagnetic wave vector . bergeron12 ; wang13 For generic, that is, incommensurate footnote_incomm nesting vectors instead, Altshuler et al. altshuler95 found strong infrared divergences in two-loop contributions to the susceptibility and concluded that the order-parameter fluctuations destroy the quantum critical point, replacing the second-order by a first-order phase transition.
The analysis of non-Fermi liquid behavior at the onset of charge and spin-density wave order with an incommensurate nesting vector in two-dimensional metals was initiated more recently. The fermion self-energy was calculated to first order (one loop) in the order parameter fluctuation propagator. holder14 The fluctuation propagator was computed from a one-loop approximation with bare particle-hole bubbles. A breakdown of Landau Fermi liquid theory was obtained at hot spots on the Fermi surface, that is, Fermi points connected by the ordering wave vector . Two qualitatively distinct cases arise according to the number of hot spots connected by .
In the simplest case, connects only a single pair of hot spots, and points typically in a high symmetry direction (axial or diagonal). The frequency dependence of the one-loop fermion self-energy at the hot spots follows a power law with an exponent in this case. holder14 The decay rate of fermionic excitations is thus larger than their excitation energy in the low energy limit. The momentum dependence yields a singular renormalization of the Fermi velocity and a flattening of the Fermi surface near the hot spots. sykora18 An important issue is the self-consistency of these results, or whether the non-Fermi liquid self-energy destroys the quantum critical point, since it has a tendency to wipe out the peak in the density wave susceptibility. It was shown that the quantum critical point can survive only with the assistance of a sufficiently strong Fermi surface flattening around the hot spots. sykora18 This scenario was subsequently supported by a systematic -expansion around the critical dimension , at least to leading order in . halbinger19
In the second case, the density wave vector connects two pairs of hot spots. The Fermi velocities are collinear within each pair, but not between the pairs. In Fig. 1 we show various possible geometries, with Fermi surfaces and hot spot pairs on the left, and the lines of all nesting vectors satisfying the condition on the right. These lines are obtained by backfolding the line consisting of all doubled Fermi wave vectors into the first Brillouin zone. Wave vectors connecting two pairs of hot spots correspond to crossing points of the lines of nesting vectors. The geometry in the top row of Fig. 1 is realized by the spin-density wave quantum critical point in the two-dimensional Hubbard model, while the -wave bond charge order instability mentioned above provides an example for the geometry in the second row. For two hot spot pairs, the imaginary part of the one-loop self-energy at the hot spots was found to be a linear function of frequency, with a distinct prefactor for positive and negative frequencies. holder14 Hence, the decay rate of fermionic excitations is proportional to their excitation energy, implying a marginal breakdown of Fermi liquid theory.






In this work we extend the analysis of fluctuation effects at the onset of density wave order with two pairs of hot spots. We compute the frequency dependence of the one-loop self-energy on the imaginary frequency axis and show that the results are consistent with the real frequency self-energy obtained previously. holder14 We also compute the momentum dependence near the hot spots. Unlike the case of a single hot spot pair, sykora18 the momentum dependence of the self-energy entails only a finite renormalization of the Fermi velocity, and the renormalized Fermi surface maintains a finite curvature at the hot spots.
The perturbative one-loop results are not self-consistent. Incorporating the one-loop self-energy in the fermion propagator can lead to a shift of the peak in the susceptibility away from the nesting vector, which spoils the quantum critical point. To include the self-energy feedback self-consistently, we explore two distinct routes. First, we compute the fluctuation propagator and the fermion self-energy from a renormalization group flow. Due to a peculiar cutoff dependence of the fermion self-energy, the self-energy feedback is incomplete in this approach and the shift of the susceptibility peak persists. Second, we obtain a fully self-consistent solution by solving the coupled integral equations for the particle-hole bubble and the fermion self-energy with self-energy feedback. Here the peak at the nesting vector survives.
The paper is structured as follows. In Sec. II we derive the order parameter susceptibility and effective interaction at the quantum critical point in one-loop approximation with bare fermion propagators. In Sec. III we evaluate the frequency and momentum dependence of the one-loop fermion self-energy near the hot spots. In Sec. IV we show that the perturbative one-loop results are not self-consistent. In Sec. V we discuss our renormalization group analysis of the quantum critical point, and in Sec. VI we present the self-consistent solution of the one-loop equations with self-energy feedback. A conclusion in Sec. VII closes the article.
II RPA susceptibility and effective interaction
We consider a one-band system of interacting fermions with a bare dispersion relation . Our calculations are based on the standard quantum many-body formalism with an imaginary frequency representation of dynamical quantities.negele87 The bare fermion propagator is given by
(1) |
where is the frequency variable, and is the single-particle energy relative to the chemical potential . At zero temperature the Matsubara frequency is a continuous variable.
We assume that, in mean-field theory, the system undergoes a charge or spin density-wave instability with an incommensurate and nested () modulation vector at a QCP which can be reached by tuning a suitable parameter such as electron density or interaction strength. Approaching the QCP from the normal metallic regime, the instability is signalled by a diverging RPA susceptibility
(2) |
where is the coupling constant parametrizing the bare interaction in the instability channel, and is the bare susceptibility
(3) |
is the spin-multiplicity ( for spin- fermions), and is a form factor related to the internal structure of the density-wave order parameter. For an order parameter with -wave symmetry, is symmetric under rotations and reflections. In the following we assume for definiteness. A generalization to form factors with other symmetries such as -wave symmetry is straightforward. Eq. (2) holds also for the spin susceptibility in the normal (symmetric) phase of spin-rotation invariant systems, where all components of the spin susceptibility are equal.


The bare susceptibility exhibits a square-root singularity on the line defined by all vectors in -space. To parametrize momenta near a momentum , we introduce relative coordinates normal and tangential to the line at , which we denote by and , respectively (see Fig. 2). For momenta near and low frequencies, the bare susceptibility can be expanded as altshuler95 ; holder14
(4) |
where and are positive constants, is a real constant, and
(5) |
The energy-momentum relation is given by
(6) |
where is the Fermi velocity at the hot spots on the Fermi surface connected by , and parametrizes the Fermi surface curvature at these points ( is the radius of curvature). Hence, is the oriented distance of from the -line. The prefactor of the square-root term is determined by the Fermi velocity and curvature near as
(7) |
while the other constants and receive contributions from everywhere. For fermions with a parabolic dispersion in the continuum and a constant form factor, and vanish, as can be seen from Stern’s exact analytic formula for in this case. stern67 The constant vanishes also at high symmetry points for lattice fermions, especially if points along an axis or a diagonal in the Brillouin zone. In this case a correction of order may become relevant. sykora18
Throughout this article we consider the case where the ordering wave vector connects two pairs of hot spots and on the Fermi surface. The wave vector is then a crossing point of two lines, as shown in Fig. 1 (right panel). Let and be the normal and tangential coordinates relative to the first of the lines as introduced above. The normal and tangential coordinates relative to the second line, and , respectively, are related to the former by
(8) |
where is the angle between the Fermi surface normal vectors at and , which is also the angle between the two lines crossing at (see Fig. 2). The singularities on these two lines need to be added, such that holder14
(9) |
where and are defined relative to the first line, while is defined relative to the second -line. The last two terms in Eq. (9) describe the leading regular momentum dependence near . We have chosen the variables and to parametrize this dependence. footnote_chi_0 The hot spots are often related by a point group symmetry of the lattice, as in the first two examples in Fig. 1, so that , , and . The bare susceptibility exhibits a peak at if and only if . In the following we assume that this is the case.
At the QCP one has , so that the RPA susceptibility assumes the singular form
(10) |
To deal with the critical order parameter fluctuations, the perturbation expansion has to be organized in powers of a dynamical effective interaction. This arises naturally as a boson propagator by decoupling the bare interaction in the instability channel via a Hubbard-Stratonovich transformation.hertz76 ; millis93 Alternatively it can be obtained by an RPA resummation of particle-hole bubbles or ladders. In the simplest case of a charge-density wave instability in a spinless fermion system, the RPA effective interaction can be written as
(11) |
For a charge-density wave instability in a spin- fermion system, the effective interaction has the diagonal spin structure
(12) |
where () are the spin-indices of the ingoing (outgoing) fermions. For a spin-density wave, the effective interaction acquires a non-diagonal spin structure. In a spin-rotation invariant system of spin- fermions, it can be written as
(13) |
where is the vector formed by the three Pauli matrices . footnote3
Expanding the bare susceptibility as in Eq. (9), the effective interaction at the QCP assumes the asymptotic form
(14) |
It thus features the same singularity as the RPA susceptibility. In case that the coupling has a (regular) momentum dependence for near , the coefficients and are not determined by only, but receive additional contributions from the expansion of around .
III One-loop fermion self-energy
To leading order in the effective interaction, the fermion self-energy is given by the one-loop expression
(15) |
with for a charge density and for a spin density instability. We evaluate for low frequencies and momenta near one of the hot spots, say . The dominant contributions come from momentum transfers near , such that is situated near the antipodal hot spot .
In our derivations we assume that the Fermi surface is convex at the hot spots. Results for the case of a concave Fermi surface follow from a simple particle-hole transformation. Using normal and tangential coordinates for near , we expand the dispersion relation to leading order as , and for near .
Inserting Eq. (14) into Eq. (15), the one-loop self-energy is thus given by the integral
(16) | |||||
Note that we have not yet introduced an ultraviolet cutoff for the momentum integral. We will do so at a later stage of the evaluation, whenever needed.
III.1 Frequency dependence at hot spot
For finite frequencies () the self-energy is complex. The imaginary part of the integral in Eq. (16) diverges logarithmically for large momenta, such that an ultraviolet cutoff is needed. In the low-frequency limit, the imaginary part of the self-energy at a hot spot can be computed analytically (see Appendix A), and behaves as
(17) |
where
(18) |
and is an ultraviolet momentum cutoff. In the evaluation of , a cutoff limiting the geometric mean of and turned out to be sufficient and convenient (see Appendix A). Other choices of a cutoff would merely amount to a different factor in the argument of the logarithm.
Subtracting the zero frequency constant, the real part of the self-energy is finite, that is, it does not require any ultraviolet cutoff. A simple rescaling of variables yields
(19) |
for , where is a real function depending only on the parameter . The function is given by a double integral which doesn’t seem to be elementary (see Eq. (60) in Appendix A). However, we could show analytically that for . In Fig. 3 we plot as obtained from a numerical integration of Eq. (60.

The leading low frequency behavior of the one-loop self-energy, Eqs. (17) and (19), exhibits a remarkable cancellation. Although the tangential momentum dependence of and the corresponding Fermi surface curvature were crucial in deriving these results, the mass , which determines the curvature radius, drops out. Moreover, the parameters characterizing the fermion dispersion near do not appear in the asymptotic expressions.
In Ref. holder14, the one-loop self-energy was evaluated on the real frequency axis. The result obtained in that paper can be written in the following form footnote4
(20) |
where is the sign of , and
(21) |
For large , these coefficients have the simple form and . We can relate these results for the one-loop self-energy to our results on the imaginary frequency axis by using the Kramers-Kronig-type relation for the self-energy in the complex frequency plane
(22) |
where is an arbitrary complex number (but not real), and we have dropped the momentum argument. Inserting the real-frequency behavior (20), one finds
(23) |
For large , this implies in agreement with Eq. (19). For , the imaginary part of the integral in Eq. (22) diverges logarithmically. Restricting the frequency integration by an ultraviolet cutoff one obtains
(24) |
This is consistent with Eq. (17) if . In Ref. holder14, this latter relation was reported only for the limit , while it is actually true for any . Note that the momentum cutoff in Eq. (17) is not equivalent to the frequency cutoff . Comparing Eqs. (17) and (24) suggests that a momentum cutoff corresponds to a frequency cutoff . Anyway, numerical prefactors such as in the argument of the logarithm can also be absorbed in the subleading correction of order .
The above results have been derived for the case where the Fermi surface is convex at the hot spots. For a concave Fermi surface, Eq. (17) for remains the same, in Eq. (19) for there is a sign change (from minus to plus) on the right hand side, and the expressions for and in Eq. (21) are exchanged.
The logarithm in Eq. (17) implies a logarithmic divergence in the one-loop contribution to the inverse quasi-particle weight, . footnote_Z Hence, Landau quasi-particles do not seem to exist at the hot spots. Moreover, the self-energy exhibits an anomalous real part proportional to . This contribution is directly related to the asymmetry of the real frequency self-energy with respect to a sign change of .
III.2 Momentum dependence near hot spot
We now discuss the momentum dependence of the one-loop self-energy near the hot spot at zero frequency, which we denote by , where and are the normal and tangential momentum coordinates, respectively, relative to the hot spot, and the frequency argument has been dropped.


The leading normal momentum dependence for small has the form
(25) |
where are positive numbers depending on and on the sign of . The derivation of this result and an expression for in the form of a two-fold integral is provided in Appendix A. No ultraviolet cutoff is needed here. In Fig. 4 we show the dependence of on . One can see that for all . For small , one finds and , while for large both coefficients are of order . Due to the positive sign of , the self-energy enhances the bare Fermi velocity of the system. The enhancement is stronger for than for , leading to a kink in the dispersion relation at the hot spot. The above results have been derived for a convex Fermi surface. For a concave Fermi surface one obtains the same results with and exchanged.
The leading tangential momentum dependence for small has the form
(26) |
where is a dimensionless function of , and is an ultraviolet cutoff on and . The derivation of this result is provided in Appendix A. The linear dependence of the self-energy entails a tilt of the Fermi surface at the hot spot. A deformation of the Fermi surface due to interactions is a generic phenomenon, not restricted to critical behavior. Since the hot spots and their antipodes on the Fermi surface are related by point group symmetries, the tilt does not spoil the nesting condition of collinear Fermi velocities. Hence, the critical behavior is not affected by this tilt.
To analyze the contribution of order , we have evaluated the second tangential momentum derivative of the self-energy at . Applying this derivative to Eq. (16) and pulling it under the integral leads to an ill-defined expression due to multiple poles for zero frequency and momenta on the Fermi surface. The singularity can be regularized by introducing an infrared frequency cutoff . Qualitatively, this regularization corresponds to a finite temperature , since the lowest fermionic Matsubara frequencies are given by for . The resulting integral is convergent for any , including , and it does not require any ultraviolet cutoff. A numerical evaluation of the integral shows that the regularized integral converges to a finite number even in the limit (taken after integration). Hence, there is only a finite renormalization of the quadratic dependence of the fermionic dispersion. The mass and the Fermi surface curvature thus remain finite.
IV Self-consistency check
The self-energy obtained from the perturbative one-loop calculation modifies the fermion propagator substantially at the hot spots. Since the effective interaction and the fermion self-energy were computed with bare fermion propagators, we need to check whether the results remain qualitatively the same when the bare propagator is replaced by the interacting propagator with the non-Fermi liquid form
(27) |
for low frequencies and momenta near a hot spot. From the Dyson equation and Eqs. (17) and (19) one can read off the renormalization factors
(28) | |||||
(29) |
The dispersion relation in Eq. (27) has a renormalized Fermi velocity following from the momentum dependence of the self-energy in Eq. (25),
(30) |
with a distinct renormalization for momenta inside and outside the Fermi surface, and a renormalized mass .
Eq. (27) describes the frequency dependence for only for hot spot momenta. For momenta near a hot spot the frequency dependence is the same only above a momentum-dependent crossover scale, below which Fermi liquid behavior with a large finite and a vanishing is recovered. The crossover scale is of order in the normal direction, and of order in the tangential direction. Approximating the fermion propagator by Eq. (27) in quantities that involve momentum integrals of will lead to quantitative errors, but it is not expected to affect the qualitative behavior.
In a self-consistent one-loop calculation, the effective interaction is still given by the RPA form
(31) |
as in Eq. (11), where the particle-hole bubble
(32) |
is now being computed with the interacting propagator instead of the bare one. Vertex corrections do not play an important role here, since there are no singular one-loop vertex corrections for incommensurate density waves. altshuler95 For ordering wave vectors distinct from half a reciprocal lattice vector, there is no coalescence of divergences of propagators in the vertex correction.
To extract the leading momentum and frequency dependence of for momenta near and low frequencies, we expand the dispersion to linear order in and to quadratic order in as previously. We first discard the (finite) renormalizations of the dispersion and replace by the bare dispersion , focusing thus on the changes imposed by the frequency dependence of the self-energy. The momentum integrals in Eq. (32) can then be computed analytically. In Appendix B we derive the result
(33) | |||||
where . For a numerical evaluation of the frequency integral in Eq. (33), we extend Eq. (27) to large frequencies by the ansatz and , where and are constants, and is an arbitrary ultraviolet cutoff. In this way the propagator has the correct large frequency asymptotics for . Eqs. (28) and (29) imply that the ratio is given by .

In Fig. 5 we plot the static () particle-hole bubble as a function of for and various choices of . One can see that exhibits a peak at only if . Since the behavior for small is determined by low frequency contributions, the criterion for a peak is independent of the choice of and depends only on the ratio . Hence, a peak is obtained if and only if this ratio is larger than . However, the perturbative result yields which is smaller than one. The one-loop fermion self-energy computed with bare propagators thus removes the peak in the particle-hole bubble at the nesting vector, and thus the corresponding peak in the susceptibility and in the effective interaction . The quantum critical point seems thus spoiled by the self-energy feedback, and the perturbative calculation is not even qualitatively consistent.
We now analyze the effect of the Fermi velocity renormalization (30) generated by the momentum dependence of the self-energy. Although this is a finite renormalization, it leads to a qualitative change since the renormalization factors differ for particles inside and outside the Fermi sea. We compute for momenta near and low frequencies as previously, but now with the renormalized Fermi velocity as obtained from the one-loop self-energy. We still ignore the finite renormalization of the mass. In this case only the integration in Eq. (32) can be performed analytically by residues, but the remaining two integrals can be easily carried out numerically.

In Fig. 6 we show the resulting static particle-hole bubble as a function of for various choices of the parameter . The -factors in the fermion propagators assume the -dependent values described by Eqs. (28) and (29) at low frequencies. The renormalized Fermi velocity is given by Eq. (30) for all momenta. One can see that that now exhibits a sharp peak at for sufficiently large values of . The peak at is again destroyed for small , but this time due to a change of slope for positive . The critical value for the stability of the peak at is . Note that this value depends on the choice of and . For and we find . In any case, the Fermi velocity renormalization with a larger renormalization factor for particles inside the Fermi sea helps in stabilizing the QCP with a nested wave vector. Indeed, taking only the Fermi velocity renormalization into account, and ignoring the frequency dependence of the self-energy, we found that the peak at is stabilized for any choice of .
While the QCP seems stabilized at least for sufficiently large , the one-loop self-energy feedback leads to a steeper, non-linear momentum dependence of for negative . This, in turn, reduces the singular contributions to the self-energy, so that the perturbative one-loop result with bare propagators is not self-consistent anyway.
In the following two sections we try to incorporate the self-energy feedback self-consistently, first by a renormalization group flow, and then by a self-consistent solution of the coupled integral equations for the particle-hole bubble and the fermion self-energy. In both sections we focus on the frequency dependence of the self-energy, discarding its momentum dependence for simplicity.
V Renormalization group analysis
A systematic way of dealing with low energy singularities and the corresponding divergences in perturbation theory is provided by the renormalization group, where self-consistency can in principle be achieved by solving a set of flow equations, that is, ordinary differential equations, instead of solving non-linear integral equations. In this section we derive and solve a flow equation for the fermion self-energy with self-energy feedback on the right hand side, and we compare the results to those from perturbation theory. To access the full frequency dependence of the self-energy, we use a functional renormalization group (fRG) framework. metzner12 Flow equations based on the fRG have already been applied to other quantum critical points in two-dimensional interacting fermion systems, for example, at the onset of nematic drukier12 and (non-nested) antiferromagnetic maier16 order in metals, and at the onset of superfluidity in semi-metals. obert11
V.1 Flow equation
To capture order parameter fluctuation effects efficiently, it is convenient to decouple the two-fermion interaction by introducing a bosonic order parameter field via a Hubbard-Stratonovich transformation. stratonovich58 ; hubbard59 The system is then described by a coupled fermion-boson theory with a bare fermion propagator , a bare boson progagator , and a constant Yukawa interaction which couples the fermionic charge or spin density linearly to the boson field. The bare Yukawa coupling is equal to one.
The fRG is based on a scale-by-scale evaluation of the functional integral representing the partition function and correlation functions of the system. metzner12 A flow is generated by letting the bare propagator depend on a flow parameter , usually an infrared cutoff. The corresponding scale-dependent effective action interpolates smoothly between the bare action of the system and the final effective action , from which the grand canonical potential, the self-energy, and higher order vertex functions can be obtained. The flow of is governed by an exact functional flow equation. wetterich93
We impose a sharp frequency cutoff on the fermion propagator, that is,
(34) |
This suppresses all contributions to the functional integral at the initial cutoff , and it regularizes the Fermi surface singularity at until . No cutoff is needed for the bare boson propagator , since it is bounded, and also the full boson propagator remains finite as long as .
The exact functional flow equation for leads to an infinite hierarchy of flow equations for the self-energies (for fermions and bosons) and vertex functions of any order. metzner12 We truncate this hierarchy at the leading order, that is, we keep only the self-energies and the Yukawa vertex. For incommensurate density waves, the Yukawa vertex receives no singular contributions because of the absence of coalescent divergences in the vertex correction. altshuler95 Neglecting its flow altogether does therefore not affect the qualitative behavior of the self-energies. We thus fix the Yukawa coupling at its initial value (one).
The scale-dependent self-energies are related to the full and bare propagators by the usual Dyson equations, that is and , where and are the fermionic and bosonic self-energies, respectively. The flow equation for the fermion self-energy reads
(35) |
where is the single-scale propagator metzner12
(36) |
The flow equation for the boson self-energy has the form
(37) | |||||
These flow equations can be derived by inserting the truncation described above into the exact hierarchy of flow equations. metzner12 Note that they have the same form as a -derivative of the perturbative expressions Eqs. (15) and (32), with full and scale-dependent propagators, and the proviso that on the right hand sides the derivative acts only on the explicit cutoff dependence of .
Due to the -function in the single-scale propagator , the frequency integral in Eq. (35) can be easily carried out, leading to
(38) |
We focus on the frequency dependence of the self-energy at a hot spot as described by the function . Neglecting the -dependence of the self-energy on the right hand side of the flow equation, and absorbing the constant by a shift of the chemical potential (keeping the Fermi surface fixed), we obtain the simplified flow equation
(39) |
The singular contributions to the fermion self-energy are due to momentum transfers near . We therefore can use the coordinates and as in the perturbative calculation, and expand . An ultraviolet cutoff restricts the momentum integral to and .
To compute the boson self-energy , we partition the -integration domain in Eq. (37) in regions close to the hot spots and (so that is close to and , respectively), and regions far from the hot spots. The singular part of is entirely due to the former regions. We denote the contribution from near as and the contribution from near as . The total self-energy can thus be written as , where the last contribution comes from momenta far from and and is regular even for . For near , the regular part can be expanded to linear order in and as ). Its frequency dependence is irrelevant compared to the singular terms.
The flow equations for and containing the singular contributions can be simplified by expanding the dispersion around the hot spots. For example, for near , we expand and . Inserting this expansion, the momentum integral in the flow equation (37) is convergent without the need of an ultraviolet cutoff. For a momentum-independent self-energy, the momentum integral can be performed analytically. Extending the integration region of and to infinity does not affect the low frequency behavior, and allows for an easy evaluation via the residue theorem. The calculation is basically the same as the one leading to Eq. (81) in Appendix B, and yields
(40) | |||||
where . The flow equation for has the same form with replaced by .
For large , the flow generates only smooth contributions. Hence, we do not start the flow at , but rather at some finite . In the regular part of we discard the contributions from and insert fixed (scale-independent) parameters for , , and . By this procedure we just miss some finite renormalizations. Hence, the scale-dependent boson propagator has the final form
(41) |
where , and the flow of and is determined by Eq. (40). To tune to the quantum critical point, the renormalized coupling has to be chosen such that
(42) |
V.2 Results
We now show results for the fermion self-energy and the boson propagator as obtained from a numerical solution of the flow equations. We choose a fixed set of parameters , and . The angle between the Fermi surface normal vectors at and is chosen as , such that and . footnote_num For the initial value of the frequency cutoff we choose , and the momentum integrals for the fermion self-energy are cut off by . Other choices of the cutoffs lead to the same qualitative behavior.
The coupling constant is tuned to the quantum critical point, that is, we try to choose it such that Eq. (42) is satisfied at the end of the flow. However, it turns out that the effective interaction diverges at a wave vector slightly away from the nesting vector (see below). Hence, we show results where . The corresponding coupling constant is .


In Fig. 7 we show the result for the fermion self-energy at a hot spot as a function of frequency. We plot the ratio to reveal the deviations from a linear frequency dependence. We also show the result obtained from flow equations with bare fermion propagators, that is, without fermion self-energy feedback (here the ordering wave vector remains ), as well as the result from plain one-loop perturbation theory. The latter can be obtained from the flow equation for the fermion self-energy by neglecting the self-energy feedback and by inserting the fully integrated bare particle-hole bubble into the flow equation for the fermion self-energy.
The perturbative result agrees with the asymptotic low frequency analysis presented in Sec. III. The real part of is linear in frequency at low frequencies, and the imaginary part is proportional to , in agreement with Eqs. (17) and (19). For the ratio tends to a finite value near in agreement with the asymptotic result for . The asymptotic behavior of the real part is approached rather slowly, because subleading corrections to the leading low frequency behavior are suppressed only by a low power of frequency.
The self-energy obtained from the flow equation (with self-energy feedback) does not exhibit any simple scaling behavior, neither at intermediate nor at the lowest accessible frequencies. The imaginary part is only roughly proportional to , where the prefactor is significantly smaller than in perturbation theory, and it starts decreasing for frequencies below instead of tending to a constant. exhibits a pronounced frequency dependence both at high and low frequencies. There is a minimum at . We have checked that the change of trends in both real and imaginary parts at is not an artifact of the final infrared cutoff at which we have to stop the flow before running into numerical instabilities.
Discarding the self-energy feedback on the right hand side of the flow equations has a significant effect on the frequency dependence. Both real and imaginary parts of the self-energy obtained from such a simplified flow are proportional to over a wide frequency range down to the lowest accessible frequencies. Hence, the change of trend at observed above is obviously due to the self-energy feedback.

In Fig. 8 we show the momentum dependence of the inverse fluctuation propagator at three stages of the flow (, , and ) for as a function of . One can see that vanishes at a small negative value of , while it is negative everywhere else. It is negative also for momenta with (not shown in the figure). The fluctuation propagator thus diverges at a wave vector . The QCP is thus spoiled, and the low energy behavior will ultimately be governed by a different universality class with a non-nested ordering wave vector. Since the shift of the ordering wave vector is very small, its effect on the fermion self-energy will be visible only at very low frequencies. We do not explore this in more detail, since there are reasons to believe that the shift of is an artifact of the approximate flow equations, as we will now explain.


The real part of the self-energy exhibits a very peculiar cutoff dependence. In Fig. 9 we show how the self-energy evolves in the course of the flow. converges to its final value (for ) at a scale of order . By contrast, converges much slower, roughly at scales . For the real part of the self-energy is very far from , it even has the opposite sign.
On the right hand side of the flow equations the self-energy is evaluated for frequencies at or near . Hence, the real part of the self-energy inserted on the right hand side of the flow equations differs drastically from the real part of the final self-energy (for ), it even has the wrong sign. This is a very unusual situation, and it seems that such a self-energy feedback works against self-consistency rather than implementing it. Choosing a smooth instead of a sharp cutoff doesn’t help. At present we do not know how to improve the truncation of the exact fRG flow equation to avoid this problem.
VI Self-consistent solution
We now present a self-consistent solution of the coupled one-loop equations for the fermion and boson self-energies. Momentum dependences will be approximated in a similar spirit as in the previous section. We will also introduce an infrared frequency cutoff as in the fRG approach. Here, this cutoff is not used for setting up differential flow equations. It is merely a technical device to achieve a self-consistent solution by iteration. We will see that the nesting QCP persists in the self-consistent solution.
VI.1 Self-consistent equations
The self-consistent one-loop equations are given by
(43) |
where is the full propagator, and Eq. (31) for the effective interaction with the particle-hole bubble from Eq. (32). Together these equations form a coupled set of non-linear integral equations which are very hard to solve with the accuracy needed to resolve the low frequency behavior. We therefore focus on the momentum region near the hot spots, and we simplify the equations by expanding the momentum dependences in close analogy to what we did in the previous sections.
We define , neglect the -dependence of the self-energy on the right hand side of Eq. (43), and absorb the constant by a shift of the chemical potential. Introducing a sharp infrared frequency cutoff , we then obtain
(44) |
where for and zero else. An ultraviolet momentum cutoff restricts the momentum integration to and . The dispersion is expanded as previously around the hot spot, that is, .
The particle-hole bubble is again decomposed as , where the first two contributions come from momenta close to and , respectively, and the last (regular) contribution from momenta far from and . For near , the regular part can be expanded to linear order in and as ). Its frequency dependence is irrelevant compared to the singular terms.
The singular contributions and are computed by expanding the dispersion around the hot spots and extending the momentum integrals to infinity. Eq. (81) yields
(45) |
has the same form with replaced by . The boson propagator is then given by Eq. (41) as in the fRG approach.
It is instructive to compare the integral equations (44) and (45) to the flow equations (39) and (40), respectively. The latter could be obtained from the former by applying a -derivative which acts only on the integration boundary imposed by on the right hand side, not on the -dependent integrands.
We solve the coupled integral equations by reducing gradually from to the smallest accessible values, and updating the fermion self-energy at each step. Due to the gradual adaption of the self-energy in this procedure a converged solution can be achieved, while by a direct iteration without infrared cutoff it is difficult to reach the self-consistent attractor. The frequency cutoff is thus a useful device to obtain a self-consistent solution. Ideally one would like to reduce it to zero at the end of the calculation. We managed to reduce it to before running into numerical instabilities.
VI.2 Results
We now show results for the fermion self-energy and the fluctuation propagator as obtained from a numerical solution of the self-consistent equations. We choose the same fixed set of parameters as in the fRG calculation in the previous section, that is, , , and . footnote_num The ultraviolet cutoffs are also the same: and . The coupling constant is tuned to the QCP, that is, it is chosen such that diverges for , but not earlier. The critical coupling is , which differs only slightly from the critical coupling obtained from the fRG.


In Fig. 10 we show the self-consistent result for the fermion self-energy at a hot spot as a function of frequency, plotting the ratio , for various choices of the infrared cutoff . At first sight the results look similar to the fRG results (see Fig. 9), except for the overall size of the real part, which is smaller by a factor of four compared to the self-energy obtained from the flow equations. The convergence of for fixed and is again very slow for the real part. However, in contrast to the fRG flow, the self-energy feedback now involves all frequencies, not only those near the scale . Hence, in the limit , the cutoff dependence of the feedback disappears.
The converged imaginary part is proportional to in a broad frequency range . It is not clear whether the decreasing slope at the lowest frequencies (below ) is exclusively due to the finite cutoff. The real part also seems proportional to in a certain frequency range, which is however limited to frequencies above . For each , exhibits a minimum at a frequency much larger than . The position of the minimum decreases with decreasing , but the ratio increases. Hence, it is not clear whether approaches zero or a finite value for . In other words, we do not know whether the minimum of persists for . In view of the slow convergence of the real part for , and our limitation to cutoffs , we believe that our results for the real part are converged only for frequencies , while for the imaginary part the results seem converged for .

In Fig. 11 we show the momentum dependence of the inverse fluctuation propagator for as a function of , for various choices of the infrared cutoff . One can see that a well defined minimum is formed at for . The point is a minimum with respect to deviations in all directions. For (not shown) the curves are much steeper than for . Hence, the QCP with a nested ordering wave vector is stabilized by self-consistently computed one-loop fluctuations. The slope of the inverse fluctuation propagator as a function of increases for small and small . The data in Fig. 11 suggest that the slope actually diverges for . In any case it is dominated by the hot spot contributions and to the fluctuation propagator, while the regular contribution is comparatively small for small . Hence, it looks as if the latter becomes irrelevant in the low energy, low momentum limit. The self-consistent solution thus seems to exhibit a higher degree of universality than the perturbative one-loop result, since the asymptotic behavior is determined entirely by the hot spot region, and the dependence on the parameters and disappears.
VII Conclusion
We have analyzed quantum fluctuation effects at the onset of charge or spin density wave order with an incommensurate wave vector in two-dimensional metals – for the case where connects two pairs of hot spots on the Fermi surface. This type of QCP is realized, for example, by the spin density wave instability of the Hubbard model at finite doping, schulz90 ; igoshev10 and by the onset of -wave bond charge order generated by antiferromagnetic fluctuations in spin-fermion models for cuprates. metlitski10_af2 ; sachdev13 ; holder12
We first computed the effective fluctuation propagator and the fermion self-energy at the QCP in a one-loop approximation without self-energy feedback. The marginal violation of Fermi liquid theory discovered in Ref. holder14, was thereby confirmed. As a function of the imaginary (Matsubara) frequency , the real part of the self-energy at the hot spots is proportional to for small , while the imaginary part is proportional to . This corresponds to an asymmetric linear frequency dependence of the imaginary part of the self-energy on the real frequency axis, with distinct coefficients for positive and negative frequencies. Unlike the case of the QCP with a single hot spot pair, sykora18 the momentum dependence of the self-energy at the hot spots leads only to finite renormalizations of the Fermi velocity and the Fermi surface curvature. The Fermi velocity is enhanced by distinct renormalization factors for momenta inside and outside the Fermi sea, leading to a kink in the renormalized fermion dispersion at the hot spots.
Going beyond the leading order perturbation expansion we found that the one-loop result computed with bare propagators is not self-consistent. The particle-hole bubble with propagators dressed by the one-loop self-energy differs strongly from the bare particle-hole bubble. Taking only the singular frequency dependence of the self-energy into account, the peak of the particle-hole bubble at the vector is shifted away from the line to a generic incommensurate wave vector. Hence, the QCP with a nested ordering wave vector, which is naturally favored in mean-field theory, seems to be spoiled by fluctuations. Including also the kink in the fermion dispersion generated by the momentum dependence of the self-energy, the peak remains pinned at the nesting vector at least for some choices of parameters, but its shape in the vicinity of always differs qualitatively from the peak of the bare bubble.
We attempted to achieve self-consistency by performing a one-loop functional renormalization group calculation with self-energy feedback on the propagators. In this approach the ordering vector again turned out to be shifted away from the nesting vector , albeit only by a small amount. We believe that this result is an artifact of the extremely slow convergence of the fermion self-energy as a function of the flow parameter (an infrared frequency cutoff), which prevents a proper feedback of the self-energy in the flow equations.
Obtaining a fully self-consistent solution of the coupled one-loop integral equations with self-energy feedback and a high resolution at low energies is numerically challenging. Fortunately, the fRG calculation provided valuable hints on how to expand around the singular points, and how to use a slowly decreasing infrared cutoff as a technical device to converge to a self-consistent solution. In this solution the QCP is not destroyed by fluctuations, at least for our choice of parameters. The peak in the fluctuation propagator (and the susceptibility) is even getting more pronounced, and its low energy structure is determined exclusively by low energy fluctuations in the hot spot region. While we cannot reach arbitrarily low frequencies, we get converged results over three frequency decades for the real part of the fermion self-energy, and five decades for the imaginary part (for imaginary frequencies). In that regime the real part is close to the result from one-loop perturbation theory, while the imaginary part follows the same behavior, but with a significantly reduced prefactor. The marginal violation of Fermi liquid theory with a logarithmically vanishing quasi-particle weight obtained from the plain one-loop expansion is thus confirmed by the self-consistent calculation. The results on the imaginary frequency axis are consistent with a roughly linear frequency dependence of the imaginary part of the self-energy on the real frequency axis, with a steeper slope for negative (positive) frequencies, if the Fermi surface is convex (concave) at the hot spots.
In summary, we have established a new universality class for quantum critical behavior in two-dimensional metals – at the onset of density wave order with an incommensurate nesting vector connecting two pairs of hot spots on the Fermi surface. It is worthwhile to further explore the critical behavior at and near this QCP. While we could reach rather low energy scales by our numerical solution of the self-consistent equations, one may try to extract the ultimate low energy behavior analytically. There are no relevant one-loop vertex corrections, but the role of higher order (two loop and beyond) corrections remains to be analyzed. One could also extend the analysis to the quantum critical regime at finite temperatures, look for secondary instabilities such as pairing, and study transport properties.
Acknowledgements.
We are grateful to Pietro Bonetti, Lukas Debbeler, Johannes Mitscherling, and Demetrio Vilardi for valuable discussions.Appendix A Computation of one-loop self-energy
In this appendix we derive the asymptotic results for the fermion self-energy, starting from the one-loop integral Eq. (16). To simplify the equations we set the global prefactor equal to one in the course of the derivation, and restore it in the results presented in the main text.
A.1 Frequency dependence at hot spot
For , that is, at the hot spot, one has and Eq. (16) reduces to
(46) | |||||
The form of the fermion propagator indicates the following scaling of the integration variables in the low-frequency limit :
(47) |
Hence, , such that . The contributions to the denominator of the effective interaction in Eq. (46) thus scale as follows: , , , for and for . Hence, contributions from are subleading compared to the contributions from . In the latter region the largest terms in the denominator are of order , and the terms of higher order in are negligible. Assuming for definiteness, and keeping only the leading terms, the self-energy can be written as
(48) | |||||
Note that we have expressed the kinetic energy in the fermion propagator in terms of and instead of and . It is now convenient to use as integration variable instead of . The Jacobian for this substitution is .
Introducing dimensionless variables via , , and , and using one obtains
(49) |
where . The integrand depends only on a single dimensionless parameter,
(50) |
We recall that needs to be positive to have a peak in the susceptibility and effective interaction, if . For , needs to be negative and the self-energy has the same form with .
A.1.1 Imaginary part
The imaginary part of the integral in Eq. (49) diverges logarithmically. The divergence is due to contibutions from the regime with , where . The leading contribution to the imaginary part of the self-energy can thus be written as
(51) |
Introducing new momentum space variables and , and using
(52) |
one finds
(53) |
The second term in the bracket is odd in and therefore yields no contribution to the integral. The remaining integral is elementary. The integrations over and are convergent (without any UV cutoff), yielding
(54) | |||||
The -integral diverges logarithmically for large . Restricting by a UV cutoff leads to a cutoff for the variable . With this cutoff, the -integral yields
(55) |
for .
A.1.2 Real part
To evaluate the real part of the self-energy at the hot spot, , we start from the expression (49). The integral for the real part diverges linearly in the ultraviolet regime. To obtain a finite result one needs to subtract the self-energy at , which is given by the same expression with replaced by zero.
The integration can be carried out analytically by a keyhole contour integration around the negative real axis. To this end, we rewrite Eq. (49) in the form
(56) |
where
(57) |
The contour integration yields
(58) |
The real part of the self-energy at the hot spot can thus be written as
(59) |
where the function is given by the double integral
(60) |
The remaining two integrations seem to be difficult to handle analytically. However, a simple analytic result can be obtained for the real part in the limit of large . In this limit, in Eq. (58) can be set to zero except in the argument of the logarithm in the last term.
For , the real parts of the first two terms in Eq. (58) integrate to zero. To see this, we temporarily introduce ultraviolet cutoffs for the integration variables and . The integration of the first term yields
(61) | |||||
Obviously diverges linearly for . However, this divergence is canceled upon subtracting , corresponding to the subtraction of the self-energy at , and the difference even vanishes in the limit . For example, for one finds
(62) |
By the same reasoning, the second contribution vanishes, too.
The real part of the third contribution has the form
for large . Substituting by , and subtracting the zero frequency constant yields
(64) | |||||
The integral over can be extended to the entire real axis (with a compensation by a factor ), since the integrand is symmetric in . The integration can then be easily evaluated by the residue theorem, yielding
(65) |
The -integral can now be performed using an integration by parts, yielding
(66) |
Inserting this into Eq. (56) we obtain
(67) |
for large .
A.2 Momentum dependence near hot spot
We now evaluate the momentum dependence of the self-energy near the hot spot at zero frequency. The momentum is parametrized by the normal and tangential coordinates relative to the hot spot, and , respectively.
We first analyze the normal momentum dependence for , which we denote by , where the frequency argument has been dropped. Eq. (16) yields
(68) | |||||
The form of the fermion propagator indicates the following scaling of the integration variables in the limit :
(69) |
Following the same arguments as for the frequency dependence, the expression (68) can be approximated by
(70) |
for small , where , and is the same as in Eq. (50). The dimensionless integration variables are defined by , , and . The integral can now be performed by a keyhole contour integration around the negative real axis, yielding
(71) |
where
(72) |
After subtracting the constant , the remaining integrations are convergent, that is, no ultraviolet cutoff is needed. One thus obtains
(73) |
where are positive numbers depending only on and the sign of . An analytic evaluation of the remaining two integrations in Eq. (71) seems difficult, but a numerical evaluation can be done with high accuracy.
We now turn to the tangential momentum dependence of the self-energy for . Eq. (16) yields
(74) | |||||
The form of the fermion propagator indicates the following scaling of the integration variables in the limit :
(75) |
Following the same arguments as for the frequency dependence, the expression (74) can be approximated by
(76) |
for small , where , and is the same as in Eq. (50). The dimensionless integration variables are defined by , , and . The integral can again be performed by a keyhole contour integration around the negative real axis, yielding
(77) |
where
(78) |
The integral in Eq. (77) remains ultraviolet divergent even after subtracting . Restricting and by an ultraviolet cutoff , a numerical evaluation of the integals yields
(79) |
where for small and for large . Note that a cutoff on the variables and entails a cutoff on the rescaled variables and . The square-root-type UV divergence of the integral in Eq. (77) thus yields a factor proportional to , which reduces the nominally quadratic dependence of to a linear dependence.
Appendix B Computation of particle-hole bubble
In the following we evaluate the particle-hole bubble defined in Eq. (32) with a propagator of the form (27). Using radial and tangential momentum coordinates and expanding the fermion dispersion yields
(80) |
Here we have introduced the short-hand notation , where . The and integrations can be performed via the residue theorem,
(81) | |||||
Subtracting one thus obtains
(82) |
The remaining integral can be easily carried out numerically.
For the bare bubble one obtains the same expression with . The -integration is then convergent (without UV cutoff) and elementary, yielding the familiar singular contribution
(83) |
References
- (1) H. v. Löhneysen, A. Rosch, M. Vojta, and P. Wölfle, Fermi liquid instabilities at magnetic quantum phase transitions, Rev. Mod. Phys. 79, 1015 (2007).
- (2) J.A. Hertz, Quantum critical phenomena, Phys. Rev. B 14, 1165 (1976).
- (3) A.J. Millis, Effect of a nonzero temperature on quantum critical points in itinerant fermion systems, Phys. Rev. B 48, 7183 (1993).
- (4) A. Abanov and A. V. Chubukov, Anomalous scaling at the quantum critical point in itinerant antiferromagnets, Phys. Rev. Lett. 93, 255702 (2004).
- (5) A. Abanov, A. V. Chubukov, and J. Schmalian, Quantum critical theory of the spin-fermion model and its applications to cuprates: normal state analysis, Adv. Phys. 52, 119 (2003).
- (6) M. A. Metlitski and S. Sachdev, Quantum phase transitions ofmetals in two spatial dimensions. II. Spin density wave order, Phys. Rev. B 82, 075128 (2010).
- (7) S.-S. Lee, Recent developments in non-Fermi liquid theory, Annu. Rev. Condens. Matter Phys. 9, 227 (2018).
- (8) E. Berg, M. A. Metlitski, and S. Sachdev, Sign-problem-free quantum Monte Carlo at the onset of antiferromagnetism in metals, Science 338, 1606 (2012).
- (9) This notion of nesting must not be confused with the stronger condition of perfect nesting, where a momentum shift maps extended Fermi surface pieces on top of each other. Perfect nesting occurs only for special band structures and electron densities.
- (10) H. J. Schulz, Incommensurate antiferromagnetism in the two-dimensional Hubbard model, Phys. Rev. Lett. 64, 1445 (1990).
- (11) P. A. Igoshev, M. A. Timirgazin, A. A. Katanin, A. K. Arzhnikov, and V. Yu. Irkhin, Incommensurate magnetic order and phase separation in the two-dimensional Hubbard model with nearest- and next-nearest-neighbor hopping, Phys. Rev. B 81, 094407 (2010).
- (12) M. A. Metlitski and S. Sachdev, Instabilities near the onset of spin density wave order in metals, New J. Phys. 12, 105007 (2010).
- (13) S. Sachdev and R. La Plata, Bond Order in Two-Dimensional Metals with Antiferromagnetic Exchange Interactions, Phys. Rev. Lett. 111, 027202 (2013).
- (14) T. Holder and W. Metzner, Incommensurate nematic fluctuations in two-dimensional metals, Phys. Rev. B 85, 165130 (2012).
- (15) M. Punk, Nematic fluctuations and their wave vector in two-dimensional metals, Phys. Rev. B 91, 115131 (2015).
- (16) B. L. Altshuler, L. B. Ioffe, and A. J. Millis, Critical behavior of the density-wave phase transition in a two-dimensional Fermi liquid, Phys. Rev. B 52, 5563 (1995).
- (17) D. Bergeron, D. Chowdhury, M. Punk, S. Sachdev, and A.-M. S. Tremblay, Breakdown of Fermi liquid behavior at the spin-density wave quantum critical point: The case of electron-doped cuprates, Phys. Rev. B 86, 155123 (2012).
- (18) Y. Wang and A. Chubukov, Quantum-critical pairing in electron-doped cuprates, Phys. Rev. B 88, 024516 (2013).
- (19) Commensurate wave vectors can be written as a linear combination of reciprocal lattice vectors with rational coefficients, while incommensurate wave vectors cannot. Obviously there are infinitly many possible rational coefficients, but the quantitative effects of commensurability decrease rapidly with the size of their denominators.
- (20) T. Holder and W. Metzner, Non-Fermi-liquid behavior at the onset of incommensurate charge- or spin-density wave order in two dimensions, Phys. Rev. B 90, 161106(R) (2014).
- (21) J. Sýkora, T. Holder, and W. Metzner, Fluctuation effects at the onset of the density wave order with one pair of hot spots in two-dimensional metals, Phys. Rev. B 97, 155159 (2018).
- (22) J. Halbinger, D. Pimenov, and M. Punk, Incommensurate density wave quantum criticality in two-dimensional metals, Phys. Rev. B 99, 195102 (2019).
- (23) J. W. Negele and H. Orland, Quantum Many-Particle Systems (Addison-Wesley, Reading, MA, 1987).
- (24) F. Stern, Polarizability of a Two-Dimensional Electron Gas, Phys. Rev. Lett. 18, 546 (1967).
- (25) In Ref. holder14, the regular contributions to were parametrized by a sum of four terms involving normal and tangential coordinates of both lines. Two of the terms are however redundant, since two parameters are sufficient to capture the leading (linear) momentum dependence.
- (26) As illustrative examples, the structure of the effective interaction is derived for a spinless fermion model and for the Hubbard model in Ref. sykora18, .
- (27) Our parameter corresponds to in Ref. holder14, , and our angle corresponds to the angle . Moreover, was tacidly assumed to be positive in Ref. holder14, .
- (28) Our definition of follows the usual convention in quantum field theory. In textbooks on Fermi liquid theory, however, the quasi-particle weight itself (not its inverse) is frequently denoted by .
- (29) For a review on the fRG for interacting fermion systems, see W. Metzner, M. Salmhofer, C. Honerkamp, V. Meden, and K. Schönhammer, Functional renormalization group approach to correlated fermion systems, Rev. Mod. Phys. 84, 299 (2012).
- (30) C. Drukier, L. Bartosch, A. Isidori, and P. Kopietz, Functional renormalization group approach to the Ising-nematic quantum critical point of two-dimensional metals, Phys. Rev. B 85, 245120 (2012).
- (31) S. A. Maier and P. Strack, Universality in antiferromagnetic strange metals, Phys. Rev. B 93, 165114 (2016).
- (32) B. Obert, S. Takei, and W. Metzner, Anomalous criticality near semimetal-to-superfluid quantum phase transition in a two-dimensional Dirac cone model, Ann. Phys. (Berlin) 523, 621 (2011).
- (33) R. L. Stratonovich, On a Method of Calculating Quantum Distribution Functions, Soviet Physics Doklady 2, 416 (1958).
- (34) J. Hubbard, Calculation of Partition Functions, Phys. Rev. Lett. 3, 77 (1959).
- (35) C. Wetterich, Exact evolution equation for the effective potential, Phys. Lett. B 301, 90 (1993).
- (36) In the numerical solution of the flow equations from Sec. V, and of the self-consistent integral equations from Sec. VI, we actally parametrize the regular contributions to the inverse fluctuation propagator as , and choose . For , this is equivalent to with up to subleading corrections, since for and .