Depinning in the quenched Kardar-Parisi-Zhang class II: Field theory
Abstract
There are two main universality classes for depinning of elastic interfaces in disordered media: quenched Edwards-Wilkinson (qEW), and quenched Kardar-Parisi-Zhang (qKPZ). The first class is relevant as long as the elastic force between two neighboring sites on the interface is purely harmonic, and invariant under tilting. The second class applies when the elasticity is non-linear, or the surface grows preferentially in its normal direction. It encompasses fluid imbibition, the Tang-Leschorn cellular automaton of 1992 (TL92), depinning with anharmonic elasticity (aDep), and qKPZ. While the field theory is well developed for qEW, there is no consistent theory for qKPZ. The aim of this paper is to construct this field theory within the Functional renormalization group (FRG) framework, based on large-scale numerical simulations in dimensions , and , presented in a companion paper. In order to measure the effective force correlator and coupling constants, the driving force is derived from a confining potential with curvature . We show, that contrary to common belief this is allowed in the presence of a KPZ term. The ensuing field theory becomes massive, and can no longer be Cole-Hopf transformed. In exchange, it possesses an IR attractive stable fixed point at a finite KPZ non-linearity . Since there is neither elasticity nor a KPZ term in dimension , qEW and qKPZ merge there. As a result, the two universality classes are distinguished by terms linear in . This allows us to build a consistent field theory in dimension , which loses some of its predictive powers in higher dimensions.
I Introduction

Disordered elastic manifolds exhibit universal critical behavior when driven slowly, known as depinning. There are two main universality classes, each associated with a stochastic differential equation of evolution: quenched Edwards-Wilkinson (qEW) Wiese2021 , and quenched Kardar-Parisi-Zhang (qKPZ) TangKardarDhar1995 . The first class is relevant as long as the elasticity of the interface is purely harmonic, and invariant under tilting. This description is valid in a variety of situations such as magnetic domain walls in the presence of disorder a.k.a. the Barkhausen effect Barkhausen1919 ; DurinZapperi2006b ; DurinBohnCorreaSommerDoussalWiese2016 ; terBurgBohnDurinSommerWiese2021 , vortex lattices ScheidlVinokur1998 ; BucheliWagnerGeshkenbeinLarkinBlatter1998 , charge-density waves NarayanDSFisher1992b , and DNA unzipping WieseBercyMelkonyanBizebard2019 . While these systems have short-ranged elasticity, this framework can readily be adapted to describe systems with long-range (LR) elasticity such as contact-line depinning LeDoussalWieseRaphaelGolestanian2004 ; BachasLeDoussalWiese2006 ; LeDoussalWieseMoulinetRolley2009 , earthquakes DSFisher1998 ; FisherDahmenRamanathanBenZion1997 and knitting PoinclouxAdda-BediaLechenault2018b ; PoinclouxAdda-BediaLechenault2018 .
The second class is relevant when the elasticity is non-linear, or the surface grows preferentially in its normal direction. It encompasses fluid imbibition BuldyrevBarabasiCasertaHavlinStanleyVicsek1992 , the Tang-Leschorn cellular automaton of 1992 (TL92) TangLeschhorn1992 or its variants AmaralBarabasiBuldyrevHarringtonHavlinSadr-LahijanyStanley1995 , depinning with anharmonic elasticity (aDep) RossoKrauth2001b , and qKPZ TangKardarDhar1995 . That all these models are in the same universality class is non-trivial, but is now firmly established MukerjeeBonachelaMunozWiese2022 . This so-called qKPZ class has been observed for magnetic domain walls MoonKimYooChoHwangKahngMinShinChoe2013 ; Diaz-PardoMoisanAlbornozLemaitreCurialeJeudy2019 , in growing bacterial colonies HuergoMuzzioPasqualeGonzalezBolzanArvia2014 and chemical reaction fronts AtisDubeySalinTalonLeDoussalWiese2014 .
While the field theory for qEW is well established NarayanDSFisher1992a ; NarayanDSFisher1993a ; LeschhornNattermannStepanowTang1997 ; NattermannStepanowTangLeschhorn1992 ; ChauveLeDoussalWiese2000a ; LeDoussalWieseChauve2002 ; LeDoussalWieseChauve2003 ; Wiese2021 , building a field theory for qKPZ is a challenge. It has previously been attempted in Ref. LeDoussalWiese2002 . In that work, the running coupling constant for the non-linearity goes to infinity. The first question one needed to clarify was whether this is true, or an artifact of the Functional Renormalization Group (FRG) treatment. In Ref. MukerjeeBonachelaMunozWiese2022 we measured in a numerical simulation the effective action of three models: qKPZ, TL92, and aDep. For we found that all three possess an effective long-distance behavior fully described by the terms in the qKPZ equation,
(1) | |||||
The disorder forces are quenched Gaussian random variables with variance
(2) |
is the microscopic disorder-force correlator, assumed to decay rapidly for short-range (SR) disorder. In higher dimensions the same conclusions were reached, although with larger uncertainties.
Motivated by these findings, we reconsider the field theory corresponding to Eq. (1). There are two key observations. First of all, for , three universality classes merge: qKPZ, qEW with short-ranged, and qEW with LR interactions. This is visualized on Fig. 1. The second key observation is that the way we drive the system is important. In fact, we drive with a force that derives from a confining potential, the term in Eq. (1). That allows us to measure the effective force correlator defined via
(3) | |||||
(4) | |||||
(5) |
In this protocol, is increased in steps. One then waits until the interface stops, which defines . Its center-of-mass position is , and its fluctuations define .
This driving force appears in the effective action Eq. (II.1) as a mass term, as compared to LeDoussalWiese2002 , which considered a massless theory. Their motivation was that a massive term breaks Galilean invariance, and this is something “you do not want for the KPZ equation”. We believe that this is not a problem here, for two reasons: First of all, Galilean invariance is already broken by the quenched disorder , even after disorder averaging. Second, even if the driving breaks Galilean invariance, this should only affect large-scale properties, but not small-scale ones (small with respect to the correlation length), and especially not critical properties.
There is a prize to pay for introducing a massive term: one loses the Cole-Hopf transformation, a transformation that allows to map the KPZ equation to a simpler stochastic heat equation with multiplicative noise (see section III.6). This the authors of Ref. LeDoussalWiese2002 were not ready to give up, as it complicates perturbation theory. As we will see below, it breaks the non-renormalization of , allowing us to find a fixed point for the latter. As both the massive and the massless scheme give, at least in 1-loop order, the same results close to the upper critical dimension , what we will present below is not a systematic -expansion. Rather, if we suppose we know the FRG fixed point for qEW, then our scheme allows us to control qKPZ perturbatively in , in an expansion around the qEW fixed point. While the latter is known analytically in LeDoussalWiese2008a , what we use here is the 1-loop fixed point, obtained via the expansion. The latter is actually quite good even down to : It predicts a roughness exponent of , as compared to the best numerical value of GrassbergerDharMohanty2016 ; ShapiraWieseUnpublished . The FRG correlator is, even quantitatively, rather well approximated by its 1-loop value terBurgBohnDurinSommerWiese2021 ; Wiese2021 . We restrict ourselves here to 1-loop order, which has the benefit of greater transparency. Preliminary calculations show that extension to 2-loop order is straightforward though cumbersome. The method we present below allows us to compute analytically the different critical exponents as well as the full force correlator, and present quantitative agreement with the numerical simulations.
This paper is organized as follows: In the next section II we first define the field theory and review perturbation theory (section II.1). We then summarize scaling arguments described in detail in the companion paper MukerjeeBonachelaMunozWiese2022 (section II.2). The effective force correlator is defined in section II.3, and the relation to directed percolation in section II.4, followed by a discussion of the effective action measured in simulations, section II.5. Section III is dedicated to the field theory. We start with a reminder on the generation of the KPZ term from an anharmonic elasticity (section III.1). All 1-loop contributions are given in section III.2, with details relegated to appendix A. Section III.3 establishes the flow equations. Necessary conditions for their solution are derived in section III.4, followed by an analytical solution in section III.5, first giving the scheme (section III.5.1), and then explicit values in to (sections III.5.2 to III.5.4). Tables summarize our findings in section III.5.6. We comment on the Cole-Hopf transformation (section III.6), and present in layman terms physical insights from our work section IV, before concluding in section V.
II Model and Phenomenology
II.1 Model, action and perturbation theory
The Martin-Siggia-Rose MSR action corresponding to Eq. (1) reads
The field is an auxiliary field introduced to enforce Eq. (1) and called the response field. The last term is obtained by averaging over , using its variance (2). Perturbation theory is constructed by expanding around the free theory obtained by setting and in Eq. (II.1). The free response function is the response of the field to an additional force acting at , ,
(7) | |||||
In Fourier space it reads
(8) | |||||
Graphically this is represented by an arrow from to . The (microscopic) disorder is represented by two dots connected by a dashed line, whereas the KPZ vertex is a dot with two incoming lines with bars for the derivatives, and one outgoing one. Examples for diagrams correcting the disorder are given on Fig. 3. For an introduction into functional perturbation theory we refer to section 3 of Wiese2021 . Note that the disorder is corrected by the KPZ force, and what we loosely call the renormalized disorder is more precisely the renormalized force correlator, which contains contributions from the KPZ term (see Section II.3 for a detailed discussion). Non-trivial correlations necessitate at least one “disorder” vertex . As an example, the leading order to the equal-time 2-point function is
(9) | |||||
The arrows represent the response function , the dotted line the effective force correlator . We assume an upper critical dimension of as in qEW. Simulations show that in the interface is still rough MukerjeeBonachelaMunozWiese2022 , so the upper critical dimension is above . Noting that physical realizations can only be constructed in integer dimensions, the remaining open question is whether the 4-dimensional system is at its upper critical dimension, or potentially above, see section III.5.4. Note that an interface in anharmonic depinning is less rough than in qEW; this excludes an upper critical dimension larger than four.
II.2 Scaling and anomalous exponents
Scaling arguments were given in the companion paper MukerjeeBonachelaMunozWiese2022 . We recall the main results here. The static 2-point function is defined as
(10) |
The average is taken over different disorder configurations (there are no thermal fluctuaions). is the standard roughness exponent. In contrast to qEW, there is a new exponent . The reason is that the elasticity renormalizes and thus its anomalous dimension gives rise to another exponent. The quantity in Eq. (10) is the correlation length created by the confining potential. Every length parallel to the interface scales as or , whereas in the perpendicular direction it scales as . To estimate , we take in Eq. (10), obtaining . As a consequence
(11) |
Note that as for qEW. Fig. 2 shows a scaling collapse of the 2-point function with these scalings.
Define , and to be the anomalous dimensions of , and in units of ,
(12) | |||||
(13) | |||||
(14) |
In order to relate them to the standard scaling exponents , and , we first need to define . It is given by the temporal spread of the perturbation in the surface in the 2-point function as
(15) |
With these definitions at the fixed point we can derive
(16) | ||||
(17) | ||||
(18) |
The first relation is obtained from , implying . The second follows from . The last one is obtained from , implying .
II.3 The renormalized correlator
In Eq. (3) we had defined the renormalized (effective) force correlator as
(19) |
The definition of is given in Eq. (4). This is the same definition as the one used for qEW LeDoussalWiese2006a ; MiddletonLeDoussalWiese2006 ; RossoLeDoussalWiese2006a ; BonachelaAlavaMunoz2008 . Integrating the equation of motion (1) over space for a configuration at rest yields
(20) |
Thus the correlator in Eq. (19) measures fluctuations of the total force. Only for qEW () this equals the force exerted by the disorder. To be specific, let us define
(21) | |||||
(22) |
A configuration at rest then has
(23) |
Our goal is to compare observables with objects in the field theory. What is calculated there is the effective action, or more precisely its 2-time contribution. (In the statics this would be the 2-replica term.) It is the sum of all connected 2-time diagrams, i.e. with two external fields. To 1-loop order, these are shown in Fig. 3. The 2-point function is obtained to all orders by contracting the 2-time contribution to the effective action with two response functions. While in real space this is a convolution, in momentum and frequency space this is simply a multiplication with the response function . According to Eq. (19) it is to be evaluated at momentum and frequency . Recall that the response function is the response of the observable to a small uniform kick in force at . Since the center of mass follows the center of the driving parabola ,
(24) |
Thus a uniform kick leads to a response for the center of mass according to . As a result, the integrated response function is given by
(25) |
This is equivalent to .
We finally need to remember the field-theoretic definition of the effective action : It is obtained from the corresponding expectation values by amputation of the response function, which is equivalent to dividing by the response function (in Fourier representation). Due to Eq. (25) this is nothing but multiplication with , once for each of the two external fields . This gives the factor of in Eq. (19), and Eq. (19) is nothing but the 2-time contribution to the effective action , equivalent to the renormalized force correlator . It is the mode of the full effective force correlator in the field theory for depinning.
Having established that Eq. (19) is the proper definition of the renormalized , it is still instructive to study the correlations of all three forces appearing in Eq. (20). To this aim, let us define in addition to Eq. (19)
(26) | |||||
(27) | |||||
(28) |
A measurement of these quantities is shown below in Fig. 10.
Let us finally give the scaling dimensions,
(29) |
The scaling of the argument of is given by
(30) |
These scalings are reflected in the FRG flow equations derived below in Eq. (55).
II.4 Link to directed percolation, exponents given in the literature, and other relations
For TL92 in , the scaling of a blocked interface at depinning is given by directed percolation TangLeschhorn1992 ; AmaralBarabasiBuldyrevHarringtonHavlinSadr-LahijanyStanley1995 ; BarabasiGrinsteinMunoz1996 ; Hinrichsen2000 ; AraujoGrassbergerKahngSchrenkZiff2014 ; Dhar2017 . In table 1 we summarize the exponents obtained this way, which guide us in the construction and tests of the FRG. Details are given in MukerjeeBonachelaMunozWiese2022 .
In dimensions directed percolation paths are 1-dimensional, whereas the interface is -dimensional. As a result, the mapping to DP no longer exists, and one has to introduce directed surfaces BarabasiGrinsteinMunoz1996 . The exponents we find in and are summarized on table 3 (page 3).
II.5 The effective action in simulations
To guide our field-theoretical work, we first checked in dimension that the scaling exponents given in table 1 account for the measured values of and given in Eqs. (12)-(13). To this aim, a novel algorithm was designed MukerjeeBonachelaMunozWiese2022 to measure and by imposing a spatial modulation in the background-field configuration . The simulations were performed for three different models, all in the qKPZ universality class: the cellular automaton TL92 TangLeschhorn1992 , anharmonic depinning RossoKrauth2001b ; MukerjeeBonachelaMunozWiese2022 , and a direct simulation of Eq. (1) MukerjeeBonachelaMunozWiese2022 . The best results were achieved for anharmonic depinning, thanks to an efficient algorithm for its evolution RossoKrauth2001b .
With the novel algorithm designed in MukerjeeBonachelaMunozWiese2022 , we measured the effective couplings and , as a function of . In Fig. 4 (left) we show their flow as a function of . To be specific, what we measure (left), and what is predicted from DP via table 1 (right) is
(31) | |||||
(32) |
This confirms our scaling analysis and allows us to measure as shown on Fig. 4 the dimensionless amplitude
(33) |
The ideas behind this definition is that the KPZ term has one field more than the elastic term. Thus the ratio has the inverse dimension of a field, which is compensated by the first ratio. That converges to the same value for two different models gives strong evidence that qKPZ is the effective theory, and that a fixed point of the renormalization-group flow is reached. In , this ratio reads
(34) |
The last points to verify is that we can measure the effective-force correlator , that different models in the qKPZ class have the same , and that this function is close to, but distinct from the one for qEW. This is shown in Fig. 12.
III Field theory
Now that we verified that all models have a fixed point represented by the qKPZ equation, and that we have the correct scaling dimensions for every variable, we can confidently construct their field theory.
III.1 Reminder: Generation of KPZ term from anharmonic elasticity
Let us remind how anharmonic elastic terms generate a KPZ term at depinning LeDoussalWiese2002 : To this purpose consider a standard elastic energy, supplemented by an additional anharmonic (quartic) term (setting for simplicity),
(35) |
The corresponding terms in the equation of motion read
(36) | |||||
Since the r.h.s. of Eq. (36) is a total derivative, it is surprising that a KPZ-term can be generated in the limit of a vanishing driving velocity. This puzzle was solved in Ref. LeDoussalWiese2002 , where the KPZ term arises by contracting the non-linearity with one bare disorder (we drop the index on from now on for simplicity of notation),
![]() |
(37) | ||||
As , the leading term in Eq. (37) can be written as
Integrating over and using the radial symmetry in yields
(39) |
This shows that in the FRG a KPZ term is generated from the non-linearity. As , its amplitude is positive. The integral (39) has a strong UV divergence, thus the generation of this term happens at small scales, similar to the generation of the critical force, see appendix A.3.
III.2 1-loop contributions
Here we summarize the 1-loop contributions to , , and . This is almost the same calculation as in Ref. LeDoussalWiese2002 , with a little twist: Since we work in a massive scheme, many of the cancelations in LeDoussalWiese2002 no longer exist. We remind that this change in scheme was forced upon us by our decision to measure the effective parameters of the theory, necessitating to drive with a confining potential. We believe that this is also much closer to real experiments. It is a scheme widely used for perturbative RG for the Ising model in , pioneered by G. Parisi and used up to 7 loop-order by B. Nickel and collaborators ParisiBook ; Parisi1980 ; BakerNickelGreenMeiron1976 ; NickelMeironBaker1977 ; BakerNickelMeiron1978 . As discussed above, we think of this fixed-dimension renormalization scheme as an expansion around the qEW fixed point. The diagrams from the perturbation in are given in Figs. 5-7.
We obtain the same diagrams as in LeDoussalWiese2002 but with coefficients that differ from LeDoussalWiese2002 away from the upper critical dimension. The explicit calculations are given in appendix A. Terms with numerical coefficients only (no ) are those appearing already in qEW.
(40) | |||||
(41) | |||||
(42) | |||||
(43) |
(44) | |||||
(45) | |||||
(46) | |||||
(47) |
The coefficients in the limit of used by LeDoussalWiese2002 are obtained by setting , resulting into for all , except . While this is the standard procedure followed in a dimensional expansion, it misses that in dimension the KPZ term does not exist, thus cannot correct the remaining terms: viscosity , and effective force correlator . The factors of in coefficients and reflect this physical necessity. No such constraint exists for and : since they are absent from the equation of motion (1) in , their coefficients can well be modified.
As and appear in the combination of , the important question is whether this ratio is corrected. This is indeed the case as
(48) |
Note that this term is negative, and have a power in superior to one. It will therefore stop the RG flow for at large , allowing us to close our system of equations!
A final important point to mention is that the confining potential is not renormalized. In qEW this is due to the statistical tilt symmetry (STS) Wiese2021 , which can be checked perturbatively: Since the effective force correlator contains only as a difference , no field without a time derivative can be generated. The same holds true here: since the additional KPZ vertex has additional spatial derivatives, it cannot generate a field without spatial derivatives. This property is very useful, as we can as in qEW use as an RG scale, without caveat.
Finally, the critical force is
(49) | |||||
The first contribution is negative, identical to qEW. The second is positive, and specific to qKPZ. The non-linearity reduces the force needed to depin the interface. This is derived in appendix A.3.
III.3 Flow equations
Above we calculated the perturbative corrections. We now derive the corresponding RG relations. Since is not corrected under renormalization, we use it to parameterize the flow of the remaining quantities. To this aim, first define the dimensionless field as
(50) |
We have . The integral defined in Eq. (45) is evaluated in Eq. (103) of appendix A.1,
(51) |
It scales as
(52) |
where we remind that
(53) |
The dimensionless renormalized correlator is then defined in terms of the effective force correlator , such that it absorbs as
(54) |
The explicit -dependent factor in front of is the scaling dimension given in Eq. (29). This yields the flow equation for the effective dimensionless force correlator,
(55) | ||||
Here we defined the dimensionless combination
(56) |
Its flow equation is obtained from Eq. (48) as
(57) |
It has one fixed point , and a second non-trivial fixed point at
(58) |
We can see that in the fixed point disappears as goes to infinity.
The anomalous dimension defined in Eq. (12) reads
(59) |
Using Eq. (16), we find
(60) |
Eq. (55) is still cumbersome to solve. Reinjecting Eq. (60), we obtain at the fixed point
(61) | |||||
The anomalous contribution reads
(62) |
Using Eq. (18) this yields
(63) |
We note that for the contribution of in equation (61) disappears, thus we recover the qEW fixed point. This is not the case in the massless scheme LeDoussalWiese2002 . Increasing we expect the qKPZ fixed point to smoothly move away from the qEW one. In Figure 12 we show that in dimension the shape of the measured for qEW and qKPZ are close, even though their amplitudes may be rather different. We take this as an encouraging sign to construct the FRG fixed point for qKPZ. This is the task of section III.5. Since our expansion is uncontrolled, we need to obtain additional safeguards in order to see if where our approach hold, and where it is too crude. For that, we derive constraints to be satisfied by the fixed point.
III.4 Necessary conditions for a fixed point, and bounds
III.4.1 Disorder and force correlator relevant
We now assume (as in qEW) that the effective force correlator is relevant, thus . This is satisfied in , see Table 1. There one finds . To compare, in (qEW) one gets . In qEW has .
Taking the limit of in Eq. (61), we obtain a soft bound at 1-loop order,
(64) |
When violated, the rescaling term becomes negative, and we expect the effective force correlator to disappear at large scales. Using the definition of the universal amplitude in Eq. (33), we can rewrite the bound (64) as111Note that the definition (33) for remains unchanged upon replacing all quantities by their dimensionless analogue, noted with a tilde.
(65) |
III.4.2
We expect that the effective would grow at large scales, since it describes the long distance behavior of models with stronger than harmonic elasticity. As a result we demande that (which implies ). Eq. (60) then yields
(66) |
This can be rewritten as
(67) |

III.4.3 Positive pinning force
The last condition is that the critical force at depinning needs to be negative (keeping us pinned), equivalent to a negative square bracket in Eq. (49). In terms of , this results in
(68) |
We find that in the strongest bound is for the critical force, followed by the one for and ,
(69) |
It would be interesting to continue this to 2-loop order.
III.5 Solution of the flow equations
III.5.1 Scheme
How do we solve these coupled equations (Eqs. (58)-(63) ) The procedure is adapted from the standard ansatz for qEW LeDoussalWieseChauve2002 , explained in detail in Ref. Wiese2021 :
-
(i)
Use the normalization . In practice, this corresponds to setting and in Eq. (61), and then solving the flow equations with in the code.
-
(ii)
Solve the (such rescaled) flow equation (61) for . The correct solution is the one for which decays to zero at least exponentially fast: A power-law decay, or an increase with , is not permitted by the physical initial condition.
-
(iii)
The critical that satisfies Eq. (58) in our scheme is
(70) Given , the first square root is a number; the second one is the result from step (ii) above.
It is interesting to see how the different exponents depends on that is why we solve the flow equations for different instead of plugging the value given by Eq. (58).
III.5.2
The procedure and the values obtained for different are shown for in Fig. 8. We see that slightly decreases from its qEW value of . The ratio starts at for , and then grows. The effective force correlator becomes irrelevant for . At the same time the bound (65) for (marked here as a red dashed line ) is violated. The critical respects all bounds in Eq. (69). It gives
(71) | |||||
(72) | |||||
(73) | |||||
(74) |
This can be compared to their values for (qEW), , and , and the numerically obtained values , , and . The values (71)-(73) are pretty reasonable for 1-loop estimates: For qEW in comes out smaller (1 instead of ); the same reduction applies to our prediction for in qKPZ. is about larger than the numerical value. Finally, while is too large, using the numerically known value for with the same 1-loop estimate would yield , smaller than the measured value of . (Note that the prediction of in AmaralBarabasiBuldyrevHarringtonHavlinSadr-LahijanyStanley1995 is invalidated by numerics MukerjeeBonachelaMunozWiese2022 .)

III.5.3
Relevant quantities as a function of are given on Fig. 9. Evaluation at yields
(75) | |||||
(76) | |||||
(77) | |||||
(78) |
These results violate the bound (68) on for . Supposing that this is an artifact of the 1-loop approximation, the next bound to consider is the bound (65), asking that the effective force correlator is relevant at the transition. This bound is only slightly violated. We therefore hope that the values given in Eqs. (75)-(78) are usable.
Our own numerical simulations MukerjeeBonachelaMunozWiese2022 give , for TL92, and , for anharmonic depinning. We expect the latter to be more reliable as there are less finite-size corrections. The agreement is then excellent.
For comparison we note that 1-loop qEW gives , and , while numerics gives and .
III.5.4
Relevant quantities as a function of are given on Fig. 11. At the non-trivial fixed point (70) for , we find
(79) | |||||
(80) | |||||
(81) | |||||
(82) |
These values violate all bounds, and thus need to be rejected. There are four possible conclusions:
-
(i)
since the effective force correlator is irrelevant at this fixed point, there is no qKPZ class.
-
(ii)
this fixed point is irrelevant, but there is a another fixed point not contained in our approach.
-
(iii)
our approach is too crude.
-
(iv)
our approach is crude as the fixed-point value for is too large, but providing a better value for it remains predictive.
If we believe Ref. RossoHartmannKrauth2002 , there is a distinguished fixed point for both classes, eliminating (i) while allowing for (ii). While the following option (iii) is suggestive, we can still try (iv): we use such that the effective depinning force at the fixed point is zero. Since the KPZ term grows under renormalization, it will finally render all pinned configurations unstable. This in turn reduces the generation of the KPZ term, making it less relevant. Our conjecture, which needs to be validated in numerical simulations, is that the system gets stuck at this precise point. Under this assumption we obtain
(83) | |||||
(84) | |||||
(85) | |||||
(86) |
These values are pretty much in line with the simulations for anharmonic depinning in : , . We do not know the values of and .
We remark that the behavior in calls for more investigation: for example, may be between and .
III.5.5 Force amplitude ratio
Let us now address the relative fluctuations of forces defined in Eqs. (26) to (28). At leading order in perturbation theory we can estimate from Fig. 3 (where the are defined) that
(87) | |||||
(88) | |||||
(89) |
These equations simplify upon using that . Given the similar functional forms shown in Fig. 10, let us focus on the relative amplitudes. With the universal amplitude defined in Eq. (33), we get
(90) | |||||
(91) | |||||
(92) |

In our simulations in we find
(93) | |||||
(94) | |||||
(95) |
The theory in has
(96) |
which gives , and for the three ratios in Eqs. (93) to (95). Using the measured amplitude these ratios become , and which is closer to the measured amplitudes. All these values seem pretty reasonable given the order of approximation.
III.5.6 Other quantities and summary
Other properties of derived from the FRG solution are presented in table 2. An interesting property is the curvature , defined as
(97) |
It is constructed such that an exponential decaying , which gives a straight line for , has a vanishing curvature. The definition was motivated by the observation in LeDoussalWiese2002 that the FRG flow in the massless scheme possesses an exponentially decaying subspace, protected to all orders in perturbation theory. Our simulations in MukerjeeBonachelaMunozWiese2022 showed no evidence for this subspace. Still, is a scale-free parameter which allows one to distinguish different shapes.

quantity | qKPZ FT | qKPZ sim | qEW FT | |
---|---|---|---|---|
Our results for the exponents are summarized in table 3, and in Figs. 12 and 13 for the full function , rescaled such that . They show excellent agreement between theory and simulation.

Exponent | dim | field theory | simulations |
1 | 1.2781 | ||
III.6 Cole-Hopf transformation
The Cole-Hopf transformation is defined by
(98) |
It is build to remove the non-linear term proportional to from the KPZ equation (1), and reproduced here,
(99) | |||||
The transformed equation reads
(100) | |||||
Some remarks are in order:
- (i)
- (ii)
This indicates that the Cole-Hopf transformation heavily shakes up infrared and ultraviolet properties of the theory. It may therefore not be surprising that in LeDoussalWiese2002 no fixed point was found, whereas here, with properly defined physical fields, there is an FRG fixed point. A better understanding of the Cole-Hopf transformation and its consequences are desirable. We cannot exclude that it has some bearing on the perturbative treatment FreyTaeuber1994 ; Laessig1995 ; Wiese1997c ; Wiese1998a of the KPZ equation itself, or on the mapping between the KPZ equation and the corresponding directed polymer problem LeDoussalWiese2005a ; Wiese2021 , with all that this entails.
IV Physical insights
Let us summarize the main physical insights from our work:
-
1.
Most importantly, the qKPZ class covers a wide range of microscopic models, and is universal. Strong evidence for this comes from the ability of the theory to predict not only the critical exponents, but also the effective KPZ amplitude , and the force-correlator .
-
2.
The introduction of the non-linearity facilitates depinning as compared to qEW, Eq. (49). This favors “flatter” interfaces, i.e. those for which the integrated KPZ term is smaller, reducing the roughness exponent .
-
3.
The renormalized force correlator in dimension is close in shape to the correlator of qEW. This means that all properties linked to the shape of the correlator are close: for example, the avalanches-size correlations ThieryLeDoussalWiese2016 , or the correlation length . The scaling dimension of is close to its qEW counterpart, , whereas the roughness exponents are rather different.
-
4.
To properly renormalize the qKPZ class, one needs a confining potential. The confining potential forbids large fluctuations of the interface, which on the technical level provides a clear distinction between short-distance and long-distance divergences.
V Conclusion
We revisited the qKPZ universality class. Using a careful comparison to numerical simulations in dimensions , , and , we constructed a consistent theory. The crucial ingredient is a flow-equation for the KPZ non-linearity, which is controlled by dimension . Behind this feature lies the observation that all field theories for qEW with SR or LR elasticity, as well as qKPZ merge into a single theory in dimension . Our theory has predictive powers as long as we have a sufficient knowledge of the qEW fixed point in small dimensions, and we are not too far away from . We derived several bounds, respected in low dimensions, but violated in dimension ; there we currently can only close our scheme with an adhoc assumption.
We hope that our method of first measuring the effective theory in a simulation, before attempting to build a field theory, can serve in other contexts as well. Applying our approach to other growth experiments for which no theory is available seems promising DiasYunkerYodhAraujoTelo-da-Gama2018 . We hope it will also shed light on the problems in the standard (thermal) KPZ equation in higher dimensions.
Acknowledgements.
We thank Juan A. Bonachela and Miguel A. Muñoz for stimulating discussions and collaboration on the numerical part of this project, published in MukerjeeBonachelaMunozWiese2022 .Appendix A Field-theory details
As explained in the main text, our field theory is massive, with a time integrated response function given by . All diagrams are calculated with . In appendix A.1 we first give all momentum integrals appearing in the main text or used later. In the following appendix A.2, we recalculate all diagrams in the massive scheme.
A.1 Useful momentum integrals
To calculate all integrals, we use the Feynman representation of the time integrated response,
(101) |
This lets appear a normalization factor
(102) |
The elasticity and the mass both appear in the momentum integrals, and can be taken out by a rescaling of . As an example consider
(103) | |||||
In the first step, we rescaled . In the second step . These steps assume that there are no explicit cutoffs on , and that the only cutoff is set by , and dimensional regularization is used. We then used the auxiliary integral (101), and the momentum integral (102). Below we give a complete list of all encountered integrals, after rescaling to eliminate the and dependence.
(104) | |||||
(105) | |||||
(106) | |||||
(107) | |||||
(108) | |||||
(109) |
Integral (106) is the key-integral used to define the renormalized force correlator , see Eqs. (54)-(51). It is therefore useful to express as far as possible all integrals w.r.t. to integral (106), or including the dimensions w.r.t integral (103).
A.2 Diagrams
A.2.1 The coefficient
According to LeDoussalWiese2002 , Eq. (A3)
![]() |
(110) | ||||
(111) | |||||
The relevant integral is Eq. (108), thus in Eq. (40)
(112) |
Note that it does not modify in dimension .
A.2.2 The coefficient
A.2.3 The coefficient
A.2.4 The coefficient
A.2.5 The coefficient
![]() |
(119) | ||||
![]() |
(120) |
Together their amplitude (without the factor of and ) is
(121) |
Therefore in Eq. (42),
(122) |
A.2.6 The coefficient
A.3 Depinning force
The perturbative calculation gives in absence of KPZ terms
(124) |
The new contribution induced by the KPZ term is
(125) |
(There is a combinatorial factor of from , followed by a for the number of possible contractions.) The total is
(126) | |||||
References
- (1) K.J. Wiese, Theory and experiments for disordered elastic manifolds, depinning, avalanches, and sandpiles, Rep. Prog. Phys. 85 (2022) 086502 (133pp), arXiv:2102.01215.
- (2) L.-H. Tang, M. Kardar and D. Dhar, Driven depinning in anisotropic media, Phys. Rev. Lett. 74 (1995) 920–3.
- (3) H. Barkhausen, Zwei mit Hilfe der neuen Verstärker entdeckte Erscheinungen, Phys. Ztschr. 20 (1919) 401–403.
- (4) G. Durin and S. Zapperi, The Barkhausen effect, in G. Bertotti and I. Mayergoyz, editors, The Science of Hysteresis, page 51, Amsterdam, 2006, cond-mat/0404512.
- (5) G. Durin, F. Bohn, M.A. Correa, R.L. Sommer, P. Le Doussal and K.J. Wiese, Quantitative scaling of magnetic avalanches, Phys. Rev. Lett. 117 (2016) 087201, arXiv:1601.01331.
- (6) C. ter Burg, F. Bohn, F. Durin, R.L. Sommer and K.J. Wiese, Force correlations in disordered magnets, Phys. Rev. Lett. 129 (2022) 107205, arXiv:2109.01197.
- (7) S. Scheidl and V.M. Vinokur, Driven dynamics of periodic elastic media in disorder, Phys. Rev. E 57 (1998) 2574–2593, cond-mat/9708222.
- (8) H. Bucheli, O.S. Wagner, V.B. Geshkenbein, A.I. Larkin and G. Blatter, -dimensional elastic manifolds in random media: a renormalization-group analysis, Phys. Rev. B 57 (1998) 7642–52.
- (9) O. Narayan and D.S. Fisher, Critical behavior of sliding charge-density waves in 4-epsilon dimensions, Phys. Rev. B 46 (1992) 11520–49.
- (10) K.J. Wiese, M. Bercy, L. Melkonyan and T. Bizebard, Universal force correlations in an RNA-DNA unzipping experiment, Phys. Rev. Research 2 (2020) 043385, arXiv:1909.01319.
- (11) P. Le Doussal, K.J. Wiese, E. Raphael and Ramin Golestanian, Can non-linear elasticity explain contact-line roughness at depinning?, Phys. Rev. Lett. 96 (2006) 015702, cond-mat/0411652.
- (12) C. Bachas, P. Le Doussal and K.J. Wiese, Wetting and minimal surfaces, Phys. Rev. E 75 (2007) 031601, hep-th/0606247.
- (13) P. Le Doussal, K.J. Wiese, S. Moulinet and E. Rolley, Height fluctuations of a contact line: A direct measurement of the renormalized disorder correlator, EPL 87 (2009) 56001, arXiv:0904.4156.
- (14) D.S. Fisher, Collective transport in random media: From superconductors to earthquakes, Phys. Rep. 301 (1998) 113–150.
- (15) D. Fisher, K. Dahmen, S. Ramanathan and Y. Ben-Zion, Statistics of Earthquakes in Simple Models of Heterogeneous Faults, Phys. Rev. Lett. 78 (1997) 4885–4888.
- (16) Samuel Poincloux, Mokhtar Adda-Bedia and Frédéric Lechenault, Geometry and elasticity of a knitted fabric, Phys. Rev. X 8 (2018) 021075.
- (17) Samuel Poincloux, Mokhtar Adda-Bedia and Frédéric Lechenault, Crackling dynamics in the mechanical response of knitted fabrics, Phys. Rev. Lett. 121 (2018) 058002.
- (18) S.V. Buldyrev, A.-L. Barabasi, F. Caserta, S. Havlin, H.E. Stanley and T. Vicsek, Anomalous interface roughening in porous media: experiment and model, Phys. Rev. A 45 (1992) R8313–16.
- (19) L.-H. Tang and H. Leschhorn, Pinning by directed percolation, Phys. Rev. A 45 (1992) R8309–12.
- (20) L.A.N. Amaral, A.-L. Barabási, S.V. Buldyrev, S.T. Harrington, S. Havlin, R. Sadr-Lahijany and H.E. Stanley, Avalanches and the directed percolation depinning model: Experiments, simulations, and theory, Phys. Rev. E 51 (1995) 4655–4673.
- (21) A. Rosso and W. Krauth, Origin of the roughness exponent in elastic strings at the depinning threshold, Phys. Rev. Lett. 87 (2001) 187002, cond-mat/0104198.
- (22) G. Mukerjee, J. Bonachela, M. Munoz and K.J. Wiese, Depinning in the quenched Kardar-Parisi-Zhang class I: Mappings, simulations and algorithm, (2022), arXiv:2207.08341.
- (23) K.-W. Moon, D.-H. Kim, S.-C. Cheol Yoo, C.-G. Cho, S. Hwang, B. Kahng, B.-C. Min, K.-H. Shin and S.-B. Choe, Distinct universality classes of domain wall roughness in two-dimensional films, Phys. Rev. Lett. 110 (2013) 107203.
- (24) R. Díaz Pardo, N. Moisan, L. J. Albornoz, A. Lemaître, J. Curiale and V. Jeudy, Common universal behavior of magnetic domain walls driven by spin-polarized electrical current and magnetic field, Phys. Rev. B 100 (2019) 184420.
- (25) M.A.C. Huergo, N.E. Muzzio, M.A. Pasquale, P.H.Pedro González, A.E. Bolzán and A.J. Arvia, Dynamic scaling analysis of two-dimensional cell colony fronts in a gel medium: A biological system approaching a quenched Kardar-Parisi-Zhang universality, Phys. Rev. E 90 (2014) 022706.
- (26) S. Atis, A.K. Dubey, D. Salin, L. Talon, P. Le Doussal and K.J. Wiese, Experimental evidence for three universality classes for reaction fronts in disordered flows, Phys. Rev. Lett. 114 (2015) 234502, arXiv:1410.1097.
- (27) O. Narayan and D.S. Fisher, Dynamics of sliding charge-density waves in 4-epsilon dimensions, Phys. Rev. Lett. 68 (1992) 3615–18.
- (28) O. Narayan and D.S. Fisher, Threshold critical dynamics of driven interfaces in random media, Phys. Rev. B 48 (1993) 7030–42.
- (29) H. Leschhorn, T. Nattermann, S. Stepanow and L.-H. Tang, Driven interface depinning in a disordered medium, Annalen der Physik 509 (1997) 1–34, arXiv:cond-mat/9603114.
- (30) T. Nattermann, S. Stepanow, L.-H. Tang and H. Leschhorn, Dynamics of interface depinning in a disordered medium, J. Phys. II (France) 2 (1992) 1483–8.
- (31) P. Chauve, P. Le Doussal and K.J. Wiese, Renormalization of pinned elastic systems: How does it work beyond one loop?, Phys. Rev. Lett. 86 (2001) 1785–1788, cond-mat/0006056.
- (32) P. Le Doussal, K.J. Wiese and P. Chauve, 2-loop functional renormalization group analysis of the depinning transition, Phys. Rev. B 66 (2002) 174201, cond-mat/0205108.
- (33) P. Le Doussal, K.J. Wiese and P. Chauve, Functional renormalization group and the field theory of disordered elastic systems, Phys. Rev. E 69 (2004) 026112, cond-mat/0304614.
- (34) P. Le Doussal and K.J. Wiese, Functional renormalization group for anisotropic depinning and relation to branching processes, Phys. Rev. E 67 (2003) 016121, cond-mat/0208204.
- (35) G. Parisi, Statistical Field Theory, Frontiers in Physics, Addison-Wesely, 1988.
- (36) G. Parisi, Field-theoretic approach to second-order phase transitions in two- and three-dimensional systems, J. Stat. Phys. 23 (1980) 49–82.
- (37) G.A. Baker, B.G. Nickel, M.S. Green and D.I. Meiron, Ising-model critical indices in three dimensions from the Callan-Symanzik equation, Phys. Rev. Lett. 36 (1976) 1351–1354.
- (38) B.G. Nickel, D.I. Meiron and G.A. Baker, Compilation of 2-pt and 4-pt gprahs for continuous spin models, University of Guelph Report (1977).
- (39) G.A. Baker, B.G. Nickel and D.I. Meiron, Critical indices from perturbation analysis of the Callan-Symanzik equation, Phys. Rev. B 17 (1978) 1365–1374.
- (40) P. Le Doussal and K.J. Wiese, Driven particle in a random landscape: disorder correlator, avalanche distribution and extreme value statistics of records, Phys. Rev. E 79 (2009) 051105, arXiv:0808.3217.
- (41) P. Grassberger, D. Dhar and P. K. Mohanty, Oslo model, hyperuniformity, and the quenched Edwards-Wilkinson model, Phys. Rev. E 94 (2016) 042314, arXiv:1606.02553.
- (42) A. Shapira and K.J. Wiese, unpublished.
- (43) P.C. Martin, E.D. Siggia and H.A. Rose, Statistical dynamics of classical systems, Phys. Rev. A 8 (1973) 423–437.
- (44) P. Le Doussal and K.J. Wiese, How to measure Functional RG fixed-point functions for dynamics and at depinning, EPL 77 (2007) 66001, cond-mat/0610525.
- (45) A.A. Middleton, P. Le Doussal and K.J. Wiese, Measuring functional renormalization group fixed-point functions for pinned manifolds, Phys. Rev. Lett. 98 (2007) 155701, cond-mat/0606160.
- (46) A. Rosso, P. Le Doussal and K.J. Wiese, Numerical calculation of the functional renormalization group fixed-point functions at the depinning transition, Phys. Rev. B 75 (2007) 220201, cond-mat/0610821.
- (47) J.A. Bonachela, M. Alava and M.A. Muñoz, Cusps in systems with (many) absorbing states, Phys. Rev. E 79 (2009) 050106(R), arXiv:0810.4395.
- (48) A.-L. Barabasi, G. Grinstein and M.A. Muñoz, Directed surfaces in disordered media, Phys. Rev. Lett. 76 (1996) 1481–4.
- (49) H. Hinrichsen, Non-equilibrium critical phenomena and phase transitions into absorbing states, Adv. Phys. 49 (2000) 815–958, cond-mat/0001070.
- (50) N. Araújo, P. Grassberger, B. Kahng, K. J. Schrenk and R. M. Ziff, Recent advances and open challenges in percolation, Eur. Phys. J. Spec. Top. 223 (2014) 2307–2321, arXiv:1404.5325.
- (51) D. Dhar, Directed percolation and directed animals, (2017), arXiv:1703.07541.
- (52) A. Rosso, A.K. Hartmann and W. Krauth, Depinning of elastic manifolds, Phys. Rev. E 67 (2003) 021602, cond-mat/0207288.
- (53) E. Frey and U.C. Täuber, Two-loop renormalization group analysis of the Burgers-Kardar-Parisi-Zhang equation, Phys. Rev. E 50 (1994) 1024–1045.
- (54) M. Lässig, On the renormalization of the Kardar-Parisi-Zhang equation, Nucl. Phys. B 448 (1995) 559–574, cond-mat/9501094.
- (55) K.J. Wiese, Critical discussion of the 2-loop calculations for the KPZ-equation, Phys. Rev. E 56 (1997) 5013–5017, cond-mat/9706009.
- (56) K.J. Wiese, On the perturbation expansion of the KPZ-equation, J. Stat. Phys. 93 (1998) 143–154, cond-mat/9802068.
- (57) P. Le Doussal and K.J. Wiese, 2-loop functional renormalization for elastic manifolds pinned by disorder in dimensions, Phys. Rev. E 72 (2005) 035101 (R), cond-mat/0501315.
- (58) T. Thiery, P. Le Doussal and K.J. Wiese, Universal correlations between shocks in the ground state of elastic interfaces in disordered media, Phys. Rev. E 94 (2016) 012110, arXiv:1604.05556.
- (59) C.S. Dias, P.J. Yunker, A.G. Yodh, N.A.M. Araújo and M.M. Telo da Gama, Interaction anisotropy and the KPZ to KPZQ transition in particle deposition at the edges of drying drops, Soft Matter 14 (2018) 1903–1907.