arxiv: v1 [cond-mat.stat-mech] 27 Nov 2017

Similar documents
Phase Transitions in Nonequilibrium Steady States and Power Laws and Scaling Functions of Surface Growth Processes

Growth equations for the Wolf-Villain and Das Sarma Tamborenea models of molecular-beam epitaxy

FRACTAL CONCEPT S IN SURFACE GROWT H

Stochastic dynamics of surfaces and interfaces

ANOMALIES IN SIMULATIONS OF NEAREST NEIGHBOR BALLISTIC DEPOSITION

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

Scaling during shadowing growth of isolated nanocolumns

Numerical evaluation of the upper critical dimension of percolation in scale-free networks

arxiv: v1 [physics.soc-ph] 7 Jul 2015

Interface Roughening in a Hydrodynamic Lattice- Gas Model with Surfactant

Stochastic theory of non-equilibrium wetting

arxiv:nlin/ v1 [nlin.ps] 4 Sep 2004

Spherical models of interface growth

Weak universality of the KPZ equation

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Oct 2001

Anisotropic KPZ growth in dimensions: fluctuations and covariance structure

Relativistic treatment of Verlinde s emergent force in Tsallis statistics

Stochastic modeling of photoresist development in two and three dimensions

Chapter 3 Burgers Equation

Myllys, M.; Maunuksela, J.; Alava, Mikko; Ala-Nissilä, Tapio; Timonen, J. Scaling and noise in slow combustion of paper

arxiv: v1 [nlin.ps] 9 May 2015

Solving the Schrödinger equation for the Sherrington Kirkpatrick model in a transverse field

Optimal paths on the road network as directed polymers

Investigation of film surface roughness and porosity dependence on lattice size in a porous thin film deposition process

Cluster Distribution in Mean-Field Percolation: Scaling and. Universality arxiv:cond-mat/ v1 [cond-mat.stat-mech] 6 Jun 1997.

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 29 Nov 2006

Some Topics in Stochastic Partial Differential Equations

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 13 Mar 2001

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 13 Apr 1999

Microcanonical scaling in small systems arxiv:cond-mat/ v1 [cond-mat.stat-mech] 3 Jun 2004

Interfaces with short-range correlated disorder: what we learn from the Directed Polymer

Nonequilibrium processes: Driven lattice gases, interface dynamics, and quenched-disorder effects on density profiles and currents

The one-dimensional KPZ equation and its universality

arxiv: v1 [cond-mat.stat-mech] 12 Jul 2007

Exploring Geometry Dependence of Kardar-Parisi-Zhang Interfaces

arxiv:cond-mat/ v1 22 Sep 1998

arxiv:cond-mat/ v1 [cond-mat.dis-nn] 24 Mar 2005

arxiv: v2 [cond-mat.stat-mech] 17 Feb 2014

New Insights into the Dynamics of Swarming Bacteria: A Theoretical Study

arxiv:chao-dyn/ v1 5 Mar 1996

Turbulence models of gravitational clustering

Anomalous scaling at non-thermal fixed points of Gross-Pitaevskii and KPZ turbulence

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 25 Oct 2000

PERSISTENCE AND SURVIVAL IN EQUILIBRIUM STEP FLUCTUATIONS. Chandan Dasgupta

arxiv: v1 [cond-mat.mes-hall] 16 Nov 2007

Interface depinning in a disordered medium - numerical results

Statistical studies of turbulent flows: self-similarity, intermittency, and structure visualization

Intermittency, Fractals, and β-model

Universal phenomena in random systems

Langevin equations for the run-and-tumble of swimming bacteria

Using relaxational dynamics to reduce network congestion

Quasi-Stationary Simulation: the Subcritical Contact Process

arxiv: v2 [cond-mat.stat-mech] 6 Jun 2010

A Persistence Probability Analysis in Major Financial Indices

Patterns in the Kardar Parisi Zhang equation

Roughness scaling of plasma-etched silicon surfaces

Transient Orientation and Electrooptical Properties of Axially Symmetric Macromolecules in an Electric Field of Arbitrary Strength

Local persistense and blocking in the two dimensional Blume-Capel Model arxiv:cond-mat/ v1 [cond-mat.stat-mech] 4 Apr 2004.

Evolving network with different edges

Singular Stochastic PDEs and Theory of Regularity Structures

Can two chaotic systems make an order?

Universal scaling behavior of directed percolation and the pair contact process in an external field

arxiv: v1 [cond-mat.mtrl-sci] 7 Nov 2012

arxiv:cond-mat/ v1 2 Jul 2000

Effective Field Theory of Dissipative Fluids

RELAXATION AND TRANSIENTS IN A TIME-DEPENDENT LOGISTIC MAP

Renormalization Group: non perturbative aspects and applications in statistical and solid state physics.

Modeling the combined effect of surface roughness and shear rate on slip flow of simple fluids

A measure of data collapse for scaling

Analysis of Sunspot Number Fluctuations arxiv:nlin/ v1 [nlin.ao] 16 Mar 2004

arxiv: v1 [gr-qc] 9 May 2017

Testing Extended General Relativity with Galactic Sizes and Compact Gravitational Sources

Numerical analysis of the noisy Kuramoto-Sivashinsky equation in 2 1 dimensions

Bethe Ansatz Solutions and Excitation Gap of the Attractive Bose-Hubbard Model

Regularity Structures

Taming infinities. M. Hairer. 2nd Heidelberg Laureate Forum. University of Warwick

UNIQUENESS OF SELF-SIMILAR VERY SINGULAR SOLUTION FOR NON-NEWTONIAN POLYTROPIC FILTRATION EQUATIONS WITH GRADIENT ABSORPTION

arxiv: v2 [cond-mat.stat-mech] 25 Nov 2011

Nonconservative Abelian sandpile model with the Bak-Tang-Wiesenfeld toppling rule

arxiv: v1 [gr-qc] 15 Nov 2017

Nano-patterning of surfaces by ion sputtering: Numerical study of the Kuramoto-Sivashinsky Equation

Beyond the Gaussian universality class

This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail.

Desynchronization waves in small-world networks

arxiv: v2 [gr-qc] 22 Jan 2014

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 7 Aug 1997

arxiv: v1 [quant-ph] 21 Dec 2009

Generalized Manna Sandpile Model with Height Restrictions

THE problem of phase noise and its influence on oscillators

arxiv:cond-mat/ v1 8 Jan 2004

VIII.B Equilibrium Dynamics of a Field

Dynamics of driven interfaces near isotropic percolation transition

Growth oscillations. LElTER TO THE EDITOR. Zheming Cheng and Robert Savit

arxiv: v1 [cond-mat.stat-mech] 14 Apr 2009

Open boundary conditions in stochastic transport processes with pair-factorized steady states

arxiv: v3 [cond-mat.stat-mech] 1 Dec 2016

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 8 Sep 1999

A determinantal formula for the GOE Tracy-Widom distribution

Amorphous thin-film growth: Theory compared with experiment

SUPPLEMENTARY INFORMATION

Transcription:

Growing interfaces: A brief review on the tilt method M. F. Torres and R. C. Buceta arxiv:1711.09652v1 [cond-mat.stat-mech] 27 Nov 2017 Instituto de Investigaciones Físicas de Mar del Plata (UNMdP and CONICET) and Departamento de Física, FCEyN, Universidad Nacional de Mar del Plata Funes 3350, B7602AYL Mar del Plata, Argentina (Dated: November 28, 2017) Abstract The tilt method applied to models of growing interfaces is a useful tool to characterize the nonlinearities of their associated equation. Growing interfaces with average slope m, in models and equations belonging to Kardar-Parisi-Zhang (KPZ) universality class, have average saturation velocity V sat = Υ + 1 2 Λm2 when m 1. This property is sufficient to ensure that there is a nonlinearity type square height-gradient. Usually, the constant Λ is considered equal to the nonlinear coefficient λ of the KPZ equation. In this paper, we show that the mean square heightgradient h 2 = a+b, whereb = 1 forthecontinuous KPZequation and b 1otherwise, e.g. ballistic deposition(bd) and restricted-solid-on-solid (RSOS) models. In order to find the nonlinear coefficient λ associated to each system, we establish the relationship Λ = bλ and we test it through the discrete integration of the KPZ equation. We conclude that height-gradient fluctuations as function of are constant for continuous KPZ equation and increasing or decreasing in other systems, such as BD or RSOS models, respectively. mtorres@ifimar-conicet.gob.ar rbuceta@mdp.edu.ar 1

I. INTRODUCTION The tilt method was initially proposed by Krug [1, 2] to prove that growing interface models belonging to a universality class can be characterized, in addition to the exponents and laws of scaling, by the nonlinearities present in the system. Usually, when an interface grows with nonzero average slope, helical boundary conditions are applied [3], i.e. h m (L+ 1) = ml+h m (1),whereh m isthetiltedinterfaceheight, Listhelateralsize, andm = h m is the average slope of the tilted interface. Periodic boundary condition corresponds to and interface without tilting, i.e. m = 0. It is well known, based on observation [3], that the models belonging to the KPZ universality class show a dependency between the average velocity in the saturation and the slope of the interface like: V (m) sat = V (0) sat + Λ 2 m2, (1) for m 1 and where the real constant Λ 0 and the label (m) refers to an interface with average slope m. Otherwise, if Λ = 0 the models are included in the EW universality class. The linear behaviour of V (m) sat as function of is strictly valid for m 1; otherwise, others behaviours arise. If the dependency of the average saturation velocity with the slope is different than, it indicates that the studied model does not belong to the KPZ or EW universality classes. The quadratic constant Λ has been associated with the nonlinear constant λ of the KPZ equation h t = F +ν 2 h+ λ 2 h 2 +η(x,t), (2) where h(x, t) is the interface height of the d-dimensional substratum at the position x and time t. The real constants F, ν and λ take into account the growth force, the surface relaxation intensity and the lateral growth, respectively. The noise η(x, t) is Gaussian with zero mean and covariance η(x,t)η(x,t ) = 2Dδ(x x )δ(t t ), D being the noise intensity. In this work, we show the tilt method as a powerful tool to find the nonlinear coefficient λ of a given model. Although there are other possible methods to obtain it [4 6], these methods are usually focused on the relation Γ = λ A 1/α, where A is the power-law coefficient of the second-order height-difference correlation as a function of the distance between columns, while Γ comes from the evolution equation of the interface-height average in the non-stationary regime and α is the global roughness exponent. These methods usually reproduce a value close to Λ. 2

From the KPZ equation (2), when the interface is tilted, the average velocity is V (m) KPZ = hm = F + λ hm t 2 2, (3) which is a function of the average slope m. Notice that the noise average is zero and, if the tilted system is sufficiently large or helical boundary conditions are taken, the Laplacian average is negligible or zero, respectively. In the saturation equation (3) is independent of time; thus, it is only a function of m. In this paper, we show that both in simulations and integrations of models belonging to the KPZ universality class it is verified that the mean square height-gradient (MSG) is h m 2 = h 0 2 +b, (4) where b is a positive real constant. As it is well established, b has a clear theorical meaning for the continuous KPZ equation (2), i.e. b = 1. We show here that b 1 in two discrete models belonging to KPZ universality class, the restricted-solid-on-solid (RSOS) and the ballistic deposition (BD) models. Also, we show that b 1 (but not equal) for KPZ numerical integrations by several methods. Together, we researched some of the properties of the height-gradient fluctuations of interfaces with average slope m. We show that the mean square fluctuation(msf) of the height-gradient is a function of m that increases, decreases or is constant for the BD model, the RSOS model, or the continuous KPZ equation, respectively. In Section II, we show analytically equation (4) by calculating the fluctuations of the heightgradient. We also show the relationship Λ = bλ between the nonlinear coefficient λ of the KPZ equation and the parameters Λ and b of the tilt method. In Section III, we show the validity of equation (4) and determine the parameters of the tilt method, obtained from simulations of the BD and RSOS models and integration of modified KPZ equations, all included in the KPZ universality class. Additionally, for each model and equation, we show how the interface slope affects the height-gradient fluctuations. Finally, in the Conclusions, we make a brief summary of the results achieved, discussing the highlights of this work. II. METHOD BASICS The height gradient h m of a tilted growing interface is subjected to fluctuations that depend on the height slope and the non-tilted height gradient, which in mean value must 3

be zero, such that h m = m. The most obvious proposal is h m = m+f(m, h 0 ), (5) where F is a generalized function that indicates the fluctuations of the tilted height gradient andverifies F(m, h 0 ) = 0 andf(0, h 0 ) = h 0. Asimple calculation allowsustoobtain hm 2 = + [F(m, h 0 )] 2, (6) where the MSF F 2 is an even function of m. Assuming m 1 and developing F 2 in Taylor series around m = 0, we obtain [F(m, h0 )] 2 = h 0 2 + G( h 0 ) +O(m 4 ). (7) where G isaconstant dependent onthesystem. In order tojustify equation (4), we replace equation (7) in equation (6) getting G( h0 ) = 1 b. (8) The sign of G is decisive when determining if the model fluctuation decreases or increases with the average slope of the interface. Between both behaviours, G = 0 for interfaces that evolve according to the continuous KPZ equation and the MSF is constant regardless of the tilt, i.e. F 2 = h 0 2. By replacing of equation (4) in equation (1) we obtain V (m) sat = V (0) sat Λ 2b h 0 2 + Λ 2b h m 2. (9) By equating equations (3) and (9) in the saturation we can establish the following relation between the coefficients λ = Λ b, (10) F = V (0) sat Λ 2b h 0 2. (11) Thus, equation (10) allows to obtain the nonlinear coefficient of each model in terms of the two tilt coefficients of equations (1) and(4). Only for continuous KPZ equation λ = Λ is obtained. 4

III. RESULTS FOR SEVERAL MODELS AND EQUATIONS RSOS model. The interface evolution of the (1+1)-dimensional RSOS model is established by the following rule: choosing a random column, if the height of its two firstneighbouring columns are greater or equal to the chosen one, it grows one unit. Symbolically, if h(i+1) h(i) 0 and h(i 1) h(i) 0, where i is the column chosen at random, then h(i) h(i)+1. Left plot of Figure 1 shows the saturation velocity as a function of, for several system sizes, verifying equation (1). The measured values of slope Λ, given in Table I, are close to 0.75 obtained by several authors with good precision [4, 5]. Right plot of Figure 1 shows h m 2 as a function of, which verifies equation (4), with measured slope b given in Table I. Aditionally, Table I shows the nonlinear coefficients λ of the model v sat 0.42 0.4 0.38 0.36 L =512 L =1024 L =2048 (a) L =512 L =1024 L =2048 0.46 0.44 0.42 0.4 0.38 0.36 h m 2 FIG. 1. (color online) RSOS model. Left plot: Average saturation velocity V sat as a function of for several sizes of L. The slope of the straight lines is Λ/2. Right plot: MSG h m 2 as a function of for several sizes L. The slope of the straight lines is b. The measured values Λ and b with their respective fit errors are shown in Table I. L Λ (δλ) b (δb) λ (δλ) 512 0.747(3) 0.583(2) 1.281(10) 1024 0.746(3) 0.582(1) 1.282(6) 2048 0.743(3) 0.580(1) 1.281(6) TABLE I. Coefficients corresponding to RSOS model for several sizes L of simulation. The table shows the measured values of Λ and b, and the calculated values λ, with their respective errors in parentheses. In all tables (δν) = ±δν 10 3 is (are) the error(s) of a quantity ν. calculated by equation (10) for several system sizes. Notice that the measured coefficients do not change when increasing the size of the system, so the results are valid at the thermodynamic limit. We can see that b 0.58, which indicates that the MSF decreases as the 5

slope m increases. This phenomena is understood from the fact that, when increasing the interface tilt of the RSOS model, the application of their rules decreases to such a point that growth is not possible. In addition, the MSF verifies F 2 h 0 2 0.36. Ballistic deposition model. The evolution of the (1+1)-dimensional BD model is given by the following rule: the height of the chosen column h(i) grows to max[h(i 1), h(i)+ 1,h(i+1)]. Left plot of Figure 2 shows the saturation velocity V (m) sat as a function of for several system sizes. As before, equation (1) is verified with the measured coefficient Λ given in Table II. The BD model is known for having finite-size dependencies in its exponents; however, the coefficient Λ shows slight changes when the system size increases. Our measured value Λ = 1.25 has a small departure from the value 1.30 obtained by others authors [4, 5], which is explained from the slight variations of Λ with the range of slopes taken. Right plot of Figure 2 shows the MSG h m 2 as a function of with measured slopes b given in Table II. Note that, in this model, the measured coefficients do not change when increasing the size of the system. There are important differences between this model and the RSOS model that become more evident when the interface is tilted. On the one hand, taking into account that b 6.1 the MSF increases with m. On the other hand, in the BD model the appearance of strong gradients is accompanied by very intense fluctuations with MSF F 2 h 0 2 4.0. v sat 2.24 2.22 2.2 2.18 2.16 2.14 L =512 L =1024 L =2048 L =512 L =1024 L =2048 5 4.8 4.6 4.4 4.2 0 0.05 0.1 4 0.15 h m 2 FIG. 2. (color online) BD model. Both plots: Idem to Figure 1. The measured values Λ and b with their respective fit errors are shown in Table II. Numerical integration of the modified KPZ equations. The numerical integration method of the KPZ equation (2) introduced by Dasgupta et al. allows to avoid divergences by smoothing the values of the nonlinearity h 2 in each integration step. The method proposes replacing this nonlinearity with a function f( h) = 1 c ( 1 e c h 2), (12) 6

L Λ (δλ) b (δb) λ (δλ) 512 1.254(7) 6.061(70) 0.207(4) 1024 1.255(8) 6.140(63) 0.204(4) 2048 1.253(9) 6.165(63) 0.203(4) TABLE II. Coefficients corresponding to BD model for several sizes L of simulation. The table shows the measured values of Λ and b, and the calculated λ, with their respective errors in parentheses. where c is a real positive parameter. Then, during the numerical integration the system evolves according to the smoothing function f instead of to the nonlinear term. Although the KPZ equation is modified by infinite terms, with an appropriate choice of parameter c the scaling properties are not modified. The introduced change should hold for other properties of the KPZ equation. Equation (4), consistently with the change made, must be replaced by f( h m ) = f( h 0 ) +b, (13) keeping b = 1, or equivalently λ = Λ, for the KPZ equation. However, this is not verified with precision. To show this, we integrate the modified equation with the values of the parameters mentioned in the caption of Figure 3. These elections correspond to a coupling c Λ (δλ) b (δb) λ (δλ) 0.15 7.846(9) 1.013(1) 7.745(17) 0.25 7.695(7) 0.993(1) 7.749(15) 0.50 7.326(42) 0.946(6) 7.744(93) 1.00 6.700(93) 0.865(12) 7.746(215) TABLE III. Coefficients corresponding to the integration of KPZ equation by the Dasgupta et al. method for several smoothing parameters c of equation (12). The table shows the measured values of Λ and b, and the calculated λ, with their respective errors in parentheses. constant g = 15, value for which the roughness shows a longest power law behaviour [9]. Figure 3 shows that both the saturation velocity and the average of the smoothing function f( h m ) are linear functions of, with measured coefficients Λ and b, and its respective errors, given in Table III. We observe that for several values of c, the data value λ = 0.7460, 7

v sat 0.8 0.6 0.4 0.2 c =0.15 c =0.25 c =0.5 c =1 c =0.15 c =0.25 c =0.5 c =1 0.2 0.15 0.1 0.05 f ( h m 2 ) FIG. 3. (color online) Both plots: We use the Dasgupta et al. method to control the divergences of the KPZ integration [7, 8], taking size system L = 512 and parameters λ = 7.7460, ν = 0.5, F = 0, and σ = 0.25, integration mesh x = 1 and integration step t = 0.05. These parameters correspond to a coupling constant g = 15. Left plot: Average saturation velocity V sat as a function of for several parameters c. The slope of the straight lines are Λ/2. Right plot: Average of the smoothing function f( h m ) given by equation 12) as a function of for several parameters c. The slope of the straight lines is b. The measured values Λ and b with their respective fit errors are shown in Table III. in all cases, is not located within the error bars of the measured Λ. Likewise, b = 1 is also not located within the error bars of the measured b. However, we observe that the data λ = 0.7460 is included within the error bars of the calculated λ that we show in Table III. The range of the parameter values c chosen allows to obtain roughnesses that have longer power laws with growth exponents β 0.33. To confirm that the results obtained are not dependent on the integration method, we use Torres and Buceta s method [9] of imposing restrictions on the excessive growth introduced by nonlinearities. The restriction method states that the nonlinear term h 2 can be replaced by the generalized function f( h) = min ( h 2,ε ), (14) where ε is the restriction parameter. Similarly to Dasguta et al. method, the evolution of the system does not follow a square height-gradient but the min function. In Figure 4 we calculate the saturation velocity and the average of restriction function given by equation (14) as functions of the slope m for several values of the restriction parameter ε, for which the main scaling properties of the KPZ equation are recovered. The fit coefficients Λ and b are presented in Table IV; as before, the theoretical values λ = 0.7460 and b = 1 fall outside the error bars of the measurements made. By both integration methods used above 8

v sat 0.8 0.6 0.4 0.2 ε =0.5 ε =0.75 ε =1 ε =1.25 ε =0.5 ε =0.75 ε =1 ε =1.25 0.2 0.15 0.1 0.05 min( h m 2,ε) FIG. 4. (color online) Both plots: We use the Torres and Buceta s method to control the divergences of the KPZ integration [9], taking the values given in Figure 3. Left plot: Average saturation velocity V sat as a function of for several restriction parameters ε. The slope of the straight lines are Λ/2. Right plot: Average of the restriction function min ( h m 2,ε ) as a function of for several restriction parameters ε. The slope of the straight lines is b. The measured values Λ and b with their respective fit errors are shown in Table IV. ε Λ (δλ) b (δb) λ (δλ) 0.50 7.685(74) 0.992(9) 7.747(145) 0.75 8.004(11) 1.033(1) 7.748(18) 1.00 8.075(29) 1.042(4) 7.749(57) 1.25 8.096(36) 1.045(5) 7.747(71) TABLE IV. Coefficients corresponding to the integration of KPZ equation by the Torres and Buceta s method for several restriction parameters ε. The table shows the measured values of Λ and b, and the calculated λ, with their respective errors in parentheses. we find that the tilt coefficient b is very close to 1, although it is not equal. This is explained by the fact that what is integrated is not exactly the KPZ equation. In both cases, the methods are perturbations implemented in order to control or limit the divergences proper to the numerical integration of the KPZ equation. But, again in both methods, applying the relation (10) we recover the original value of λ with great precision. IV. CONCLUSIONS From the theoretical studies of the KPZ equation it is widely known that the tilt method explains the existence of h 2 nonlinearities. In addition, it is also known that the average saturation velocity as a function of average slope m, given by equation (1), has quadratic 9

coefficient Λ = λ, where λ is the KPZ nonlinear coefficient. However, we have shown in this paper that Λ λ for several models and equations (excluding the KPZ equation) belonging to the KPZ universality class. To show this, we study the MSG h 2 when the interface grows with an average slope m. By introducing the concept of tilted height-gradient fluctuations we show equation (4), which is verified by simulation of several models (RSOS and BD) and numerical integration of KPZ equation by two alternative methods. We show that Λ = bλ, where b is the quadratic coefficient of the MSG as a function of m, which depends on each model, being b = 1 for the continuous KPZ equation and b 1 otherwise. Additionally, we show that the MSF of the height gradient as function of average slop m follows equation (7). Onthe onehand, themsf of thecontinuous KPZ equation isconstant. On the other hand, the MSF of the RSOS or BD model decreases or increases, respectively, as m increases. The numerical integration methods of the KPZ equation establish that b 1 (but close) which indicates that the modifications or restrictions to the KPZ introduced to avoid divergences makes λ Λ. We found that λ Λ/b with great precision, confirming the equation (10). Unexpectedly, the integration methods introduce fluctuations dependent on the tilt interface. This shows that the integration methods explicitly break the symmetry contained in the continuous KPZ equation, symmetry by which its interface has constant MSF. The reason for the discrepancy between the results presented in this paper and those of other methods [10, 11], in reference to the KPZ nonlinear coefficient λ, is a subject of open interest. As a final remark, the conjecture hereby proved suggests a revision of the method applied to models and equations belonging to other universality classes, such as that of Lai-Das Sarma [12, 13]. ACKNOWLEDGMENTS This work was partially supported by Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), Argentina, PIP 2014/16 No. 112-201301-00629. R.C.B. thanks C. Rabini for her suggestions on the final manuscript. [1] J. Krug, J. Phys. A. Math. Gen. 22, L769 (1989). 10

[2] J. Krug and H. Sphon, Phys. Rev. Lett. 64, 543 (1990). [3] L.-A. Barabási and H. E. Stanley, Fractal Concepts in Surface Growth (Cambridge University Press, Cambridge, 1995). [4] J. Krug, P. Meakin, and T. Halpin-Healy, Phys. Rev. A. 45, 638 (1992). [5] T. J. Oliveira, S. C. Ferreira, and S. G. Alves, Phys. Rev. E 85, 010601(R) (2012). [6] S. G. Alves, T. J. Oliveira, and S. C. Ferreira, Phys. Rev. E 90, 020103(R) (2014). [7] C. Dasgupta, S. Das Sarma, and J. M. Kim, Phys. Rev. E 54, R4552 (1996). [8] C. Dasgupta, J. M. Kim, M. Dutta, and S. Das Sarma, Phys. Rev. E 55, 2235 (1997). [9] M. F. Torres and R. C. Buceta, arxiv:1707.03011 (2017). [10] S. G. Alves and S. C. Ferreira, Phys. Rev. E 93, 052131 (2016). [11] R. C. Buceta and D. Hansmann, J. Phys. A: Math. Gen. 45, 435202 (2012). [12] Z. W. Lai and S. Das Sarma, Phys. Rev. Lett. 66, 2348 (1991). [13] S.Das SarmaandP.Tamborenea,Phys. Rev. Lett. 66, 325 (1991); P.TamboreneaandS.Das Sarma, Phys. Rev. E 48, 2575 (1993). 11