1,170 research outputs found

    Comment on " A simple model for DNA denaturation"

    Get PDF
    The replacment of mutual avoidance of polymers by a long-range interaction of the type proposed by Garel etal (Europhys. Lett. 55, 132 (2001), cond-mat/0101058) is inconsistent with the prevalent renormalization group arguments.Comment: 2 pages, Comment on Garel etal. Europhys. Lett 55, 132(2001) cond-mat/0101058. Appeared in Europhys Let

    Delocalization transition of the selective interface model: distribution of pseudo-critical temperatures

    Full text link
    According to recent progress in the finite size scaling theory of critical disordered systems, the nature of the phase transition is reflected in the distribution of pseudo-critical temperatures Tc(i,L)T_c(i,L) over the ensemble of samples (i)(i) of size LL. In this paper, we apply this analysis to the delocalization transition of an heteropolymeric chain at a selective fluid-fluid interface. The width ΔTc(L)\Delta T_c(L) and the shift [Tc()Tcav(L)][T_c(\infty)-T_c^{av}(L)] are found to decay with the same exponent L1/νRL^{-1/\nu_{R}}, where 1/νR0.261/\nu_{R} \sim 0.26. The distribution of pseudo-critical temperatures Tc(i,L)T_c(i,L) is clearly asymmetric, and is well fitted by a generalized Gumbel distribution of parameter m3m \sim 3. We also consider the free energy distribution, which can also be fitted by a generalized Gumbel distribution with a temperature dependent parameter, of order m0.7m \sim 0.7 in the critical region. Finally, the disorder averaged number of contacts with the interface scales at TcT_c like LρL^{\rho} with ρ0.261/νR\rho \sim 0.26 \sim 1/\nu_R .Comment: 9 pages,6 figure

    Statistics of low energy excitations for the directed polymer in a 1+d1+d random medium (d=1,2,3d=1,2,3)

    Full text link
    We consider a directed polymer of length LL in a random medium of space dimension d=1,2,3d=1,2,3. The statistics of low energy excitations as a function of their size ll is numerically evaluated. These excitations can be divided into bulk and boundary excitations, with respective densities ρLbulk(E=0,l)\rho^{bulk}_L(E=0,l) and ρLboundary(E=0,l)\rho^{boundary}_L(E=0,l). We find that both densities follow the scaling behavior ρLbulk,boundary(E=0,l)=L1θdRbulk,boundary(x=l/L)\rho^{bulk,boundary}_L(E=0,l) = L^{-1-\theta_d} R^{bulk,boundary}(x=l/L), where θd\theta_d is the exponent governing the energy fluctuations at zero temperature (with the well-known exact value θ1=1/3\theta_1=1/3 in one dimension). In the limit x=l/L0x=l/L \to 0, both scaling functions Rbulk(x)R^{bulk}(x) and Rboundary(x)R^{boundary}(x) behave as Rbulk,boundary(x)x1θdR^{bulk,boundary}(x) \sim x^{-1-\theta_d}, leading to the droplet power law ρLbulk,boundary(E=0,l)l1θd\rho^{bulk,boundary}_L(E=0,l)\sim l^{-1-\theta_d} in the regime 1lL1 \ll l \ll L. Beyond their common singularity near x0x \to 0, the two scaling functions Rbulk,boundary(x)R^{bulk,boundary}(x) are very different : whereas Rbulk(x)R^{bulk}(x) decays monotonically for 0<x<10<x<1, the function Rboundary(x)R^{boundary}(x) first decays for 0<x<xmin0<x<x_{min}, then grows for xmin<x<1x_{min}<x<1, and finally presents a power law singularity Rboundary(x)(1x)σdR^{boundary}(x)\sim (1-x)^{-\sigma_d} near x1x \to 1. The density of excitations of length l=Ll=L accordingly decays as ρLboundary(E=0,l=L)Lλd\rho^{boundary}_L(E=0,l=L)\sim L^{- \lambda_d} where λd=1+θdσd\lambda_d=1+\theta_d-\sigma_d. We obtain λ10.67\lambda_1 \simeq 0.67, λ20.53\lambda_2 \simeq 0.53 and λ30.39\lambda_3 \simeq 0.39, suggesting the possible relation λd=2θd\lambda_d= 2 \theta_d.Comment: 15 pages, 25 figure

    Directed polymer in a random medium of dimension 1+1 and 1+3: weights statistics in the low-temperature phase

    Full text link
    We consider the low-temperature T<TcT<T_c disorder-dominated phase of the directed polymer in a random potentiel in dimension 1+1 (where Tc=T_c=\infty) and 1+3 (where Tc<T_c<\infty). To characterize the localization properties of the polymer of length LL, we analyse the statistics of the weights wL(r)w_L(\vec r) of the last monomer as follows. We numerically compute the probability distributions P1(w)P_1(w) of the maximal weight wLmax=maxr[wL(r)]w_L^{max}= max_{\vec r} [w_L(\vec r)], the probability distribution Π(Y2)\Pi(Y_2) of the parameter Y2(L)=rwL2(r)Y_2(L)= \sum_{\vec r} w_L^2(\vec r) as well as the average values of the higher order moments Yk(L)=rwLk(r)Y_k(L)= \sum_{\vec r} w_L^k(\vec r). We find that there exists a temperature Tgap<TcT_{gap}<T_c such that (i) for T<TgapT<T_{gap}, the distributions P1(w)P_1(w) and Π(Y2)\Pi(Y_2) present the characteristic Derrida-Flyvbjerg singularities at w=1/nw=1/n and Y2=1/nY_2=1/n for n=1,2..n=1,2... In particular, there exists a temperature-dependent exponent μ(T)\mu(T) that governs the main singularities P1(w)(1w)μ(T)1P_1(w) \sim (1-w)^{\mu(T)-1} and Π(Y2)(1Y2)μ(T)1\Pi(Y_2) \sim (1-Y_2)^{\mu(T)-1} as well as the power-law decay of the moments Yk(i)ˉ1/kμ(T) \bar{Y_k(i)} \sim 1/k^{\mu(T)}. The exponent μ(T)\mu(T) grows from the value μ(T=0)=0\mu(T=0)=0 up to μ(Tgap)2\mu(T_{gap}) \sim 2. (ii) for Tgap<T<TcT_{gap}<T<T_c, the distribution P1(w)P_1(w) vanishes at some value w0(T)<1w_0(T)<1, and accordingly the moments Yk(i)ˉ\bar{Y_k(i)} decay exponentially as (w0(T))k(w_0(T))^k in kk. The histograms of spatial correlations also display Derrida-Flyvbjerg singularities for T<TgapT<T_{gap}. Both below and above TgapT_{gap}, the study of typical and averaged correlations is in full agreement with the droplet scaling theory.Comment: 13 pages, 29 figure

    Glassy phases in Random Heteropolymers with correlated sequences

    Full text link
    We develop a new analytic approach for the study of lattice heteropolymers, and apply it to copolymers with correlated Markovian sequences. According to our analysis, heteropolymers present three different dense phases depending upon the temperature, the nature of the monomer interactions, and the sequence correlations: (i) a liquid phase, (ii) a ``soft glass'' phase, and (iii) a ``frozen glass'' phase. The presence of the new intermediate ``soft glass'' phase is predicted for instance in the case of polyampholytes with sequences that favor the alternation of monomers. Our approach is based on the cavity method, a refined Bethe Peierls approximation adapted to frustrated systems. It amounts to a mean field treatment in which the nearest neighbor correlations, which are crucial in the dense phases of heteropolymers, are handled exactly. This approach is powerful and versatile, it can be improved systematically and generalized to other polymeric systems

    Probing the tails of the ground state energy distribution for the directed polymer in a random medium of dimension d=1,2,3d=1,2,3 via a Monte-Carlo procedure in the disorder

    Full text link
    In order to probe with high precision the tails of the ground-state energy distribution of disordered spin systems, K\"orner, Katzgraber and Hartmann \cite{Ko_Ka_Ha} have recently proposed an importance-sampling Monte-Carlo Markov chain in the disorder. In this paper, we combine their Monte-Carlo procedure in the disorder with exact transfer matrix calculations in each sample to measure the negative tail of ground state energy distribution Pd(E0)P_d(E_0) for the directed polymer in a random medium of dimension d=1,2,3d=1,2,3. In d=1d=1, we check the validity of the algorithm by a direct comparison with the exact result, namely the Tracy-Widom distribution. In dimensions d=2d=2 and d=3d=3, we measure the negative tail up to ten standard deviations, which correspond to probabilities of order Pd(E0)1022P_d(E_0) \sim 10^{-22}. Our results are in agreement with Zhang's argument, stating that the negative tail exponent η(d)\eta(d) of the asymptotic behavior lnPd(E0)E0η(d)\ln P_d (E_0) \sim - | E_0 |^{\eta(d)} as E0E_0 \to -\infty is directly related to the fluctuation exponent θ(d)\theta(d) (which governs the fluctuations ΔE0(L)Lθ(d)\Delta E_0(L) \sim L^{\theta(d)} of the ground state energy E0E_0 for polymers of length LL) via the simple formula η(d)=1/(1θ(d))\eta(d)=1/(1-\theta(d)). Along the paper, we comment on the similarities and differences with spin-glasses.Comment: 13 pages, 16 figure

    Protein Folding and Spin Glass

    Full text link
    We explicitly show the connection between the protein folding problem and spin glass transition. This is then used to identify appropriate quantities that are required to describe the transition. A possible way of observing the spin glass transition is proposed.Comment: Revtex3+epsf, 8 pages and one postscript figure tarred, compressed and uuencoded--appended at the end of the file. Minor TeX change

    On Heteropolymer Shape Dynamics

    Full text link
    We investigate the time evolution of the heteropolymer model introduced by Iori, Marinari and Parisi to describe some of the features of protein folding mechanisms. We study how the (folded) shape of the chain evolves in time. We find that for short times the mean square distance (squared) between chain configurations evolves according to a power law, DtνD \sim t ^\nu. We discuss the influence of the quenched disorder (represented by the randomness of the coupling constants in the Lennard-Jones potential) on value of the critical exponent. We find that ν\nu decreases from 23\frac{2}{3} to 12\frac{1}{2} when the strength of the quenched disorder increases.Comment: 12 pages, very simple LaTeX file, 6 figures not included, sorry. SCCS 33

    Adsorption of a random heteropolymer at a potential well revisited: location of transition point and design of sequences

    Full text link
    The adsorption of an ideal heteropolymer loop at a potential point well is investigated within the frameworks of a standard random matrix theory. On the basis of semi-analytical/semi-numerical approach the histogram of transition points for the ensemble of quenched heteropolymer structures with bimodal symmetric distribution of types of chain's links is constructed. It is shown that the sequences having the transition points in the tail of the histogram display the correlations between nearest-neighbor monomers.Comment: 11 pages (revtex), 3 figure

    On the multifractal statistics of the local order parameter at random critical points : application to wetting transitions with disorder

    Full text link
    Disordered systems present multifractal properties at criticality. In particular, as discovered by Ludwig (A.W.W. Ludwig, Nucl. Phys. B 330, 639 (1990)) on the case of diluted two-dimensional Potts model, the moments ρq(r)ˉ\bar{\rho^q(r)} of the local order parameter ρ(r)\rho(r) scale with a set x(q)x(q) of non-trivial exponents x(q)qx(1)x(q) \neq q x(1). In this paper, we revisit these ideas to incorporate more recent findings: (i) whenever a multifractal measure w(r)w(r) normalized over space rw(r)=1 \sum_r w(r)=1 occurs in a random system, it is crucial to distinguish between the typical values and the disorder averaged values of the generalized moments Yq=rwq(r)Y_q =\sum_r w^q(r), since they may scale with different generalized dimensions D(q)D(q) and D~(q)\tilde D(q) (ii) as discovered by Wiseman and Domany (S. Wiseman and E. Domany, Phys Rev E {\bf 52}, 3469 (1995)), the presence of an infinite correlation length induces a lack of self-averaging at critical points for thermodynamic observables, in particular for the order parameter. After this general discussion valid for any random critical point, we apply these ideas to random polymer models that can be studied numerically for large sizes and good statistics over the samples. We study the bidimensional wetting or the Poland-Scheraga DNA model with loop exponent c=1.5c=1.5 (marginal disorder) and c=1.75c=1.75 (relevant disorder). Finally, we argue that the presence of finite Griffiths ordered clusters at criticality determines the asymptotic value x(q)=dx(q \to \infty) =d and the minimal value αmin=D(q)=dx(1) \alpha_{min}=D(q \to \infty)=d-x(1) of the typical multifractal spectrum f(α)f(\alpha).Comment: 17 pages, 20 figure
    corecore