Self-Assembly and Convergence Rates of Heterogeneous Reversible Growth Processes (Extended Abstract)

Similar documents
Sampling Biased Lattice Configurations using Exponential Metrics

On Times to Compute Shapes in 2D Tile Self-Assembly

Convergence Rates of Markov Chains for Some Self-Assembly and Non-Saturated Ising Models

Coupling. 2/3/2010 and 2/5/2010

Mixing Times of Markov Chains for Self-Organizing Lists and Biased Permutations

Exponential Metrics. Lecture by Sarah Cannon and Emma Cohen. November 18th, 2014

Integer and Vector Multiplication by Using DNA

Decomposition Methods and Sampling Circuits in the Cartesian Lattice

Phase Transitions in Random Dyadic Tilings and Rectangular Dissections

Coupling AMS Short Course

Mixing Times of Markov Chains for Self-Organizing Lists and Biased Permutations

Stochastic Analysis of Reversible Self-Assembly

Active Tile Self Assembly:

Sampling Adsorbing Staircase Walks Using a New Markov Chain Decomposition Method

Molecular Self-Assembly: Models and Algorithms

Random Walks A&T and F&S 3.1.2

Randomized Algorithms

Definition A finite Markov chain is a memoryless homogeneous discrete stochastic process with a finite number of states.

The coupling method - Simons Counting Complexity Bootcamp, 2016

Sampling Good Motifs with Markov Chains

Markov chain Monte Carlo algorithms

The Markov Chain Monte Carlo Method

A Note on the Glauber Dynamics for Sampling Independent Sets

Approximately Sampling Elements with Fixed Rank in Graded Posets

Approximately Sampling Elements with Fixed Rank in Graded Posets

Markov Chain Monte Carlo The Metropolis-Hastings Algorithm

Lecture 28: April 26

Session 3A: Markov chain Monte Carlo (MCMC)

Biased random walk on percolation clusters. Noam Berger, Nina Gantert and Yuval Peres

5. Simulated Annealing 5.1 Basic Concepts. Fall 2010 Instructor: Dr. Masoud Yaghini

Propp-Wilson Algorithm (and sampling the Ising model)

Can extra updates delay mixing?

Monte Caro simulations

Monte Carlo. Lecture 15 4/9/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

Oberwolfach workshop on Combinatorics and Probability

Parallelism and time in hierarchical self-assembly

Lecture 3: September 10

CS 781 Lecture 9 March 10, 2011 Topics: Local Search and Optimization Metropolis Algorithm Greedy Optimization Hopfield Networks Max Cut Problem Nash

Flip dynamics on canonical cut and project tilings

Unpredictable Walks on the Integers

Physics 562: Statistical Mechanics Spring 2002, James P. Sethna Prelim, due Wednesday, March 13 Latest revision: March 22, 2002, 10:9

6 Markov Chain Monte Carlo (MCMC)

Lecture 6: September 22

Negative Examples for Sequential Importance Sampling of Binary Contingency Tables

Advanced Sampling Algorithms

Markov Chains CK eqns Classes Hitting times Rec./trans. Strong Markov Stat. distr. Reversibility * Markov Chains

Characterization of cutoff for reversible Markov chains

2. Variance and Covariance: We will now derive some classic properties of variance and covariance. Assume real-valued random variables X and Y.

XVI International Congress on Mathematical Physics

arxiv: v4 [cs.dc] 26 Feb 2019

ARCC WORKSHOP: SHARP THRESHOLDS FOR MIXING TIMES SCRIBES: SHARAD GOEL, ASAF NACHMIAS, PETER RALPH AND DEVAVRAT SHAH

University of Chicago Autumn 2003 CS Markov Chain Monte Carlo Methods

Clustering and Mixing Times for Segregation Models on Z 2

Approximate Counting and Markov Chain Monte Carlo

Predator-Prey Population Dynamics

Math 456: Mathematical Modeling. Tuesday, April 9th, 2018

Protein Folding Challenge and Theoretical Computer Science

25.1 Markov Chain Monte Carlo (MCMC)

144 IEEE/ACM TRANSACTIONS ON NETWORKING, VOL. 17, NO. 1, FEBRUARY A PDF f (x) is completely monotone if derivatives f of all orders exist

Not all counting problems are efficiently approximable. We open with a simple example.

MCMC notes by Mark Holder

CS155: Probability and Computing: Randomized Algorithms and Probabilistic Analysis

Guiding Monte Carlo Simulations with Machine Learning

Efficient Algorithms for Universal Portfolios

Getting lost while hiking in the Boolean wilderness

Rapid Mixing of Several Markov Chains for a Hard-Core Model

Calculus with business applications, Lehigh U, Lecture 03 notes Summer

Path Coupling and Aggregate Path Coupling

Markov Processes. Stochastic process. Markov process

Lecture 15: MCMC Sanjeev Arora Elad Hazan. COS 402 Machine Learning and Artificial Intelligence Fall 2016

Latent voter model on random regular graphs

Phase transition and spontaneous symmetry breaking

Counting independent sets up to the tree threshold

7.1 Coupling from the Past

One-Parameter Processes, Usually Functions of Time

Lecture 19: November 10

Programmable Control of Nucleation for Algorithmic Self-Assembly

CS 246 Review of Proof Techniques and Probability 01/14/19

Introduction to Design of Experiments

EECS 495: Randomized Algorithms Lecture 14 Random Walks. j p ij = 1. Pr[X t+1 = j X 0 = i 0,..., X t 1 = i t 1, X t = i] = Pr[X t+

Phase Transitions and Emergent Phenomena in Algorithms and Applications

Math 350: An exploration of HMMs through doodles.

Cheng Soon Ong & Christian Walder. Canberra February June 2018

Markov Chain Monte Carlo

Lecture 24 Nov. 20, 2014

Tensor network simulations of strongly correlated quantum systems

ICCP Project 2 - Advanced Monte Carlo Methods Choose one of the three options below

Markov chain Monte Carlo

Discrete solid-on-solid models

Universal phenomena in random systems

3.320 Lecture 18 (4/12/05)

Communities, Spectral Clustering, and Random Walks

Lecture 14: October 22

Conductance and Rapidly Mixing Markov Chains

Chapter 6 Randomization Algorithm Theory WS 2012/13 Fabian Kuhn

A Geometric Interpretation of the Metropolis Hastings Algorithm

Chapter 4 Beyond Classical Search 4.1 Local search algorithms and optimization problems

Bias-Variance Error Bounds for Temporal Difference Updates

Probabilistic Machine Learning

Computer Science 385 Analysis of Algorithms Siena College Spring Topic Notes: Limitations of Algorithms

Transcription:

Self-Assembly and Convergence Rates of Heterogeneous Reversible Growth Processes (Extended Abstract) Amanda Pascoe 1 and Dana Randall 2 1 School of Mathematics, Georgia Institute of Technology, Atlanta GA, 30332-0280. 2 School of Computer Science, Georgia Institute of Technology, Atlanta GA, 30332-0765. Abstract. In tile-based models of self-assembly, tiles arrive and attach to large a aggregate according to various rules. Likewise, tiles can break off and detach from the aggregate. The equilibrium and non-equilibrium dynamics of this process can be studied by viewing them as reversible Markov chains. We give conditions under which certain chains in this class are provably rapidly mixing, or quickly converging to equilibrium. Previous bounds on convergence times have been restricted to homogeneous chains where the rates at which tiles attach and detach are independent of the location. We generalize these models to allow rates to be location dependent, giving the first rigorous results of fast convergence to equilibrium for a heterogeneous tile-based growth models. Supported in part by NSF grants CCR-0830367 and DMS-0505505.

Self-Assembly and Convergence Rates of Heterogeneous Reversible Growth Processes (Extended Abstract) Amanda Pascoe 1 and Dana Randall 2 1 School of Mathematics, Georgia Institute of Technology, Atlanta GA, 30332-0280. 2 School of Computer Science, Georgia Institute of Technology, Atlanta GA, 30332-0765. Abstract. In tile-based models of self-assembly, tiles arrive and attach to large a aggregate according to various rules. Likewise, tiles can break off and detach from the aggregate. The equilibrium and non-equilibrium dynamics of this process can be studied by viewing them as reversible Markov chains. We give conditions under which certain chains in this class are provably rapidly mixing, or quickly converging to equilibrium. Previous bounds on convergence times have been restricted to homogeneous chains where the rates at which tiles attach and detach are independent of the location. We generalize these models to allow rates to be location dependent, giving the first rigorous results of fast convergence to equilibrium for a heterogeneous tile-based growth models. 1 Introduction In tile-based self-assembly models, tiles are constructed so that specified pairs are encouraged or discouraged to join together. This model was first introduced by Wang [11] and was shown to be a universal model of computation. Winfree [12] and others studied the model in the context of DNA computation by constructing tiles made of double-crossover strands, where the sides of each tile could be encoded with single-stranded DNA and tiles are more or less likely to form bonds according to the number of complementary base pairs on their corresponding sides. See, e.g., [5,??, 12] for more details. Here we are concerned with determining the rate of convergence to equilibrium of some simple self-assembly processes. In particular, we consider a lattice-based growth model consisting of three types of tiles: a seed tile placed at the origin, border tiles that form the left and bottom border of an n n region, and interior tiles that can be placed in the remaining part of the region. We assume that the border forms quickly after a seed tile appears, and we restrict to a growth model in which interior tiles can attach to the aggregate if their left and bottom neighbors are present, whereas they can detach if their right and upper neighbors are absent. It is easy to see that the only shapes that can be formed are tiles that are neatly packed into the lower-left corner of the region so that the upper border forms a staircase Supported in part by NSF grants CCR-0830367 and DMS-0505505.

walk that travels to the right and down (see Fig. 1). We refer to the ratio of the attach and detach rates as the bias of the growth process. Fig.1. A pair of tilings connected by a single move of the Markov chain. Variants of this growth model have been previously considered in order to study various aspects of self-assembly. For example, Baryshnikov et al. [2] consider a non-reversible version of this process when n is large to determine the limiting shape, and note that the behavior is captured by a TASEP (totally asymmetric simple exclusion process). Majumder et al. [8] initiated the study of the reversible process in the finite setting where tiles are allowed to detach as well as attach to the aggregate. It has been noted for many models that dynamics of self-assembly can be viewed as a Markov chain, and the convergence rate of the chain captures the rate that the assembly process tends towards equilibrium (see, e.g., [1, 7, 8, 6]). Majumder et al. [8] prove that the reversible growth process described above is rapidly mixing (i.e., it converges in time polynomial in n) in two and three dimensions if tiles attach at a much faster rate than they detach. In fact, the same two-dimensional process had been studied previously by Benjamini et al. [3] in the context of biased card shuffling and biased exclusion processes. They give optimal bounds on the mixing time of the chain of O(n 2 ) for all fixed values of the bias. Subsequently, Greenberg et al. [7] discovered an alternative proof establishing the optimal bound on the mixing rate for any fixed bias in two dimensions. This new proof is much simpler, and moreover it can be generalized to higher dimensional growth processes provided the bias is not too close to one. A common feature of all of these previous results is that all interior tiles are always treated identically. Thus, the rate at which a tile is allowed to attach or detach from the large aggregate is independent of the position of the tile. This is a severe shortcoming since in reality several properties of the growth processes can affect these rates, including the particular encodings on the sides of the double-crossover molecules comprising the tiles, as well as the relative densities of different tiles that can be attached in various positions. Mathematically, the problem becomes more intriguing as well. Physicists refer to this property whereby the bias is location dependent as fluctuating bias, and it has been noted that methods that allow us to analyze systems with fixed bias do not readily generalize to the fluctuating setting.

In this paper we consider the heterogeneous setting where the bias of a tile depends on its location. The state space Ω us the set of valid tilings of the n n region where the upper/right boundary is a staircase walk starting at (0, n) and ending at (n, 0). Let α(x, y) and β(x, y) be the probabilities of adding and removing a tile at position (x, y) for the heterogeneous model. We call λ x,y = α(x, y)/β(x, y) the bias at position (x, y). The self-assembly Markov chain M SA tries to add or remove individual tiles in each step and is formalized below. We give the first rigorous proofs that the chain is rapidly mixing for various settings of the {λ x,y }. In particular, we derive optimal bounds on the mixing time if λ x,y > 4 for all (x, y) and we give polynomial bounds on the mixing time when λ x,y 2. 2 The Mixing Time of the Self-Assembly Markov Chain The growth process is defined so that the rate of adding a tile at position (x, y) is α(x, y) and the rate of removing that tile is β(x, y). We are interested in the case where λ x,y = α(x, y)/β(x, y) 1 for all x, y. Since the chain is reversible, there is a unique stationary distribution, and it is straightforward to see the chain converges to (x,y) σ π(σ) = λ x,y, Z where Z = τ Ω (x,y) τ λ x,y is the normalizing constant. Our goal is to sample configurations in Ω according to the distribution π. We can now formalize the Markov chain M SA. At each time step, we pick a vertex v uniformly along the upper border of the aggregate, each with probability 1/(2n) since the border has length 2n. It will be convenient in the analysis to add a self-loop probability at each position, so we flip a coin and if the coin lands on heads we try to add a tile at v and if the coin is tales we try to remove the tile at v. To define the transition probabilities of M SA, we use the Metropolis-Hastings probabilities [9] with respect to π. The transitions are 1/(4n), if τ = σ t x,y for some tile t x,y ; 1/(4nλ x,y ), if σ = τ t x,y for some tile t x,y ; M SA (σ, τ) = 1 τ τ M SA(σ, τ ), if σ = τ; 0, otherwise. The time a Markov chain takes to converge to its stationary distribution, or the mixing time, is measured in terms of the total variation distance between the distribution at time t and the stationary distribution (see, e.g., [7]). Our goal is to show that the mixing time of the chain M SA is bounded by a polynomial in the size of the region being tiled. We can show this provided the lower bound on the bias is not too close to one. We briefly state our results and outline the two general techniques used, deferring complete proofs for the full version of the paper. We believe these approaches can be generalized to many other self-assembly models.

2.1 Coupling Using A Geometric Distance Function Theorem 1 Let λ L 1 be a lower bound on the bias and let λ U be an upper bound satisfying { ( 2 λ U := λl 1) 1 if λ L < 4 if λ L 4. If, for all x, y, we have λ L < λ x,y < λ U, then the mixing time of M SA satisfies τ(ǫ) = O(n 2 lnǫ 1 ). In particular, this theorem gives the optimal bound on the mixing time when the lower bound λ L on the bias everywhere is at least four. This result is obtained by the path coupling method (see [4]) using a carefully defined distance metric. As in [7] which deals with the homogeneous case (i.e., fixed λ), we use a geometric distance metric, with the following additional definition. For any tile v = (x, y), let diag(v) = x + y. Then we define the distance function φ to be φ(σ, ρ) = γ diag(v) 2n, v=(x,y) σ ρ where γ = 1 2 (1 + 1 λ U ). We can now verify that any pair of tilings at Hamming distance 1 (e.g., Fig 2 (a)) φ is decreasing in expectation during moves of the coupled chain. This is sufficient to bound the mixing time of the chain following the arguments in [7]. T λ v 2 T ρ t σ t X λ v 1 Fig.2. (a) Example of tilings at Hamming distance 1; (b) A staircase walk with 5 good (increasing) moves and 4 bad (decreasing) moves with respect to the highest tiling T; (c) A region R where the tiles along the upper boundary have bias greater than 2, whereas the other tiles are unrestricted. 2.2 Hitting Time to the Maximal Tiling Using a different approach based on hitting times we can show that the chain is rapidly mixing provided the bias everywhere is at least 2. Theorem 2 Suppose that for all positions (x, y), 2 λ L. Then the mixing time of M SA satisfies τ(ǫ) = O(n 3 ln(nǫ 1 )).

The proof of this theorem relies on the monotonicity of M SA with respect to the trivial coupling. In other words, if (σ t, ρ t ) are coupled and σ t ρ t, then after one step of the coupling, σ t+1 ρ t+1. This implies that the coupling time is bounded by the time to hit the highest configuration T starting from the bottom configuration B, and we can show that this will happen quickly because the distance to T is always non-increasing in expectation (see Fig. 2 (b)). In fact, something more general is true. The tile biases can vary widely everywhere, as long as the upper boundary of the region has bias at least λ L = 2 (see Fig. 2 (c)). This surprising result says that the pull of the upper boundary tiles is strong enough to ensure a fast hitting time to the highest configuration. Theorem 3 Let R be an n n region. Suppose that for every (x, y) R the bias λ x,y > 1. Define the upper border S = {(x, n) R} {(n, y) R} and suppose that for all (x, y) S, we have λ x,y > 2. Then τ(ǫ) = O(n 3 ln(nǫ 1 )). References 1. L. Adleman, Q. Cheng, A. Goel, M.D. Huang and H. Wasserman. Linear selfassemblies: Equilibria, entropies and convergence rates. 6th Int. Conference on Difference Equations and Applications, 2001. 2. Y. Baryshnikov, E.G. Coffman Jr. and B. Yimwadsana. Times to Compute Shapes in 2D Self Assembly. Proc. of the 12th Int. Meeting on DNA Computing, 2006. 3. I. Benjamini, N. Berger, C. Hoffman and E. Mossel. Mixing times of the biased card shuffling and the asymmetric exclusion process. Transactions of the American Mathematics Society 357: 3013 3029, 2005. 4. R. Bubley and M. Dyer. Faster random generation of linear extensions. Discrete Math., 201: 81 88, 1999. 5. T.-J. Fu and N.C. Seeman. DNA double-crossover molecules. Biochemistry, 32: 3211 3220, 1993. 6. S. Greenberg and D. Randall. Convergence rates of Markov chains for some selfassembly and non-saturated Ising models. Theoretical Computer Science, 410: 1417 1427, 2009. 7. S. Greenberg, A. Pascoe and D. Randall. Sampling biased lattice configurations using exponential metrics. 19th Symposium on Discrete Algorithms, 76 85, 2009. 8. U. Majumder, S. Sahu and J. Reif. Stochastic analysis of reversible assembly. Journal of Computational and Theoretical Nanoscience, 5: 1289 1305, 2008. 9. N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, and E. Teller. Equation of state calculations by fast computing machines. Journal of Chemical Physics 21: 1087 1092, 1953. 10. N.C. Seeman. DNA in a material world. Nature, 421: 427 431, 2003. 11. H. Wang. Proving theorems by pattern recognition II. Bell Systems Technical Journal, 40: 1 41, 1961. 12. E. Winfree. Simulations of Computing by Self-Assembly. 4th DIMACS Meeting on DNA Based Computers, University of Pennsylvania, 1998.