PROTEINS: Structure, Function, and Bioinformatics 61:94 –104 (2005)

Exhaustive Metropolis Monte Carlo Sampling and Analysis of Polyalanine Conformations Adopted Under the Influence of Hydrogen Bonds Alexei A. Podtelezhnikov* and David L. Wild Keck Graduate Institute of Applied Life Sciences, Claremont, California

ABSTRACT We propose a novel Metropolis Monte Carlo procedure for protein modeling and analyze the influence of hydrogen bonding on the distribution of polyalanine conformations. We use an atomistic model of the polyalanine chain with rigid and planar polypeptide bonds, and elastic ␣ carbon valence geometry. We adopt a simplified energy function in which only hard-sphere repulsion and hydrogen bonding interactions between the atoms are considered. Our Metropolis Monte Carlo procedure utilizes local crankshaft moves and is combined with parallel tempering to exhaustively sample the conformations of 16-mer polyalanine. We confirm that Flory’s isolated-pair hypothesis (the steric independence between the dihedral angles of individual amino acids) does not hold true in long polypeptide chains. In addition to 310- and ␣-helices, we identify a kink stabilized by 2 hydrogen bonds with a shared acceptor as a common structural motif. Varying the strength of hydrogen bonds, we induce the helix– coil transition in the model polypeptide chain. We compare the propensities for various hydrogen bonding patterns and determine the degree of cooperativity of hydrogen bond formation in terms of the Hill coefficient. The observed helix– coil transition is also quantified according to Zimm–Bragg theory. Proteins 2005;61:94 –104. ©

2005 Wiley-Liss, Inc.

Key words: loop-closure; rebridging; parallel tempering; polypeptide; 310-helix; ␣-helix; helix– coil transition; ␲-bulge INTRODUCTION For many years, computer simulations have been indispensable in studying protein structure.1,2 Computer simulations continue to shed light on the dynamics of protein folding, help in searching for the lowest energy conformation of a protein with unknown tertiary structure, and unravel protein function phenomena. Structural protein modeling continues to face 2 major challenges: one is the development of an efficient sampling technique for rapid search through the enormous conformational space; the other is the derivation of the energy function describing molecular interactions for the problem at hand. In this work, we propose a new exhaustive sampling technique and adopt a simplified energy function to validate it through a study of polyalanine conformations. ©

2005 WILEY-LISS, INC.

Monte Carlo (MC) simulations, along with molecular dynamics, are among the most commonly used methods of sampling conformational space. To reduce the number of degrees of freedom, MC simulations are often performed in the space of dihedral ␾–␺ angles. In an efficient Metropolis MC procedure, the conformations of just a few amino acids are perturbed locally on each step, leaving the rest of the chain intact. Metropolis MC with local moves in the dihedral space requires solving the so-called “loopclosure,” “ring-closure,” or “rebridging” problem3 (i.e., special moves have to be designed and computed at each step of the procedure). The solution of this problem is not trivial and is computationally expensive.4 –11 We propose a new Metropolis MC procedure that avoids computationally expensive “rebridging” at the price of an extra degree of freedom per amino acid in addition to the ␾–␺ angles. The important feature of our model is the flexibility of the ␣ carbon (C␣) valence geometry. With flexible C␣ valence angles, we are able to use local crankshaft moves inspired by earlier Metropolis MC studies of large-scale DNA properties.12,13 We believe that the simplicity of our Metropolis MC procedure is well worth the adding of an extra degree of freedom to our model. We also draw on parallel tempering (replica exchange)14 to speed up equilibration of the system. Identifying and calibrating the interatomic interactions that ultimately control protein folding and function remain the most challenging aspects of protein science (see review15). The first and probably the most successful model describing interatomic interactions is the hardsphere model. This model alone is responsible for the Ramachandran plots,16,17 the concept of surface accessible area,18 and tight protein packing.19 Second, hydrogen bonds play a central role in stabilizing the secondary structure of proteins.20,21 Finally, hydrophobic22 and electrostatic23 interactions are also important in determining protein structure.24,25

Grant sponsor: National Institutes of Health; Grant number: 1P01 GM63208. *Correspondence to: Alexei A. Podtelezhnikov, Keck Graduate Institute of Applied Life Sciences, 935 Watson Drive, Claremont, CA 91711. E-mail: [email protected] Received 8 December 2004; Revised 14 February 2005; Accepted 15 February 2005 Published online 27 July 2005 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/prot.20513

METROPOLIS SAMPLING OF POLYALANINE

Here, our primary focus is on the influence of hydrogen bonding on the structure of the 16-mer polyalanine. Polyalanine chains have long been considered an attractive model system with which to study polypeptide structure; see Nguyen et al.25 and references therein. Our polyalanine model accounts only for hard-sphere repulsion and hydrogen bonding. Although we do not include any explicit interactions with solution, we vary the strength of hydrogen bonds that mimics the effect of different solvent conditions. We have chosen not to take electrostatic and hydrophobic interactions into consideration to facilitate interpretability of the results. Perhaps surprisingly, even with this simplified force field, our simulations recapitulate many experimentally observed motifs of protein secondary structure. The Methods section contains a detailed description of the algorithm and the force field used in this work. Using our algorithm, we were able to perform exhaustive sampling of polyalanine conformations in continuous (off-lattice) space. We confirmed the earlier finding that Flory’s isolated-pair hypothesis does not hold true.26 We observed the predominant formation of 310- and ␣-helices in polyalanine under the influence of hydrogen bonds. The formation of ␤-hairpins was not reliably detected. Our simulated Ramachandran plots resemble the Ramachandran plots observed in actual crystal structures in the ␣-helical region, including the region around ␾ ⫽ ⫺120° and ␺ ⫽ ⫺40°.27–30 In our simulations, these (second most populated) conformations were found in the stable kink between the helices, where 2 hydrogen bonds formed with single-backbone oxygen. We were able to compare the relative propensities for different structural motifs. Varying the strength of hydrogen bonds, we observed and quantified the classical helix– coil transition31–33 in our model polypeptide. These and other observations are presented in the Results and Discussion section. METHODS Polypeptide Chain Model In our model, the primary descriptors of the polypeptide chain conformation were the orientations of the peptide bonds in the laboratory frame [Fig. 1(A)]. For a chain of N amino acids, the orientations of the peptide bonds were specified by the orthonormal triplets, (xi, yi, zi), i ⫽ 0, …, N. The vector zi pointed in the direction from C␣i (C␣ of amino acid i) to C␣i ⫹ 1. The distance between the C␣’s was fixed and equal to 3.804 Å, that is, r(C ␣i ⫹ 1) ⫽ r(C ␣i ) ⫹ 3.804z i,

(1)

i ␣

where r(C ) is the position of the C␣ in the laboratory frame, i ⫽ 1, …, N. We used Eq. (1) to calculate the positions of C␣’s iteratively. The peptide bond atoms lay on the yz planes. Specifically, the positions of the backbone atoms were calculated as follows: r(C i) ⫽ r(C i␣) ⫺ 0.534y i ⫹ 1.428zi r(O i) ⫽ r(C␣i ) ⫺ 1.745yi ⫹ 1.648zi r共Ni兲 ⫽ r(Ci␣) ⫹ 0.377yi ⫺ 1 ⫺ 1.408zi ⫺ 1

(2)

95

Fig. 1. (A) Polypeptide model. The orientations of perfectly planar and rigid peptide bonds are given by the orthonormal triplets (x, y, z), with z pointing along the C␣OC␣ direction. Other peptide bond atoms lie in the plane yz. The position of the side-chain atoms R is specified by the vectors n and c. (B) Local Metropolis moves. Two types of moves are used in this work: a crankshaft rotation around the line connecting 2 C␣ atoms in the middle of the chain, and a random rotation at the termini around a random axis passing through the C␣ atom.

r(H i) ⫽ r(C i␣) ⫹ 1.328y i ⫺ 1 ⫺ 1.648z i ⫺ 1, where the coordinates are given in angstroms. This backbone geometry corresponds to the classical average bond lengths and angles from Engh and Huber.34 Note that the terminal triplets (x0, y0, z0) and (xN, yN, zN) specified only the coordinates of N- and C-terminal atoms, rather than the complete set of the backbone atoms. Each side-chain was represented by a single pseudoatom R in our model. The length of the C␣OR bond, b, was stipulated by the amino acid type. The direction of the bond corresponded to the direction of the ␤ carbon, with the chirality of L-amino acids preserved. To derive the position of the pseudoatom R, we assumed ideal tetrahedral valence geometry at C␣, with all valence angles equal to arccos(⫺1/3). Specifically, b b 冑3 r(R i) ⫽ r(C ␣i ) ⫹ (n i ⫺ ci) ⫺ ni ⫻ ci, 2 2

(3)

where ni and ci are the unit vectors in the direction of the NOC␣ and C␣OC bonds, respectively [Fig. 1(A)]. According to Eq. (2), these 2 vectors are defined by the orientations of the corresponding peptide bonds. Thus, r(Ri) was unambiguously determined by the 2 adjacent triplets. In the case of alanine, the position of the pseudoatom R corresponded to the position of the ␤ carbon, C␤, with the C␣OC␤ bond length b ⫽ 1.531 Å.34 In our simulations, the adjacent peptide bonds (the triplets) were allowed to rotate relative to each other.

96

A.A. PODTELEZHNIKOV AND D.L. WILD

These rotations resulted in alterations in dihedral angles ␾i and ␺i. The angles were unambiguously determined by the peptide bond orientations and calculated as follows:



␾ i ⫽ arctan



pi ⫺ 1 䡠 关ni ⫻ ci] [pi ⫺ 1 ⫻ ni] 䡠 [ni ⫻ ci]



␺i ⫽ arctan



[ni ⫻ ci] 䡠 pi , [ni ⫻ ci] 䡠 [ci ⫻ pi]

(4)

where pi is the unit vector in the direction of the CiONi⫹1 bond [Fig. 1(A)]. The signs of the numerator and the denominator in Eq. (4) define the quadrant of the angles. We readily calculated the angles using the standard C library function atan2.35 The axes of rotations of adjacent peptide bonds i ⫺ 1 and i relative to each other were not constrained to vectors ni and ci. It is important to note that we did not hold fixed the valence geometry of the C␣ atoms, but rather restricted it with bending energy terms. Namely, the valence angles NOC␣OC, ␶i ⫽ arccos(⫺ ni 䡠 ci), were permitted to slightly diverge from ideal value of arccos(⫺1/3) as a result of the peptide bond rotations. We still applied Eq. (3) to calculate r(Ri) in the deformed amino acids. As follows from Eq. (3), the valence angles NOC␣OR and ROC␣OC were also flexible and coupled to the variations of the angle ␶i. To summarize, in our model, there are 3 independent degrees of freedom per amino acid of the peptide chain: the valence angle, ␶, and 2 traditional dihedral angles, ␾ and ␺. The former was firmly restrained by the bending energy. The latter 2 were controlled by other interatomic interactions.

r(B). We used values of hard-sphere atomic radii close to the lower limit of the range set in the literature16,26,37,38: r(C␣) ⫽ r(C␤) ⫽ 1.57 Å, r(C) ⫽ 1.42 Å, r(O) ⫽ 1.29 Å, r(N) ⫽ 1.29 Å. Note that in our model the backbone hydrogen atoms were excluded from the collision analysis but were important in identifying hydrogen bonds. Hydrogen bonds were introduced as nonlocal interactions between the backbone oxygen and hydrogen atoms only. The energy of the hydrogen bonds was given by square-well potentials, E ijHB ⫽ ⫺ nhH,

where nh is the number of hydrogen bonds between the amino acids i and j; H is the strength of each hydrogen bond. We considered the hydrogen bond formed when r(O, H) ⬍ 2.1 Å, ⬔OHN ⬎ 150°, and ⬔COH ⬎ 90°. Symbol ⬔ denotes the angle between the 3 atoms. The lower boundary on the separation between the atoms [r(O, H) ⬎ 1.8 Å] was implicitly set by the hard-sphere collision between oxygen and nitrogen. This definition of a hydrogen bond captures the most important characteristics of hydrogen bond geometry observed in crystal structures.39 – 44 We varied the strength of hydrogen bonds H in order to examine their importance in the formation of helices and other structural motifs in peptides. Although we left out a more detailed description of hydrogen bonds, we believe that this description is sufficient for the purposes of our simulations. To summarize, the total energy of a polypeptide chain conformation was calculated as follows:

Model Interactions

E iB ⫽

冘 N

To restrict the valence angle NOC␣OC, ␶i, to the values near the ideal tetrahedral geometry, we introduced a special local bending energy given by ε␶ 关␶ ⫺ arccos共 ⫺ 1/3兲兴2, 2 i

(5)

where ε␶ is the bending rigidity equal to 250 kcal/(mol 䡠 rad2). In our simulations at room temperature, T ⫽ 300 K, with this value of bending rigidity, the standard deviation of ␶i was approximately equal to 2.8°, in agreement with the deviation observed in crystal structures. This valence angle is the most diverse angle in crystal structures.34 Also, this angle variability was shown to play an important role in some peptide conformations.17,26,36 Since the valence angles NOC␣OR and ROC␣OC are not independent parameters in our model, we did not introduce any other bending energy terms besides Eq. (5). To mimic van der Waals repulsions between the atoms, we used hard-sphere potentials: E ijvdW ⫽ ncW,

(6)

where nc is the number of collisions between the atoms of amino acids i and j; W ⫽ 15 kcal/mol is the prohibitively large positive cost of collisions. Two atoms, separated by at least 3 chemical bonds, collided when the distance between them, r(A, B), was less then the sum of their radii, r(A) ⫹

(7)

E⫽

i⫽1

冘冘 N

E iB ⫹

i

共EijvdW ⫹ EijHB兲.

(8)

i⫽1 j⫽1

Our model force field does not include other energy terms such as electrostatic and hydrophobic interactions, and other interactions with solution. We used interatomic energy terms that were inspired by the classical force fields but greatly simplified for the purposes of our simulation. Metropolis Monte Carlo Procedure To obtain the canonical ensemble of polypeptide conformations, we developed a novel Metropolis MC procedure. Each new chain conformation was generated from the previous one by applying a move to a randomly chosen chain segment containing 1 or 2 adjacent peptide bonds. Each move was local, and the conformation of the rest of the chain was not altered. Local moves are extremely important to achieve efficient sampling of the polypeptide conformational space.5,8 There were 2 types of moves in our procedure. The move of the first type was a crankshaft rotation12 of the peptide bonds i and i ⫹ 1, where i was picked randomly. The rotation axis was the line connecting C␣’s i and i ⫹ 2 given by the difference r(C␣i ⫹ 2) ⫺ r(C␣i ) [Fig. 1(B)]. The rotation angle was uniformly distributed over the range [⫺␣0, ␣0]. We applied the rotation to the triplets (xi, yi, zi) and (xi⫹1, yi⫹1, zi⫹1) and then calculated the new

97

METROPOLIS SAMPLING OF POLYALANINE

positions of the atoms using Eqs. (1) through (3). As a result, the backbone atoms of the peptide bonds i and i ⫹ 1, and the entire amino acid i ⫹ 1 were rotated as a rigid body. Since this rotation was not necessarily around one of the NOC␣ or C␣OC bonds, this move could alter all the dihedral angles ␾i, ␺i, ␾i⫹2, and ␺i⫹2, as well as the valence angles ␶i and ␶i⫹2. It also likely changed the positions of side-chains i and i ⫹ 2. The rest of the chain and its geometry remained intact. The move of the second type involved a rigid-body rotation of 1 or 2 peptide bonds at one of the chain termini picked randomly [Fig. 1(B)]. The rotation axis passed through the C␣ and had a random direction. The rotation angle was uniformly distributed over the same range [⫺␣0, ␣0]. This move altered the conformation of the polypeptide terminus in a manner similar to that described above. Since the direction of these rotations was random (unlike the rotations of the first type), the C␣ atoms at the N- and C-terminus were able to move in space. To equalize the chances for each chain segment to be moved, moves of the first type were proposed with the probability of (N ⫺ 2)/(N ⫹ 2), while moves of the second type were attempted with the remaining probability of 4/(N ⫹ 2). In our Metropolis MC procedure, forward and reverse rotations were proposed with equal probabilities. The Jacobian of the transformation, which is a combination of rotation matrices, is equal to 1. The proposed rotations from original to trial states conserve the volume element in the space of peptide bond orientations. Therefore, to satisfy the principle of microscopic reversibility, trial rotations were accepted with the standard Metropolis probability45: P共old 3 new) ⫽ min兵1,exp(⫺␤⌬E兲其,

(9)

where ⌬E is the energy difference between the trial and the old conformations given by Eq. (8), and ␤ ⫽ 1/RT is the reciprocal product of the gas constant and the absolute temperature. The rotation amplitude, ␣0, was selected so that approximately half of the trial rotations were accepted. This Metropolis procedure would be impossible without making the C␣ valence geometry flexible in our model. Our local moves do alter the valence angles ␶. Designing the local moves that keep this angle fixed is a geometrically complex and computationally expensive problem often referred to as “loop-closure” or “rebridging” in the literature.5,7,8,11 At the price of 1 degree of freedom per amino acid, we avoided solving this problem at each step of the simulation procedure. Also, in this formulation, the acceptance function given by Eq. (9) does not require evaluation of the Jacobian correction, which is necessary in the space of dihedral angles.4,6,9,10 We believe that 3 instead of 2 degrees of freedom per amino acid is a small price to pay for the simplicity of our implementation. Parallel Tempering Scheme Besides using local moves in the Metropolis MC procedure, to further improve the efficiency of sampling, we used parallel tempering.14,46 In this method, S replicas of

our model polypeptide are simulated at different temperatures Ti, i ⫽ 0, …, S ⫺ 1, starting from room temperature and up. The systems at higher temperatures evolve faster, because they can more easily escape from local energy minima. In addition to the normal set of Metropolis moves performed in each system, swapping moves occasionally exchange the temperatures of 2 randomly chosen replicas.47 The temperature swaps are accepted with the following probability: P共swap) ⫽ min兵1,exp(⫺⌬␤⌬E兲其

(10)

where ⌬␤ ⫽ ␤i ⫺ ␤j ⫽ 1/RTi ⫺ 1/RTj. When the system at room temperature is involved in the swap, its counterpart becomes the new room temperature system that we follow closely in our simulations. Although the difference in the energies of the systems ⌬E can be very large, 2 systems with close temperatures (small ⌬␤) can still swap with high probability. As a result, the large conformational changes that happen at higher temperatures propagate to lower temperatures. To balance the swapping probabilities between all replicas, their temperatures must have equal relative separation (or be evenly spaced on a logscale). Given the size of available multiprocessor cluster or the total number of replicas S, the choice of the temperatures for the replicas is governed by 2 conflicting requirements: the large value for the highest temperature and a small spacing between the neighboring temperatures. To balance between these requirements, we used the following initial iterative assignment:



␤ i⫹1 ⫽ ␤ i 1 ⫺



log 2共S ⫹ 1兲 , 2S

(11)

with ␤0 corresponding to room temperature. In this scheme, the highest temperature grows asymptotically linearly with S, while the separation between the temperatures is approximately reciprocal to S. The frequency of the temperature swap attempts was arbitrarily set to once per 16,384 regular Metropolis steps. We verified that the room temperature replica successfully participated in a temperature swap at least 100 times in a typical simulation. Validation and Performance To validate the principle of microscopic reversibility in our Metropolis MC procedure, we tested the distribution of the backbone dihedral (torsion) angles. This distribution should be uniform in the absence of hard-sphere repulsion (W ⫽ 0) and hydrogen bonding (H ⫽ 0). This sensitive test has been used in earlier studies.4,10 Figure 2 shows the histograms of ␾ and ␺ angles obtained in the special simulation run using Eq. (4). The unbiased sampling of dihedral angles within statistical error demonstrates that the microscopic reversibility principle is indeed satisfied in our procedure. The performance of our Metropolis procedure was evaluated by measuring the average step size. In our typical simulation run with H/RT ⫽ 3, approximately 50% of the proposed rotations were accepted when we chose the

98

A.A. PODTELEZHNIKOV AND D.L. WILD

tempering improved mixing between helical and extended conformations, which is unlikely at low temperatures because of the high energetic barrier. To demonstrate the efficiency of mixing, we traced the Shannon entropy of the reduced 2-state distribution of the amino acid conformations:

冘 N

h共p,q兲 ⫽ ⫺

共p i log 2pi ⫹ qilog2qi兲,

(13)

i⫽1

Fig. 2. Distribution of dihedral angles in the simulation without hardsphere repulsion and hydrogen bonding. The solid horizontal line corresponds to the theoretical uniform probability density. The 48-bin normalized histograms for the ␾ (solid line) and ␺ (dotted line) angles were obtained from 6000 octa-peptide conformations. The deviations from the theoretical value are within the statistical error of about 3.2%.

amplitude of proposed rotations ␣0 ⫽ 14°. The standard deviation of the rotation angles in each step was equal to 4.5°. We found that the step size was controlled by the C␣ elasticity and reciprocal to the value of bending rigidity, ε␶. We also calculated that each step of our procedure resulted in an average absolute change of a single dihedral angle ␾ or ␺ by approximately 2°. This value indicates improvement in comparison with the earlier Metropolis sampling procedures, where the dihedral angle step size was less than 1°.10 We estimated the statistical inefficiency of our procedure as determined by the variance of an observable, ␴2(A), and the variance of its block averages, ␴2(An): n␴2共An兲 , 2 n 3 ⬁ ␴ 共A兲

I共A兲 ⫽ lim

(12)

where the block averages An are taken from the consecutive adjacent blocks of size n.2 The statistical inefficiency defines the size of uncorrelated blocks or how often a particular observable should be sampled in the procedure. For the 16-mer polyalanine with H/RT ⫽ 3, we calculated that I(cos ␾8) was approximately equal to 2 䡠 105. Given the accessible space of 160° for a ␾ angle, the step size of 2° in our procedure, with ␾8 being randomly altered each eighth step, we can estimate a theoretical lower bound for the inefficiency 8 䡠 (160/2)2 ⫽ 51,200. Thus, the statistical inefficiency of our procedure is consistent with the theoretical estimate. Parallel tempering played a crucial role in our Metropolis procedure. At high hydrogen bond strengths, parallel tempering provided more efficient mixing between numerous stable hydrogen bonding patterns. In addition, parallel

where the sum is taken over N amino acids. The probabilities of helical and extended conformations of the amino acid i, pi, and qi, respectively, were estimated from their frequencies in the samples collected so far. We considered any conformation inside the contiguous region with ␾ ⰻ [0°, 120°] and ␺ 僆 [⫺120°, 60°] as helical. All the other conformations were considered as extended (i.e., by definition pi ⫹ qi ⫽ 1). Figure 3(A) demonstrates the Shannon entropy trace in the special simulations of 16-mer polyalanine with H/RT ⫽ 3. The total of 65 million Metropolis steps were performed using 1, 4, 8, or 16 parallel replicas. The simulation with a single replica failed to reach equilibrium altogether. The simulations with multiple replicas all reach equilibrium with the Shannon entropy of approximately 15.6. Increasing the number of replicas resulted in faster equilibration. Figure 3(B) shows which of the 16 replicas was evolving at room temperature at a given Metropolis step. The successful temperature swaps with a room temperature replica occurred 68 times in the course of this simulation, or about 13.6% of the time. To summarize our method in its entirety, we simulated up to S ⫽ 32 parallel replicas of our model system at temperatures given by Eq. (11). We attempted 1 temperature swap move per 16,384 regular Metropolis steps. We started each computer experiment from extended polypeptide conformations. In the case of the strongest hydrogen bonding considered in this work (H/RT ⫽ 8), after initial 524,288 “burn-in” steps, we collected 2000 conformations separated by 524,288 Metropolis steps. Thus, we carried out more than 1 billion Metropolis steps in our longest computer experiment. The convergence of our procedure was validated by comparison of the results of 2 independent runs. On a Sun Netra X1 Cluster Grid, we achieved a performance of 20 million Metropolis steps per hour for each replica. Our longest simulation required 52 h to complete. Our program was written in standard C. We used MPI (Message Passing Interface) for parallel communications between replicas and relied on the Sun MPI implementation. RESULTS AND DISCUSSION Ramachandran Plot Analysis In this work, we simulated an equilibrium ensemble of 16-mer polyalanine conformations. In a helical conformation, the amino acids in the middle of the chain form 2 hydrogen bonds, while the terminal amino acids can only have 1 hydrogen bond. We disregarded 4 terminal amino acids at each end of the chain. We collected 2000 model conformations and examined the combined distribution of

METROPOLIS SAMPLING OF POLYALANINE

99

aries of the regions are blurry, because the valence angle NOC␣OC is allowed to slightly deviate from the ideal tetrahedral values in our simulations (see Methods section). The distribution of the ␾ and ␺ angles is almost uniform within each region away from the boundaries. This indicates that the chain adopts random coil conformations when there are no hydrogen bonds formed. We also compared the observed distribution with the distribution for the middle amino acid in model trialanine. In trialanine, the probability density was identical in both extended and compact helical conformations (data not shown). In the long model chains, we found that the extended conformations were about 70% more likely than the compact helical ones. This dissimilarity between the distributions contradicts Flory’s isolated-pair hypothesis (the steric independence between the dihedral angles of individual amino acids). This hypothesis was also proven invalid in earlier studies.26 As the strength of the hydrogen bond grows, the region corresponding to the 310 and ␣-helices becomes more and more populated, eventually overcoming the bias toward the extended conformations and taking over the overwhelming majority of the conformations (Fig. 4). The formation of ␤-hairpins, although theoretically possible, was not reliably detected in our simulations. Our simulations show that the right-handed helical structures are more favorable, as they provide saturated hydrogen bonding more readily. Hydrogen Bond Donors

Fig. 3. (A) Shannon entropy trace in the simulations with 1, 4, 8, and 16 replicas. Different lines correspond to the different number of available replicas as specified in the legend. (B) Temperature swaps in the course of the simulation with 16 replicas. The panel shows which replica is evolving at room temperature at a given step of the simulation.

the dihedral angles ␾ and ␺ for the 8 middle amino acids. Figure 4 shows the simulated distribution of the dihedral angles (Ramachandran plots) for 6 different values of the hydrogen bond strength, H from Eq. (7). The panels are labeled by the H/RT values in the top right corner. The highest value of the hydrogen bond strength in the figure corresponds to 4.8 kcal/mol. When H/RT ⫽ 0.0, the conformations of our model polypeptide chain are only defined by hard-sphere clashes between the atoms. Not surprisingly, on the simulated Ramachandran plot (Fig. 4, top left) we observed 2 major sterically allowed regions that correspond to the extended conformations and the compact helical ones. The bound-

Because the values of ␾ and ␺ angles for 310 and ␣-helix are close, it is impossible to distinguish between these structures using Ramachandran plots and precisely quantify the number of amino acids in each conformation. To analyze the distribution between the structures for different values of hydrogen bond strengths, we determined the patterns of hydrogen bonding between backbone atoms. In forming a hydrogen bond, each amino acid in the chain can be a donor (of hydrogen) or an acceptor (at oxygen). The conditions under which we considered a hydrogen bond to be formed are given in the Methods section [see Eq. (7)]. In our simulations, we observed that the patterns of hydrogen bonding were asymmetric for donors versus acceptors. Namely, each amino acid could be a donor in only 1 hydrogen bond, while it could be an acceptor in 1, 2, or 3 hydrogen bonds. This asymmetry was due to the fact that, in our model, oxygen had a finite hard-sphere radius, while we did not assign any radius to hydrogen atoms. Therefore, hydrogens from multiple donors could come close enough to form several hydrogen bonds with a single acceptor. We routinely observed “double” and, more rarely, “triple” hydrogen bonds. When the hydrogen bonds were weak, the model polyalanine chain largely adopted random coil conformations, where traditional secondary structure assignment48 is impossible. Since each amino acid can be a donor in only 1 hydrogen bond, it is more discriminatory to analyze hydrogen-donating patterns rather than hydrogen-accepting patterns. Based on this information, we were able to

100

A.A. PODTELEZHNIKOV AND D.L. WILD

Fig. 4. Simulated Ramachandran plots. The combined distribution of ␾ and ␺ angles for the amino acids in the middle of the chains was obtained for different values of hydrogen bond strengths. The panels are labeled at their top-right corners by the value of H/RT. The total of 16,000 data points is shown in each panel.

101

METROPOLIS SAMPLING OF POLYALANINE

TABLE I. Number of Donors in Different Hydrogen Bond Patterns Depending on the Hydrogen Bond Strength H/RT

None

i 3 i ⫺ 3 (310)

i 3 i ⫺ 4 (␣)

i 3 i ⫺ 5 (␲)

Other

0.0 1.0 2.0 3.0 3.5 4.0 4.5 5.0 6.0 7.0 8.0

15,980 15,946 15,748 14,767 12,285 10,532 7217 3259 1328 848 369

13 45 193 834 2638 3950 6282 9192 11,205 11,541 11,607

3 3 41 301 908 1397 2031 3233 3333 3531 3931

0 1 7 63 109 67 461 309 105 67 64

4 5 11 35 60 54 9 7 29 15 29

TABLE II. Number of Acceptors in Different Hydrogen Bond Patterns Depending on the Hydrogen Bond Strength H/RT

None

i 4 i ⫹ 3 (310)

i⫹3 i 4 i⫹4 (310,␣)

i 4 i ⫹ 4 (␣)

i 4 i ⫹ 5 (␲*)

Other

0.0 1.0 2.0 3.0 3.5 4.0 4.5 5.0 6.0 7.0 8.0

15,987 15,948 15,764 15,249 13,693 12,482 9556 5051 3376 2764 2723

11 39 176 503 1387 2195 3996 7227 8658 9118 8594

0 4 23 145 687 987 1233 2053 2020 2218 2572

1 4 23 44 91 145 339 1100 1303 1359 1714

0 1 7 36 87 120 776 541 635 499 355

1 4 7 23 55 71 100 28 8 42 42

*Majority of acceptors participating in ␲-like hydrogen bonds also participated 310- and/or ␣- like hydrogen bonds. The number of acceptors participating in ␲-like hydrogen bonds includes acceptors participating in multiple hydrogen bonds.

reasonably assign a helix type to a given amino acid. This definition is more suitable and convenient for our simulations. We counted the number of donors forming different types of hydrogen bonds among the middle 8 amino acids in the ensemble of 2000 conformations of 16-mer polyalanine. The results for different hydrogen bond strengths are presented in Table I. The overwhelming majority of formed hydrogen bonds could be classified as 310-, ␣-, or ␲-helical. The data shown in Table I demonstrate that the 310 helix is the most favorable structure of polyalanine adopted under influence of hydrogen bonds. It is worth noting that, independent of the hydrogen bond strength, there are roughly 3 times as many amino acids in the 310-helical conformation as in the ␣-helical one. We did not detect any appreciable dependence of the proportion on the selected criteria of hydrogen bond formation. Since all the hydrogen bond types have the same strength in our model, the 3-to-1 ratio of probabilities translates to a difference in entropy for these hydrogen bonds of S(310) ⫺ S(␣) ⬇ R. This entropy difference is close to an earlier estimate obtained using rejection MC sampling.49 The formation of 310 helices in alanine-based peptides has also been detected experimentally.50,51 The experimental evidence and our simulations suggest that both 310 and ␣-helices can coexist and interchange. The experimental estimates of the exact proportion of these conformations vary in the literature.

Hydrogen Bond Acceptors Unusual hydrogen bonding patterns, where bifurcated and double hydrogen bonds are observed, are common in helices52,53 and helix capping motifs.54 Our model only permits multiple hydrogen bonds with a shared acceptor, and we observed such bonds during our simulations. We counted the number of acceptors forming different single or multiple hydrogen bonds among the 8 middle amino acids of 16-mer polyalanine. The results are presented in Table II. While acceptors preferably form 310-like hydrogen bonds, we found that the next most common type of hydrogen i⫹3 bond was the double bond i 4 i⫹3 . When it was formed, the amino acid i ⫹ 3 adopted the conformation with ␾ ⫽ ⫺120° and ␺ ⫽ ⫺40°. The increased population in this region is clearly visible on the Ramachandran plots (see Fig. 4). The amino acid i ⫹ 2 often deviated towards ␾ ⫽ ⫺100° and ␺ ⫽ 10°, which is also visible on the Ramachandran plots. In some instances, the amino acid i ⫹ 3 terminated one helix and initiated the other, forming a kink. The axis of the helix at the kink deflects by approximately 45°. The structure of the kink between two 310 helices is presented in Figure 5. We also detected similar kinks between ␣-helices. To our knowledge, this is the first computational observation of the spontaneous formation of such stable kinks. Two previous analyses of Ramachandran plots for

102

A.A. PODTELEZHNIKOV AND D.L. WILD

Fig. 5. Double hydrogen-bonded kink between 310 helices. The 2 hydrogen bonds with a shared acceptor are shown by dotted lines on the stereo pair.

the proteins in the current Protein Data Bank55 (PDB) indicate that these structures are common in real proteins: at the C-cap of the helices (see Fig. 2D in Ho et al.29) and in the intrahelical ␲-bulges (see Figure 2C in Cartailler and Luecke30). While some acceptors in our model form multiple hydrogen bonds, others do not find a matching donor even at very high hydrogen bond strengths. As a result, at H/RT ⫽ 8.0, a significant number (17%) of acceptors remained unbonded, while almost all donors found their match. This is another indication of asymmetry in hydrogen bonding between backbone atoms. Helix–Coil Transition In our simulations, we observed the classical helix– coil transition32 with changes in hydrogen bond strength. Figure 6 shows the decrease in the fraction of the amino acids in the non-hydrogen-bonded conformations with the increase of hydrogen bond strength. Our data demonstrate the cooperative transition from random coil to hydrogen bonded structures between H ⫽ 2.1 kcal/mol to H ⫽ 3.1 kcal/mol. The midpoint of the transition with a 50% fraction of amino acids in a hydrogen-bonded conformation corresponded to H ⫽ 2.6 kcal/mol. Both the midpoint and the width of our simulated transition are in agreement with earlier MC studies.26 In the framework of Zimm–Bragg theory, the helix– coil transition is described in terms of the propagation parameter, s, and the entropic coast of the helix nucleation, ␴.31,32,56 The fraction of random coil residues in this theory is given by 1 1⫺s . ␪⫽ ⫹ 2 2 冑共1 ⫺ s兲 2 ⫹ 4s␴

(14)

The elongation parameter for our model can be expressed as s ⫽ s0exp(H/RT) and determines the midpoint of the helix– coil transition. The parameter ␴ describes the cooperativity level and determines the width of the transition. We were able to obtain an excellent fit of the observed transition by Eq. (14) with 2 adjustable parameters s0 ⫽

Fig. 6. Helix– coil transition under the influence of hydrogen bond strength. The fraction, ␪, of donors (filled symbols) and acceptors (hollow symbols) that do not participate in hydrogen bonding is shown as a function of the hydrogen bond strength H/RT. The fitting procedure is described in the text.

0.013 and ␴ ⫽ 0.3. The reviews of experimentally determined values of s(0°C) range from 1.3 to 2.2 for alaninebased peptides.25,33 In the framework of our model, this corresponds to a hydrogen bond strength H between 2.5 and 2.8 kcal/mol. There is no consensus regarding the experimentally determined values of ␴, which have been reported to range from 0.002 to 0.01.25,33,57 There is even less agreement between simulated values. Our estimate of ␴ agrees with the values determined in earlier simulations; see Nguyen et al.25 and references therein. We have observed that the hydrogen bonding criteria influence the helix– coil transition parameters (data not shown). We therefore believe that it is possible to select the hydrogen bonding criteria that will bring the ␴ values in better agreement with experimental observations. Such calibration of hydrogen bonding criteria will be the subject of our future work. Alternatively, the sigmoid curve for the cooperative helix– coil transition can be characterized by modified Hill equation ␪⫽

1 , K nH ⫹ 1

(15)

where the equilibrium constant of hydrogen bond formation, K, can be expressed as K ⫽ K0exp(H/RT) and the cooperativity parameter nH is the Hill coefficient,32 which can be interpreted as the average number of simultaneously formed hydrogen bonds. We were able to fit our simulation data for hydrogen bond donors by Eq. (15) with 2 adjustable parameters K0 ⫽ 0.013 and nH ⬇ 1.7. This means that, in our model, during the helix– coil transition, approximately 2 hydrogen bonds are formed or broken cooperatively. The fraction of hydrogen bond acceptors

METROPOLIS SAMPLING OF POLYALANINE

(Fig. 6) was fitted in a similar fashion. The quality of fitting by both methods was comparable.

13.

CONCLUSIONS In this work, we used a simplified model of the polypeptide chain featuring only hard-sphere repulsions between the atoms and hydrogen bonding defined by the angular and distance constraints, similar to those used by Pauling et al. more than 50 years ago.20 Using our efficient Metropolis MC procedure, we were able to perform exhaustive sampling of the polyalanine conformations. Our longest computer experiments were carried out in the reasonable time of about 52 h. Our atomistic model permitted the formation of 310 and ␣-helices, kinks between them, and other structural motifs with unusual hydrogen bonding patterns. We simulated helix– coil transition and were able to assess propensities for the different structures under the influence of hydrogen bonding. The secondary structures observed in our simulations were dominated by 310 helices, in agreement with experimental observations.50 It is remarkable that in the framework of this simplified model, our simulated Ramachandran plots capture all the important characteristics of helical conformations. For the first time, to our knowledge, we report the spontaneous formation of stable kinks between the helices, which are common in the crystal structures of real proteins.29,30

14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.

ACKNOWLEDGMENTS We are grateful to John Huelsenbeck for his advice on implementing parallel tempering. We thank the Institute of Pure and Applied Mathematics at UCLA, where part of this work was conducted. REFERENCES 1. McCammon JA, Harvey SC. Dynamics of proteins and nucleic acids. New York: Cambridge University Press; 1987. 234 p. 2. Leach AR. Molecular modelling: principles and applications. New York: Prentice-Hall; 2001. 744 p. 3. Go N, Scheraga HA. Ring closure and local conformational deformations of chain molecules. Macromolecules 1970;3:178 –187. 4. Dodd LR, Boone TD, Theodorou DN. A concerted rotation algorithm for atomistic Monte Carlo simulation of polymer melts and glasses. Mol Phys 1993;78:961–996. 5. Elofsson A, Le Grand SM, Eisenberg D. Local moves: an efficient algorithm for simulation of protein folding. Proteins 1995;23:73– 82. 6. Hoffmann D, Knapp EW. Polypeptide folding with off-lattice Monte Carlo dynamics: the method. Eur Biophys J 1996;24:387– 403. 7. Wedemeyer WJ, Scheraga HA. Exact analytical loop closure in proteins using polynomial equations. J Comput Chem 1999;20: 819 – 844. 8. Wu MHG, Deem MW. Analytical rebridging Monte Carlo: application to cis/trans isomerization in proline-containing, cyclic peptides. J Chem Phys 1999;111:6625– 6632. 9. Mezei M. Efficient Monte Carlo sampling for long molecular chains using local moves, tested on a solvated lipid bilayer. J Chem Phys 2003;118:3874 –3879. 10. Ulmschneider JP, Jorgensen WL. Monte Carlo backbone sampling for polypeptides with variable bond angles and dihedral angles using concerted rotations and a Gaussian bias. J Chem Phys 2003;118:4261– 4271. 11. Coutsias EA, Seok C, Jacobson MP, Dill KA. A kinematic view of loop closure. J Comput Chem 2004;25:510 –528. 12. Klenin KV, Vologodskii AV, Anshelevich VV, Dykhne AM, Frank-

27. 28.

29. 30. 31. 32. 33. 34. 35. 36. 37. 38.

39. 40. 41.

103

Kamenetskii MD. Computer simulation of DNA supercoiling. J Mol Biol 1991;217:413– 419. Podtelezhnikov AA, Mao C, Seeman NC, Vologodskii A. Multimerization– cyclization of DNA fragments as a method of conformational analysis. Biophys J 2000;79:2692–2704. Geyer CJ. Markov chain Monte Carlo maximum likelihood. In: Keramidas EM, editor. Computing science and statistics. New York: American Statistical Association; 1991. p 156 –163. Rose GD, Wolfenden R. Hydrogen bonding, hydrophobicity, packing, and protein folding. Annu Rev Biophys Biomol Struct 1993;22: 381– 415. Ramachandran GN, Ramakrishnan C, Sasisekharan V. Stereochemistry of polypeptide chain configurations. J Mol Biol 1963;7: 95–99. Ramachandran GN, Sasisekharan V. Conformation of polypeptides and proteins. Adv Protein Chem 1968;23:283– 438. Lee B, Richards FM. The interpretation of protein structures: estimation of static accessibility. J Mol Biol 1971;55:379 – 400. Richards FM. Areas, volumes, packing and protein structure. Annu Rev Biophys Bioeng 1977;6:151–176. Pauling L, Corey RB, Branson HR. The structure of proteins: two hydrogen-bonded helical configurations of the polypeptide chain. Proc Natl Acad Sci USA 1951;37:205–211. Pauling L, Corey RB. The pleated sheet, a new layer configuration of polypeptide chains. Proc Natl Acad Sci USA 1951;37:251–256. Chothia C. The nature of the accessible and buried surfaces in proteins. J Mol Biol 1976;105:1–12. Honig B, Nicholls A. Classical electrostatics in biology and chemistry. Science 1995;268:1144 –1149. Mezei M, Fleming PJ, Srinivasan R, Rose GD. Polyproline II helix is the preferred conformation for unfolded polyalanine in water. Proteins 2004;55:502–507. Nguyen HD, Marchut AJ, Hall CK. Solvent effects on the conformational transition of a model polyalanine peptide. Protein Sci 2004;13:2909 –2924. Pappu RV, Srinivasan R, Rose GD. The Flory isolated-pair hypothesis is not valid for polypeptide chains: implications for protein folding. Proc Natl Acad Sci USA 2000;97:12565–12570. Gunasekaran K, Ramakrishnan C, Balaram P. Disallowed Ramachandran conformations of amino acid residues in protein structures. J Mol Biol 1996;264:191–198. Lovell SC, Davis IW, Arendall WB III, de Bakker PI, Word JM, Prisant MG, Richardson JS, Richardson DC. Structure validation by C-alpha geometry: phi,psi and C-beta deviation. Proteins 2003;50:437– 450. Ho BK, Thomas A, Brasseur R. Revisiting the Ramachandran plot: hard-sphere repulsion, electrostatics, and H-bonding in the alpha-helix. Protein Sci 2003;12:2508 –2522. Cartailler JP, Luecke H. Structural and functional characterization of pi bulges and other short intrahelical deformations. Structure (Camb) 2004;12:133–144. Zimm BH, Bragg JK. Theory of the phase transition between helix and random coil in polypeptide chains. J Chem Phys 1959;31:526 – 535. Cantor CR, Schimmel PR. Biophysical chemistry. San Francisco: W. H. Freeman; 1980. Scheraga HA, Vila JA, Ripoll DR. Helix– coil transitions revisited. Biophys Chem 2002;101/102:255–265. Engh RA, Huber R. Accurate bond and angle parameters for X-Ray protein–structure refinement. Acta Crystallogr A 1991;47: 392– 400. Schildt H. C/C⫹⫹ programmer’s reference. New York: McGrawHill/Osborne; 2003. 358 p. Karplus PA. Experimentally observed conformation-dependent geometry and hidden strain in proteins. Protein Sci 1996;5:1406 – 1420. Hopfinger AJ. Conformational properties of macromolecules. New York: Academic Press; 1973. 339 p. Word JM, Lovell SC, LaBean TH, Taylor HC, Zalis ME, Presley BK, Richardson JS, Richardson DC. Visualizing and quantifying molecular goodness-of-fit: small-probe contact dots with explicit hydrogen atoms. J Mol Biol 1999;285:1711–1733. Baker EN, Hubbard RE. Hydrogen bonding in globular proteins. Prog Biophys Mol Biol 1984;44:97–179. Stickle DF, Presta LG, Dill KA, Rose GD. Hydrogen bonding in globular proteins. J Mol Biol 1992;226:1143–1159. Lommerse JPM, Price SL, Taylor R. Hydrogen bonding of car-

104

42. 43.

44.

45. 46. 47. 48. 49.

A.A. PODTELEZHNIKOV AND D.L. WILD

bonyl, ether, and ester oxygen atoms with alkanol hydroxyl groups. J Comput Chem 1997;18:757–774. Fabiola F, Bertram R, Korostelev A, Chapman MS. An improved hydrogen bond potential: impact on medium resolution protein structures. Protein Sci 2002;11:1415–1423. Kortemme T, Morozov AV, Baker D. An orientation-dependent hydrogen bonding potential improves prediction of specificity and structure for proteins and protein–protein complexes. J Mol Biol 2003;326:1239 –1259. Morozov AV, Kortemme T, Tsemekhman K, Baker D. Close agreement between the orientation dependence of hydrogen bonds observed in protein structures and quantum mechanical calculations. Proc Natl Acad Sci USA 2004;101:6946 – 6951. Metropolis N, Rosenbluth AW, Rosenbluth MN, Teller AH, Teller E. Equation of state calculations by fast computing machines. J Chem Phys 1953;21:1087–1092. Sugita Y, Okamoto Y. Replica-exchange molecular dynamics method for protein folding. Chem Phys Lett 1999;314:141–151. Altekar G, Dwarkadas S, Huelsenbeck JP, Ronquist F. Parallel metropolis coupled Markov chain Monte Carlo for Bayesian phylogenetic inference. Bioinformatics 2004;20:407– 415. Kabsch W, Sander C. Dictionary of protein secondary structure: pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 1983;22:2577–2637. Rakhmaninova AB, Mironov AA. [Change in entropy of the free

50. 51.

52. 53. 54. 55.

56. 57.

polypeptide chain during formation of hydrogen bonds]. Mol Biol (Mosk) 2001;35:451– 461. Miick SM, Martinez GV, Fiori WR, Todd AP, Millhauser GL. Short alanine-based peptides may form 3(10)-helices and not alphahelices in aqueous solution. Nature 1992;359:653– 655. Millhauser GL, Stenland CJ, Hanson P, Bolin KA, vandeVen FJM. Estimating the relative populations of 3(10)-helix and alpha-helix in Ala-rich peptides: a hydrogen exchange and high field NMR study. J Mol Biol 1997;267:963–974. Preissner R, Egner U, Saenger W. Occurrence of bifurcated three-center hydrogen bonds in proteins. FEBS Lett 1991;288:192– 196. Fain AV, Berezovskii IN, Chekhov VO, Ukrainskii DL, Esipova NG. [Double and bifurcated hydrogen bonds in alpha-helices of globular proteins]. Biofizika 2001;46:969 –977. Aurora R, Rose GD. Helix capping. Protein Sci 1998;7:21–38. Bernstein FC, Koetzle TF, Williams GJ, Meyer EF Jr, Brice MD, Rodgers JR, Kennard O, Shimanouchi T, Tasumi M. The Protein Data Bank: a computer-based archival file for macromolecular structures. Eur J Biochem 1977;80:319 –324. Grosberg AY, Khokhlov AR. Statistical physics of macromolecules. New York: AIP Press; 1994. 350 p. Holtzer A. Application of old and new values of alpha-helix propensities to the helix-coil transition of poly(L-glutamic acid). J Am Chem Soc 1994;116:10837–10838.

Exhaustive Metropolis Monte Carlo sampling and ...

Jul 27, 2005 - is combined with parallel tempering to exhaustively sample the conformations ... pensable in studying protein structure.1,2 Computer simu- lations continue to shed ..... room temperature replica occurred 68 times in the course.

249KB Sizes 1 Downloads 205 Views

Recommend Documents

lecture 9: monte carlo sampling and integration - GitHub
analysis if needed. • There is also a concept of quasi-random numbers, which attempt to make the Monte Carlo integrals converge faster than N-1/2: e.g. Sobol ... Note that the weights can be >1 and the method is not very useful when the values>>1.

Monte Carlo Simulation
You are going to use simulation elsewhere in the .... If we use Monte Carlo simulation to price a. European ...... Do not put all of your “business logic” in your GUI.

a monte carlo study
Mar 22, 2005 - We confirm this result using simulated data for a wide range of specifications by ...... Federal Reserve Bank of Kansas City and University of Missouri. ... Clements M.P., Krolzig H.$M. (1998), lA Comparison of the Forecast ...

Sequential Monte Carlo multiple testing
Oct 13, 2011 - can be reproduced through a Galaxy Pages document at: ... Then, in Section 3, we show on both simulated and real data that this method can ...

Introduction to Monte Carlo Simulation
Crystal Ball Global Business Unit ... Simulation is the application of models to predict future outcomes ... As an experimenter increases the number of cases to.

Sequential Monte Carlo multiple testing
Oct 13, 2011 - An example of such a local analysis is the study of how the relation ... and then perform a statistical test of a null hypothesis H0 versus. ∗To whom ... resampling risk (Gandy, 2009), and prediction of P-values using. Random ...

Bayes and Big Data: The Consensus Monte Carlo ... - Semantic Scholar
Oct 31, 2013 - posterior distribution based on very large data sets. When the ... and Jordan (2011) extend the bootstrap to distributed data with the “bag of little ...

Bayes and Big Data: The Consensus Monte Carlo ... - Rob McCulloch
Oct 31, 2013 - The number of distinct configurations of xij in each domain is small. ...... within around 11,000 advertisers using a particular Google advertising.

Hamiltonian Monte Carlo for Hierarchical Models
Dec 3, 2013 - eigenvalues, which encode the direction and magnitudes of the local deviation from isotropy. data, latent mean µ set to zero, and a log-normal ...

Introduction to Monte Carlo Simulation - PDFKUL.COM
Monte Carlo Simulation Steps. • Following are the important steps for Monte Carlo simulation: 1. Deterministic model generation. 2. Input distribution identification. 3. Random number generation. 4. Analysis and decision making ..... perform output

Bayes and Big Data: The Consensus Monte Carlo ... - Semantic Scholar
Oct 31, 2013 - and Jordan (2011) extend the bootstrap to distributed data with the “bag of little .... Taken together, equations (4) and (5) suggest the algorithm in Figure 2. Although ...... Propagation algorithms for variational Bayesian learning

Bayes and Big Data: The Consensus Monte Carlo ... - Semantic Scholar
Oct 31, 2013 - the parameters, or reduction of the sufficient statistics, or a long ... The idea behind consensus Monte Carlo is to break the data into groups ...

Statistical Modeling for Monte Carlo Simulation using Hspice - CiteSeerX
To enable Monte Carlo methods, a statistical model is needed. This is a model ..... However, it is difficult to determine the correlation without a lot of statistical data. The best case .... [3] HSPICE Simulation and Analysis User Guide. March 2005.

Sonification of Markov chain Monte Carlo simulations
This paper illustrates the use of sonification as a tool for monitor- ... tional visualization methods to understand the important features of Ф. When. , however ...

Using the Direct Simulation Monte Carlo Approach for ...
The viability of using the Direct Simulation Monte Carlo (DSMC) approach to study the blast-impact ... by computing load definition for two model geometries - a box and an 'I' shaped beam. ... On the other hand, particle methods do not make the conti

A Non-Resampling Sequential Monte Carlo Detector for ... - IEEE Xplore
SMC methods are traditionally built on the techniques of sequential importance sampling (SIS) and resampling. In this paper, we apply the SMC methodology.

Monte Carlo simulations of interacting anyon chains - Semantic Scholar
Apr 8, 2010 - ... Field Laboratory, Florida State University, Tallahassee, FL 32310, USA ..... Mod. Phys. 80 (2008) 1083. [6] G. Rumer, Gцttingen Nachr. Tech.