ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/langevin/langevin.bib
Revision: 2746
Committed: Wed May 3 16:44:46 2006 UTC (18 years, 3 months ago) by tim
File size: 63317 byte(s)
Log Message:
initial draft

File Contents

# User Rev Content
1 tim 2746 This file was created with JabRef 2.0.1.
2     Encoding: GBK
3    
4     @ARTICLE{2003,
5     author = {J. G. {de la Torre} and H. E. Sanchez and A. Ortega and J. G. Hernandez
6     and M. X. Fernandes and F. G. Diaz and M. C. L. Martinez},
7     title = {Calculation of the solution properties of flexible macromolecules:
8     methods and applications},
9     journal = {European Biophysics Journal with Biophysics Letters},
10     year = {2003},
11     volume = {32},
12     pages = {477-486},
13     number = {5},
14     month = {Aug},
15     abstract = {While the prediction of hydrodynamic properties of rigid particles
16     is nowadays feasible using simple and efficient computer programs,
17     the calculation of such properties and, in general, the dynamic
18     behavior of flexible macromolecules has not reached a similar situation.
19     Although the theories are available, usually the computational work
20     is done using solutions specific for each problem. We intend to
21     develop computer programs that would greatly facilitate the task
22     of predicting solution behavior of flexible macromolecules. In this
23     paper, we first present an overview of the two approaches that are
24     most practical: the Monte Carlo rigid-body treatment, and the Brownian
25     dynamics simulation technique. The Monte Carlo procedure is based
26     on the calculation of properties for instantaneous conformations
27     of the macromolecule that are regarded as if they were instantaneously
28     rigid. We describe how a Monte Carlo program can be interfaced to
29     the programs in the HYDRO suite for rigid particles, and provide
30     an example of such calculation, for a hypothetical particle: a protein
31     with two domains connected by a flexible linker. We also describe
32     briefly the essentials of Brownian dynamics, and propose a general
33     mechanical model that includes several kinds of intramolecular interactions,
34     such as bending, internal rotation, excluded volume effects, etc.
35     We provide an example of the application of this methodology to
36     the dynamics of a semiflexible, wormlike DNA.},
37     annote = {724XK Times Cited:6 Cited References Count:64},
38     issn = {0175-7571},
39     uri = {<Go to ISI>://000185513400011},
40     }
41    
42     @ARTICLE{Alakent2005,
43     author = {B. Alakent and M. C. Camurdan and P. Doruker},
44     title = {Hierarchical structure of the energy landscape of proteins revisited
45     by time series analysis. II. Investigation of explicit solvent effects},
46     journal = {Journal of Chemical Physics},
47     year = {2005},
48     volume = {123},
49     pages = {-},
50     number = {14},
51     month = {Oct 8},
52     abstract = {Time series analysis tools are employed on the principal modes obtained
53     from the C-alpha trajectories from two independent molecular-dynamics
54     simulations of alpha-amylase inhibitor (tendamistat). Fluctuations
55     inside an energy minimum (intraminimum motions), transitions between
56     minima (interminimum motions), and relaxations in different hierarchical
57     energy levels are investigated and compared with those encountered
58     in vacuum by using different sampling window sizes and intervals.
59     The low-frequency low-indexed mode relationship, established in
60     vacuum, is also encountered in water, which shows the reliability
61     of the important dynamics information offered by principal components
62     analysis in water. It has been shown that examining a short data
63     collection period (100 ps) may result in a high population of overdamped
64     modes, while some of the low-frequency oscillations (< 10 cm(-1))
65     can be captured in water by using a longer data collection period
66     (1200 ps). Simultaneous analysis of short and long sampling window
67     sizes gives the following picture of the effect of water on protein
68     dynamics. Water makes the protein lose its memory: future conformations
69     are less dependent on previous conformations due to the lowering
70     of energy barriers in hierarchical levels of the energy landscape.
71     In short-time dynamics (< 10 ps), damping factors extracted from
72     time series model parameters are lowered. For tendamistat, the friction
73     coefficient in the Langevin equation is found to be around 40-60
74     cm(-1) for the low-indexed modes, compatible with literature. The
75     fact that water has increased the friction and that on the other
76     hand has lubrication effect at first sight contradicts. However,
77     this comes about because water enhances the transitions between
78     minima and forces the protein to reduce its already inherent inability
79     to maintain oscillations observed in vacuum. Some of the frequencies
80     lower than 10 cm(-1) are found to be overdamped, while those higher
81     than 20 cm(-1) are slightly increased. As for the long-time dynamics
82     in water, it is found that random-walk motion is maintained for
83     approximately 200 ps (about five times of that in vacuum) in the
84     low-indexed modes, showing the lowering of energy barriers between
85     the higher-level minima.},
86     annote = {973OH Times Cited:1 Cited References Count:33},
87     issn = {0021-9606},
88     uri = {<Go to ISI>://000232532000064},
89     }
90    
91     @ARTICLE{Allison1991,
92     author = {S. A. Allison},
93     title = {A Brownian Dynamics Algorithm for Arbitrary Rigid Bodies - Application
94     to Polarized Dynamic Light-Scattering},
95     journal = {Macromolecules},
96     year = {1991},
97     volume = {24},
98     pages = {530-536},
99     number = {2},
100     month = {Jan 21},
101     abstract = {A Brownian dynamics algorithm is developed to simulate dynamics experiments
102     of rigid macromolecules. It is applied to polarized dynamic light
103     scattering from rodlike sturctures and from a model of a DNA fragment
104     (762 base pairs). A number of rod cases are examined in which the
105     translational anisotropy is increased form zero to a large value.
106     Simulated first cumulants as well as amplitudes and lifetimes of
107     the dynamic form factor are compared with predictions of analytic
108     theories and found to be in very good agreement with them. For DNA
109     fragments 762 base pairs in length or longer, translational anisotropy
110     does not contribute significantly to dynamic light scattering. In
111     a comparison of rigid and flexible simulations on semistiff models
112     of this fragment, it is shown directly that flexing contributes
113     to the faster decay processes probed by light scattering and that
114     the flexible model studies are in good agreement with experiment.},
115     annote = {Eu814 Times Cited:8 Cited References Count:32},
116     issn = {0024-9297},
117     uri = {<Go to ISI>://A1991EU81400029},
118     }
119    
120     @ARTICLE{Auerbach2005,
121     author = {A. Auerbach},
122     title = {Gating of acetylcholine receptor channels: Brownian motion across
123     a broad transition state},
124     journal = {Proceedings of the National Academy of Sciences of the United States
125     of America},
126     year = {2005},
127     volume = {102},
128     pages = {1408-1412},
129     number = {5},
130     month = {Feb 1},
131     abstract = {Acetylcholine receptor channels (AChRs) are proteins that switch between
132     stable #closed# and #open# conformations. In patch clamp recordings,
133     diliganded AChR gating appears to be a simple, two-state reaction.
134     However, mutagenesis studies indicate that during gating dozens
135     of residues across the protein move asynchronously and are organized
136     into rigid body gating domains (#blocks#). Moreover, there is an
137     upper limit to the apparent channel opening rate constant. These
138     observations suggest that the gating reaction has a broad, corrugated
139     transition state region, with the maximum opening rate reflecting,
140     in part, the mean first-passage time across this ensemble. Simulations
141     reveal that a flat, isotropic energy profile for the transition
142     state can account for many of the essential features of AChR gating.
143     With this mechanism, concerted, local structural transitions that
144     occur on the broad transition state ensemble give rise to fractional
145     measures of reaction progress (Phi values) determined by rate-equilibrium
146     free energy relationship analysis. The results suggest that the
147     coarse-grained AChR gating conformational change propagates through
148     the protein with dynamics that are governed by the Brownian motion
149     of individual gating blocks.},
150     annote = {895QF Times Cited:9 Cited References Count:33},
151     issn = {0027-8424},
152     uri = {<Go to ISI>://000226877300030},
153     }
154    
155     @ARTICLE{Banerjee2004,
156     author = {D. Banerjee and B. C. Bag and S. K. Banik and D. S. Ray},
157     title = {Solution of quantum Langevin equation: Approximations, theoretical
158     and numerical aspects},
159     journal = {Journal of Chemical Physics},
160     year = {2004},
161     volume = {120},
162     pages = {8960-8972},
163     number = {19},
164     month = {May 15},
165     abstract = {Based on a coherent state representation of noise operator and an
166     ensemble averaging procedure using Wigner canonical thermal distribution
167     for harmonic oscillators, a generalized quantum Langevin equation
168     has been recently developed [Phys. Rev. E 65, 021109 (2002); 66,
169     051106 (2002)] to derive the equations of motion for probability
170     distribution functions in c-number phase-space. We extend the treatment
171     to explore several systematic approximation schemes for the solutions
172     of the Langevin equation for nonlinear potentials for a wide range
173     of noise correlation, strength and temperature down to the vacuum
174     limit. The method is exemplified by an analytic application to harmonic
175     oscillator for arbitrary memory kernel and with the help of a numerical
176     calculation of barrier crossing, in a cubic potential to demonstrate
177     the quantum Kramers' turnover and the quantum Arrhenius plot. (C)
178     2004 American Institute of Physics.},
179     annote = {816YY Times Cited:8 Cited References Count:35},
180     issn = {0021-9606},
181     uri = {<Go to ISI>://000221146400009},
182     }
183    
184     @ARTICLE{Barth1998,
185     author = {E. Barth and T. Schlick},
186     title = {Overcoming stability limitations in biomolecular dynamics. I. Combining
187     force splitting via extrapolation with Langevin dynamics in LN},
188     journal = {Journal of Chemical Physics},
189     year = {1998},
190     volume = {109},
191     pages = {1617-1632},
192     number = {5},
193     month = {Aug 1},
194     abstract = {We present an efficient new method termed LN for propagating biomolecular
195     dynamics according to the Langevin equation that arose fortuitously
196     upon analysis of the range of harmonic validity of our normal-mode
197     scheme LIN. LN combines force linearization with force splitting
198     techniques and disposes of LIN'S computationally intensive minimization
199     (anharmonic correction) component. Unlike the competitive multiple-timestepping
200     (MTS) schemes today-formulated to be symplectic and time-reversible-LN
201     merges the slow and fast forces via extrapolation rather than impulses;
202     the Langevin heat bath prevents systematic energy drifts. This combination
203     succeeds in achieving more significant speedups than these MTS methods
204     which are Limited by resonance artifacts to an outer timestep less
205     than some integer multiple of half the period of the fastest motion
206     (around 4-5 fs for biomolecules). We show that LN achieves very
207     good agreement with small-timestep solutions of the Langevin equation
208     in terms of thermodynamics (energy means and variances), geometry,
209     and dynamics (spectral densities) for two proteins in vacuum and
210     a large water system. Significantly, the frequency of updating the
211     slow forces extends to 48 fs or more, resulting in speedup factors
212     exceeding 10. The implementation of LN in any program that employs
213     force-splitting computations is straightforward, with only partial
214     second-derivative information required, as well as sparse Hessian/vector
215     multiplication routines. The linearization part of LN could even
216     be replaced by direct evaluation of the fast components. The application
217     of LN to biomolecular dynamics is well suited for configurational
218     sampling, thermodynamic, and structural questions. (C) 1998 American
219     Institute of Physics.},
220     annote = {105HH Times Cited:29 Cited References Count:49},
221     issn = {0021-9606},
222     uri = {<Go to ISI>://000075066300006},
223     }
224    
225     @ARTICLE{Batcho2001,
226     author = {P. F. Batcho and T. Schlick},
227     title = {Special stability advantages of position-Verlet over velocity-Verlet
228     in multiple-time step integration},
229     journal = {Journal of Chemical Physics},
230     year = {2001},
231     volume = {115},
232     pages = {4019-4029},
233     number = {9},
234     month = {Sep 1},
235     abstract = {We present an analysis for a simple two-component harmonic oscillator
236     that compares the use of position-Verlet to velocity-Verlet for
237     multiple-time step integration. The numerical stability analysis
238     based on the impulse-Verlet splitting shows that position-Verlet
239     has enhanced stability, in terms of the largest allowable time step,
240     for cases where an ample separation of time scales exists. Numerical
241     investigations confirm the advantages of the position-Verlet scheme
242     when used for the fastest time scales of the system. Applications
243     to a biomolecule. a solvated protein, for both Newtonian and Langevin
244     dynamics echo these trends over large outer time-step regimes. (C)
245     2001 American Institute of Physics.},
246     annote = {469KV Times Cited:6 Cited References Count:30},
247     issn = {0021-9606},
248     uri = {<Go to ISI>://000170813800005},
249     }
250    
251     @ARTICLE{Beard2003,
252     author = {D. A. Beard and T. Schlick},
253     title = {Unbiased rotational moves for rigid-body dynamics},
254     journal = {Biophysical Journal},
255     year = {2003},
256     volume = {85},
257     pages = {2973-2976},
258     number = {5},
259     month = {Nov 1},
260     abstract = {We introduce an unbiased protocol for performing rotational moves
261     in rigid-body dynamics simulations. This approach - based on the
262     analytic solution for the rotational equations of motion for an
263     orthogonal coordinate system at constant angular velocity - removes
264     deficiencies that have been largely ignored in Brownian dynamics
265     simulations, namely errors for finite rotations that result from
266     applying the noncommuting rotational matrices in an arbitrary order.
267     Our algorithm should thus replace standard approaches to rotate
268     local coordinate frames in Langevin and Brownian dynamics simulations.},
269     annote = {736UA Times Cited:0 Cited References Count:11},
270     issn = {0006-3495},
271     uri = {<Go to ISI>://000186190500018},
272     }
273    
274     @ARTICLE{Beloborodov1998,
275     author = {I. S. Beloborodov and V. Y. Orekhov and A. S. Arseniev},
276     title = {Effect of coupling between rotational and translational Brownian
277     motions on NMR spin relaxation: Consideration using green function
278     of rigid body diffusion},
279     journal = {Journal of Magnetic Resonance},
280     year = {1998},
281     volume = {132},
282     pages = {328-329},
283     number = {2},
284     month = {Jun},
285     abstract = {Using the Green function of arbitrary rigid Brownian diffusion (Goldstein,
286     Biopolymers 33, 409-436, 1993), it was analytically shown that coupling
287     between translation and rotation diffusion degrees of freedom does
288     not affect the correlation functions relevant to the NMR intramolecular
289     relaxation. It follows that spectral densities usually used for
290     the anisotropic rotation diffusion (Woessner, J. Chem. Phys. 37,
291     647-654, 1962) can be regarded as exact in respect to the rotation-translation
292     coupling for the spin system connected with a rigid body. (C) 1998
293     Academic Press.},
294     annote = {Zu605 Times Cited:2 Cited References Count:6},
295     issn = {1090-7807},
296     uri = {<Go to ISI>://000074214800017},
297     }
298    
299     @ARTICLE{Berkov2005,
300     author = {D. V. Berkov and N. L. Gorn},
301     title = {Stochastic dynamic simulations of fast remagnetization processes:
302     recent advances and applications},
303     journal = {Journal of Magnetism and Magnetic Materials},
304     year = {2005},
305     volume = {290},
306     pages = {442-448},
307     month = {Apr},
308     abstract = {Numerical simulations of fast remagnetization processes using stochastic
309     dynamics are widely used to study various magnetic systems. In this
310     paper, we first address several crucial methodological problems
311     of such simulations: (i) the influence of finite-element discretization
312     on simulated dynamics, (ii) choice between Ito and Stratonovich
313     stochastic calculi by the solution of micromagnetic stochastic equations
314     of motion and (iii) non-trivial correlation properties of the random
315     (thermal) field. Next, we discuss several examples to demonstrate
316     the great potential of the Langevin dynamics for studying fast remagnetization
317     processes in technically relevant applications: we present numerical
318     analysis of equilibrium magnon spectra in patterned structures,
319     study thermal noise effects on the magnetization dynamics of nanoelements
320     in pulsed fields and show some results for a remagnetization dynamics
321     induced by a spin-polarized current. (c) 2004 Elsevier B.V. All
322     rights reserved.},
323     annote = {Part 1 Sp. Iss. SI 922KU Times Cited:2 Cited References Count:25},
324     issn = {0304-8853},
325     uri = {<Go to ISI>://000228837600109},
326     }
327    
328     @ARTICLE{Berkov2005a,
329     author = {D. V. Berkov and N. L. Gorn},
330     title = {Magnetization precession due to a spin-polarized current in a thin
331     nanoelement: Numerical simulation study},
332     journal = {Physical Review B},
333     year = {2005},
334     volume = {72},
335     pages = {-},
336     number = {9},
337     month = {Sep},
338     abstract = {In this paper a detailed numerical study (in frames of the Slonczewski
339     formalism) of magnetization oscillations driven by a spin-polarized
340     current through a thin elliptical nanoelement is presented. We show
341     that a sophisticated micromagnetic model, where a polycrystalline
342     structure of a nanoelement is taken into account, can explain qualitatively
343     all most important features of the magnetization oscillation spectra
344     recently observed experimentally [S. I. Kiselev , Nature 425, 380
345     (2003)], namely, existence of several equidistant spectral bands,
346     sharp onset and abrupt disappearance of magnetization oscillations
347     with increasing current, absence of the out-of-plane regime predicted
348     by a macrospin model, and the relation between frequencies of so-called
349     small-angle and quasichaotic oscillations. However, a quantitative
350     agreement with experimental results (especially concerning the frequency
351     of quasichaotic oscillations) could not be achieved in the region
352     of reasonable parameter values, indicating that further model refinement
353     is necessary for a complete understanding of the spin-driven magnetization
354     precession even in this relatively simple experimental situation.},
355     annote = {969IT Times Cited:2 Cited References Count:55},
356     issn = {1098-0121},
357     uri = {<Go to ISI>://000232228500058},
358     }
359    
360     @ARTICLE{Berkov2002,
361     author = {D. V. Berkov and N. L. Gorn and P. Gornert},
362     title = {Magnetization dynamics in nanoparticle systems: Numerical simulation
363     using Langevin dynamics},
364     journal = {Physica Status Solidi a-Applied Research},
365     year = {2002},
366     volume = {189},
367     pages = {409-421},
368     number = {2},
369     month = {Feb 16},
370     abstract = {We report on recent progress achieved by the development of numerical
371     methods based on the stochastic (Langevin) dynamics applied to systems
372     of interacting magnetic nanoparticles. The method enables direct
373     simulations of the trajectories of magnetic moments taking into
374     account (i) all relevant interactions, (ii) precession dynamics,
375     and (iii) temperature fluctuations included via the random (thermal)
376     field. We present several novel results obtained using new methods
377     developed for the solution of the Langevin equations. In particular,
378     we have investigated magnetic nanodots and disordered granular systems
379     of single-domain magnetic particles. For the first case we have
380     calculated the spectrum and the spatial distribution of spin excitations.
381     For the second system the complex ac susceptibility chi(omega, T)
382     for various particle concentrations and particle anisotropies were
383     computed and compared with numerous experimental results.},
384     annote = {526TF Times Cited:4 Cited References Count:37},
385     issn = {0031-8965},
386     uri = {<Go to ISI>://000174145200026},
387     }
388    
389     @ARTICLE{Bernal1980,
390     author = {J.M. Bernal and J. G. {de la Torre}},
391     title = {Transport Properties and Hydrodynamic Centers of Rigid Macromolecules
392     with Arbitrary Shape},
393     journal = {Biopolymers},
394     year = {1980},
395     volume = {19},
396     pages = {751-766},
397     }
398    
399     @ARTICLE{Brunger1984,
400     author = {A. Brunger and C. L. Brooks and M. Karplus},
401     title = {Stochastic Boundary-Conditions for Molecular-Dynamics Simulations
402     of St2 Water},
403     journal = {Chemical Physics Letters},
404     year = {1984},
405     volume = {105},
406     pages = {495-500},
407     number = {5},
408     annote = {Sm173 Times Cited:143 Cited References Count:22},
409     issn = {0009-2614},
410     uri = {<Go to ISI>://A1984SM17300007},
411     }
412    
413     @ARTICLE{Chin2004,
414     author = {S. A. Chin},
415     title = {Dynamical multiple-time stepping methods for overcoming resonance
416     instabilities},
417     journal = {Journal of Chemical Physics},
418     year = {2004},
419     volume = {120},
420     pages = {8-13},
421     number = {1},
422     month = {Jan 1},
423     abstract = {Current molecular dynamics simulations of biomolecules using multiple
424     time steps to update the slowly changing force are hampered by instabilities
425     beginning at time steps near the half period of the fastest vibrating
426     mode. These #resonance# instabilities have became a critical barrier
427     preventing the long time simulation of biomolecular dynamics. Attempts
428     to tame these instabilities by altering the slowly changing force
429     and efforts to damp them out by Langevin dynamics do not address
430     the fundamental cause of these instabilities. In this work, we trace
431     the instability to the nonanalytic character of the underlying spectrum
432     and show that a correct splitting of the Hamiltonian, which renders
433     the spectrum analytic, restores stability. The resulting Hamiltonian
434     dictates that in addition to updating the momentum due to the slowly
435     changing force, one must also update the position with a modified
436     mass. Thus multiple-time stepping must be done dynamically. (C)
437     2004 American Institute of Physics.},
438     annote = {757TK Times Cited:1 Cited References Count:22},
439     issn = {0021-9606},
440     uri = {<Go to ISI>://000187577400003},
441     }
442    
443     @ARTICLE{Cui2003,
444     author = {B. X. Cui and M. Y. Shen and K. F. Freed},
445     title = {Folding and misfolding of the papillomavirus E6 interacting peptide
446     E6ap},
447     journal = {Proceedings of the National Academy of Sciences of the United States
448     of America},
449     year = {2003},
450     volume = {100},
451     pages = {7087-7092},
452     number = {12},
453     month = {Jun 10},
454     abstract = {All-atom Langevin dynamics simulations have been performed to study
455     the folding pathways of the 18-residue binding domain fragment E6ap
456     of the human papillomavirus E6 interacting peptide. Six independent
457     folding trajectories, with a total duration of nearly 2 mus, all
458     lead to the same native state in which the E6ap adopts a fluctuating
459     a-helix structure in the central portion (Ser-4-Leu-13) but with
460     very flexible N and C termini. Simulations starting from different
461     core configurations exhibit the E6ap folding dynamics as either
462     a two- or three-state folder with an intermediate misfolded state.
463     The essential leucine hydrophobic core (Leu-9, Leu-12, and Leu-13)
464     is well conserved in the native-state structure but absent in the
465     intermediate structure, suggesting that the leucine core is not
466     only essential for the binding activity of E6ap but also important
467     for the stability of the native structure. The free energy landscape
468     reveals a significant barrier between the basins separating the
469     native and misfolded states. We also discuss the various underlying
470     forces that drive the peptide into its native state.},
471     annote = {689LC Times Cited:3 Cited References Count:48},
472     issn = {0027-8424},
473     uri = {<Go to ISI>://000183493500037},
474     }
475    
476     @ARTICLE{Denisov2003,
477     author = {S. I. Denisov and T. V. Lyutyy and K. N. Trohidou},
478     title = {Magnetic relaxation in finite two-dimensional nanoparticle ensembles},
479     journal = {Physical Review B},
480     year = {2003},
481     volume = {67},
482     pages = {-},
483     number = {1},
484     month = {Jan 1},
485     abstract = {We study the slow phase of thermally activated magnetic relaxation
486     in finite two-dimensional ensembles of dipolar interacting ferromagnetic
487     nanoparticles whose easy axes of magnetization are perpendicular
488     to the distribution plane. We develop a method to numerically simulate
489     the magnetic relaxation for the case that the smallest heights of
490     the potential barriers between the equilibrium directions of the
491     nanoparticle magnetic moments are much larger than the thermal energy.
492     Within this framework, we analyze in detail the role that the correlations
493     of the nanoparticle magnetic moments and the finite size of the
494     nanoparticle ensemble play in magnetic relaxation.},
495     annote = {642XH Times Cited:11 Cited References Count:31},
496     issn = {1098-0121},
497     uri = {<Go to ISI>://000180830400056},
498     }
499    
500     @ARTICLE{Derreumaux1998,
501     author = {P. Derreumaux and T. Schlick},
502     title = {The loop opening/closing motion of the enzyme triosephosphate isomerase},
503     journal = {Biophysical Journal},
504     year = {1998},
505     volume = {74},
506     pages = {72-81},
507     number = {1},
508     month = {Jan},
509     abstract = {To explore the origin of the large-scale motion of triosephosphate
510     isomerase's flexible loop (residues 166 to 176) at the active site,
511     several simulation protocols are employed both for the free enzyme
512     in vacuo and for the free enzyme with some solvent modeling: high-temperature
513     Langevin dynamics simulations, sampling by a #dynamics##driver#
514     approach, and potential-energy surface calculations. Our focus is
515     on obtaining the energy barrier to the enzyme's motion and establishing
516     the nature of the loop movement. Previous calculations did not determine
517     this energy barrier and the effect of solvent on the barrier. High-temperature
518     molecular dynamics simulations and crystallographic studies have
519     suggested a rigid-body motion with two hinges located at both ends
520     of the loop; Brownian dynamics simulations at room temperature pointed
521     to a very flexible behavior. The present simulations and analyses
522     reveal that although solute/solvent hydrogen bonds play a crucial
523     role in lowering the energy along the pathway, there still remains
524     a high activation barrier, This finding clearly indicates that,
525     if the loop opens and closes in the absence of a substrate at standard
526     conditions (e.g., room temperature, appropriate concentration of
527     isomerase), the time scale for transition is not in the nanosecond
528     but rather the microsecond range. Our results also indicate that
529     in the context of spontaneous opening in the free enzyme, the motion
530     is of rigid-body type and that the specific interaction between
531     residues Ala(176) and Tyr(208) plays a crucial role in the loop
532     opening/closing mechanism.},
533     annote = {Zl046 Times Cited:30 Cited References Count:29},
534     issn = {0006-3495},
535     uri = {<Go to ISI>://000073393400009},
536     }
537    
538     @ARTICLE{Edwards2005,
539     author = {S. A. Edwards and D. R. M. Williams},
540     title = {Stretching a single diblock copolymer in a selective solvent: Langevin
541     dynamics simulations},
542     journal = {Macromolecules},
543     year = {2005},
544     volume = {38},
545     pages = {10590-10595},
546     number = {25},
547     month = {Dec 13},
548     abstract = {Using the Langevin dynamics technique, we have carried out simulations
549     of a single-chain flexible diblock copolymer. The polymer consists
550     of two blocks of equal length, one very poorly solvated and the
551     other close to theta-conditions. We study what happens when such
552     a polymer is stretched, for a range of different stretching speeds,
553     and correlate our observations with features in the plot of force
554     vs extension. We find that at slow speeds this force profile does
555     not increase monotonically, in disagreement with earlier predictions,
556     and that at high speeds there is a strong dependence on which end
557     of the polymer is pulled, as well as a high level of hysteresis.},
558     annote = {992EC Times Cited:0 Cited References Count:13},
559     issn = {0024-9297},
560     uri = {<Go to ISI>://000233866200035},
561     }
562    
563     @ARTICLE{Ermak1978,
564     author = {D. L. Ermak and J. A. Mccammon},
565     title = {Brownian Dynamics with Hydrodynamic Interactions},
566     journal = {Journal of Chemical Physics},
567     year = {1978},
568     volume = {69},
569     pages = {1352-1360},
570     number = {4},
571     annote = {Fp216 Times Cited:785 Cited References Count:42},
572     issn = {0021-9606},
573     uri = {<Go to ISI>://A1978FP21600004},
574     }
575    
576     @ARTICLE{Fernandes2002,
577     author = {M. X. Fernandes and J. G. {de la Torre}},
578     title = {Brownian dynamics simulation of rigid particles of arbitrary shape
579     in external fields},
580     journal = {Biophysical Journal},
581     year = {2002},
582     volume = {83},
583     pages = {3039-3048},
584     number = {6},
585     month = {Dec},
586     abstract = {We have developed a Brownian dynamics simulation algorithm to generate
587     Brownian trajectories of an isolated, rigid particle of arbitrary
588     shape in the presence of electric fields or any other external agents.
589     Starting from the generalized diffusion tensor, which can be calculated
590     with the existing HYDRO software, the new program BROWNRIG (including
591     a case-specific subprogram for the external agent) carries out a
592     simulation that is analyzed later to extract the observable dynamic
593     properties. We provide a variety of examples of utilization of this
594     method, which serve as tests of its performance, and also illustrate
595     its applicability. Examples include free diffusion, transport in
596     an electric field, and diffusion in a restricting environment.},
597     annote = {633AD Times Cited:2 Cited References Count:43},
598     issn = {0006-3495},
599     uri = {<Go to ISI>://000180256300012},
600     }
601    
602     @ARTICLE{Gelin1999,
603     author = {M. F. Gelin},
604     title = {Inertial effects in the Brownian dynamics with rigid constraints},
605     journal = {Macromolecular Theory and Simulations},
606     year = {1999},
607     volume = {8},
608     pages = {529-543},
609     number = {6},
610     month = {Nov},
611     abstract = {To investigate the influence of inertial effects on the dynamics of
612     an assembly of beads subjected to rigid constraints and placed in
613     a buffer medium, a convenient method to introduce suitable generalized
614     coordinates is presented. Without any restriction on the nature
615     of the soft forces involved (both stochastic and deterministic),
616     pertinent Langevin equations are derived. Provided that the Brownian
617     forces are Gaussian and Markovian, the corresponding Fokker-Planck
618     equation (FPE) is obtained in the complete phase space of generalized
619     coordinates and momenta. The correct short time behavior for correlation
620     functions (CFs) of generalized coordinates is established, and the
621     diffusion equation with memory (DEM) is deduced from the FPE in
622     the high friction Limit. The DEM is invoked to perform illustrative
623     calculations in two dimensions of the orientational CFs for once
624     broken nonrigid rods immobilized on a surface. These calculations
625     reveal that the CFs under certain conditions exhibit an oscillatory
626     behavior, which is irreproducible within the standard diffusion
627     equation. Several methods are considered for the approximate solution
628     of the DEM, and their application to three dimensional DEMs is discussed.},
629     annote = {257MM Times Cited:2 Cited References Count:82},
630     issn = {1022-1344},
631     uri = {<Go to ISI>://000083785700002},
632     }
633    
634     @ARTICLE{Gray2003,
635     author = {J. J. Gray and S. Moughon and C. Wang and O. Schueler-Furman and
636     B. Kuhlman and C. A. Rohl and D. Baker},
637     title = {Protein-protein docking with simultaneous optimization of rigid-body
638     displacement and side-chain conformations},
639     journal = {Journal of Molecular Biology},
640     year = {2003},
641     volume = {331},
642     pages = {281-299},
643     number = {1},
644     month = {Aug 1},
645     abstract = {Protein-protein docking algorithms provide a means to elucidate structural
646     details for presently unknown complexes. Here, we present and evaluate
647     a new method to predict protein-protein complexes from the coordinates
648     of the unbound monomer components. The method employs a low-resolution,
649     rigid-body, Monte Carlo search followed by simultaneous optimization
650     of backbone displacement and side-chain conformations using Monte
651     Carlo minimization. Up to 10(5) independent simulations are carried
652     out, and the resulting #decoys# are ranked using an energy function
653     dominated by van der Waals interactions, an implicit solvation model,
654     and an orientation-dependent hydrogen bonding potential. Top-ranking
655     decoys are clustered to select the final predictions. Small-perturbation
656     studies reveal the formation of binding funnels in 42 of 54 cases
657     using coordinates derived from the bound complexes and in 32 of
658     54 cases using independently determined coordinates of one or both
659     monomers. Experimental binding affinities correlate with the calculated
660     score function and explain the predictive success or failure of
661     many targets. Global searches using one or both unbound components
662     predict at least 25% of the native residue-residue contacts in 28
663     of the 32 cases where binding funnels exist. The results suggest
664     that the method may soon be useful for generating models of biologically
665     important complexes from the structures of the isolated components,
666     but they also highlight the challenges that must be met to achieve
667     consistent and accurate prediction of protein-protein interactions.
668     (C) 2003 Elsevier Ltd. All rights reserved.},
669     annote = {704QL Times Cited:48 Cited References Count:60},
670     issn = {0022-2836},
671     uri = {<Go to ISI>://000184351300022},
672     }
673    
674     @ARTICLE{Hao1993,
675     author = {M. H. Hao and M. R. Pincus and S. Rackovsky and H. A. Scheraga},
676     title = {Unfolding and Refolding of the Native Structure of Bovine Pancreatic
677     Trypsin-Inhibitor Studied by Computer-Simulations},
678     journal = {Biochemistry},
679     year = {1993},
680     volume = {32},
681     pages = {9614-9631},
682     number = {37},
683     month = {Sep 21},
684     abstract = {A new procedure for studying the folding and unfolding of proteins,
685     with an application to bovine pancreatic trypsin inhibitor (BPTI),
686     is reported. The unfolding and refolding of the native structure
687     of the protein are characterized by the dimensions of the protein,
688     expressed in terms of the three principal radii of the structure
689     considered as an ellipsoid. A dynamic equation, describing the variations
690     of the principal radii on the unfolding path, and a numerical procedure
691     to solve this equation are proposed. Expanded and distorted conformations
692     are refolded to the native structure by a dimensional-constraint
693     energy minimization procedure. A unique and reproducible unfolding
694     pathway for an intermediate of BPTI lacking the [30,51] disulfide
695     bond is obtained. The resulting unfolded conformations are extended;
696     they contain near-native local structure, but their longest principal
697     radii are more than 2.5 times greater than that of the native structure.
698     The most interesting finding is that the majority of expanded conformations,
699     generated under various conditions, can be refolded closely to the
700     native structure, as measured by the correct overall chain fold,
701     by the rms deviations from the native structure of only 1.9-3.1
702     angstrom, and by the energy differences of about 10 kcal/mol from
703     the native structure. Introduction of the [30,51] disulfide bond
704     at this stage, followed by minimization, improves the closeness
705     of the refolded structures to the native structure, reducing the
706     rms deviations to 0.9-2.0 angstrom. The unique refolding of these
707     expanded structures over such a large conformational space implies
708     that the folding is strongly dictated by the interactions in the
709     amino acid sequence of BPTI. The simulations indicate that, under
710     conditions that favor a compact structure as mimicked by the volume
711     constraints in our algorithm; the expanded conformations have a
712     strong tendency to move toward the native structure; therefore,
713     they probably would be favorable folding intermediates. The results
714     presented here support a general model for protein folding, i.e.,
715     progressive formation of partially folded structural units, followed
716     by collapse to the compact native structure. The general applicability
717     of the procedure is also discussed.},
718     annote = {Ly294 Times Cited:27 Cited References Count:57},
719     issn = {0006-2960},
720     uri = {<Go to ISI>://A1993LY29400014},
721     }
722    
723     @ARTICLE{Hinsen2000,
724     author = {K. Hinsen and A. J. Petrescu and S. Dellerue and M. C. Bellissent-Funel
725     and G. R. Kneller},
726     title = {Harmonicity in slow protein dynamics},
727     journal = {Chemical Physics},
728     year = {2000},
729     volume = {261},
730     pages = {25-37},
731     number = {1-2},
732     month = {Nov 1},
733     abstract = {The slow dynamics of proteins around its native folded state is usually
734     described by diffusion in a strongly anharmonic potential. In this
735     paper, we try to understand the form and origin of the anharmonicities,
736     with the principal aim of gaining a better understanding of the
737     principal motion types, but also in order to develop more efficient
738     numerical methods for simulating neutron scattering spectra of large
739     proteins. First, we decompose a molecular dynamics (MD) trajectory
740     of 1.5 ns for a C-phycocyanin dimer surrounded by a layer of water
741     into three contributions that we expect to be independent: the global
742     motion of the residues, the rigid-body motion of the sidechains
743     relative to the backbone, and the internal deformations of the sidechains.
744     We show that they are indeed almost independent by verifying the
745     factorization of the incoherent intermediate scattering function.
746     Then, we show that the global residue motions, which include all
747     large-scale backbone motions, can be reproduced by a simple harmonic
748     model which contains two contributions: a short-time vibrational
749     term, described by a standard normal mode calculation in a local
750     minimum, and a long-time diffusive term, described by Brownian motion
751     in an effective harmonic potential. The potential and the friction
752     constants were fitted to the MD data. The major anharmonic contribution
753     to the incoherent intermediate scattering function comes from the
754     rigid-body diffusion of the sidechains. This model can be used to
755     calculate scattering functions for large proteins and for long-time
756     scales very efficiently, and thus provides a useful complement to
757     MD simulations, which are best suited for detailed studies on smaller
758     systems or for shorter time scales. (C) 2000 Elsevier Science B.V.
759     All rights reserved.},
760     annote = {Sp. Iss. SI 368MT Times Cited:16 Cited References Count:31},
761     issn = {0301-0104},
762     uri = {<Go to ISI>://000090121700003},
763     }
764    
765     @ARTICLE{Izaguirre2001,
766     author = {J. A. Izaguirre and D. P. Catarello and J. M. Wozniak and R. D. Skeel},
767     title = {Langevin stabilization of molecular dynamics},
768     journal = {Journal of Chemical Physics},
769     year = {2001},
770     volume = {114},
771     pages = {2090-2098},
772     number = {5},
773     month = {Feb 1},
774     abstract = {In this paper we show the possibility of using very mild stochastic
775     damping to stabilize long time step integrators for Newtonian molecular
776     dynamics. More specifically, stable and accurate integrations are
777     obtained for damping coefficients that are only a few percent of
778     the natural decay rate of processes of interest, such as the velocity
779     autocorrelation function. Two new multiple time stepping integrators,
780     Langevin Molly (LM) and Brunger-Brooks-Karplus-Molly (BBK-M), are
781     introduced in this paper. Both use the mollified impulse method
782     for the Newtonian term. LM uses a discretization of the Langevin
783     equation that is exact for the constant force, and BBK-M uses the
784     popular Brunger-Brooks-Karplus integrator (BBK). These integrators,
785     along with an extrapolative method called LN, are evaluated across
786     a wide range of damping coefficient values. When large damping coefficients
787     are used, as one would for the implicit modeling of solvent molecules,
788     the method LN is superior, with LM closely following. However, with
789     mild damping of 0.2 ps(-1), LM produces the best results, allowing
790     long time steps of 14 fs in simulations containing explicitly modeled
791     flexible water. With BBK-M and the same damping coefficient, time
792     steps of 12 fs are possible for the same system. Similar results
793     are obtained for a solvated protein-DNA simulation of estrogen receptor
794     ER with estrogen response element ERE. A parallel version of BBK-M
795     runs nearly three times faster than the Verlet-I/r-RESPA (reversible
796     reference system propagator algorithm) when using the largest stable
797     time step on each one, and it also parallelizes well. The computation
798     of diffusion coefficients for flexible water and ER/ERE shows that
799     when mild damping of up to 0.2 ps-1 is used the dynamics are not
800     significantly distorted. (C) 2001 American Institute of Physics.},
801     annote = {397CQ Times Cited:14 Cited References Count:36},
802     issn = {0021-9606},
803     uri = {<Go to ISI>://000166676100020},
804     }
805    
806     @ARTICLE{Klimov1997,
807     author = {D. K. Klimov and D. Thirumalai},
808     title = {Viscosity dependence of the folding rates of proteins},
809     journal = {Physical Review Letters},
810     year = {1997},
811     volume = {79},
812     pages = {317-320},
813     number = {2},
814     month = {Jul 14},
815     abstract = {The viscosity (eta) dependence of the folding rates for four sequences
816     (the native state of three sequences is a beta sheet, while the
817     fourth forms an alpha helix) is calculated for off-lattice models
818     of proteins. Assuming that the dynamics is given by the Langevin
819     equation, we show that the folding rates increase linearly at low
820     viscosities eta, decrease as 1/eta at large eta, and have a maximum
821     at intermediate values. The Kramers' theory of barrier crossing
822     provides a quantitative fit of the numerical results. By mapping
823     the simulation results to real proteins we estimate that for optimized
824     sequences the time scale for forming a four turn alpha-helix topology
825     is about 500 ns, whereas for beta sheet it is about 10 mu s.},
826     annote = {Xk293 Times Cited:77 Cited References Count:17},
827     issn = {0031-9007},
828     uri = {<Go to ISI>://A1997XK29300035},
829     }
830    
831     @ARTICLE{Liwo2005,
832     author = {A. Liwo and M. Khalili and H. A. Scheraga},
833     title = {Ab initio simulations of protein folding pathways by molecular dynamics
834     with the united-residue (UNRES) model of polypeptide chains},
835     journal = {Febs Journal},
836     year = {2005},
837     volume = {272},
838     pages = {359-360},
839     month = {Jul},
840     annote = {Suppl. 1 005MG Times Cited:0 Cited References Count:0},
841     issn = {1742-464X},
842     uri = {<Go to ISI>://000234826102043},
843     }
844    
845     @ARTICLE{Mielke2004,
846     author = {S. P. Mielke and W. H. Fink and V. V. Krishnan and N. Gronbech-Jensen
847     and C. J. Benham},
848     title = {Transcription-driven twin supercoiling of a DNA loop: A Brownian
849     dynamics study},
850     journal = {Journal of Chemical Physics},
851     year = {2004},
852     volume = {121},
853     pages = {8104-8112},
854     number = {16},
855     month = {Oct 22},
856     abstract = {The torque generated by RNA polymerase as it tracks along double-stranded
857     DNA can potentially induce long-range structural deformations integral
858     to mechanisms of biological significance in both prokaryotes and
859     eukaryotes. In this paper, we introduce a dynamic computer model
860     for investigating this phenomenon. Duplex DNA is represented as
861     a chain of hydrodynamic beads interacting through potentials of
862     linearly elastic stretching, bending, and twisting, as well as excluded
863     volume. The chain, linear when relaxed, is looped to form two open
864     but topologically constrained subdomains. This permits the dynamic
865     introduction of torsional stress via a centrally applied torque.
866     We simulate by Brownian dynamics the 100 mus response of a 477-base
867     pair B-DNA template to the localized torque generated by the prokaryotic
868     transcription ensemble. Following a sharp rise at early times, the
869     distributed twist assumes a nearly constant value in both subdomains,
870     and a succession of supercoiling deformations occurs as superhelical
871     stress is increasingly partitioned to writhe. The magnitude of writhe
872     surpasses that of twist before also leveling off when the structure
873     reaches mechanical equilibrium with the torsional load. Superhelicity
874     is simultaneously right handed in one subdomain and left handed
875     in the other, as predicted by the #transcription-induced##twin-supercoiled-domain#
876     model [L. F. Liu and J. C. Wang, Proc. Natl. Acad. Sci. U.S.A. 84,
877     7024 (1987)]. The properties of the chain at the onset of writhing
878     agree well with predictions from theory, and the generated stress
879     is ample for driving secondary structural transitions in physiological
880     DNA. (C) 2004 American Institute of Physics.},
881     annote = {861ZF Times Cited:3 Cited References Count:34},
882     issn = {0021-9606},
883     uri = {<Go to ISI>://000224456500064},
884     }
885    
886     @ARTICLE{Naess2001,
887     author = {S. N. Naess and H. M. Adland and A. Mikkelsen and A. Elgsaeter},
888     title = {Brownian dynamics simulation of rigid bodies and segmented polymer
889     chains. Use of Cartesian rotation vectors as the generalized coordinates
890     describing angular orientations},
891     journal = {Physica A},
892     year = {2001},
893     volume = {294},
894     pages = {323-339},
895     number = {3-4},
896     month = {May 15},
897     abstract = {The three Eulerian angles constitute the classical choice of generalized
898     coordinates used to describe the three degrees of rotational freedom
899     of a rigid body, but it has long been known that this choice yields
900     singular equations of motion. The latter is also true when Eulerian
901     angles are used in Brownian dynamics analyses of the angular orientation
902     of single rigid bodies and segmented polymer chains. Starting from
903     kinetic theory we here show that by instead employing the three
904     components of Cartesian rotation vectors as the generalized coordinates
905     describing angular orientation, no singularity appears in the configuration
906     space diffusion equation and the associated Brownian dynamics algorithm.
907     The suitability of Cartesian rotation vectors in Brownian dynamics
908     simulations of segmented polymer chains with spring-like or ball-socket
909     joints is discussed. (C) 2001 Elsevier Science B.V. All rights reserved.},
910     annote = {433TA Times Cited:7 Cited References Count:19},
911     issn = {0378-4371},
912     uri = {<Go to ISI>://000168774800005},
913     }
914    
915     @ARTICLE{Noguchi2002,
916     author = {H. Noguchi and M. Takasu},
917     title = {Structural changes of pulled vesicles: A Brownian dynamics simulation},
918     journal = {Physical Review E},
919     year = {2002},
920     volume = {65},
921     pages = {-},
922     number = {5},
923     month = {may},
924     abstract = {We Studied the structural changes of bilayer vesicles induced by mechanical
925     forces using a Brownian dynamics simulation. Two nanoparticles,
926     which interact repulsively with amphiphilic molecules, are put inside
927     a vesicle. The position of one nanoparticle is fixed, and the other
928     is moved by a constant force as in optical-trapping experiments.
929     First, the pulled vesicle stretches into a pear or tube shape. Then
930     the inner monolayer in the tube-shaped region is deformed, and a
931     cylindrical structure is formed between two vesicles. After stretching
932     the cylindrical region, fission occurs near the moved vesicle. Soon
933     after this the cylindrical region shrinks. The trapping force similar
934     to 100 pN is needed to induce the formation of the cylindrical structure
935     and fission.},
936     annote = {Part 1 568PX Times Cited:5 Cited References Count:39},
937     issn = {1063-651X},
938     uri = {<Go to ISI>://000176552300084},
939     }
940    
941     @ARTICLE{Noguchi2001,
942     author = {H. Noguchi and M. Takasu},
943     title = {Fusion pathways of vesicles: A Brownian dynamics simulation},
944     journal = {Journal of Chemical Physics},
945     year = {2001},
946     volume = {115},
947     pages = {9547-9551},
948     number = {20},
949     month = {Nov 22},
950     abstract = {We studied the fusion dynamics of vesicles using a Brownian dynamics
951     simulation. Amphiphilic molecules spontaneously form vesicles with
952     a bilayer structure. Two vesicles come into contact and form a stalk
953     intermediate, in which a necklike structure only connects the outer
954     monolayers, as predicted by the stalk hypothesis. We have found
955     a new pathway of pore opening from stalks at high temperature: the
956     elliptic stalk bends and contact between the ends of the arc-shaped
957     stalk leads to pore opening. On the other hand, we have clarified
958     that the pore-opening process at low temperature agrees with the
959     modified stalk model: a pore is induced by contact between the inner
960     monolayers inside the stalk. (C) 2001 American Institute of Physics.},
961     annote = {491UW Times Cited:48 Cited References Count:25},
962     issn = {0021-9606},
963     uri = {<Go to ISI>://000172129300049},
964     }
965    
966     @ARTICLE{Palacios1998,
967     author = {J. L. Garcia-Palacios and F. J. Lazaro},
968     title = {Langevin-dynamics study of the dynamical properties of small magnetic
969     particles},
970     journal = {Physical Review B},
971     year = {1998},
972     volume = {58},
973     pages = {14937-14958},
974     number = {22},
975     month = {Dec 1},
976     abstract = {The stochastic Landau-Lifshitz-Gilbert equation of motion for a classical
977     magnetic moment is numerically solved (properly observing the customary
978     interpretation of it as a Stratonovich stochastic differential equation),
979     in order to study the dynamics of magnetic nanoparticles. The corresponding
980     Langevin-dynamics approach allows for the study of the fluctuating
981     trajectories of individual magnetic moments, where we have encountered
982     remarkable phenomena in the overbarrier rotation process, such as
983     crossing-back or multiple crossing of the potential barrier, rooted
984     in the gyromagnetic nature of the system. Concerning averaged quantities,
985     we study the linear dynamic response of the archetypal ensemble
986     of noninteracting classical magnetic moments with axially symmetric
987     magnetic anisotropy. The results are compared with different analytical
988     expressions used to model the relaxation of nanoparticle ensembles,
989     assessing their accuracy. It has been found that, among a number
990     of heuristic expressions for the linear dynamic susceptibility,
991     only the simple formula proposed by Shliomis and Stepanov matches
992     the coarse features of the susceptibility reasonably. By comparing
993     the numerical results with the asymptotic formula of Storonkin {Sov.
994     Phys. Crystallogr. 30, 489 (1985) [Kristallografiya 30, 841 (1985)]},
995     the effects of the intra-potential-well relaxation modes on the
996     low-temperature longitudinal dynamic response have been assessed,
997     showing their relatively small reflection in the susceptibility
998     curves but their dramatic influence on the phase shifts. Comparison
999     of the numerical results with the exact zero-damping expression
1000     for the transverse susceptibility by Garanin, Ishchenko, and Panina
1001     {Theor. Math. Phys. (USSR) 82, 169 (1990) [Teor. Mat. Fit. 82, 242
1002     (1990)]}, reveals a sizable contribution of the spread of the precession
1003     frequencies of the magnetic moment in the anisotropy field to the
1004     dynamic response at intermediate-to-high temperatures. [S0163-1829
1005     (98)00446-9].},
1006     annote = {146XW Times Cited:66 Cited References Count:45},
1007     issn = {0163-1829},
1008     uri = {<Go to ISI>://000077460000052},
1009     }
1010    
1011     @ARTICLE{Pastor1988,
1012     author = {R. W. Pastor and B. R. Brooks and A. Szabo},
1013     title = {An Analysis of the Accuracy of Langevin and Molecular-Dynamics Algorithms},
1014     journal = {Molecular Physics},
1015     year = {1988},
1016     volume = {65},
1017     pages = {1409-1419},
1018     number = {6},
1019     month = {Dec 20},
1020     annote = {T1302 Times Cited:61 Cited References Count:26},
1021     issn = {0026-8976},
1022     uri = {<Go to ISI>://A1988T130200011},
1023     }
1024    
1025     @ARTICLE{Recio2004,
1026     author = {J. Fernandez-Recio and M. Totrov and R. Abagyan},
1027     title = {Identification of protein-protein interaction sites from docking
1028     energy landscapes},
1029     journal = {Journal of Molecular Biology},
1030     year = {2004},
1031     volume = {335},
1032     pages = {843-865},
1033     number = {3},
1034     month = {Jan 16},
1035     abstract = {Protein recognition is one of the most challenging and intriguing
1036     problems in structural biology. Despite all the available structural,
1037     sequence and biophysical information about protein-protein complexes,
1038     the physico-chemical patterns, if any, that make a protein surface
1039     likely to be involved in protein-protein interactions, remain elusive.
1040     Here, we apply protein docking simulations and analysis of the interaction
1041     energy landscapes to identify protein-protein interaction sites.
1042     The new protocol for global docking based on multi-start global
1043     energy optimization of an allatom model of the ligand, with detailed
1044     receptor potentials and atomic solvation parameters optimized in
1045     a training set of 24 complexes, explores the conformational space
1046     around the whole receptor without restrictions. The ensembles of
1047     the rigid-body docking solutions generated by the simulations were
1048     subsequently used to project the docking energy landscapes onto
1049     the protein surfaces. We found that highly populated low-energy
1050     regions consistently corresponded to actual binding sites. The procedure
1051     was validated on a test set of 21 known protein-protein complexes
1052     not used in the training set. As much as 81% of the predicted high-propensity
1053     patch residues were located correctly in the native interfaces.
1054     This approach can guide the design of mutations on the surfaces
1055     of proteins, provide geometrical details of a possible interaction,
1056     and help to annotate protein surfaces in structural proteomics.
1057     (C) 2003 Elsevier Ltd. All rights reserved.},
1058     annote = {763GQ Times Cited:21 Cited References Count:59},
1059     issn = {0022-2836},
1060     uri = {<Go to ISI>://000188066900016},
1061     }
1062    
1063     @ARTICLE{Sandu1999,
1064     author = {A. Sandu and T. Schlick},
1065     title = {Masking resonance artifacts in force-splitting methods for biomolecular
1066     simulations by extrapolative Langevin dynamics},
1067     journal = {Journal of Computational Physics},
1068     year = {1999},
1069     volume = {151},
1070     pages = {74-113},
1071     number = {1},
1072     month = {May 1},
1073     abstract = {Numerical resonance artifacts have become recognized recently as a
1074     limiting factor to increasing the timestep in multiple-timestep
1075     (MTS) biomolecular dynamics simulations. At certain timesteps correlated
1076     to internal motions (e.g., 5 fs, around half the period of the fastest
1077     bond stretch, T-min), visible inaccuracies or instabilities can
1078     occur. Impulse-MTS schemes are vulnerable to these resonance errors
1079     since large energy pulses are introduced to the governing dynamics
1080     equations when the slow forces are evaluated. We recently showed
1081     that such resonance artifacts can be masked significantly by applying
1082     extrapolative splitting to stochastic dynamics. Theoretical and
1083     numerical analyses of force-splitting integrators based on the Verlet
1084     discretization are reported here for linear models to explain these
1085     observations and to suggest how to construct effective integrators
1086     for biomolecular dynamics that balance stability with accuracy.
1087     Analyses for Newtonian dynamics demonstrate the severe resonance
1088     patterns of the Impulse splitting, with this severity worsening
1089     with the outer timestep. Delta t: Constant Extrapolation is generally
1090     unstable, but the disturbances do not grow with Delta t. Thus. the
1091     stochastic extrapolative combination can counteract generic instabilities
1092     and largely alleviate resonances with a sufficiently strong Langevin
1093     heat-bath coupling (gamma), estimates for which are derived here
1094     based on the fastest and slowest motion periods. These resonance
1095     results generally hold for nonlinear test systems: a water tetramer
1096     and solvated protein. Proposed related approaches such as Extrapolation/Correction
1097     and Midpoint Extrapolation work better than Constant Extrapolation
1098     only for timesteps less than T-min/2. An effective extrapolative
1099     stochastic approach for biomolecules that balances long-timestep
1100     stability with good accuracy for the fast subsystem is then applied
1101     to a biomolecule using a three-class partitioning: the medium forces
1102     are treated by Midpoint Extrapolation via position Verlet, and the
1103     slow forces are incorporated by Constant Extrapolation. The resulting
1104     algorithm (LN) performs well on a solvated protein system in terms
1105     of thermodynamic properties and yields an order of magnitude speedup
1106     with respect to single-timestep Langevin trajectories. Computed
1107     spectral density functions also show how the Newtonian modes can
1108     be approximated by using a small gamma in the range Of 5-20 ps(-1).
1109     (C) 1999 Academic Press.},
1110     annote = {194FM Times Cited:14 Cited References Count:32},
1111     issn = {0021-9991},
1112     uri = {<Go to ISI>://000080181500004},
1113     }
1114    
1115     @ARTICLE{Shen2002,
1116     author = {M. Y. Shen and K. F. Freed},
1117     title = {Long time dynamics of met-enkephalin: Comparison of explicit and
1118     implicit solvent models},
1119     journal = {Biophysical Journal},
1120     year = {2002},
1121     volume = {82},
1122     pages = {1791-1808},
1123     number = {4},
1124     month = {Apr},
1125     abstract = {Met-enkephalin is one of the smallest opiate peptides. Yet, its dynamical
1126     structure and receptor docking mechanism are still not well understood.
1127     The conformational dynamics of this neuron peptide in liquid water
1128     are studied here by using all-atom molecular dynamics (MID) and
1129     implicit water Langevin dynamics (LD) simulations with AMBER potential
1130     functions and the three-site transferable intermolecular potential
1131     (TIP3P) model for water. To achieve the same simulation length in
1132     physical time, the full MID simulations require 200 times as much
1133     CPU time as the implicit water LID simulations. The solvent hydrophobicity
1134     and dielectric behavior are treated in the implicit solvent LD simulations
1135     by using a macroscopic solvation potential, a single dielectric
1136     constant, and atomic friction coefficients computed using the accessible
1137     surface area method with the TIP3P model water viscosity as determined
1138     here from MID simulations for pure TIP3P water. Both the local and
1139     the global dynamics obtained from the implicit solvent LD simulations
1140     agree very well with those from the explicit solvent MD simulations.
1141     The simulations provide insights into the conformational restrictions
1142     that are associated with the bioactivity of the opiate peptide dermorphin
1143     for the delta-receptor.},
1144     annote = {540MH Times Cited:36 Cited References Count:45},
1145     issn = {0006-3495},
1146     uri = {<Go to ISI>://000174932400010},
1147     }
1148    
1149     @ARTICLE{Shillcock2005,
1150     author = {J. C. Shillcock and R. Lipowsky},
1151     title = {Tension-induced fusion of bilayer membranes and vesicles},
1152     journal = {Nature Materials},
1153     year = {2005},
1154     volume = {4},
1155     pages = {225-228},
1156     number = {3},
1157     month = {Mar},
1158     annote = {901QJ Times Cited:9 Cited References Count:23},
1159     issn = {1476-1122},
1160     uri = {<Go to ISI>://000227296700019},
1161     }
1162    
1163     @ARTICLE{Skeel2002,
1164     author = {R. D. Skeel and J. A. Izaguirre},
1165     title = {An impulse integrator for Langevin dynamics},
1166     journal = {Molecular Physics},
1167     year = {2002},
1168     volume = {100},
1169     pages = {3885-3891},
1170     number = {24},
1171     month = {Dec 20},
1172     abstract = {The best simple method for Newtonian molecular dynamics is indisputably
1173     the leapfrog Stormer-Verlet method. The appropriate generalization
1174     to simple Langevin dynamics is unclear. An analysis is presented
1175     comparing an 'impulse method' (kick; fluctuate; kick), the 1982
1176     method of van Gunsteren and Berendsen, and the Brunger-Brooks-Karplus
1177     (BBK) method. It is shown how the impulse method and the van Gunsteren-Berendsen
1178     methods can be implemented as efficiently as the BBK method. Other
1179     considerations suggest that the impulse method is the best basic
1180     method for simple Langevin dynamics, with the van Gunsteren-Berendsen
1181     method a close contender.},
1182     annote = {633RX Times Cited:8 Cited References Count:22},
1183     issn = {0026-8976},
1184     uri = {<Go to ISI>://000180297200014},
1185     }
1186    
1187     @ARTICLE{Skeel1997,
1188     author = {R. D. Skeel and G. H. Zhang and T. Schlick},
1189     title = {A family of symplectic integrators: Stability, accuracy, and molecular
1190     dynamics applications},
1191     journal = {Siam Journal on Scientific Computing},
1192     year = {1997},
1193     volume = {18},
1194     pages = {203-222},
1195     number = {1},
1196     month = {Jan},
1197     abstract = {The following integration methods for special second-order ordinary
1198     differential equations are studied: leapfrog, implicit midpoint,
1199     trapezoid, Stormer-Verlet, and Cowell-Numerov. We show that all
1200     are members, or equivalent to members, of a one-parameter family
1201     of schemes. Some methods have more than one common form, and we
1202     discuss a systematic enumeration of these forms. We also present
1203     a stability and accuracy analysis based on the idea of ''modified
1204     equations'' and a proof of symplecticness. It follows that Cowell-Numerov
1205     and ''LIM2'' (a method proposed by Zhang and Schlick) are symplectic.
1206     A different interpretation of the values used by these integrators
1207     leads to higher accuracy and better energy conservation. Hence,
1208     we suggest that the straightforward analysis of energy conservation
1209     is misleading.},
1210     annote = {We981 Times Cited:30 Cited References Count:35},
1211     issn = {1064-8275},
1212     uri = {<Go to ISI>://A1997WE98100012},
1213     }
1214    
1215     @ARTICLE{Tao2005,
1216     author = {Y. G. Tao and W. K. {den Otter} and J. T. Padding and J. K. G. Dhont
1217     and W. J. Briels},
1218     title = {Brownian dynamics simulations of the self- and collective rotational
1219     diffusion coefficients of rigid long thin rods},
1220     journal = {Journal of Chemical Physics},
1221     year = {2005},
1222     volume = {122},
1223     pages = {-},
1224     number = {24},
1225     month = {Jun 22},
1226     abstract = {Recently a microscopic theory for the dynamics of suspensions of long
1227     thin rigid rods was presented, confirming and expanding the well-known
1228     theory by Doi and Edwards [The Theory of Polymer Dynamics (Clarendon,
1229     Oxford, 1986)] and Kuzuu [J. Phys. Soc. Jpn. 52, 3486 (1983)]. Here
1230     this theory is put to the test by comparing it against computer
1231     simulations. A Brownian dynamics simulation program was developed
1232     to follow the dynamics of the rods, with a length over a diameter
1233     ratio of 60, on the Smoluchowski time scale. The model accounts
1234     for excluded volume interactions between rods, but neglects hydrodynamic
1235     interactions. The self-rotational diffusion coefficients D-r(phi)
1236     of the rods were calculated by standard methods and by a new, more
1237     efficient method based on calculating average restoring torques.
1238     Collective decay of orientational order was calculated by means
1239     of equilibrium and nonequilibrium simulations. Our results show
1240     that, for the currently accessible volume fractions, the decay times
1241     in both cases are virtually identical. Moreover, the observed decay
1242     of diffusion coefficients with volume fraction is much quicker than
1243     predicted by the theory, which is attributed to an oversimplification
1244     of dynamic correlations in the theory. (c) 2005 American Institute
1245     of Physics.},
1246     annote = {943DN Times Cited:3 Cited References Count:26},
1247     issn = {0021-9606},
1248     uri = {<Go to ISI>://000230332400077},
1249     }
1250    
1251     @ARTICLE{Tuckerman1992,
1252     author = {M. Tuckerman and B. J. Berne and G. J. Martyna},
1253     title = {Reversible Multiple Time Scale Molecular-Dynamics},
1254     journal = {Journal of Chemical Physics},
1255     year = {1992},
1256     volume = {97},
1257     pages = {1990-2001},
1258     number = {3},
1259     month = {Aug 1},
1260     abstract = {The Trotter factorization of the Liouville propagator is used to generate
1261     new reversible molecular dynamics integrators. This strategy is
1262     applied to derive reversible reference system propagator algorithms
1263     (RESPA) that greatly accelerate simulations of systems with a separation
1264     of time scales or with long range forces. The new algorithms have
1265     all of the advantages of previous RESPA integrators but are reversible,
1266     and more stable than those methods. These methods are applied to
1267     a set of paradigmatic systems and are shown to be superior to earlier
1268     methods. It is shown how the new RESPA methods are related to predictor-corrector
1269     integrators. Finally, we show how these methods can be used to accelerate
1270     the integration of the equations of motion of systems with Nose
1271     thermostats.},
1272     annote = {Je891 Times Cited:680 Cited References Count:19},
1273     issn = {0021-9606},
1274     uri = {<Go to ISI>://A1992JE89100044},
1275     }
1276