ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/dissertation.bib
(Generate patch)

Comparing trunk/tengDissertation/dissertation.bib (file contents):
Revision 2796 by tim, Mon Jun 5 21:00:46 2006 UTC vs.
Revision 2797 by tim, Tue Jun 6 02:12:34 2006 UTC

# Line 3 | Line 3 | Encoding: GBK
3  
4   @ARTICLE{Torre2003,
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},
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},
8 >    methods and applications},
9    journal = {European Biophysics Journal with Biophysics Letters},
10    year = {2003},
11    volume = {32},
# Line 13 | Line 13 | Encoding: GBK
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.},
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},
# Line 42 | Line 42 | Encoding: GBK
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},
45 >    by time series analysis. II. Investigation of explicit solvent effects},
46    journal = {Journal of Chemical Physics},
47    year = {2005},
48    volume = {123},
# Line 50 | Line 50 | Encoding: GBK
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.},
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},
# Line 99 | Line 99 | Encoding: GBK
99   @ARTICLE{Allison1991,
100    author = {S. A. Allison},
101    title = {A Brownian Dynamics Algorithm for Arbitrary Rigid Bodies - Application
102 <        to Polarized Dynamic Light-Scattering},
102 >    to Polarized Dynamic Light-Scattering},
103    journal = {Macromolecules},
104    year = {1991},
105    volume = {24},
# Line 107 | Line 107 | Encoding: GBK
107    number = {2},
108    month = {Jan 21},
109    abstract = {A Brownian dynamics algorithm is developed to simulate dynamics experiments
110 <        of rigid macromolecules. It is applied to polarized dynamic light
111 <        scattering from rodlike sturctures and from a model of a DNA fragment
112 <        (762 base pairs). A number of rod cases are examined in which the
113 <        translational anisotropy is increased form zero to a large value.
114 <        Simulated first cumulants as well as amplitudes and lifetimes of
115 <        the dynamic form factor are compared with predictions of analytic
116 <        theories and found to be in very good agreement with them. For DNA
117 <        fragments 762 base pairs in length or longer, translational anisotropy
118 <        does not contribute significantly to dynamic light scattering. In
119 <        a comparison of rigid and flexible simulations on semistiff models
120 <        of this fragment, it is shown directly that flexing contributes
121 <        to the faster decay processes probed by light scattering and that
122 <        the flexible model studies are in good agreement with experiment.},
110 >    of rigid macromolecules. It is applied to polarized dynamic light
111 >    scattering from rodlike sturctures and from a model of a DNA fragment
112 >    (762 base pairs). A number of rod cases are examined in which the
113 >    translational anisotropy is increased form zero to a large value.
114 >    Simulated first cumulants as well as amplitudes and lifetimes of
115 >    the dynamic form factor are compared with predictions of analytic
116 >    theories and found to be in very good agreement with them. For DNA
117 >    fragments 762 base pairs in length or longer, translational anisotropy
118 >    does not contribute significantly to dynamic light scattering. In
119 >    a comparison of rigid and flexible simulations on semistiff models
120 >    of this fragment, it is shown directly that flexing contributes
121 >    to the faster decay processes probed by light scattering and that
122 >    the flexible model studies are in good agreement with experiment.},
123    annote = {Eu814 Times Cited:8 Cited References Count:32},
124    issn = {0024-9297},
125    uri = {<Go to ISI>://A1991EU81400029},
# Line 128 | Line 128 | Encoding: GBK
128   @ARTICLE{Andersen1983,
129    author = {H. C. Andersen},
130    title = {Rattle - a Velocity Version of the Shake Algorithm for Molecular-Dynamics
131 <        Calculations},
131 >    Calculations},
132    journal = {Journal of Computational Physics},
133    year = {1983},
134    volume = {52},
# Line 142 | Line 142 | Encoding: GBK
142   @ARTICLE{Auerbach2005,
143    author = {A. Auerbach},
144    title = {Gating of acetylcholine receptor channels: Brownian motion across
145 <        a broad transition state},
145 >    a broad transition state},
146    journal = {Proceedings of the National Academy of Sciences of the United States
147 <        of America},
147 >    of America},
148    year = {2005},
149    volume = {102},
150    pages = {1408-1412},
151    number = {5},
152    month = {Feb 1},
153    abstract = {Acetylcholine receptor channels (AChRs) are proteins that switch between
154 <        stable #closed# and #open# conformations. In patch clamp recordings,
155 <        diliganded AChR gating appears to be a simple, two-state reaction.
156 <        However, mutagenesis studies indicate that during gating dozens
157 <        of residues across the protein move asynchronously and are organized
158 <        into rigid body gating domains (#blocks#). Moreover, there is an
159 <        upper limit to the apparent channel opening rate constant. These
160 <        observations suggest that the gating reaction has a broad, corrugated
161 <        transition state region, with the maximum opening rate reflecting,
162 <        in part, the mean first-passage time across this ensemble. Simulations
163 <        reveal that a flat, isotropic energy profile for the transition
164 <        state can account for many of the essential features of AChR gating.
165 <        With this mechanism, concerted, local structural transitions that
166 <        occur on the broad transition state ensemble give rise to fractional
167 <        measures of reaction progress (Phi values) determined by rate-equilibrium
168 <        free energy relationship analysis. The results suggest that the
169 <        coarse-grained AChR gating conformational change propagates through
170 <        the protein with dynamics that are governed by the Brownian motion
171 <        of individual gating blocks.},
154 >    stable #closed# and #open# conformations. In patch clamp recordings,
155 >    diliganded AChR gating appears to be a simple, two-state reaction.
156 >    However, mutagenesis studies indicate that during gating dozens
157 >    of residues across the protein move asynchronously and are organized
158 >    into rigid body gating domains (#blocks#). Moreover, there is an
159 >    upper limit to the apparent channel opening rate constant. These
160 >    observations suggest that the gating reaction has a broad, corrugated
161 >    transition state region, with the maximum opening rate reflecting,
162 >    in part, the mean first-passage time across this ensemble. Simulations
163 >    reveal that a flat, isotropic energy profile for the transition
164 >    state can account for many of the essential features of AChR gating.
165 >    With this mechanism, concerted, local structural transitions that
166 >    occur on the broad transition state ensemble give rise to fractional
167 >    measures of reaction progress (Phi values) determined by rate-equilibrium
168 >    free energy relationship analysis. The results suggest that the
169 >    coarse-grained AChR gating conformational change propagates through
170 >    the protein with dynamics that are governed by the Brownian motion
171 >    of individual gating blocks.},
172    annote = {895QF Times Cited:9 Cited References Count:33},
173    issn = {0027-8424},
174    uri = {<Go to ISI>://000226877300030},
# Line 177 | Line 177 | Encoding: GBK
177   @ARTICLE{Baber1995,
178    author = {J. Baber and J. F. Ellena and D. S. Cafiso},
179    title = {Distribution of General-Anesthetics in Phospholipid-Bilayers Determined
180 <        Using H-2 Nmr and H-1-H-1 Noe Spectroscopy},
180 >    Using H-2 Nmr and H-1-H-1 Noe Spectroscopy},
181    journal = {Biochemistry},
182    year = {1995},
183    volume = {34},
# Line 185 | Line 185 | Encoding: GBK
185    number = {19},
186    month = {May 16},
187    abstract = {The effect of the general anesthetics halothane, enflurane, and isoflurane
188 <        on hydrocarbon chain packing in palmitoyl(d(31))oleoylphosphatidylcholine
189 <        membranes in the liquid crystalline phase was investigated using
190 <        H-2 NMR. Upon the addition of the anesthetics, the first five methylene
191 <        units near the interface generally show a very small increase in
192 <        segmental order, while segments deeper within the bilayer show a
193 <        small decrease in segmental order. From the H-2 NMR results, the
194 <        chain length for the perdeuterated palmitoyl chain in the absence
195 <        of anesthetic was found to be 12.35 Angstrom. Upon the addition
196 <        of halothane enflurane, or isoflurane, the acyl chain undergoes
197 <        slight contractions of 0.11, 0.20, or 0.16 Angstrom, respectively,
198 <        at 50 mol % anesthetic. A simple model was used to estimate the
199 <        relative amounts of anesthetic located near the interface and deeper
200 <        in the bilayer hydrocarbon region, and only a slight preference
201 <        for an interfacial location was observed. Intermolecular H-1-H-1
202 <        nuclear Overhauser effects (NOEs) were measured between phospholipid
203 <        and halothane protons. These NOEs are consistent with the intramembrane
204 <        location of the anesthetics suggested by the H-2 NMR data. In addition,
205 <        the NOE data indicate that anesthetics prefer the interfacial and
206 <        hydrocarbon regions of the membrane and are not found in high concentrations
207 <        in the phospholipid headgroup.},
188 >    on hydrocarbon chain packing in palmitoyl(d(31))oleoylphosphatidylcholine
189 >    membranes in the liquid crystalline phase was investigated using
190 >    H-2 NMR. Upon the addition of the anesthetics, the first five methylene
191 >    units near the interface generally show a very small increase in
192 >    segmental order, while segments deeper within the bilayer show a
193 >    small decrease in segmental order. From the H-2 NMR results, the
194 >    chain length for the perdeuterated palmitoyl chain in the absence
195 >    of anesthetic was found to be 12.35 Angstrom. Upon the addition
196 >    of halothane enflurane, or isoflurane, the acyl chain undergoes
197 >    slight contractions of 0.11, 0.20, or 0.16 Angstrom, respectively,
198 >    at 50 mol % anesthetic. A simple model was used to estimate the
199 >    relative amounts of anesthetic located near the interface and deeper
200 >    in the bilayer hydrocarbon region, and only a slight preference
201 >    for an interfacial location was observed. Intermolecular H-1-H-1
202 >    nuclear Overhauser effects (NOEs) were measured between phospholipid
203 >    and halothane protons. These NOEs are consistent with the intramembrane
204 >    location of the anesthetics suggested by the H-2 NMR data. In addition,
205 >    the NOE data indicate that anesthetics prefer the interfacial and
206 >    hydrocarbon regions of the membrane and are not found in high concentrations
207 >    in the phospholipid headgroup.},
208    annote = {Qz716 Times Cited:38 Cited References Count:37},
209    issn = {0006-2960},
210    uri = {<Go to ISI>://A1995QZ71600035},
# Line 213 | Line 213 | Encoding: GBK
213   @ARTICLE{Banerjee2004,
214    author = {D. Banerjee and B. C. Bag and S. K. Banik and D. S. Ray},
215    title = {Solution of quantum Langevin equation: Approximations, theoretical
216 <        and numerical aspects},
216 >    and numerical aspects},
217    journal = {Journal of Chemical Physics},
218    year = {2004},
219    volume = {120},
# Line 221 | Line 221 | Encoding: GBK
221    number = {19},
222    month = {May 15},
223    abstract = {Based on a coherent state representation of noise operator and an
224 <        ensemble averaging procedure using Wigner canonical thermal distribution
225 <        for harmonic oscillators, a generalized quantum Langevin equation
226 <        has been recently developed [Phys. Rev. E 65, 021109 (2002); 66,
227 <        051106 (2002)] to derive the equations of motion for probability
228 <        distribution functions in c-number phase-space. We extend the treatment
229 <        to explore several systematic approximation schemes for the solutions
230 <        of the Langevin equation for nonlinear potentials for a wide range
231 <        of noise correlation, strength and temperature down to the vacuum
232 <        limit. The method is exemplified by an analytic application to harmonic
233 <        oscillator for arbitrary memory kernel and with the help of a numerical
234 <        calculation of barrier crossing, in a cubic potential to demonstrate
235 <        the quantum Kramers' turnover and the quantum Arrhenius plot. (C)
236 <        2004 American Institute of Physics.},
224 >    ensemble averaging procedure using Wigner canonical thermal distribution
225 >    for harmonic oscillators, a generalized quantum Langevin equation
226 >    has been recently developed [Phys. Rev. E 65, 021109 (2002); 66,
227 >    051106 (2002)] to derive the equations of motion for probability
228 >    distribution functions in c-number phase-space. We extend the treatment
229 >    to explore several systematic approximation schemes for the solutions
230 >    of the Langevin equation for nonlinear potentials for a wide range
231 >    of noise correlation, strength and temperature down to the vacuum
232 >    limit. The method is exemplified by an analytic application to harmonic
233 >    oscillator for arbitrary memory kernel and with the help of a numerical
234 >    calculation of barrier crossing, in a cubic potential to demonstrate
235 >    the quantum Kramers' turnover and the quantum Arrhenius plot. (C)
236 >    2004 American Institute of Physics.},
237    annote = {816YY Times Cited:8 Cited References Count:35},
238    issn = {0021-9606},
239    uri = {<Go to ISI>://000221146400009},
# Line 251 | Line 251 | Encoding: GBK
251   @ARTICLE{Barth1998,
252    author = {E. Barth and T. Schlick},
253    title = {Overcoming stability limitations in biomolecular dynamics. I. Combining
254 <        force splitting via extrapolation with Langevin dynamics in LN},
254 >    force splitting via extrapolation with Langevin dynamics in LN},
255    journal = {Journal of Chemical Physics},
256    year = {1998},
257    volume = {109},
# Line 259 | Line 259 | Encoding: GBK
259    number = {5},
260    month = {Aug 1},
261    abstract = {We present an efficient new method termed LN for propagating biomolecular
262 <        dynamics according to the Langevin equation that arose fortuitously
263 <        upon analysis of the range of harmonic validity of our normal-mode
264 <        scheme LIN. LN combines force linearization with force splitting
265 <        techniques and disposes of LIN'S computationally intensive minimization
266 <        (anharmonic correction) component. Unlike the competitive multiple-timestepping
267 <        (MTS) schemes today-formulated to be symplectic and time-reversible-LN
268 <        merges the slow and fast forces via extrapolation rather than impulses;
269 <        the Langevin heat bath prevents systematic energy drifts. This combination
270 <        succeeds in achieving more significant speedups than these MTS methods
271 <        which are Limited by resonance artifacts to an outer timestep less
272 <        than some integer multiple of half the period of the fastest motion
273 <        (around 4-5 fs for biomolecules). We show that LN achieves very
274 <        good agreement with small-timestep solutions of the Langevin equation
275 <        in terms of thermodynamics (energy means and variances), geometry,
276 <        and dynamics (spectral densities) for two proteins in vacuum and
277 <        a large water system. Significantly, the frequency of updating the
278 <        slow forces extends to 48 fs or more, resulting in speedup factors
279 <        exceeding 10. The implementation of LN in any program that employs
280 <        force-splitting computations is straightforward, with only partial
281 <        second-derivative information required, as well as sparse Hessian/vector
282 <        multiplication routines. The linearization part of LN could even
283 <        be replaced by direct evaluation of the fast components. The application
284 <        of LN to biomolecular dynamics is well suited for configurational
285 <        sampling, thermodynamic, and structural questions. (C) 1998 American
286 <        Institute of Physics.},
262 >    dynamics according to the Langevin equation that arose fortuitously
263 >    upon analysis of the range of harmonic validity of our normal-mode
264 >    scheme LIN. LN combines force linearization with force splitting
265 >    techniques and disposes of LIN'S computationally intensive minimization
266 >    (anharmonic correction) component. Unlike the competitive multiple-timestepping
267 >    (MTS) schemes today-formulated to be symplectic and time-reversible-LN
268 >    merges the slow and fast forces via extrapolation rather than impulses;
269 >    the Langevin heat bath prevents systematic energy drifts. This combination
270 >    succeeds in achieving more significant speedups than these MTS methods
271 >    which are Limited by resonance artifacts to an outer timestep less
272 >    than some integer multiple of half the period of the fastest motion
273 >    (around 4-5 fs for biomolecules). We show that LN achieves very
274 >    good agreement with small-timestep solutions of the Langevin equation
275 >    in terms of thermodynamics (energy means and variances), geometry,
276 >    and dynamics (spectral densities) for two proteins in vacuum and
277 >    a large water system. Significantly, the frequency of updating the
278 >    slow forces extends to 48 fs or more, resulting in speedup factors
279 >    exceeding 10. The implementation of LN in any program that employs
280 >    force-splitting computations is straightforward, with only partial
281 >    second-derivative information required, as well as sparse Hessian/vector
282 >    multiplication routines. The linearization part of LN could even
283 >    be replaced by direct evaluation of the fast components. The application
284 >    of LN to biomolecular dynamics is well suited for configurational
285 >    sampling, thermodynamic, and structural questions. (C) 1998 American
286 >    Institute of Physics.},
287    annote = {105HH Times Cited:29 Cited References Count:49},
288    issn = {0021-9606},
289    uri = {<Go to ISI>://000075066300006},
# Line 292 | Line 292 | Encoding: GBK
292   @ARTICLE{Batcho2001,
293    author = {P. F. Batcho and T. Schlick},
294    title = {Special stability advantages of position-Verlet over velocity-Verlet
295 <        in multiple-time step integration},
295 >    in multiple-time step integration},
296    journal = {Journal of Chemical Physics},
297    year = {2001},
298    volume = {115},
# Line 300 | Line 300 | Encoding: GBK
300    number = {9},
301    month = {Sep 1},
302    abstract = {We present an analysis for a simple two-component harmonic oscillator
303 <        that compares the use of position-Verlet to velocity-Verlet for
304 <        multiple-time step integration. The numerical stability analysis
305 <        based on the impulse-Verlet splitting shows that position-Verlet
306 <        has enhanced stability, in terms of the largest allowable time step,
307 <        for cases where an ample separation of time scales exists. Numerical
308 <        investigations confirm the advantages of the position-Verlet scheme
309 <        when used for the fastest time scales of the system. Applications
310 <        to a biomolecule. a solvated protein, for both Newtonian and Langevin
311 <        dynamics echo these trends over large outer time-step regimes. (C)
312 <        2001 American Institute of Physics.},
303 >    that compares the use of position-Verlet to velocity-Verlet for
304 >    multiple-time step integration. The numerical stability analysis
305 >    based on the impulse-Verlet splitting shows that position-Verlet
306 >    has enhanced stability, in terms of the largest allowable time step,
307 >    for cases where an ample separation of time scales exists. Numerical
308 >    investigations confirm the advantages of the position-Verlet scheme
309 >    when used for the fastest time scales of the system. Applications
310 >    to a biomolecule. a solvated protein, for both Newtonian and Langevin
311 >    dynamics echo these trends over large outer time-step regimes. (C)
312 >    2001 American Institute of Physics.},
313    annote = {469KV Times Cited:6 Cited References Count:30},
314    issn = {0021-9606},
315    uri = {<Go to ISI>://000170813800005},
# Line 318 | Line 318 | Encoding: GBK
318   @ARTICLE{Bates2005,
319    author = {M. A. Bates and G. R. Luckhurst},
320    title = {Biaxial nematic phases and V-shaped molecules: A Monte Carlo simulation
321 <        study},
321 >    study},
322    journal = {Physical Review E},
323    year = {2005},
324    volume = {72},
# Line 326 | Line 326 | Encoding: GBK
326    number = {5},
327    month = {Nov},
328    abstract = {Inspired by recent claims that compounds composed of V-shaped molecules
329 <        can exhibit the elusive biaxial nematic phase, we have developed
330 <        a generic simulation model for such systems. This contains the features
331 <        of the molecule that are essential to its liquid crystal behavior,
332 <        namely the anisotropies of the two arms and the angle between them.
333 <        The behavior of the model has been investigated using Monte Carlo
334 <        simulations for a wide range of these structural parameters. This
335 <        allows us to establish the relationship between the V-shaped molecule
336 <        and its ability to form a biaxial nematic phase. Of particular importance
337 <        are the criteria of geometry and the relative anisotropy necessary
338 <        for the system to exhibit a Landau point, at which the biaxial nematic
339 <        is formed directly from the isotropic phase. The simulations have
340 <        also been used to determine the orientational order parameters for
341 <        a selection of molecular axes. These are especially important because
342 <        they reveal the phase symmetry and are connected to the experimental
343 <        determination of this. The simulation results show that, whereas
344 <        some positions are extremely sensitive to the phase biaxiality,
345 <        others are totally blind to this.},
329 >    can exhibit the elusive biaxial nematic phase, we have developed
330 >    a generic simulation model for such systems. This contains the features
331 >    of the molecule that are essential to its liquid crystal behavior,
332 >    namely the anisotropies of the two arms and the angle between them.
333 >    The behavior of the model has been investigated using Monte Carlo
334 >    simulations for a wide range of these structural parameters. This
335 >    allows us to establish the relationship between the V-shaped molecule
336 >    and its ability to form a biaxial nematic phase. Of particular importance
337 >    are the criteria of geometry and the relative anisotropy necessary
338 >    for the system to exhibit a Landau point, at which the biaxial nematic
339 >    is formed directly from the isotropic phase. The simulations have
340 >    also been used to determine the orientational order parameters for
341 >    a selection of molecular axes. These are especially important because
342 >    they reveal the phase symmetry and are connected to the experimental
343 >    determination of this. The simulation results show that, whereas
344 >    some positions are extremely sensitive to the phase biaxiality,
345 >    others are totally blind to this.},
346    annote = {Part 1 988LQ Times Cited:0 Cited References Count:38},
347    issn = {1539-3755},
348    uri = {<Go to ISI>://000233603100030},
# Line 358 | Line 358 | Encoding: GBK
358    number = {5},
359    month = {Nov 1},
360    abstract = {We introduce an unbiased protocol for performing rotational moves
361 <        in rigid-body dynamics simulations. This approach - based on the
362 <        analytic solution for the rotational equations of motion for an
363 <        orthogonal coordinate system at constant angular velocity - removes
364 <        deficiencies that have been largely ignored in Brownian dynamics
365 <        simulations, namely errors for finite rotations that result from
366 <        applying the noncommuting rotational matrices in an arbitrary order.
367 <        Our algorithm should thus replace standard approaches to rotate
368 <        local coordinate frames in Langevin and Brownian dynamics simulations.},
361 >    in rigid-body dynamics simulations. This approach - based on the
362 >    analytic solution for the rotational equations of motion for an
363 >    orthogonal coordinate system at constant angular velocity - removes
364 >    deficiencies that have been largely ignored in Brownian dynamics
365 >    simulations, namely errors for finite rotations that result from
366 >    applying the noncommuting rotational matrices in an arbitrary order.
367 >    Our algorithm should thus replace standard approaches to rotate
368 >    local coordinate frames in Langevin and Brownian dynamics simulations.},
369    annote = {736UA Times Cited:0 Cited References Count:11},
370    issn = {0006-3495},
371    uri = {<Go to ISI>://000186190500018},
# Line 374 | Line 374 | Encoding: GBK
374   @ARTICLE{Beloborodov1998,
375    author = {I. S. Beloborodov and V. Y. Orekhov and A. S. Arseniev},
376    title = {Effect of coupling between rotational and translational Brownian
377 <        motions on NMR spin relaxation: Consideration using green function
378 <        of rigid body diffusion},
377 >    motions on NMR spin relaxation: Consideration using green function
378 >    of rigid body diffusion},
379    journal = {Journal of Magnetic Resonance},
380    year = {1998},
381    volume = {132},
# Line 383 | Line 383 | Encoding: GBK
383    number = {2},
384    month = {Jun},
385    abstract = {Using the Green function of arbitrary rigid Brownian diffusion (Goldstein,
386 <        Biopolymers 33, 409-436, 1993), it was analytically shown that coupling
387 <        between translation and rotation diffusion degrees of freedom does
388 <        not affect the correlation functions relevant to the NMR intramolecular
389 <        relaxation. It follows that spectral densities usually used for
390 <        the anisotropic rotation diffusion (Woessner, J. Chem. Phys. 37,
391 <        647-654, 1962) can be regarded as exact in respect to the rotation-translation
392 <        coupling for the spin system connected with a rigid body. (C) 1998
393 <        Academic Press.},
386 >    Biopolymers 33, 409-436, 1993), it was analytically shown that coupling
387 >    between translation and rotation diffusion degrees of freedom does
388 >    not affect the correlation functions relevant to the NMR intramolecular
389 >    relaxation. It follows that spectral densities usually used for
390 >    the anisotropic rotation diffusion (Woessner, J. Chem. Phys. 37,
391 >    647-654, 1962) can be regarded as exact in respect to the rotation-translation
392 >    coupling for the spin system connected with a rigid body. (C) 1998
393 >    Academic Press.},
394    annote = {Zu605 Times Cited:2 Cited References Count:6},
395    issn = {1090-7807},
396    uri = {<Go to ISI>://000074214800017},
# Line 399 | Line 399 | Encoding: GBK
399   @ARTICLE{Berardi1996,
400    author = {R. Berardi and S. Orlandi and C. Zannoni},
401    title = {Antiphase structures in polar smectic liquid crystals and their molecular
402 <        origin},
402 >    origin},
403    journal = {Chemical Physics Letters},
404    year = {1996},
405    volume = {261},
# Line 407 | Line 407 | Encoding: GBK
407    number = {3},
408    month = {Oct 18},
409    abstract = {We demonstrate that the overall molecular dipole organization in a
410 <        smectic liquid crystal formed of polar molecules can be strongly
411 <        influenced by the position of the dipole in the molecule. We study
412 <        by large scale Monte Carlo simulations systems of attractive-repulsive
413 <        ''Gay-Berne'' elongated ellipsoids with an axial dipole at the center
414 <        or near the end of the molecule and we show that monolayer smectic
415 <        liquid crystals and modulated antiferroelectric bilayer stripe domains
416 <        similar to the experimentally observed ''antiphase'' structures
417 <        are obtained in the two cases.},
410 >    smectic liquid crystal formed of polar molecules can be strongly
411 >    influenced by the position of the dipole in the molecule. We study
412 >    by large scale Monte Carlo simulations systems of attractive-repulsive
413 >    ''Gay-Berne'' elongated ellipsoids with an axial dipole at the center
414 >    or near the end of the molecule and we show that monolayer smectic
415 >    liquid crystals and modulated antiferroelectric bilayer stripe domains
416 >    similar to the experimentally observed ''antiphase'' structures
417 >    are obtained in the two cases.},
418    annote = {Vn637 Times Cited:49 Cited References Count:26},
419    issn = {0009-2614},
420    uri = {<Go to ISI>://A1996VN63700023},
# Line 423 | Line 423 | Encoding: GBK
423   @ARTICLE{Berkov2005,
424    author = {D. V. Berkov and N. L. Gorn},
425    title = {Magnetization precession due to a spin-polarized current in a thin
426 <        nanoelement: Numerical simulation study},
426 >    nanoelement: Numerical simulation study},
427    journal = {Physical Review B},
428    year = {2005},
429    volume = {72},
# Line 431 | Line 431 | Encoding: GBK
431    number = {9},
432    month = {Sep},
433    abstract = {In this paper a detailed numerical study (in frames of the Slonczewski
434 <        formalism) of magnetization oscillations driven by a spin-polarized
435 <        current through a thin elliptical nanoelement is presented. We show
436 <        that a sophisticated micromagnetic model, where a polycrystalline
437 <        structure of a nanoelement is taken into account, can explain qualitatively
438 <        all most important features of the magnetization oscillation spectra
439 <        recently observed experimentally [S. I. Kiselev , Nature 425, 380
440 <        (2003)], namely, existence of several equidistant spectral bands,
441 <        sharp onset and abrupt disappearance of magnetization oscillations
442 <        with increasing current, absence of the out-of-plane regime predicted
443 <        by a macrospin model, and the relation between frequencies of so-called
444 <        small-angle and quasichaotic oscillations. However, a quantitative
445 <        agreement with experimental results (especially concerning the frequency
446 <        of quasichaotic oscillations) could not be achieved in the region
447 <        of reasonable parameter values, indicating that further model refinement
448 <        is necessary for a complete understanding of the spin-driven magnetization
449 <        precession even in this relatively simple experimental situation.},
434 >    formalism) of magnetization oscillations driven by a spin-polarized
435 >    current through a thin elliptical nanoelement is presented. We show
436 >    that a sophisticated micromagnetic model, where a polycrystalline
437 >    structure of a nanoelement is taken into account, can explain qualitatively
438 >    all most important features of the magnetization oscillation spectra
439 >    recently observed experimentally [S. I. Kiselev , Nature 425, 380
440 >    (2003)], namely, existence of several equidistant spectral bands,
441 >    sharp onset and abrupt disappearance of magnetization oscillations
442 >    with increasing current, absence of the out-of-plane regime predicted
443 >    by a macrospin model, and the relation between frequencies of so-called
444 >    small-angle and quasichaotic oscillations. However, a quantitative
445 >    agreement with experimental results (especially concerning the frequency
446 >    of quasichaotic oscillations) could not be achieved in the region
447 >    of reasonable parameter values, indicating that further model refinement
448 >    is necessary for a complete understanding of the spin-driven magnetization
449 >    precession even in this relatively simple experimental situation.},
450    annote = {969IT Times Cited:2 Cited References Count:55},
451    issn = {1098-0121},
452    uri = {<Go to ISI>://000232228500058},
# Line 455 | Line 455 | Encoding: GBK
455   @ARTICLE{Berkov2005a,
456    author = {D. V. Berkov and N. L. Gorn},
457    title = {Stochastic dynamic simulations of fast remagnetization processes:
458 <        recent advances and applications},
458 >    recent advances and applications},
459    journal = {Journal of Magnetism and Magnetic Materials},
460    year = {2005},
461    volume = {290},
462    pages = {442-448},
463    month = {Apr},
464    abstract = {Numerical simulations of fast remagnetization processes using stochastic
465 <        dynamics are widely used to study various magnetic systems. In this
466 <        paper, we first address several crucial methodological problems
467 <        of such simulations: (i) the influence of finite-element discretization
468 <        on simulated dynamics, (ii) choice between Ito and Stratonovich
469 <        stochastic calculi by the solution of micromagnetic stochastic equations
470 <        of motion and (iii) non-trivial correlation properties of the random
471 <        (thermal) field. Next, we discuss several examples to demonstrate
472 <        the great potential of the Langevin dynamics for studying fast remagnetization
473 <        processes in technically relevant applications: we present numerical
474 <        analysis of equilibrium magnon spectra in patterned structures,
475 <        study thermal noise effects on the magnetization dynamics of nanoelements
476 <        in pulsed fields and show some results for a remagnetization dynamics
477 <        induced by a spin-polarized current. (c) 2004 Elsevier B.V. All
478 <        rights reserved.},
465 >    dynamics are widely used to study various magnetic systems. In this
466 >    paper, we first address several crucial methodological problems
467 >    of such simulations: (i) the influence of finite-element discretization
468 >    on simulated dynamics, (ii) choice between Ito and Stratonovich
469 >    stochastic calculi by the solution of micromagnetic stochastic equations
470 >    of motion and (iii) non-trivial correlation properties of the random
471 >    (thermal) field. Next, we discuss several examples to demonstrate
472 >    the great potential of the Langevin dynamics for studying fast remagnetization
473 >    processes in technically relevant applications: we present numerical
474 >    analysis of equilibrium magnon spectra in patterned structures,
475 >    study thermal noise effects on the magnetization dynamics of nanoelements
476 >    in pulsed fields and show some results for a remagnetization dynamics
477 >    induced by a spin-polarized current. (c) 2004 Elsevier B.V. All
478 >    rights reserved.},
479    annote = {Part 1 Sp. Iss. SI 922KU Times Cited:2 Cited References Count:25},
480    issn = {0304-8853},
481    uri = {<Go to ISI>://000228837600109},
# Line 484 | Line 484 | Encoding: GBK
484   @ARTICLE{Berkov2002,
485    author = {D. V. Berkov and N. L. Gorn and P. Gornert},
486    title = {Magnetization dynamics in nanoparticle systems: Numerical simulation
487 <        using Langevin dynamics},
487 >    using Langevin dynamics},
488    journal = {Physica Status Solidi a-Applied Research},
489    year = {2002},
490    volume = {189},
# Line 492 | Line 492 | Encoding: GBK
492    number = {2},
493    month = {Feb 16},
494    abstract = {We report on recent progress achieved by the development of numerical
495 <        methods based on the stochastic (Langevin) dynamics applied to systems
496 <        of interacting magnetic nanoparticles. The method enables direct
497 <        simulations of the trajectories of magnetic moments taking into
498 <        account (i) all relevant interactions, (ii) precession dynamics,
499 <        and (iii) temperature fluctuations included via the random (thermal)
500 <        field. We present several novel results obtained using new methods
501 <        developed for the solution of the Langevin equations. In particular,
502 <        we have investigated magnetic nanodots and disordered granular systems
503 <        of single-domain magnetic particles. For the first case we have
504 <        calculated the spectrum and the spatial distribution of spin excitations.
505 <        For the second system the complex ac susceptibility chi(omega, T)
506 <        for various particle concentrations and particle anisotropies were
507 <        computed and compared with numerous experimental results.},
495 >    methods based on the stochastic (Langevin) dynamics applied to systems
496 >    of interacting magnetic nanoparticles. The method enables direct
497 >    simulations of the trajectories of magnetic moments taking into
498 >    account (i) all relevant interactions, (ii) precession dynamics,
499 >    and (iii) temperature fluctuations included via the random (thermal)
500 >    field. We present several novel results obtained using new methods
501 >    developed for the solution of the Langevin equations. In particular,
502 >    we have investigated magnetic nanodots and disordered granular systems
503 >    of single-domain magnetic particles. For the first case we have
504 >    calculated the spectrum and the spatial distribution of spin excitations.
505 >    For the second system the complex ac susceptibility chi(omega, T)
506 >    for various particle concentrations and particle anisotropies were
507 >    computed and compared with numerous experimental results.},
508    annote = {526TF Times Cited:4 Cited References Count:37},
509    issn = {0031-8965},
510    uri = {<Go to ISI>://000174145200026},
# Line 513 | Line 513 | Encoding: GBK
513   @ARTICLE{Bernal1980,
514    author = {J.M. Bernal and J. G. {de la Torre}},
515    title = {Transport Properties and Hydrodynamic Centers of Rigid Macromolecules
516 <        with Arbitrary Shape},
516 >    with Arbitrary Shape},
517    journal = {Biopolymers},
518    year = {1980},
519    volume = {19},
# Line 523 | Line 523 | Encoding: GBK
523   @ARTICLE{Brunger1984,
524    author = {A. Brunger and C. L. Brooks and M. Karplus},
525    title = {Stochastic Boundary-Conditions for Molecular-Dynamics Simulations
526 <        of St2 Water},
526 >    of St2 Water},
527    journal = {Chemical Physics Letters},
528    year = {1984},
529    volume = {105},
# Line 537 | Line 537 | Encoding: GBK
537   @ARTICLE{Budd1999,
538    author = {C. J. Budd and G. J. Collins and W. Z. Huang and R. D. Russell},
539    title = {Self-similar numerical solutions of the porous-medium equation using
540 <        moving mesh methods},
540 >    moving mesh methods},
541    journal = {Philosophical Transactions of the Royal Society of London Series
542 <        a-Mathematical Physical and Engineering Sciences},
542 >    a-Mathematical Physical and Engineering Sciences},
543    year = {1999},
544    volume = {357},
545    pages = {1047-1077},
546    number = {1754},
547    month = {Apr 15},
548    abstract = {This paper examines a synthesis of adaptive mesh methods with the
549 <        use of symmetry to study a partial differential equation. In particular,
550 <        it considers methods which admit discrete self-similar solutions,
551 <        examining the convergence of these to the true self-similar solution
552 <        as well as their stability. Special attention is given to the nonlinear
553 <        diffusion equation describing flow in a porous medium.},
549 >    use of symmetry to study a partial differential equation. In particular,
550 >    it considers methods which admit discrete self-similar solutions,
551 >    examining the convergence of these to the true self-similar solution
552 >    as well as their stability. Special attention is given to the nonlinear
553 >    diffusion equation describing flow in a porous medium.},
554    annote = {199EE Times Cited:4 Cited References Count:14},
555    issn = {1364-503X},
556    uri = {<Go to ISI>://000080466800005},
# Line 566 | Line 566 | Encoding: GBK
566    number = {21},
567    month = {Dec 1},
568    abstract = {Fluids of hard bent-core molecules have been studied using theory
569 <        and computer simulation. The molecules are composed of two hard
570 <        spherocylinders, with length-to-breadth ratio L/D, joined by their
571 <        ends at an angle 180 degrees - gamma. For L/D = 2 and gamma = 0,10,20
572 <        degrees, the simulations show isotropic, nematic, smectic, and solid
573 <        phases. For L/D = 2 and gamma = 30 degrees, only isotropic, nematic,
574 <        and solid phases are in evidence, which suggests that there is a
575 <        nematic-smectic-solid triple point at an angle in the range 20 degrees
576 <        < gamma < 30 degrees. In all of the orientationally ordered fluid
577 <        phases the order is purely uniaxial. For gamma = 10 degrees and
578 <        20 degrees, at the studied densities, the solid is also uniaxially
579 <        ordered, whilst for gamma = 30 degrees the solid layers are biaxially
580 <        ordered. For L/D = 2 and gamma = 60 degrees and 90 degrees we find
581 <        no spontaneous orientational ordering. This is shown to be due to
582 <        the interlocking of dimer pairs which precludes alignment. We find
583 <        similar results for L/D = 9.5 and gamma = 72 degrees, where an isotropic-biaxial
584 <        nematic transition is predicted by Onsager theory. Simulations in
585 <        the biaxial nematic phase show it to be at least mechanically stable
586 <        with respect to the isotropic phase, however. We have compared the
587 <        quasi-exact simulation results in the isotropic phase with the predicted
588 <        equations of state from three theories: the virial expansion containing
589 <        the second and third virial coefficients; the Parsons-Lee equation
590 <        of state; an application of Wertheim's theory of associating fluids
591 <        in the limit of infinite attractive association energy. For all
592 <        of the molecule elongations and geometries we have simulated, the
593 <        Wertheim theory proved to be the most accurate. Interestingly, the
594 <        isotropic equation of state is virtually independent of the dimer
595 <        bond angle-a feature that is also reflected in the lack of variation
596 <        with angle of the calculated second and third virial coefficients.
597 <        (C) 1999 American Institute of Physics. [S0021-9606(99)50445-5].},
569 >    and computer simulation. The molecules are composed of two hard
570 >    spherocylinders, with length-to-breadth ratio L/D, joined by their
571 >    ends at an angle 180 degrees - gamma. For L/D = 2 and gamma = 0,10,20
572 >    degrees, the simulations show isotropic, nematic, smectic, and solid
573 >    phases. For L/D = 2 and gamma = 30 degrees, only isotropic, nematic,
574 >    and solid phases are in evidence, which suggests that there is a
575 >    nematic-smectic-solid triple point at an angle in the range 20 degrees
576 >    < gamma < 30 degrees. In all of the orientationally ordered fluid
577 >    phases the order is purely uniaxial. For gamma = 10 degrees and
578 >    20 degrees, at the studied densities, the solid is also uniaxially
579 >    ordered, whilst for gamma = 30 degrees the solid layers are biaxially
580 >    ordered. For L/D = 2 and gamma = 60 degrees and 90 degrees we find
581 >    no spontaneous orientational ordering. This is shown to be due to
582 >    the interlocking of dimer pairs which precludes alignment. We find
583 >    similar results for L/D = 9.5 and gamma = 72 degrees, where an isotropic-biaxial
584 >    nematic transition is predicted by Onsager theory. Simulations in
585 >    the biaxial nematic phase show it to be at least mechanically stable
586 >    with respect to the isotropic phase, however. We have compared the
587 >    quasi-exact simulation results in the isotropic phase with the predicted
588 >    equations of state from three theories: the virial expansion containing
589 >    the second and third virial coefficients; the Parsons-Lee equation
590 >    of state; an application of Wertheim's theory of associating fluids
591 >    in the limit of infinite attractive association energy. For all
592 >    of the molecule elongations and geometries we have simulated, the
593 >    Wertheim theory proved to be the most accurate. Interestingly, the
594 >    isotropic equation of state is virtually independent of the dimer
595 >    bond angle-a feature that is also reflected in the lack of variation
596 >    with angle of the calculated second and third virial coefficients.
597 >    (C) 1999 American Institute of Physics. [S0021-9606(99)50445-5].},
598    annote = {255TC Times Cited:24 Cited References Count:38},
599    issn = {0021-9606},
600    uri = {<Go to ISI>://000083685400056},
# Line 610 | Line 610 | Encoding: GBK
610    number = {11},
611    month = {Nov},
612    abstract = {A review is presented of molecular and mesoscopic computer simulations
613 <        of liquid crystalline systems. Molecular simulation approaches applied
614 <        to such systems are described, and the key findings for bulk phase
615 <        behaviour are reported. Following this, recently developed lattice
616 <        Boltzmann approaches to the mesoscale modelling of nemato-dynanics
617 <        are reviewed. This paper concludes with a discussion of possible
618 <        areas for future development in this field.},
613 >    of liquid crystalline systems. Molecular simulation approaches applied
614 >    to such systems are described, and the key findings for bulk phase
615 >    behaviour are reported. Following this, recently developed lattice
616 >    Boltzmann approaches to the mesoscale modelling of nemato-dynanics
617 >    are reviewed. This paper concludes with a discussion of possible
618 >    areas for future development in this field.},
619    annote = {989TU Times Cited:2 Cited References Count:258},
620    issn = {0034-4885},
621    uri = {<Go to ISI>://000233697600004},
# Line 624 | Line 624 | Encoding: GBK
624   @ARTICLE{Carrasco1999,
625    author = {B. Carrasco and J. G. {de la Torre}},
626    title = {Hydrodynamic properties of rigid particles: Comparison of different
627 <        modeling and computational procedures},
627 >    modeling and computational procedures},
628    journal = {Biophysical Journal},
629    year = {1999},
630    volume = {76},
# Line 632 | Line 632 | Encoding: GBK
632    number = {6},
633    month = {Jun},
634    abstract = {The hydrodynamic properties of rigid particles are calculated from
635 <        models composed of spherical elements (beads) using theories developed
636 <        by Kirkwood, Bloomfield, and their coworkers. Bead models have usually
637 <        been built in such a way that the beads fill the volume occupied
638 <        by the particles. Sometimes the beads are few and of varying sizes
639 <        (bead models in the strict sense), and other times there are many
640 <        small beads (filling models). Because hydrodynamic friction takes
641 <        place at the molecular surface, another possibility is to use shell
642 <        models, as originally proposed by Bloomfield. In this work, we have
643 <        developed procedures to build models of the various kinds, and we
644 <        describe the theory and methods for calculating their hydrodynamic
645 <        properties, including approximate methods that may be needed to
646 <        treat models with a very large number of elements. By combining
647 <        the various possibilities of model building and hydrodynamic calculation,
648 <        several strategies can be designed. We have made a quantitative
649 <        comparison of the performance of the various strategies by applying
650 <        them to some test cases, for which the properties are known a priori.
651 <        We provide guidelines and computational tools for bead modeling.},
635 >    models composed of spherical elements (beads) using theories developed
636 >    by Kirkwood, Bloomfield, and their coworkers. Bead models have usually
637 >    been built in such a way that the beads fill the volume occupied
638 >    by the particles. Sometimes the beads are few and of varying sizes
639 >    (bead models in the strict sense), and other times there are many
640 >    small beads (filling models). Because hydrodynamic friction takes
641 >    place at the molecular surface, another possibility is to use shell
642 >    models, as originally proposed by Bloomfield. In this work, we have
643 >    developed procedures to build models of the various kinds, and we
644 >    describe the theory and methods for calculating their hydrodynamic
645 >    properties, including approximate methods that may be needed to
646 >    treat models with a very large number of elements. By combining
647 >    the various possibilities of model building and hydrodynamic calculation,
648 >    several strategies can be designed. We have made a quantitative
649 >    comparison of the performance of the various strategies by applying
650 >    them to some test cases, for which the properties are known a priori.
651 >    We provide guidelines and computational tools for bead modeling.},
652    annote = {200TT Times Cited:46 Cited References Count:57},
653    issn = {0006-3495},
654    uri = {<Go to ISI>://000080556700016},
# Line 657 | Line 657 | Encoding: GBK
657   @ARTICLE{Chandra1999,
658    author = {A. Chandra and T. Ichiye},
659    title = {Dynamical properties of the soft sticky dipole model of water: Molecular
660 <        dynamics simulations},
660 >    dynamics simulations},
661    journal = {Journal of Chemical Physics},
662    year = {1999},
663    volume = {111},
# Line 665 | Line 665 | Encoding: GBK
665    number = {6},
666    month = {Aug 8},
667    abstract = {Dynamical properties of the soft sticky dipole (SSD) model of water
668 <        are calculated by means of molecular dynamics simulations. Since
669 <        this is not a simple point model, the forces and torques arising
670 <        from the SSD potential are derived here. Simulations are carried
671 <        out in the microcanonical ensemble employing the Ewald method for
672 <        the electrostatic interactions. Various time correlation functions
673 <        and dynamical quantities associated with the translational and rotational
674 <        motion of water molecules are evaluated and compared with those
675 <        of two other commonly used models of liquid water, namely the transferable
676 <        intermolecular potential-three points (TIP3P) and simple point charge/extended
677 <        (SPC/E) models, and also with experiments. The dynamical properties
678 <        of the SSD water model are found to be in good agreement with the
679 <        experimental results and appear to be better than the TIP3P and
680 <        SPC/E models in most cases, as has been previously shown for its
681 <        thermodynamic, structural, and dielectric properties. Also, molecular
682 <        dynamics simulations of the SSD model are found to run much faster
683 <        than TIP3P, SPC/E, and other multisite models. (C) 1999 American
684 <        Institute of Physics. [S0021-9606(99)51430-X].},
668 >    are calculated by means of molecular dynamics simulations. Since
669 >    this is not a simple point model, the forces and torques arising
670 >    from the SSD potential are derived here. Simulations are carried
671 >    out in the microcanonical ensemble employing the Ewald method for
672 >    the electrostatic interactions. Various time correlation functions
673 >    and dynamical quantities associated with the translational and rotational
674 >    motion of water molecules are evaluated and compared with those
675 >    of two other commonly used models of liquid water, namely the transferable
676 >    intermolecular potential-three points (TIP3P) and simple point charge/extended
677 >    (SPC/E) models, and also with experiments. The dynamical properties
678 >    of the SSD water model are found to be in good agreement with the
679 >    experimental results and appear to be better than the TIP3P and
680 >    SPC/E models in most cases, as has been previously shown for its
681 >    thermodynamic, structural, and dielectric properties. Also, molecular
682 >    dynamics simulations of the SSD model are found to run much faster
683 >    than TIP3P, SPC/E, and other multisite models. (C) 1999 American
684 >    Institute of Physics. [S0021-9606(99)51430-X].},
685    annote = {221EN Times Cited:14 Cited References Count:66},
686    issn = {0021-9606},
687    uri = {<Go to ISI>://000081711200038},
# Line 711 | Line 711 | Encoding: GBK
711    number = {1-2},
712    month = {Jan},
713    abstract = {We investigate the asymptotic behavior of systems of nonlinear differential
714 <        equations and introduce a family of mixed methods from combinations
715 <        of explicit Runge-Kutta methods. These methods have better stability
716 <        behavior than traditional Runge-Kutta methods and generally extend
717 <        the range of validity of the calculated solutions. These methods
718 <        also give a way of determining if the numerical solutions are real
719 <        or spurious. Emphasis is put on examples coming from mathematical
720 <        models in ecology. (C) 2002 IMACS. Published by Elsevier Science
721 <        B.V. All rights reserved.},
714 >    equations and introduce a family of mixed methods from combinations
715 >    of explicit Runge-Kutta methods. These methods have better stability
716 >    behavior than traditional Runge-Kutta methods and generally extend
717 >    the range of validity of the calculated solutions. These methods
718 >    also give a way of determining if the numerical solutions are real
719 >    or spurious. Emphasis is put on examples coming from mathematical
720 >    models in ecology. (C) 2002 IMACS. Published by Elsevier Science
721 >    B.V. All rights reserved.},
722    annote = {633ZD Times Cited:0 Cited References Count:9},
723    issn = {0168-9274},
724    uri = {<Go to ISI>://000180314200002},
# Line 727 | Line 727 | Encoding: GBK
727   @ARTICLE{Cheung2004,
728    author = {D. L. Cheung and S. J. Clark and M. R. Wilson},
729    title = {Calculation of flexoelectric coefficients for a nematic liquid crystal
730 <        by atomistic simulation},
730 >    by atomistic simulation},
731    journal = {Journal of Chemical Physics},
732    year = {2004},
733    volume = {121},
# Line 735 | Line 735 | Encoding: GBK
735    number = {18},
736    month = {Nov 8},
737    abstract = {Equilibrium molecular dynamics calculations have been performed for
738 <        the liquid crystal molecule n-4-(trans-4-n-pentylcyclohexyl)benzonitrile
739 <        (PCH5) using a fully atomistic model. Simulation data have been
740 <        obtained for a series of temperatures in the nematic phase. The
741 <        simulation data have been used to calculate the flexoelectric coefficients
742 <        e(s) and e(b) using the linear response formalism of Osipov and
743 <        Nemtsov [M. A. Osipov and V. B. Nemtsov, Sov. Phys. Crstallogr.
744 <        31, 125 (1986)]. The temperature and order parameter dependence
745 <        of e(s) and e(b) are examined, as are separate contributions from
746 <        different intermolecular interactions. Values of e(s) and e(b) calculated
747 <        from simulation are consistent with those found from experiment.
748 <        (C) 2004 American Institute of Physics.},
738 >    the liquid crystal molecule n-4-(trans-4-n-pentylcyclohexyl)benzonitrile
739 >    (PCH5) using a fully atomistic model. Simulation data have been
740 >    obtained for a series of temperatures in the nematic phase. The
741 >    simulation data have been used to calculate the flexoelectric coefficients
742 >    e(s) and e(b) using the linear response formalism of Osipov and
743 >    Nemtsov [M. A. Osipov and V. B. Nemtsov, Sov. Phys. Crstallogr.
744 >    31, 125 (1986)]. The temperature and order parameter dependence
745 >    of e(s) and e(b) are examined, as are separate contributions from
746 >    different intermolecular interactions. Values of e(s) and e(b) calculated
747 >    from simulation are consistent with those found from experiment.
748 >    (C) 2004 American Institute of Physics.},
749    annote = {866UM Times Cited:4 Cited References Count:61},
750    issn = {0021-9606},
751    uri = {<Go to ISI>://000224798900053},
# Line 761 | Line 761 | Encoding: GBK
761    number = {1-2},
762    month = {Apr 15},
763    abstract = {Equilibrium molecular dynamics calculations have been performed for
764 <        the liquid crystal molecule n-4-(trans-4-npentylcyclohexyl)benzonitrile
765 <        (PCH5) using a fully atomistic model. Simulation data has been obtained
766 <        for a series of temperatures in the nematic phase. The rotational
767 <        viscosity co-efficient gamma(1), has been calculated using the angular
768 <        velocity correlation function of the nematic director, n, the mean
769 <        squared diffusion of n and statistical mechanical methods based
770 <        on the rotational diffusion co-efficient. We find good agreement
771 <        between the first two methods and experimental values. (C) 2002
772 <        Published by Elsevier Science B.V.},
764 >    the liquid crystal molecule n-4-(trans-4-npentylcyclohexyl)benzonitrile
765 >    (PCH5) using a fully atomistic model. Simulation data has been obtained
766 >    for a series of temperatures in the nematic phase. The rotational
767 >    viscosity co-efficient gamma(1), has been calculated using the angular
768 >    velocity correlation function of the nematic director, n, the mean
769 >    squared diffusion of n and statistical mechanical methods based
770 >    on the rotational diffusion co-efficient. We find good agreement
771 >    between the first two methods and experimental values. (C) 2002
772 >    Published by Elsevier Science B.V.},
773    annote = {547KF Times Cited:8 Cited References Count:31},
774    issn = {0009-2614},
775    uri = {<Go to ISI>://000175331000020},
# Line 778 | Line 778 | Encoding: GBK
778   @ARTICLE{Chin2004,
779    author = {S. A. Chin},
780    title = {Dynamical multiple-time stepping methods for overcoming resonance
781 <        instabilities},
781 >    instabilities},
782    journal = {Journal of Chemical Physics},
783    year = {2004},
784    volume = {120},
# Line 786 | Line 786 | Encoding: GBK
786    number = {1},
787    month = {Jan 1},
788    abstract = {Current molecular dynamics simulations of biomolecules using multiple
789 <        time steps to update the slowly changing force are hampered by instabilities
790 <        beginning at time steps near the half period of the fastest vibrating
791 <        mode. These #resonance# instabilities have became a critical barrier
792 <        preventing the long time simulation of biomolecular dynamics. Attempts
793 <        to tame these instabilities by altering the slowly changing force
794 <        and efforts to damp them out by Langevin dynamics do not address
795 <        the fundamental cause of these instabilities. In this work, we trace
796 <        the instability to the nonanalytic character of the underlying spectrum
797 <        and show that a correct splitting of the Hamiltonian, which renders
798 <        the spectrum analytic, restores stability. The resulting Hamiltonian
799 <        dictates that in addition to updating the momentum due to the slowly
800 <        changing force, one must also update the position with a modified
801 <        mass. Thus multiple-time stepping must be done dynamically. (C)
802 <        2004 American Institute of Physics.},
789 >    time steps to update the slowly changing force are hampered by instabilities
790 >    beginning at time steps near the half period of the fastest vibrating
791 >    mode. These #resonance# instabilities have became a critical barrier
792 >    preventing the long time simulation of biomolecular dynamics. Attempts
793 >    to tame these instabilities by altering the slowly changing force
794 >    and efforts to damp them out by Langevin dynamics do not address
795 >    the fundamental cause of these instabilities. In this work, we trace
796 >    the instability to the nonanalytic character of the underlying spectrum
797 >    and show that a correct splitting of the Hamiltonian, which renders
798 >    the spectrum analytic, restores stability. The resulting Hamiltonian
799 >    dictates that in addition to updating the momentum due to the slowly
800 >    changing force, one must also update the position with a modified
801 >    mass. Thus multiple-time stepping must be done dynamically. (C)
802 >    2004 American Institute of Physics.},
803    annote = {757TK Times Cited:1 Cited References Count:22},
804    issn = {0021-9606},
805    uri = {<Go to ISI>://000187577400003},
# Line 808 | Line 808 | Encoding: GBK
808   @ARTICLE{Cook2000,
809    author = {M. J. Cook and M. R. Wilson},
810    title = {Simulation studies of dipole correlation in the isotropic liquid
811 <        phase},
811 >    phase},
812    journal = {Liquid Crystals},
813    year = {2000},
814    volume = {27},
# Line 816 | Line 816 | Encoding: GBK
816    number = {12},
817    month = {Dec},
818    abstract = {The Kirkwood correlation factor g(1) determines the preference for
819 <        local parallel or antiparallel dipole association in the isotropic
820 <        phase. Calamitic mesogens with longitudinal dipole moments and Kirkwood
821 <        factors greater than 1 have an enhanced effective dipole moment
822 <        along the molecular long axis. This leads to higher values of Delta
823 <        epsilon in the nematic phase. This paper describes state-of-the-art
824 <        molecular dynamics simulations of two calamitic mesogens 4-(trans-4-n-pentylcyclohexyl)benzonitrile
825 <        (PCH5) and 4-(trans-4-n-pentylcyclohexyl) chlorobenzene (PCH5-Cl)
826 <        in the isotropic liquid phase using an all-atom force field and
827 <        taking long range electrostatics into account using an Ewald summation.
828 <        Using this methodology, PCH5 is seen to prefer antiparallel dipole
829 <        alignment with a negative g(1) and PCH5-Cl is seen to prefer parallel
830 <        dipole alignment with a positive g(1); this is in accordance with
831 <        experimental dielectric measurements. Analysis of the molecular
832 <        dynamics trajectories allows an assessment of why these molecules
833 <        behave differently.},
819 >    local parallel or antiparallel dipole association in the isotropic
820 >    phase. Calamitic mesogens with longitudinal dipole moments and Kirkwood
821 >    factors greater than 1 have an enhanced effective dipole moment
822 >    along the molecular long axis. This leads to higher values of Delta
823 >    epsilon in the nematic phase. This paper describes state-of-the-art
824 >    molecular dynamics simulations of two calamitic mesogens 4-(trans-4-n-pentylcyclohexyl)benzonitrile
825 >    (PCH5) and 4-(trans-4-n-pentylcyclohexyl) chlorobenzene (PCH5-Cl)
826 >    in the isotropic liquid phase using an all-atom force field and
827 >    taking long range electrostatics into account using an Ewald summation.
828 >    Using this methodology, PCH5 is seen to prefer antiparallel dipole
829 >    alignment with a negative g(1) and PCH5-Cl is seen to prefer parallel
830 >    dipole alignment with a positive g(1); this is in accordance with
831 >    experimental dielectric measurements. Analysis of the molecular
832 >    dynamics trajectories allows an assessment of why these molecules
833 >    behave differently.},
834    annote = {376BF Times Cited:10 Cited References Count:16},
835    issn = {0267-8292},
836    uri = {<Go to ISI>://000165437800002},
# Line 839 | Line 839 | Encoding: GBK
839   @ARTICLE{Cui2003,
840    author = {B. X. Cui and M. Y. Shen and K. F. Freed},
841    title = {Folding and misfolding of the papillomavirus E6 interacting peptide
842 <        E6ap},
842 >    E6ap},
843    journal = {Proceedings of the National Academy of Sciences of the United States
844 <        of America},
844 >    of America},
845    year = {2003},
846    volume = {100},
847    pages = {7087-7092},
848    number = {12},
849    month = {Jun 10},
850    abstract = {All-atom Langevin dynamics simulations have been performed to study
851 <        the folding pathways of the 18-residue binding domain fragment E6ap
852 <        of the human papillomavirus E6 interacting peptide. Six independent
853 <        folding trajectories, with a total duration of nearly 2 mus, all
854 <        lead to the same native state in which the E6ap adopts a fluctuating
855 <        a-helix structure in the central portion (Ser-4-Leu-13) but with
856 <        very flexible N and C termini. Simulations starting from different
857 <        core configurations exhibit the E6ap folding dynamics as either
858 <        a two- or three-state folder with an intermediate misfolded state.
859 <        The essential leucine hydrophobic core (Leu-9, Leu-12, and Leu-13)
860 <        is well conserved in the native-state structure but absent in the
861 <        intermediate structure, suggesting that the leucine core is not
862 <        only essential for the binding activity of E6ap but also important
863 <        for the stability of the native structure. The free energy landscape
864 <        reveals a significant barrier between the basins separating the
865 <        native and misfolded states. We also discuss the various underlying
866 <        forces that drive the peptide into its native state.},
851 >    the folding pathways of the 18-residue binding domain fragment E6ap
852 >    of the human papillomavirus E6 interacting peptide. Six independent
853 >    folding trajectories, with a total duration of nearly 2 mus, all
854 >    lead to the same native state in which the E6ap adopts a fluctuating
855 >    a-helix structure in the central portion (Ser-4-Leu-13) but with
856 >    very flexible N and C termini. Simulations starting from different
857 >    core configurations exhibit the E6ap folding dynamics as either
858 >    a two- or three-state folder with an intermediate misfolded state.
859 >    The essential leucine hydrophobic core (Leu-9, Leu-12, and Leu-13)
860 >    is well conserved in the native-state structure but absent in the
861 >    intermediate structure, suggesting that the leucine core is not
862 >    only essential for the binding activity of E6ap but also important
863 >    for the stability of the native structure. The free energy landscape
864 >    reveals a significant barrier between the basins separating the
865 >    native and misfolded states. We also discuss the various underlying
866 >    forces that drive the peptide into its native state.},
867    annote = {689LC Times Cited:3 Cited References Count:48},
868    issn = {0027-8424},
869    uri = {<Go to ISI>://000183493500037},
# Line 879 | Line 879 | Encoding: GBK
879    number = {1},
880    month = {Jan 1},
881    abstract = {We study the slow phase of thermally activated magnetic relaxation
882 <        in finite two-dimensional ensembles of dipolar interacting ferromagnetic
883 <        nanoparticles whose easy axes of magnetization are perpendicular
884 <        to the distribution plane. We develop a method to numerically simulate
885 <        the magnetic relaxation for the case that the smallest heights of
886 <        the potential barriers between the equilibrium directions of the
887 <        nanoparticle magnetic moments are much larger than the thermal energy.
888 <        Within this framework, we analyze in detail the role that the correlations
889 <        of the nanoparticle magnetic moments and the finite size of the
890 <        nanoparticle ensemble play in magnetic relaxation.},
882 >    in finite two-dimensional ensembles of dipolar interacting ferromagnetic
883 >    nanoparticles whose easy axes of magnetization are perpendicular
884 >    to the distribution plane. We develop a method to numerically simulate
885 >    the magnetic relaxation for the case that the smallest heights of
886 >    the potential barriers between the equilibrium directions of the
887 >    nanoparticle magnetic moments are much larger than the thermal energy.
888 >    Within this framework, we analyze in detail the role that the correlations
889 >    of the nanoparticle magnetic moments and the finite size of the
890 >    nanoparticle ensemble play in magnetic relaxation.},
891    annote = {642XH Times Cited:11 Cited References Count:31},
892    issn = {1098-0121},
893    uri = {<Go to ISI>://000180830400056},
# Line 903 | Line 903 | Encoding: GBK
903    number = {1},
904    month = {Jan},
905    abstract = {To explore the origin of the large-scale motion of triosephosphate
906 <        isomerase's flexible loop (residues 166 to 176) at the active site,
907 <        several simulation protocols are employed both for the free enzyme
908 <        in vacuo and for the free enzyme with some solvent modeling: high-temperature
909 <        Langevin dynamics simulations, sampling by a #dynamics##driver#
910 <        approach, and potential-energy surface calculations. Our focus is
911 <        on obtaining the energy barrier to the enzyme's motion and establishing
912 <        the nature of the loop movement. Previous calculations did not determine
913 <        this energy barrier and the effect of solvent on the barrier. High-temperature
914 <        molecular dynamics simulations and crystallographic studies have
915 <        suggested a rigid-body motion with two hinges located at both ends
916 <        of the loop; Brownian dynamics simulations at room temperature pointed
917 <        to a very flexible behavior. The present simulations and analyses
918 <        reveal that although solute/solvent hydrogen bonds play a crucial
919 <        role in lowering the energy along the pathway, there still remains
920 <        a high activation barrier, This finding clearly indicates that,
921 <        if the loop opens and closes in the absence of a substrate at standard
922 <        conditions (e.g., room temperature, appropriate concentration of
923 <        isomerase), the time scale for transition is not in the nanosecond
924 <        but rather the microsecond range. Our results also indicate that
925 <        in the context of spontaneous opening in the free enzyme, the motion
926 <        is of rigid-body type and that the specific interaction between
927 <        residues Ala(176) and Tyr(208) plays a crucial role in the loop
928 <        opening/closing mechanism.},
906 >    isomerase's flexible loop (residues 166 to 176) at the active site,
907 >    several simulation protocols are employed both for the free enzyme
908 >    in vacuo and for the free enzyme with some solvent modeling: high-temperature
909 >    Langevin dynamics simulations, sampling by a #dynamics##driver#
910 >    approach, and potential-energy surface calculations. Our focus is
911 >    on obtaining the energy barrier to the enzyme's motion and establishing
912 >    the nature of the loop movement. Previous calculations did not determine
913 >    this energy barrier and the effect of solvent on the barrier. High-temperature
914 >    molecular dynamics simulations and crystallographic studies have
915 >    suggested a rigid-body motion with two hinges located at both ends
916 >    of the loop; Brownian dynamics simulations at room temperature pointed
917 >    to a very flexible behavior. The present simulations and analyses
918 >    reveal that although solute/solvent hydrogen bonds play a crucial
919 >    role in lowering the energy along the pathway, there still remains
920 >    a high activation barrier, This finding clearly indicates that,
921 >    if the loop opens and closes in the absence of a substrate at standard
922 >    conditions (e.g., room temperature, appropriate concentration of
923 >    isomerase), the time scale for transition is not in the nanosecond
924 >    but rather the microsecond range. Our results also indicate that
925 >    in the context of spontaneous opening in the free enzyme, the motion
926 >    is of rigid-body type and that the specific interaction between
927 >    residues Ala(176) and Tyr(208) plays a crucial role in the loop
928 >    opening/closing mechanism.},
929    annote = {Zl046 Times Cited:30 Cited References Count:29},
930    issn = {0006-3495},
931    uri = {<Go to ISI>://000073393400009},
# Line 941 | Line 941 | Encoding: GBK
941    number = {15},
942    month = {Oct 15},
943    abstract = {Rigid body molecular models possess symplectic structure and time-reversal
944 <        symmetry. Standard numerical integration methods destroy both properties,
945 <        introducing nonphysical dynamical behavior such as numerically induced
946 <        dissipative states and drift in the energy during long term simulations.
947 <        This article describes the construction, implementation, and practical
948 <        application of fast explicit symplectic-reversible integrators for
949 <        multiple rigid body molecular simulations, These methods use a reduction
950 <        to Euler equations for the free rigid body, together with a symplectic
951 <        splitting technique. In every time step, the orientational dynamics
952 <        of each rigid body is integrated by a sequence of planar rotations.
953 <        Besides preserving the symplectic and reversible structures of the
954 <        flow, this scheme accurately conserves the total angular momentum
955 <        of a system of interacting rigid bodies. Excellent energy conservation
956 <        fan be obtained relative to traditional methods, especially in long-time
957 <        simulations. The method is implemented in a research code, ORIENT
958 <        and compared with a quaternion/extrapolation scheme for the TIP4P
959 <        model of water. Our experiments show that the symplectic-reversible
960 <        scheme is far superior to the more traditional quaternion method.
961 <        (C) 1997 American Institute of Physics.},
944 >    symmetry. Standard numerical integration methods destroy both properties,
945 >    introducing nonphysical dynamical behavior such as numerically induced
946 >    dissipative states and drift in the energy during long term simulations.
947 >    This article describes the construction, implementation, and practical
948 >    application of fast explicit symplectic-reversible integrators for
949 >    multiple rigid body molecular simulations, These methods use a reduction
950 >    to Euler equations for the free rigid body, together with a symplectic
951 >    splitting technique. In every time step, the orientational dynamics
952 >    of each rigid body is integrated by a sequence of planar rotations.
953 >    Besides preserving the symplectic and reversible structures of the
954 >    flow, this scheme accurately conserves the total angular momentum
955 >    of a system of interacting rigid bodies. Excellent energy conservation
956 >    fan be obtained relative to traditional methods, especially in long-time
957 >    simulations. The method is implemented in a research code, ORIENT
958 >    and compared with a quaternion/extrapolation scheme for the TIP4P
959 >    model of water. Our experiments show that the symplectic-reversible
960 >    scheme is far superior to the more traditional quaternion method.
961 >    (C) 1997 American Institute of Physics.},
962    annote = {Ya587 Times Cited:35 Cited References Count:32},
963    issn = {0021-9606},
964    uri = {<Go to ISI>://A1997YA58700024},
# Line 967 | Line 967 | Encoding: GBK
967   @ARTICLE{Edwards2005,
968    author = {S. A. Edwards and D. R. M. Williams},
969    title = {Stretching a single diblock copolymer in a selective solvent: Langevin
970 <        dynamics simulations},
970 >    dynamics simulations},
971    journal = {Macromolecules},
972    year = {2005},
973    volume = {38},
# Line 975 | Line 975 | Encoding: GBK
975    number = {25},
976    month = {Dec 13},
977    abstract = {Using the Langevin dynamics technique, we have carried out simulations
978 <        of a single-chain flexible diblock copolymer. The polymer consists
979 <        of two blocks of equal length, one very poorly solvated and the
980 <        other close to theta-conditions. We study what happens when such
981 <        a polymer is stretched, for a range of different stretching speeds,
982 <        and correlate our observations with features in the plot of force
983 <        vs extension. We find that at slow speeds this force profile does
984 <        not increase monotonically, in disagreement with earlier predictions,
985 <        and that at high speeds there is a strong dependence on which end
986 <        of the polymer is pulled, as well as a high level of hysteresis.},
978 >    of a single-chain flexible diblock copolymer. The polymer consists
979 >    of two blocks of equal length, one very poorly solvated and the
980 >    other close to theta-conditions. We study what happens when such
981 >    a polymer is stretched, for a range of different stretching speeds,
982 >    and correlate our observations with features in the plot of force
983 >    vs extension. We find that at slow speeds this force profile does
984 >    not increase monotonically, in disagreement with earlier predictions,
985 >    and that at high speeds there is a strong dependence on which end
986 >    of the polymer is pulled, as well as a high level of hysteresis.},
987    annote = {992EC Times Cited:0 Cited References Count:13},
988    issn = {0024-9297},
989    uri = {<Go to ISI>://000233866200035},
# Line 992 | Line 992 | Encoding: GBK
992   @ARTICLE{Egberts1988,
993    author = {E. Egberts and H. J. C. Berendsen},
994    title = {Molecular-Dynamics Simulation of a Smectic Liquid-Crystal with Atomic
995 <        Detail},
995 >    Detail},
996    journal = {Journal of Chemical Physics},
997    year = {1988},
998    volume = {89},
# Line 1020 | Line 1020 | Encoding: GBK
1020   @ARTICLE{Fennell2004,
1021    author = {C. J. Fennell and J. D. Gezelter},
1022    title = {On the structural and transport properties of the soft sticky dipole
1023 <        and related single-point water models},
1023 >    and related single-point water models},
1024    journal = {Journal of Chemical Physics},
1025    year = {2004},
1026    volume = {120},
# Line 1028 | Line 1028 | Encoding: GBK
1028    number = {19},
1029    month = {May 15},
1030    abstract = {The density maximum and temperature dependence of the self-diffusion
1031 <        constant were investigated for the soft sticky dipole (SSD) water
1032 <        model and two related reparametrizations of this single-point model.
1033 <        A combination of microcanonical and isobaric-isothermal molecular
1034 <        dynamics simulations was used to calculate these properties, both
1035 <        with and without the use of reaction field to handle long-range
1036 <        electrostatics. The isobaric-isothermal simulations of the melting
1037 <        of both ice-I-h and ice-I-c showed a density maximum near 260 K.
1038 <        In most cases, the use of the reaction field resulted in calculated
1039 <        densities which were significantly lower than experimental densities.
1040 <        Analysis of self-diffusion constants shows that the original SSD
1041 <        model captures the transport properties of experimental water very
1042 <        well in both the normal and supercooled liquid regimes. We also
1043 <        present our reparametrized versions of SSD for use both with the
1044 <        reaction field or without any long-range electrostatic corrections.
1045 <        These are called the SSD/RF and SSD/E models, respectively. These
1046 <        modified models were shown to maintain or improve upon the experimental
1047 <        agreement with the structural and transport properties that can
1048 <        be obtained with either the original SSD or the density-corrected
1049 <        version of the original model (SSD1). Additionally, a novel low-density
1050 <        ice structure is presented which appears to be the most stable ice
1051 <        structure for the entire SSD family. (C) 2004 American Institute
1052 <        of Physics.},
1031 >    constant were investigated for the soft sticky dipole (SSD) water
1032 >    model and two related reparametrizations of this single-point model.
1033 >    A combination of microcanonical and isobaric-isothermal molecular
1034 >    dynamics simulations was used to calculate these properties, both
1035 >    with and without the use of reaction field to handle long-range
1036 >    electrostatics. The isobaric-isothermal simulations of the melting
1037 >    of both ice-I-h and ice-I-c showed a density maximum near 260 K.
1038 >    In most cases, the use of the reaction field resulted in calculated
1039 >    densities which were significantly lower than experimental densities.
1040 >    Analysis of self-diffusion constants shows that the original SSD
1041 >    model captures the transport properties of experimental water very
1042 >    well in both the normal and supercooled liquid regimes. We also
1043 >    present our reparametrized versions of SSD for use both with the
1044 >    reaction field or without any long-range electrostatic corrections.
1045 >    These are called the SSD/RF and SSD/E models, respectively. These
1046 >    modified models were shown to maintain or improve upon the experimental
1047 >    agreement with the structural and transport properties that can
1048 >    be obtained with either the original SSD or the density-corrected
1049 >    version of the original model (SSD1). Additionally, a novel low-density
1050 >    ice structure is presented which appears to be the most stable ice
1051 >    structure for the entire SSD family. (C) 2004 American Institute
1052 >    of Physics.},
1053    annote = {816YY Times Cited:5 Cited References Count:39},
1054    issn = {0021-9606},
1055    uri = {<Go to ISI>://000221146400032},
# Line 1058 | Line 1058 | Encoding: GBK
1058   @ARTICLE{Fernandes2002,
1059    author = {M. X. Fernandes and J. G. {de la Torre}},
1060    title = {Brownian dynamics simulation of rigid particles of arbitrary shape
1061 <        in external fields},
1061 >    in external fields},
1062    journal = {Biophysical Journal},
1063    year = {2002},
1064    volume = {83},
# Line 1066 | Line 1066 | Encoding: GBK
1066    number = {6},
1067    month = {Dec},
1068    abstract = {We have developed a Brownian dynamics simulation algorithm to generate
1069 <        Brownian trajectories of an isolated, rigid particle of arbitrary
1070 <        shape in the presence of electric fields or any other external agents.
1071 <        Starting from the generalized diffusion tensor, which can be calculated
1072 <        with the existing HYDRO software, the new program BROWNRIG (including
1073 <        a case-specific subprogram for the external agent) carries out a
1074 <        simulation that is analyzed later to extract the observable dynamic
1075 <        properties. We provide a variety of examples of utilization of this
1076 <        method, which serve as tests of its performance, and also illustrate
1077 <        its applicability. Examples include free diffusion, transport in
1078 <        an electric field, and diffusion in a restricting environment.},
1069 >    Brownian trajectories of an isolated, rigid particle of arbitrary
1070 >    shape in the presence of electric fields or any other external agents.
1071 >    Starting from the generalized diffusion tensor, which can be calculated
1072 >    with the existing HYDRO software, the new program BROWNRIG (including
1073 >    a case-specific subprogram for the external agent) carries out a
1074 >    simulation that is analyzed later to extract the observable dynamic
1075 >    properties. We provide a variety of examples of utilization of this
1076 >    method, which serve as tests of its performance, and also illustrate
1077 >    its applicability. Examples include free diffusion, transport in
1078 >    an electric field, and diffusion in a restricting environment.},
1079    annote = {633AD Times Cited:2 Cited References Count:43},
1080    issn = {0006-3495},
1081    uri = {<Go to ISI>://000180256300012},
# Line 1084 | Line 1084 | Encoding: GBK
1084   @ARTICLE{Gay1981,
1085    author = {J. G. Gay and B. J. Berne},
1086    title = {Modification of the Overlap Potential to Mimic a Linear Site-Site
1087 <        Potential},
1087 >    Potential},
1088    journal = {Journal of Chemical Physics},
1089    year = {1981},
1090    volume = {74},
# Line 1105 | Line 1105 | Encoding: GBK
1105    number = {6},
1106    month = {Nov},
1107    abstract = {To investigate the influence of inertial effects on the dynamics of
1108 <        an assembly of beads subjected to rigid constraints and placed in
1109 <        a buffer medium, a convenient method to introduce suitable generalized
1110 <        coordinates is presented. Without any restriction on the nature
1111 <        of the soft forces involved (both stochastic and deterministic),
1112 <        pertinent Langevin equations are derived. Provided that the Brownian
1113 <        forces are Gaussian and Markovian, the corresponding Fokker-Planck
1114 <        equation (FPE) is obtained in the complete phase space of generalized
1115 <        coordinates and momenta. The correct short time behavior for correlation
1116 <        functions (CFs) of generalized coordinates is established, and the
1117 <        diffusion equation with memory (DEM) is deduced from the FPE in
1118 <        the high friction Limit. The DEM is invoked to perform illustrative
1119 <        calculations in two dimensions of the orientational CFs for once
1120 <        broken nonrigid rods immobilized on a surface. These calculations
1121 <        reveal that the CFs under certain conditions exhibit an oscillatory
1122 <        behavior, which is irreproducible within the standard diffusion
1123 <        equation. Several methods are considered for the approximate solution
1124 <        of the DEM, and their application to three dimensional DEMs is discussed.},
1108 >    an assembly of beads subjected to rigid constraints and placed in
1109 >    a buffer medium, a convenient method to introduce suitable generalized
1110 >    coordinates is presented. Without any restriction on the nature
1111 >    of the soft forces involved (both stochastic and deterministic),
1112 >    pertinent Langevin equations are derived. Provided that the Brownian
1113 >    forces are Gaussian and Markovian, the corresponding Fokker-Planck
1114 >    equation (FPE) is obtained in the complete phase space of generalized
1115 >    coordinates and momenta. The correct short time behavior for correlation
1116 >    functions (CFs) of generalized coordinates is established, and the
1117 >    diffusion equation with memory (DEM) is deduced from the FPE in
1118 >    the high friction Limit. The DEM is invoked to perform illustrative
1119 >    calculations in two dimensions of the orientational CFs for once
1120 >    broken nonrigid rods immobilized on a surface. These calculations
1121 >    reveal that the CFs under certain conditions exhibit an oscillatory
1122 >    behavior, which is irreproducible within the standard diffusion
1123 >    equation. Several methods are considered for the approximate solution
1124 >    of the DEM, and their application to three dimensional DEMs is discussed.},
1125    annote = {257MM Times Cited:2 Cited References Count:82},
1126    issn = {1022-1344},
1127    uri = {<Go to ISI>://000083785700002},
# Line 1138 | Line 1138 | Encoding: GBK
1138  
1139   @ARTICLE{Gray2003,
1140    author = {J. J. Gray and S. Moughon and C. Wang and O. Schueler-Furman and
1141 <        B. Kuhlman and C. A. Rohl and D. Baker},
1141 >    B. Kuhlman and C. A. Rohl and D. Baker},
1142    title = {Protein-protein docking with simultaneous optimization of rigid-body
1143 <        displacement and side-chain conformations},
1143 >    displacement and side-chain conformations},
1144    journal = {Journal of Molecular Biology},
1145    year = {2003},
1146    volume = {331},
# Line 1148 | Line 1148 | Encoding: GBK
1148    number = {1},
1149    month = {Aug 1},
1150    abstract = {Protein-protein docking algorithms provide a means to elucidate structural
1151 <        details for presently unknown complexes. Here, we present and evaluate
1152 <        a new method to predict protein-protein complexes from the coordinates
1153 <        of the unbound monomer components. The method employs a low-resolution,
1154 <        rigid-body, Monte Carlo search followed by simultaneous optimization
1155 <        of backbone displacement and side-chain conformations using Monte
1156 <        Carlo minimization. Up to 10(5) independent simulations are carried
1157 <        out, and the resulting #decoys# are ranked using an energy function
1158 <        dominated by van der Waals interactions, an implicit solvation model,
1159 <        and an orientation-dependent hydrogen bonding potential. Top-ranking
1160 <        decoys are clustered to select the final predictions. Small-perturbation
1161 <        studies reveal the formation of binding funnels in 42 of 54 cases
1162 <        using coordinates derived from the bound complexes and in 32 of
1163 <        54 cases using independently determined coordinates of one or both
1164 <        monomers. Experimental binding affinities correlate with the calculated
1165 <        score function and explain the predictive success or failure of
1166 <        many targets. Global searches using one or both unbound components
1167 <        predict at least 25% of the native residue-residue contacts in 28
1168 <        of the 32 cases where binding funnels exist. The results suggest
1169 <        that the method may soon be useful for generating models of biologically
1170 <        important complexes from the structures of the isolated components,
1171 <        but they also highlight the challenges that must be met to achieve
1172 <        consistent and accurate prediction of protein-protein interactions.
1173 <        (C) 2003 Elsevier Ltd. All rights reserved.},
1151 >    details for presently unknown complexes. Here, we present and evaluate
1152 >    a new method to predict protein-protein complexes from the coordinates
1153 >    of the unbound monomer components. The method employs a low-resolution,
1154 >    rigid-body, Monte Carlo search followed by simultaneous optimization
1155 >    of backbone displacement and side-chain conformations using Monte
1156 >    Carlo minimization. Up to 10(5) independent simulations are carried
1157 >    out, and the resulting #decoys# are ranked using an energy function
1158 >    dominated by van der Waals interactions, an implicit solvation model,
1159 >    and an orientation-dependent hydrogen bonding potential. Top-ranking
1160 >    decoys are clustered to select the final predictions. Small-perturbation
1161 >    studies reveal the formation of binding funnels in 42 of 54 cases
1162 >    using coordinates derived from the bound complexes and in 32 of
1163 >    54 cases using independently determined coordinates of one or both
1164 >    monomers. Experimental binding affinities correlate with the calculated
1165 >    score function and explain the predictive success or failure of
1166 >    many targets. Global searches using one or both unbound components
1167 >    predict at least 25% of the native residue-residue contacts in 28
1168 >    of the 32 cases where binding funnels exist. The results suggest
1169 >    that the method may soon be useful for generating models of biologically
1170 >    important complexes from the structures of the isolated components,
1171 >    but they also highlight the challenges that must be met to achieve
1172 >    consistent and accurate prediction of protein-protein interactions.
1173 >    (C) 2003 Elsevier Ltd. All rights reserved.},
1174    annote = {704QL Times Cited:48 Cited References Count:60},
1175    issn = {0022-2836},
1176    uri = {<Go to ISI>://000184351300022},
# Line 1186 | Line 1186 | Encoding: GBK
1186    number = {5174},
1187    month = {Aug 12},
1188    abstract = {Some of the recently developed fast summation methods that have arisen
1189 <        in scientific computing are described. These methods require an
1190 <        amount of work proportional to N or N log N to evaluate all pairwise
1191 <        interactions in an ensemble of N particles. Traditional methods,
1192 <        by contrast, require an amount of work proportional to N-2. AS a
1193 <        result, large-scale simulations can be carried out using only modest
1194 <        computer resources. In combination with supercomputers, it is possible
1195 <        to address questions that were previously out of reach. Problems
1196 <        from diffusion, gravitation, and wave propagation are considered.},
1189 >    in scientific computing are described. These methods require an
1190 >    amount of work proportional to N or N log N to evaluate all pairwise
1191 >    interactions in an ensemble of N particles. Traditional methods,
1192 >    by contrast, require an amount of work proportional to N-2. AS a
1193 >    result, large-scale simulations can be carried out using only modest
1194 >    computer resources. In combination with supercomputers, it is possible
1195 >    to address questions that were previously out of reach. Problems
1196 >    from diffusion, gravitation, and wave propagation are considered.},
1197    annote = {Pb499 Times Cited:99 Cited References Count:44},
1198    issn = {0036-8075},
1199    uri = {<Go to ISI>://A1994PB49900031},
# Line 1223 | Line 1223 | Encoding: GBK
1223    number = {4},
1224    month = {Jun},
1225    abstract = {Backward error analysis is a useful tool for the study of numerical
1226 <        approximations to ordinary differential equations. The numerical
1227 <        solution is formally interpreted as the exact solution of a perturbed
1228 <        differential equation, given as a formal and usually divergent series
1229 <        in powers of the step size. For a rigorous analysis, this series
1230 <        has to be truncated. In this article we study the influence of this
1231 <        truncation to the difference between the numerical solution and
1232 <        the exact solution of the perturbed differential equation. Results
1233 <        on the long-time behaviour of numerical solutions are obtained in
1234 <        this way. We present applications to the numerical phase portrait
1235 <        near hyperbolic equilibrium points, to asymptotically stable periodic
1236 <        orbits and Hopf bifurcation, and to energy conservation and approximation
1237 <        of invariant tori in Hamiltonian systems.},
1226 >    approximations to ordinary differential equations. The numerical
1227 >    solution is formally interpreted as the exact solution of a perturbed
1228 >    differential equation, given as a formal and usually divergent series
1229 >    in powers of the step size. For a rigorous analysis, this series
1230 >    has to be truncated. In this article we study the influence of this
1231 >    truncation to the difference between the numerical solution and
1232 >    the exact solution of the perturbed differential equation. Results
1233 >    on the long-time behaviour of numerical solutions are obtained in
1234 >    this way. We present applications to the numerical phase portrait
1235 >    near hyperbolic equilibrium points, to asymptotically stable periodic
1236 >    orbits and Hopf bifurcation, and to energy conservation and approximation
1237 >    of invariant tori in Hamiltonian systems.},
1238    annote = {Xj488 Times Cited:50 Cited References Count:19},
1239    issn = {0029-599X},
1240    uri = {<Go to ISI>://A1997XJ48800002},
# Line 1243 | Line 1243 | Encoding: GBK
1243   @ARTICLE{Hao1993,
1244    author = {M. H. Hao and M. R. Pincus and S. Rackovsky and H. A. Scheraga},
1245    title = {Unfolding and Refolding of the Native Structure of Bovine Pancreatic
1246 <        Trypsin-Inhibitor Studied by Computer-Simulations},
1246 >    Trypsin-Inhibitor Studied by Computer-Simulations},
1247    journal = {Biochemistry},
1248    year = {1993},
1249    volume = {32},
# Line 1251 | Line 1251 | Encoding: GBK
1251    number = {37},
1252    month = {Sep 21},
1253    abstract = {A new procedure for studying the folding and unfolding of proteins,
1254 <        with an application to bovine pancreatic trypsin inhibitor (BPTI),
1255 <        is reported. The unfolding and refolding of the native structure
1256 <        of the protein are characterized by the dimensions of the protein,
1257 <        expressed in terms of the three principal radii of the structure
1258 <        considered as an ellipsoid. A dynamic equation, describing the variations
1259 <        of the principal radii on the unfolding path, and a numerical procedure
1260 <        to solve this equation are proposed. Expanded and distorted conformations
1261 <        are refolded to the native structure by a dimensional-constraint
1262 <        energy minimization procedure. A unique and reproducible unfolding
1263 <        pathway for an intermediate of BPTI lacking the [30,51] disulfide
1264 <        bond is obtained. The resulting unfolded conformations are extended;
1265 <        they contain near-native local structure, but their longest principal
1266 <        radii are more than 2.5 times greater than that of the native structure.
1267 <        The most interesting finding is that the majority of expanded conformations,
1268 <        generated under various conditions, can be refolded closely to the
1269 <        native structure, as measured by the correct overall chain fold,
1270 <        by the rms deviations from the native structure of only 1.9-3.1
1271 <        angstrom, and by the energy differences of about 10 kcal/mol from
1272 <        the native structure. Introduction of the [30,51] disulfide bond
1273 <        at this stage, followed by minimization, improves the closeness
1274 <        of the refolded structures to the native structure, reducing the
1275 <        rms deviations to 0.9-2.0 angstrom. The unique refolding of these
1276 <        expanded structures over such a large conformational space implies
1277 <        that the folding is strongly dictated by the interactions in the
1278 <        amino acid sequence of BPTI. The simulations indicate that, under
1279 <        conditions that favor a compact structure as mimicked by the volume
1280 <        constraints in our algorithm; the expanded conformations have a
1281 <        strong tendency to move toward the native structure; therefore,
1282 <        they probably would be favorable folding intermediates. The results
1283 <        presented here support a general model for protein folding, i.e.,
1284 <        progressive formation of partially folded structural units, followed
1285 <        by collapse to the compact native structure. The general applicability
1286 <        of the procedure is also discussed.},
1254 >    with an application to bovine pancreatic trypsin inhibitor (BPTI),
1255 >    is reported. The unfolding and refolding of the native structure
1256 >    of the protein are characterized by the dimensions of the protein,
1257 >    expressed in terms of the three principal radii of the structure
1258 >    considered as an ellipsoid. A dynamic equation, describing the variations
1259 >    of the principal radii on the unfolding path, and a numerical procedure
1260 >    to solve this equation are proposed. Expanded and distorted conformations
1261 >    are refolded to the native structure by a dimensional-constraint
1262 >    energy minimization procedure. A unique and reproducible unfolding
1263 >    pathway for an intermediate of BPTI lacking the [30,51] disulfide
1264 >    bond is obtained. The resulting unfolded conformations are extended;
1265 >    they contain near-native local structure, but their longest principal
1266 >    radii are more than 2.5 times greater than that of the native structure.
1267 >    The most interesting finding is that the majority of expanded conformations,
1268 >    generated under various conditions, can be refolded closely to the
1269 >    native structure, as measured by the correct overall chain fold,
1270 >    by the rms deviations from the native structure of only 1.9-3.1
1271 >    angstrom, and by the energy differences of about 10 kcal/mol from
1272 >    the native structure. Introduction of the [30,51] disulfide bond
1273 >    at this stage, followed by minimization, improves the closeness
1274 >    of the refolded structures to the native structure, reducing the
1275 >    rms deviations to 0.9-2.0 angstrom. The unique refolding of these
1276 >    expanded structures over such a large conformational space implies
1277 >    that the folding is strongly dictated by the interactions in the
1278 >    amino acid sequence of BPTI. The simulations indicate that, under
1279 >    conditions that favor a compact structure as mimicked by the volume
1280 >    constraints in our algorithm; the expanded conformations have a
1281 >    strong tendency to move toward the native structure; therefore,
1282 >    they probably would be favorable folding intermediates. The results
1283 >    presented here support a general model for protein folding, i.e.,
1284 >    progressive formation of partially folded structural units, followed
1285 >    by collapse to the compact native structure. The general applicability
1286 >    of the procedure is also discussed.},
1287    annote = {Ly294 Times Cited:27 Cited References Count:57},
1288    issn = {0006-2960},
1289    uri = {<Go to ISI>://A1993LY29400014},
# Line 1291 | Line 1291 | Encoding: GBK
1291  
1292   @ARTICLE{Hinsen2000,
1293    author = {K. Hinsen and A. J. Petrescu and S. Dellerue and M. C. Bellissent-Funel
1294 <        and G. R. Kneller},
1294 >    and G. R. Kneller},
1295    title = {Harmonicity in slow protein dynamics},
1296    journal = {Chemical Physics},
1297    year = {2000},
# Line 1300 | Line 1300 | Encoding: GBK
1300    number = {1-2},
1301    month = {Nov 1},
1302    abstract = {The slow dynamics of proteins around its native folded state is usually
1303 <        described by diffusion in a strongly anharmonic potential. In this
1304 <        paper, we try to understand the form and origin of the anharmonicities,
1305 <        with the principal aim of gaining a better understanding of the
1306 <        principal motion types, but also in order to develop more efficient
1307 <        numerical methods for simulating neutron scattering spectra of large
1308 <        proteins. First, we decompose a molecular dynamics (MD) trajectory
1309 <        of 1.5 ns for a C-phycocyanin dimer surrounded by a layer of water
1310 <        into three contributions that we expect to be independent: the global
1311 <        motion of the residues, the rigid-body motion of the sidechains
1312 <        relative to the backbone, and the internal deformations of the sidechains.
1313 <        We show that they are indeed almost independent by verifying the
1314 <        factorization of the incoherent intermediate scattering function.
1315 <        Then, we show that the global residue motions, which include all
1316 <        large-scale backbone motions, can be reproduced by a simple harmonic
1317 <        model which contains two contributions: a short-time vibrational
1318 <        term, described by a standard normal mode calculation in a local
1319 <        minimum, and a long-time diffusive term, described by Brownian motion
1320 <        in an effective harmonic potential. The potential and the friction
1321 <        constants were fitted to the MD data. The major anharmonic contribution
1322 <        to the incoherent intermediate scattering function comes from the
1323 <        rigid-body diffusion of the sidechains. This model can be used to
1324 <        calculate scattering functions for large proteins and for long-time
1325 <        scales very efficiently, and thus provides a useful complement to
1326 <        MD simulations, which are best suited for detailed studies on smaller
1327 <        systems or for shorter time scales. (C) 2000 Elsevier Science B.V.
1328 <        All rights reserved.},
1303 >    described by diffusion in a strongly anharmonic potential. In this
1304 >    paper, we try to understand the form and origin of the anharmonicities,
1305 >    with the principal aim of gaining a better understanding of the
1306 >    principal motion types, but also in order to develop more efficient
1307 >    numerical methods for simulating neutron scattering spectra of large
1308 >    proteins. First, we decompose a molecular dynamics (MD) trajectory
1309 >    of 1.5 ns for a C-phycocyanin dimer surrounded by a layer of water
1310 >    into three contributions that we expect to be independent: the global
1311 >    motion of the residues, the rigid-body motion of the sidechains
1312 >    relative to the backbone, and the internal deformations of the sidechains.
1313 >    We show that they are indeed almost independent by verifying the
1314 >    factorization of the incoherent intermediate scattering function.
1315 >    Then, we show that the global residue motions, which include all
1316 >    large-scale backbone motions, can be reproduced by a simple harmonic
1317 >    model which contains two contributions: a short-time vibrational
1318 >    term, described by a standard normal mode calculation in a local
1319 >    minimum, and a long-time diffusive term, described by Brownian motion
1320 >    in an effective harmonic potential. The potential and the friction
1321 >    constants were fitted to the MD data. The major anharmonic contribution
1322 >    to the incoherent intermediate scattering function comes from the
1323 >    rigid-body diffusion of the sidechains. This model can be used to
1324 >    calculate scattering functions for large proteins and for long-time
1325 >    scales very efficiently, and thus provides a useful complement to
1326 >    MD simulations, which are best suited for detailed studies on smaller
1327 >    systems or for shorter time scales. (C) 2000 Elsevier Science B.V.
1328 >    All rights reserved.},
1329    annote = {Sp. Iss. SI 368MT Times Cited:16 Cited References Count:31},
1330    issn = {0301-0104},
1331    uri = {<Go to ISI>://000090121700003},
# Line 1341 | Line 1341 | Encoding: GBK
1341    number = {4},
1342    month = {Oct},
1343    abstract = {Evidence has been found for the existence water at the protein-lipid
1344 <        hydrophobic interface ot the membrane proteins, gramicidin and apocytochrome
1345 <        C, using two related fluorescence spectroscopic approaches. The
1346 <        first approach exploited the fact that the presence of water in
1347 <        the excited state solvent cage of a fluorophore increases the rate
1348 <        of decay. For 1,6-diphenyl-1,3,5-hexatriene (DPH) and 1-palmitoyl-2-[[2-[4-(6-phenyl-trans-1,3,5-hexatrienyl)
1349 <        phenyl]ethyl]carbonyl]-3-sn-PC (DPH-PC), where the fluorophores
1350 <        are located in the hydrophobic core of the lipid bilayer, the introduction
1351 <        of gramicidin reduced the fluorescence lifetime, indicative of an
1352 <        increased presence of water in the bilayer. Since a high protein:lipid
1353 <        ratio was used, the fluorophores were forced to be adjacent to the
1354 <        protein hydrophobic surface, hence the presence of water in this
1355 <        region could be inferred. Cholesterol is known to reduce the water
1356 <        content of lipid bilayers and this effect was maintained at the
1357 <        protein-lipid interface with both gramicidin and apocytochrome C,
1358 <        again suggesting hydration in this region. The second approach was
1359 <        to use the fluorescence enhancement induced by exchanging deuterium
1360 <        oxide (D2O) for H2O. Both the fluorescence intensities of trimethylammonium-DPH,
1361 <        located in the lipid head group region, and of the gramicidin intrinsic
1362 <        tryptophans were greater in a D2O buffer compared with H2O, showing
1363 <        that the fluorophores were exposed to water in the bilayer at the
1364 <        protein-lipid interface. In the presence of cholesterol the fluorescence
1365 <        intensity ratio of D2O to H2O decreased, indicating a removal of
1366 <        water by the cholesterol, in keeping with the lifetime data. Altered
1367 <        hydration at the protein-lipid interface could affect conformation,
1368 <        thereby offering a new route by which membrane protein functioning
1369 <        may be modified.},
1344 >    hydrophobic interface ot the membrane proteins, gramicidin and apocytochrome
1345 >    C, using two related fluorescence spectroscopic approaches. The
1346 >    first approach exploited the fact that the presence of water in
1347 >    the excited state solvent cage of a fluorophore increases the rate
1348 >    of decay. For 1,6-diphenyl-1,3,5-hexatriene (DPH) and 1-palmitoyl-2-[[2-[4-(6-phenyl-trans-1,3,5-hexatrienyl)
1349 >    phenyl]ethyl]carbonyl]-3-sn-PC (DPH-PC), where the fluorophores
1350 >    are located in the hydrophobic core of the lipid bilayer, the introduction
1351 >    of gramicidin reduced the fluorescence lifetime, indicative of an
1352 >    increased presence of water in the bilayer. Since a high protein:lipid
1353 >    ratio was used, the fluorophores were forced to be adjacent to the
1354 >    protein hydrophobic surface, hence the presence of water in this
1355 >    region could be inferred. Cholesterol is known to reduce the water
1356 >    content of lipid bilayers and this effect was maintained at the
1357 >    protein-lipid interface with both gramicidin and apocytochrome C,
1358 >    again suggesting hydration in this region. The second approach was
1359 >    to use the fluorescence enhancement induced by exchanging deuterium
1360 >    oxide (D2O) for H2O. Both the fluorescence intensities of trimethylammonium-DPH,
1361 >    located in the lipid head group region, and of the gramicidin intrinsic
1362 >    tryptophans were greater in a D2O buffer compared with H2O, showing
1363 >    that the fluorophores were exposed to water in the bilayer at the
1364 >    protein-lipid interface. In the presence of cholesterol the fluorescence
1365 >    intensity ratio of D2O to H2O decreased, indicating a removal of
1366 >    water by the cholesterol, in keeping with the lifetime data. Altered
1367 >    hydration at the protein-lipid interface could affect conformation,
1368 >    thereby offering a new route by which membrane protein functioning
1369 >    may be modified.},
1370    annote = {Ju251 Times Cited:55 Cited References Count:44},
1371    issn = {0006-3495},
1372    uri = {<Go to ISI>://A1992JU25100002},
# Line 1383 | Line 1383 | Encoding: GBK
1383   @ARTICLE{Huh2004,
1384    author = {Y. Huh and N. M. Cann},
1385    title = {Discrimination in isotropic, nematic, and smectic phases of chiral
1386 <        calamitic molecules: A computer simulation study},
1386 >    calamitic molecules: A computer simulation study},
1387    journal = {Journal of Chemical Physics},
1388    year = {2004},
1389    volume = {121},
# Line 1391 | Line 1391 | Encoding: GBK
1391    number = {20},
1392    month = {Nov 22},
1393    abstract = {Racemic fluids of chiral calamitic molecules are investigated with
1394 <        molecular dynamics simulations. In particular, the phase behavior
1395 <        as a function of density is examined for eight racemates. The relationship
1396 <        between chiral discrimination and orientational order in the phase
1397 <        is explored. We find that the transition from the isotropic phase
1398 <        to a liquid crystal phase is accompanied by an increase in chiral
1399 <        discrimination, as measured by differences in radial distributions.
1400 <        Among ordered phases, discrimination is largest for smectic phases
1401 <        with a significant preference for heterochiral contact within the
1402 <        layers. (C) 2004 American Institute of Physics.},
1394 >    molecular dynamics simulations. In particular, the phase behavior
1395 >    as a function of density is examined for eight racemates. The relationship
1396 >    between chiral discrimination and orientational order in the phase
1397 >    is explored. We find that the transition from the isotropic phase
1398 >    to a liquid crystal phase is accompanied by an increase in chiral
1399 >    discrimination, as measured by differences in radial distributions.
1400 >    Among ordered phases, discrimination is largest for smectic phases
1401 >    with a significant preference for heterochiral contact within the
1402 >    layers. (C) 2004 American Institute of Physics.},
1403    annote = {870FJ Times Cited:0 Cited References Count:63},
1404    issn = {0021-9606},
1405    uri = {<Go to ISI>://000225042700059},
# Line 1415 | Line 1415 | Encoding: GBK
1415    number = {5},
1416    month = {Feb 1},
1417    abstract = {In this paper we show the possibility of using very mild stochastic
1418 <        damping to stabilize long time step integrators for Newtonian molecular
1419 <        dynamics. More specifically, stable and accurate integrations are
1420 <        obtained for damping coefficients that are only a few percent of
1421 <        the natural decay rate of processes of interest, such as the velocity
1422 <        autocorrelation function. Two new multiple time stepping integrators,
1423 <        Langevin Molly (LM) and Brunger-Brooks-Karplus-Molly (BBK-M), are
1424 <        introduced in this paper. Both use the mollified impulse method
1425 <        for the Newtonian term. LM uses a discretization of the Langevin
1426 <        equation that is exact for the constant force, and BBK-M uses the
1427 <        popular Brunger-Brooks-Karplus integrator (BBK). These integrators,
1428 <        along with an extrapolative method called LN, are evaluated across
1429 <        a wide range of damping coefficient values. When large damping coefficients
1430 <        are used, as one would for the implicit modeling of solvent molecules,
1431 <        the method LN is superior, with LM closely following. However, with
1432 <        mild damping of 0.2 ps(-1), LM produces the best results, allowing
1433 <        long time steps of 14 fs in simulations containing explicitly modeled
1434 <        flexible water. With BBK-M and the same damping coefficient, time
1435 <        steps of 12 fs are possible for the same system. Similar results
1436 <        are obtained for a solvated protein-DNA simulation of estrogen receptor
1437 <        ER with estrogen response element ERE. A parallel version of BBK-M
1438 <        runs nearly three times faster than the Verlet-I/r-RESPA (reversible
1439 <        reference system propagator algorithm) when using the largest stable
1440 <        time step on each one, and it also parallelizes well. The computation
1441 <        of diffusion coefficients for flexible water and ER/ERE shows that
1442 <        when mild damping of up to 0.2 ps-1 is used the dynamics are not
1443 <        significantly distorted. (C) 2001 American Institute of Physics.},
1418 >    damping to stabilize long time step integrators for Newtonian molecular
1419 >    dynamics. More specifically, stable and accurate integrations are
1420 >    obtained for damping coefficients that are only a few percent of
1421 >    the natural decay rate of processes of interest, such as the velocity
1422 >    autocorrelation function. Two new multiple time stepping integrators,
1423 >    Langevin Molly (LM) and Brunger-Brooks-Karplus-Molly (BBK-M), are
1424 >    introduced in this paper. Both use the mollified impulse method
1425 >    for the Newtonian term. LM uses a discretization of the Langevin
1426 >    equation that is exact for the constant force, and BBK-M uses the
1427 >    popular Brunger-Brooks-Karplus integrator (BBK). These integrators,
1428 >    along with an extrapolative method called LN, are evaluated across
1429 >    a wide range of damping coefficient values. When large damping coefficients
1430 >    are used, as one would for the implicit modeling of solvent molecules,
1431 >    the method LN is superior, with LM closely following. However, with
1432 >    mild damping of 0.2 ps(-1), LM produces the best results, allowing
1433 >    long time steps of 14 fs in simulations containing explicitly modeled
1434 >    flexible water. With BBK-M and the same damping coefficient, time
1435 >    steps of 12 fs are possible for the same system. Similar results
1436 >    are obtained for a solvated protein-DNA simulation of estrogen receptor
1437 >    ER with estrogen response element ERE. A parallel version of BBK-M
1438 >    runs nearly three times faster than the Verlet-I/r-RESPA (reversible
1439 >    reference system propagator algorithm) when using the largest stable
1440 >    time step on each one, and it also parallelizes well. The computation
1441 >    of diffusion coefficients for flexible water and ER/ERE shows that
1442 >    when mild damping of up to 0.2 ps-1 is used the dynamics are not
1443 >    significantly distorted. (C) 2001 American Institute of Physics.},
1444    annote = {397CQ Times Cited:14 Cited References Count:36},
1445    issn = {0021-9606},
1446    uri = {<Go to ISI>://000166676100020},
1447   }
1448  
1449 @ARTICLE{Gray2003,
1450  author = {J.~J Gray,S. Moughon, C. Wang },
1451  title = {Protein-protein docking with simultaneous optimization of rigid-body
1452        displacement and side-chain conformations},
1453  journal = {jmb},
1454  year = {2003},
1455  volume = {331},
1456  pages = {281-299},
1457 }
1458
1449   @ARTICLE{Torre1977,
1450    author = {Jose Garcia De La Torre, V.A. Bloomfield},
1451    title = {Hydrodynamic properties of macromolecular complexes. I. Translation},
# Line 1468 | Line 1458 | Encoding: GBK
1458   @ARTICLE{Kane2000,
1459    author = {C. Kane and J. E. Marsden and M. Ortiz and M. West},
1460    title = {Variational integrators and the Newmark algorithm for conservative
1461 <        and dissipative mechanical systems},
1461 >    and dissipative mechanical systems},
1462    journal = {International Journal for Numerical Methods in Engineering},
1463    year = {2000},
1464    volume = {49},
# Line 1476 | Line 1466 | Encoding: GBK
1466    number = {10},
1467    month = {Dec 10},
1468    abstract = {The purpose of this work is twofold. First, we demonstrate analytically
1469 <        that the classical Newmark family as well as related integration
1470 <        algorithms are variational in the sense of the Veselov formulation
1471 <        of discrete mechanics. Such variational algorithms are well known
1472 <        to be symplectic and momentum preserving and to often have excellent
1473 <        global energy behaviour. This analytical result is verified through
1474 <        numerical examples and is believed to be one of the primary reasons
1475 <        that this class of algorithms performs so well. Second, we develop
1476 <        algorithms for mechanical systems with forcing, and in particular,
1477 <        for dissipative systems. In this case, we develop integrators that
1478 <        are based on a discretization of the Lagrange d'Alembert principle
1479 <        as well as on a variational formulation of dissipation. It is demonstrated
1480 <        that these types of structured integrators have good numerical behaviour
1481 <        in terms of obtaining the correct amounts by which the energy changes
1482 <        over the integration run. Copyright (C) 2000 John Wiley & Sons,
1483 <        Ltd.},
1469 >    that the classical Newmark family as well as related integration
1470 >    algorithms are variational in the sense of the Veselov formulation
1471 >    of discrete mechanics. Such variational algorithms are well known
1472 >    to be symplectic and momentum preserving and to often have excellent
1473 >    global energy behaviour. This analytical result is verified through
1474 >    numerical examples and is believed to be one of the primary reasons
1475 >    that this class of algorithms performs so well. Second, we develop
1476 >    algorithms for mechanical systems with forcing, and in particular,
1477 >    for dissipative systems. In this case, we develop integrators that
1478 >    are based on a discretization of the Lagrange d'Alembert principle
1479 >    as well as on a variational formulation of dissipation. It is demonstrated
1480 >    that these types of structured integrators have good numerical behaviour
1481 >    in terms of obtaining the correct amounts by which the energy changes
1482 >    over the integration run. Copyright (C) 2000 John Wiley & Sons,
1483 >    Ltd.},
1484    annote = {373CJ Times Cited:30 Cited References Count:41},
1485    issn = {0029-5981},
1486    uri = {<Go to ISI>://000165270600004},
# Line 1506 | Line 1496 | Encoding: GBK
1496    number = {2},
1497    month = {Jul 14},
1498    abstract = {The viscosity (eta) dependence of the folding rates for four sequences
1499 <        (the native state of three sequences is a beta sheet, while the
1500 <        fourth forms an alpha helix) is calculated for off-lattice models
1501 <        of proteins. Assuming that the dynamics is given by the Langevin
1502 <        equation, we show that the folding rates increase linearly at low
1503 <        viscosities eta, decrease as 1/eta at large eta, and have a maximum
1504 <        at intermediate values. The Kramers' theory of barrier crossing
1505 <        provides a quantitative fit of the numerical results. By mapping
1506 <        the simulation results to real proteins we estimate that for optimized
1507 <        sequences the time scale for forming a four turn alpha-helix topology
1508 <        is about 500 ns, whereas for beta sheet it is about 10 mu s.},
1499 >    (the native state of three sequences is a beta sheet, while the
1500 >    fourth forms an alpha helix) is calculated for off-lattice models
1501 >    of proteins. Assuming that the dynamics is given by the Langevin
1502 >    equation, we show that the folding rates increase linearly at low
1503 >    viscosities eta, decrease as 1/eta at large eta, and have a maximum
1504 >    at intermediate values. The Kramers' theory of barrier crossing
1505 >    provides a quantitative fit of the numerical results. By mapping
1506 >    the simulation results to real proteins we estimate that for optimized
1507 >    sequences the time scale for forming a four turn alpha-helix topology
1508 >    is about 500 ns, whereas for beta sheet it is about 10 mu s.},
1509    annote = {Xk293 Times Cited:77 Cited References Count:17},
1510    issn = {0031-9007},
1511    uri = {<Go to ISI>://A1997XK29300035},
# Line 1531 | Line 1521 | Encoding: GBK
1521    number = {7},
1522    month = {Aug 15},
1523    abstract = {Rigid-body molecular dynamics simulations typically are performed
1524 <        in a quaternion representation. The nonseparable form of the Hamiltonian
1525 <        in quaternions prevents the use of a standard leapfrog (Verlet)
1526 <        integrator, so nonsymplectic Runge-Kutta, multistep, or extrapolation
1527 <        methods are generally used, This is unfortunate since symplectic
1528 <        methods like Verlet exhibit superior energy conservation in long-time
1529 <        integrations. In this article, we describe an alternative method,
1530 <        which we call RSHAKE (for rotation-SHAKE), in which the entire rotation
1531 <        matrix is evolved (using the scheme of McLachlan and Scovel [J.
1532 <        Nonlin. Sci, 16 233 (1995)]) in tandem with the particle positions.
1533 <        We employ a fast approximate Newton solver to preserve the orthogonality
1534 <        of the rotation matrix. We test our method on a system of soft-sphere
1535 <        dipoles and compare with quaternion evolution using a 4th-order
1536 <        predictor-corrector integrator, Although the short-time error of
1537 <        the quaternion algorithm is smaller for fixed time step than that
1538 <        for RSHAKE, the quaternion scheme exhibits an energy drift which
1539 <        is not observed in simulations with RSHAKE, hence a fixed energy
1540 <        tolerance can be achieved by using a larger time step, The superiority
1541 <        of RSHAKE increases with system size. (C) 1997 American Institute
1542 <        of Physics.},
1524 >    in a quaternion representation. The nonseparable form of the Hamiltonian
1525 >    in quaternions prevents the use of a standard leapfrog (Verlet)
1526 >    integrator, so nonsymplectic Runge-Kutta, multistep, or extrapolation
1527 >    methods are generally used, This is unfortunate since symplectic
1528 >    methods like Verlet exhibit superior energy conservation in long-time
1529 >    integrations. In this article, we describe an alternative method,
1530 >    which we call RSHAKE (for rotation-SHAKE), in which the entire rotation
1531 >    matrix is evolved (using the scheme of McLachlan and Scovel [J.
1532 >    Nonlin. Sci, 16 233 (1995)]) in tandem with the particle positions.
1533 >    We employ a fast approximate Newton solver to preserve the orthogonality
1534 >    of the rotation matrix. We test our method on a system of soft-sphere
1535 >    dipoles and compare with quaternion evolution using a 4th-order
1536 >    predictor-corrector integrator, Although the short-time error of
1537 >    the quaternion algorithm is smaller for fixed time step than that
1538 >    for RSHAKE, the quaternion scheme exhibits an energy drift which
1539 >    is not observed in simulations with RSHAKE, hence a fixed energy
1540 >    tolerance can be achieved by using a larger time step, The superiority
1541 >    of RSHAKE increases with system size. (C) 1997 American Institute
1542 >    of Physics.},
1543    annote = {Xq332 Times Cited:11 Cited References Count:18},
1544    issn = {0021-9606},
1545    uri = {<Go to ISI>://A1997XQ33200046},
# Line 1558 | Line 1548 | Encoding: GBK
1548   @ARTICLE{Lansac2001,
1549    author = {Y. Lansac and M. A. Glaser and N. A. Clark},
1550    title = {Microscopic structure and dynamics of a partial bilayer smectic liquid
1551 <        crystal},
1551 >    crystal},
1552    journal = {Physical Review E},
1553    year = {2001},
1554    volume = {6405},
# Line 1566 | Line 1556 | Encoding: GBK
1556    number = {5},
1557    month = {Nov},
1558    abstract = {Cyanobiphenyls (nCB's) represent a useful and intensively studied
1559 <        class of mesogens. Many of the peculiar properties of nCB's (e.g.,
1560 <        the occurence of the partial bilayer smectic-A(d) phase) are thought
1561 <        to be a manifestation of short-range antiparallel association of
1562 <        neighboring molecules, resulting from strong dipole-dipole interactions
1563 <        between cyano groups. To test and extend existing models of microscopic
1564 <        ordering in nCB's, we carry out large-scale atomistic simulation
1565 <        studies of the microscopic structure and dynamics of the Sm-A(d)
1566 <        phase of 4-octyl-4'-cyanobiphenyl (8CB). We compute a variety of
1567 <        thermodynamic, structural, and dynamical properties for this material,
1568 <        and make a detailed comparison of our results with experimental
1569 <        measurements in order to validate our molecular model. Semiquantitative
1570 <        agreement with experiment is found: the smectic layer spacing and
1571 <        mass density are well reproduced, translational diffusion constants
1572 <        are similar to experiment, but the orientational ordering of alkyl
1573 <        chains is overestimated. This simulation provides a detailed picture
1574 <        of molecular conformation, smectic layer structure, and intermolecular
1575 <        correlations in Sm-A(d) 8CB, and demonstrates that pronounced short-range
1576 <        antiparallel association of molecules arising from dipole-dipole
1577 <        interactions plays a dominant role in determining the molecular-scale
1578 <        structure of 8CB.},
1559 >    class of mesogens. Many of the peculiar properties of nCB's (e.g.,
1560 >    the occurence of the partial bilayer smectic-A(d) phase) are thought
1561 >    to be a manifestation of short-range antiparallel association of
1562 >    neighboring molecules, resulting from strong dipole-dipole interactions
1563 >    between cyano groups. To test and extend existing models of microscopic
1564 >    ordering in nCB's, we carry out large-scale atomistic simulation
1565 >    studies of the microscopic structure and dynamics of the Sm-A(d)
1566 >    phase of 4-octyl-4'-cyanobiphenyl (8CB). We compute a variety of
1567 >    thermodynamic, structural, and dynamical properties for this material,
1568 >    and make a detailed comparison of our results with experimental
1569 >    measurements in order to validate our molecular model. Semiquantitative
1570 >    agreement with experiment is found: the smectic layer spacing and
1571 >    mass density are well reproduced, translational diffusion constants
1572 >    are similar to experiment, but the orientational ordering of alkyl
1573 >    chains is overestimated. This simulation provides a detailed picture
1574 >    of molecular conformation, smectic layer structure, and intermolecular
1575 >    correlations in Sm-A(d) 8CB, and demonstrates that pronounced short-range
1576 >    antiparallel association of molecules arising from dipole-dipole
1577 >    interactions plays a dominant role in determining the molecular-scale
1578 >    structure of 8CB.},
1579    annote = {Part 1 496QF Times Cited:10 Cited References Count:60},
1580    issn = {1063-651X},
1581    uri = {<Go to ISI>://000172406900063},
# Line 1601 | Line 1591 | Encoding: GBK
1591    number = {1},
1592    month = {Jan},
1593    abstract = {Recently, a new class of smectic liquid crystal phases characterized
1594 <        by the spontaneous formation of macroscopic chiral domains from
1595 <        achiral bent-core molecules has been discovered. We have carried
1596 <        out Monte Carlo simulations of a minimal hard spherocylinder dimer
1597 <        model to investigate the role of excluded volume interactions in
1598 <        determining the phase behavior of bent-core materials and to probe
1599 <        the molecular origins of polar and chiral symmetry breaking. We
1600 <        present the phase diagram of hard spherocylinder dimers of length-diameter
1601 <        ratio of 5 as a function of pressure or density and dimer opening
1602 <        angle psi. With decreasing psi, a transition from a nonpolar to
1603 <        a polar smectic A phase is observed near psi=167degrees, and the
1604 <        nematic phase becomes thermodynamically unstable for psi<135degrees.
1605 <        Free energy calculations indicate that the antipolar smectic A (SmAP(A))
1606 <        phase is more stable than the polar smectic A phase (SmAP(F)). No
1607 <        chiral smectic or biaxial nematic phases were found.},
1594 >    by the spontaneous formation of macroscopic chiral domains from
1595 >    achiral bent-core molecules has been discovered. We have carried
1596 >    out Monte Carlo simulations of a minimal hard spherocylinder dimer
1597 >    model to investigate the role of excluded volume interactions in
1598 >    determining the phase behavior of bent-core materials and to probe
1599 >    the molecular origins of polar and chiral symmetry breaking. We
1600 >    present the phase diagram of hard spherocylinder dimers of length-diameter
1601 >    ratio of 5 as a function of pressure or density and dimer opening
1602 >    angle psi. With decreasing psi, a transition from a nonpolar to
1603 >    a polar smectic A phase is observed near psi=167degrees, and the
1604 >    nematic phase becomes thermodynamically unstable for psi<135degrees.
1605 >    Free energy calculations indicate that the antipolar smectic A (SmAP(A))
1606 >    phase is more stable than the polar smectic A phase (SmAP(F)). No
1607 >    chiral smectic or biaxial nematic phases were found.},
1608    annote = {Part 1 646CM Times Cited:15 Cited References Count:38},
1609    issn = {1063-651X},
1610    uri = {<Go to ISI>://000181017300042},
# Line 1632 | Line 1622 | Encoding: GBK
1622   @ARTICLE{Leimkuhler1999,
1623    author = {B. Leimkuhler},
1624    title = {Reversible adaptive regularization: perturbed Kepler motion and classical
1625 <        atomic trajectories},
1625 >    atomic trajectories},
1626    journal = {Philosophical Transactions of the Royal Society of London Series
1627 <        a-Mathematical Physical and Engineering Sciences},
1627 >    a-Mathematical Physical and Engineering Sciences},
1628    year = {1999},
1629    volume = {357},
1630    pages = {1101-1133},
1631    number = {1754},
1632    month = {Apr 15},
1633    abstract = {Reversible and adaptive integration methods based on Kustaanheimo-Stiefel
1634 <        regularization and modified Sundman transformations are applied
1635 <        to simulate general perturbed Kepler motion and to compute classical
1636 <        trajectories of atomic systems (e.g. Rydberg atoms). The new family
1637 <        of reversible adaptive regularization methods also conserves angular
1638 <        momentum and exhibits superior energy conservation and numerical
1639 <        stability in long-time integrations. The schemes are appropriate
1640 <        for scattering, for astronomical calculations of escape time and
1641 <        long-term stability, and for classical and semiclassical studies
1642 <        of atomic dynamics. The components of an algorithm for trajectory
1643 <        calculations are described. Numerical experiments illustrate the
1644 <        effectiveness of the reversible approach.},
1634 >    regularization and modified Sundman transformations are applied
1635 >    to simulate general perturbed Kepler motion and to compute classical
1636 >    trajectories of atomic systems (e.g. Rydberg atoms). The new family
1637 >    of reversible adaptive regularization methods also conserves angular
1638 >    momentum and exhibits superior energy conservation and numerical
1639 >    stability in long-time integrations. The schemes are appropriate
1640 >    for scattering, for astronomical calculations of escape time and
1641 >    long-term stability, and for classical and semiclassical studies
1642 >    of atomic dynamics. The components of an algorithm for trajectory
1643 >    calculations are described. Numerical experiments illustrate the
1644 >    effectiveness of the reversible approach.},
1645    annote = {199EE Times Cited:11 Cited References Count:48},
1646    issn = {1364-503X},
1647    uri = {<Go to ISI>://000080466800007},
# Line 1667 | Line 1657 | Encoding: GBK
1657  
1658   @ARTICLE{Levelut1981,
1659    author = {A. M. Levelut and R. J. Tarento and F. Hardouin and M. F. Achard
1660 <        and G. Sigaud},
1660 >    and G. Sigaud},
1661    title = {Number of Sa Phases},
1662    journal = {Physical Review A},
1663    year = {1981},
# Line 1682 | Line 1672 | Encoding: GBK
1672   @ARTICLE{Lieb1982,
1673    author = {W. R. Lieb and M. Kovalycsik and R. Mendelsohn},
1674    title = {Do Clinical-Levels of General-Anesthetics Affect Lipid Bilayers -
1675 <        Evidence from Raman-Scattering},
1675 >    Evidence from Raman-Scattering},
1676    journal = {Biochimica Et Biophysica Acta},
1677    year = {1982},
1678    volume = {688},
# Line 1695 | Line 1685 | Encoding: GBK
1685  
1686   @ARTICLE{Link1997,
1687    author = {D. R. Link and G. Natale and R. Shao and J. E. Maclennan and N. A.
1688 <        Clark and E. Korblova and D. M. Walba},
1688 >    Clark and E. Korblova and D. M. Walba},
1689    title = {Spontaneous formation of macroscopic chiral domains in a fluid smectic
1690 <        phase of achiral molecules},
1690 >    phase of achiral molecules},
1691    journal = {Science},
1692    year = {1997},
1693    volume = {278},
# Line 1705 | Line 1695 | Encoding: GBK
1695    number = {5345},
1696    month = {Dec 12},
1697    abstract = {A smectic liquid-crystal phase made from achiral molecules with bent
1698 <        cores was found to have fluid layers that exhibit two spontaneous
1699 <        symmetry-breaking instabilities: polar molecular orientational ordering
1700 <        about the layer normal and molecular tilt. These instabilities combine
1701 <        to form a chiral layer structure with a handedness that depends
1702 <        on the sign of the tilt. The bulk states are either antiferroelectric-racemic,
1703 <        with the layer polar direction and handedness alternating in sign
1704 <        from layer to layer, or antiferroelectric-chiral, which is of uniform
1705 <        layer handedness. Both states exhibit an electric field-induced
1706 <        transition from antiferroelectric to ferroelectric.},
1698 >    cores was found to have fluid layers that exhibit two spontaneous
1699 >    symmetry-breaking instabilities: polar molecular orientational ordering
1700 >    about the layer normal and molecular tilt. These instabilities combine
1701 >    to form a chiral layer structure with a handedness that depends
1702 >    on the sign of the tilt. The bulk states are either antiferroelectric-racemic,
1703 >    with the layer polar direction and handedness alternating in sign
1704 >    from layer to layer, or antiferroelectric-chiral, which is of uniform
1705 >    layer handedness. Both states exhibit an electric field-induced
1706 >    transition from antiferroelectric to ferroelectric.},
1707    annote = {Yl002 Times Cited:407 Cited References Count:25},
1708    issn = {0036-8075},
1709    uri = {<Go to ISI>://A1997YL00200028},
# Line 1722 | Line 1712 | Encoding: GBK
1712   @ARTICLE{Liwo2005,
1713    author = {A. Liwo and M. Khalili and H. A. Scheraga},
1714    title = {Ab initio simulations of protein folding pathways by molecular dynamics
1715 <        with the united-residue (UNRES) model of polypeptide chains},
1715 >    with the united-residue (UNRES) model of polypeptide chains},
1716    journal = {Febs Journal},
1717    year = {2005},
1718    volume = {272},
# Line 1736 | Line 1726 | Encoding: GBK
1726   @ARTICLE{Luty1994,
1727    author = {B. A. Luty and M. E. Davis and I. G. Tironi and W. F. Vangunsteren},
1728    title = {A Comparison of Particle-Particle, Particle-Mesh and Ewald Methods
1729 <        for Calculating Electrostatic Interactions in Periodic Molecular-Systems},
1729 >    for Calculating Electrostatic Interactions in Periodic Molecular-Systems},
1730    journal = {Molecular Simulation},
1731    year = {1994},
1732    volume = {14},
1733    pages = {11-20},
1734    number = {1},
1735    abstract = {We compare the Particle-Particle Particle-Mesh (PPPM) and Ewald methods
1736 <        for calculating electrostatic interactions in periodic molecular
1737 <        systems. A brief comparison of the theories shows that the methods
1738 <        are very similar differing mainly in the technique which is used
1739 <        to perform the ''k-space'' or mesh calculation. Because the PPPM
1740 <        utilizes the highly efficient numerical Fast Fourier Transform (FFT)
1741 <        method it requires significantly less computational effort than
1742 <        the Ewald method and scale's almost linearly with system size.},
1736 >    for calculating electrostatic interactions in periodic molecular
1737 >    systems. A brief comparison of the theories shows that the methods
1738 >    are very similar differing mainly in the technique which is used
1739 >    to perform the ''k-space'' or mesh calculation. Because the PPPM
1740 >    utilizes the highly efficient numerical Fast Fourier Transform (FFT)
1741 >    method it requires significantly less computational effort than
1742 >    the Ewald method and scale's almost linearly with system size.},
1743    annote = {Qf464 Times Cited:50 Cited References Count:20},
1744    issn = {0892-7022},
1745    uri = {<Go to ISI>://A1994QF46400002},
# Line 1767 | Line 1757 | Encoding: GBK
1757   @ARTICLE{Marsden1998,
1758    author = {J. E. Marsden and G. W. Patrick and S. Shkoller},
1759    title = {Multisymplectic geometry, variational integrators, and nonlinear
1760 <        PDEs},
1760 >    PDEs},
1761    journal = {Communications in Mathematical Physics},
1762    year = {1998},
1763    volume = {199},
# Line 1775 | Line 1765 | Encoding: GBK
1765    number = {2},
1766    month = {Dec},
1767    abstract = {This paper presents a geometric-variational approach to continuous
1768 <        and discrete mechanics and field theories. Using multisymplectic
1769 <        geometry, we show that the existence of the fundamental geometric
1770 <        structures as well as their preservation along solutions can be
1771 <        obtained directly from the variational principle. In particular,
1772 <        we prove that a unique multisymplectic structure is obtained by
1773 <        taking the derivative of an action function, and use this structure
1774 <        to prove covariant generalizations of conservation of symplecticity
1775 <        and Noether's theorem. Natural discretization schemes for PDEs,
1776 <        which have these important preservation properties, then follow
1777 <        by choosing a discrete action functional. In the case of mechanics,
1778 <        we recover the variational symplectic integrators of Veselov type,
1779 <        while for PDEs we obtain covariant spacetime integrators which conserve
1780 <        the corresponding discrete multisymplectic form as well as the discrete
1781 <        momentum mappings corresponding to symmetries. We show that the
1782 <        usual notion of symplecticity along an infinite-dimensional space
1783 <        of fields can be naturally obtained by making a spacetime split.
1784 <        All of the aspects of our method are demonstrated with a nonlinear
1785 <        sine-Gordon equation, including computational results and a comparison
1786 <        with other discretization schemes.},
1768 >    and discrete mechanics and field theories. Using multisymplectic
1769 >    geometry, we show that the existence of the fundamental geometric
1770 >    structures as well as their preservation along solutions can be
1771 >    obtained directly from the variational principle. In particular,
1772 >    we prove that a unique multisymplectic structure is obtained by
1773 >    taking the derivative of an action function, and use this structure
1774 >    to prove covariant generalizations of conservation of symplecticity
1775 >    and Noether's theorem. Natural discretization schemes for PDEs,
1776 >    which have these important preservation properties, then follow
1777 >    by choosing a discrete action functional. In the case of mechanics,
1778 >    we recover the variational symplectic integrators of Veselov type,
1779 >    while for PDEs we obtain covariant spacetime integrators which conserve
1780 >    the corresponding discrete multisymplectic form as well as the discrete
1781 >    momentum mappings corresponding to symmetries. We show that the
1782 >    usual notion of symplecticity along an infinite-dimensional space
1783 >    of fields can be naturally obtained by making a spacetime split.
1784 >    All of the aspects of our method are demonstrated with a nonlinear
1785 >    sine-Gordon equation, including computational results and a comparison
1786 >    with other discretization schemes.},
1787    annote = {154RH Times Cited:88 Cited References Count:36},
1788    issn = {0010-3616},
1789    uri = {<Go to ISI>://000077902200006},
# Line 1811 | Line 1801 | Encoding: GBK
1801   @ARTICLE{McLachlan1998a,
1802    author = {R. I. McLachlan and G. R. W. Quispel},
1803    title = {Generating functions for dynamical systems with symmetries, integrals,
1804 <        and differential invariants},
1804 >    and differential invariants},
1805    journal = {Physica D},
1806    year = {1998},
1807    volume = {112},
# Line 1819 | Line 1809 | Encoding: GBK
1809    number = {1-2},
1810    month = {Jan 15},
1811    abstract = {We give a survey and some new examples of generating functions for
1812 <        systems with symplectic structure, systems with a first integral,
1813 <        systems that preserve volume, and systems with symmetries and/or
1814 <        time-reversing symmetries. Both ODEs and maps are treated, and we
1815 <        discuss how generating functions may be used in the structure-preserving
1816 <        numerical integration of ODEs with the above properties.},
1812 >    systems with symplectic structure, systems with a first integral,
1813 >    systems that preserve volume, and systems with symmetries and/or
1814 >    time-reversing symmetries. Both ODEs and maps are treated, and we
1815 >    discuss how generating functions may be used in the structure-preserving
1816 >    numerical integration of ODEs with the above properties.},
1817    annote = {Yt049 Times Cited:7 Cited References Count:26},
1818    issn = {0167-2789},
1819    uri = {<Go to ISI>://000071558900021},
# Line 1839 | Line 1829 | Encoding: GBK
1829    number = {2},
1830    month = {Apr},
1831    abstract = {We consider properties of flows, the relationships between them, and
1832 <        whether numerical integrators can be made to preserve these properties.
1833 <        This is done in the context of automorphisms and antiautomorphisms
1834 <        of a certain group generated by maps associated to vector fields.
1835 <        This new framework unifies several known constructions. We also
1836 <        use the concept of #covariance# of a numerical method with respect
1837 <        to a group of coordinate transformations. The main application is
1838 <        to explore the relationship between spatial symmetries, reversing
1839 <        symmetries, and time symmetry of flows and numerical integrators.},
1832 >    whether numerical integrators can be made to preserve these properties.
1833 >    This is done in the context of automorphisms and antiautomorphisms
1834 >    of a certain group generated by maps associated to vector fields.
1835 >    This new framework unifies several known constructions. We also
1836 >    use the concept of #covariance# of a numerical method with respect
1837 >    to a group of coordinate transformations. The main application is
1838 >    to explore the relationship between spatial symmetries, reversing
1839 >    symmetries, and time symmetry of flows and numerical integrators.},
1840    annote = {Zc449 Times Cited:14 Cited References Count:33},
1841    issn = {0036-1429},
1842    uri = {<Go to ISI>://000072580500010},
# Line 1862 | Line 1852 | Encoding: GBK
1852    number = {1},
1853    month = {Feb},
1854    abstract = {In this paper we revisit the Moser-Veselov description of the free
1855 <        rigid body in body coordinates, which, in the 3 x 3 case, can be
1856 <        implemented as an explicit, second-order, integrable approximation
1857 <        of the continuous solution. By backward error analysis, we study
1858 <        the modified vector field which is integrated exactly by the discrete
1859 <        algorithm. We deduce that the discrete Moser-Veselov (DMV) is well
1860 <        approximated to higher order by time reparametrizations of the continuous
1861 <        equations (modified vector field). We use the modified vector field
1862 <        to scale the initial data of the DMV to improve the order of the
1863 <        approximation and show the equivalence of the DMV and the RATTLE
1864 <        algorithm. Numerical integration with these preprocessed initial
1865 <        data is several orders of magnitude more accurate than the original
1866 <        DMV and RATTLE approach.},
1855 >    rigid body in body coordinates, which, in the 3 x 3 case, can be
1856 >    implemented as an explicit, second-order, integrable approximation
1857 >    of the continuous solution. By backward error analysis, we study
1858 >    the modified vector field which is integrated exactly by the discrete
1859 >    algorithm. We deduce that the discrete Moser-Veselov (DMV) is well
1860 >    approximated to higher order by time reparametrizations of the continuous
1861 >    equations (modified vector field). We use the modified vector field
1862 >    to scale the initial data of the DMV to improve the order of the
1863 >    approximation and show the equivalence of the DMV and the RATTLE
1864 >    algorithm. Numerical integration with these preprocessed initial
1865 >    data is several orders of magnitude more accurate than the original
1866 >    DMV and RATTLE approach.},
1867    annote = {911NS Times Cited:0 Cited References Count:14},
1868    issn = {1615-3375},
1869    uri = {<Go to ISI>://000228011900003},
# Line 1882 | Line 1872 | Encoding: GBK
1872   @ARTICLE{Memmer2002,
1873    author = {R. Memmer},
1874    title = {Liquid crystal phases of achiral banana-shaped molecules: a computer
1875 <        simulation study},
1875 >    simulation study},
1876    journal = {Liquid Crystals},
1877    year = {2002},
1878    volume = {29},
# Line 1890 | Line 1880 | Encoding: GBK
1880    number = {4},
1881    month = {Apr},
1882    abstract = {The phase behaviour of achiral banana-shaped molecules was studied
1883 <        by computer simulation. The banana-shaped molecules were described
1884 <        by model intermolecular interactions based on the Gay-Berne potential.
1885 <        The characteristic molecular structure was considered by joining
1886 <        two calamitic Gay-Berne particles through a bond to form a biaxial
1887 <        molecule of point symmetry group C-2v with a suitable bending angle.
1888 <        The dependence on temperature of systems of N=1024 rigid banana-shaped
1889 <        molecules with bending angle phi=140degrees has been studied by
1890 <        means of Monte Carlo simulations in the isobaric-isothermal ensemble
1891 <        (NpT). On cooling an isotropic system, two phase transitions characterized
1892 <        by phase transition enthalpy, entropy and relative volume change
1893 <        have been observed. For the first time by computer simulation of
1894 <        a many-particle system of banana-shaped molecules, at low temperature
1895 <        an untilted smectic phase showing a global phase biaxiality and
1896 <        a spontaneous local polarization in the layers, i.e. a local polar
1897 <        arrangement of the steric dipoles, with an antiferroelectric-like
1898 <        superstructure could be proven, a phase structure which recently
1899 <        has been discovered experimentally. Additionally, at intermediate
1900 <        temperature a nematic-like phase has been proved, whereas close
1901 <        to the transition to the smectic phase hints of a spontaneous achiral
1902 <        symmetry breaking have been determined. Here, in the absence of
1903 <        a layered structure a helical superstructure has been formed. All
1904 <        phases have been characterized by visual representations of selected
1905 <        configurations, scalar and pseudoscalar correlation functions, and
1906 <        order parameters.},
1883 >    by computer simulation. The banana-shaped molecules were described
1884 >    by model intermolecular interactions based on the Gay-Berne potential.
1885 >    The characteristic molecular structure was considered by joining
1886 >    two calamitic Gay-Berne particles through a bond to form a biaxial
1887 >    molecule of point symmetry group C-2v with a suitable bending angle.
1888 >    The dependence on temperature of systems of N=1024 rigid banana-shaped
1889 >    molecules with bending angle phi=140degrees has been studied by
1890 >    means of Monte Carlo simulations in the isobaric-isothermal ensemble
1891 >    (NpT). On cooling an isotropic system, two phase transitions characterized
1892 >    by phase transition enthalpy, entropy and relative volume change
1893 >    have been observed. For the first time by computer simulation of
1894 >    a many-particle system of banana-shaped molecules, at low temperature
1895 >    an untilted smectic phase showing a global phase biaxiality and
1896 >    a spontaneous local polarization in the layers, i.e. a local polar
1897 >    arrangement of the steric dipoles, with an antiferroelectric-like
1898 >    superstructure could be proven, a phase structure which recently
1899 >    has been discovered experimentally. Additionally, at intermediate
1900 >    temperature a nematic-like phase has been proved, whereas close
1901 >    to the transition to the smectic phase hints of a spontaneous achiral
1902 >    symmetry breaking have been determined. Here, in the absence of
1903 >    a layered structure a helical superstructure has been formed. All
1904 >    phases have been characterized by visual representations of selected
1905 >    configurations, scalar and pseudoscalar correlation functions, and
1906 >    order parameters.},
1907    annote = {531HT Times Cited:12 Cited References Count:37},
1908    issn = {0267-8292},
1909    uri = {<Go to ISI>://000174410500001},
# Line 1930 | Line 1920 | Encoding: GBK
1920  
1921   @ARTICLE{Mielke2004,
1922    author = {S. P. Mielke and W. H. Fink and V. V. Krishnan and N. Gronbech-Jensen
1923 <        and C. J. Benham},
1923 >    and C. J. Benham},
1924    title = {Transcription-driven twin supercoiling of a DNA loop: A Brownian
1925 <        dynamics study},
1925 >    dynamics study},
1926    journal = {Journal of Chemical Physics},
1927    year = {2004},
1928    volume = {121},
# Line 1940 | Line 1930 | Encoding: GBK
1930    number = {16},
1931    month = {Oct 22},
1932    abstract = {The torque generated by RNA polymerase as it tracks along double-stranded
1933 <        DNA can potentially induce long-range structural deformations integral
1934 <        to mechanisms of biological significance in both prokaryotes and
1935 <        eukaryotes. In this paper, we introduce a dynamic computer model
1936 <        for investigating this phenomenon. Duplex DNA is represented as
1937 <        a chain of hydrodynamic beads interacting through potentials of
1938 <        linearly elastic stretching, bending, and twisting, as well as excluded
1939 <        volume. The chain, linear when relaxed, is looped to form two open
1940 <        but topologically constrained subdomains. This permits the dynamic
1941 <        introduction of torsional stress via a centrally applied torque.
1942 <        We simulate by Brownian dynamics the 100 mus response of a 477-base
1943 <        pair B-DNA template to the localized torque generated by the prokaryotic
1944 <        transcription ensemble. Following a sharp rise at early times, the
1945 <        distributed twist assumes a nearly constant value in both subdomains,
1946 <        and a succession of supercoiling deformations occurs as superhelical
1947 <        stress is increasingly partitioned to writhe. The magnitude of writhe
1948 <        surpasses that of twist before also leveling off when the structure
1949 <        reaches mechanical equilibrium with the torsional load. Superhelicity
1950 <        is simultaneously right handed in one subdomain and left handed
1951 <        in the other, as predicted by the #transcription-induced##twin-supercoiled-domain#
1952 <        model [L. F. Liu and J. C. Wang, Proc. Natl. Acad. Sci. U.S.A. 84,
1953 <        7024 (1987)]. The properties of the chain at the onset of writhing
1954 <        agree well with predictions from theory, and the generated stress
1955 <        is ample for driving secondary structural transitions in physiological
1956 <        DNA. (C) 2004 American Institute of Physics.},
1933 >    DNA can potentially induce long-range structural deformations integral
1934 >    to mechanisms of biological significance in both prokaryotes and
1935 >    eukaryotes. In this paper, we introduce a dynamic computer model
1936 >    for investigating this phenomenon. Duplex DNA is represented as
1937 >    a chain of hydrodynamic beads interacting through potentials of
1938 >    linearly elastic stretching, bending, and twisting, as well as excluded
1939 >    volume. The chain, linear when relaxed, is looped to form two open
1940 >    but topologically constrained subdomains. This permits the dynamic
1941 >    introduction of torsional stress via a centrally applied torque.
1942 >    We simulate by Brownian dynamics the 100 mus response of a 477-base
1943 >    pair B-DNA template to the localized torque generated by the prokaryotic
1944 >    transcription ensemble. Following a sharp rise at early times, the
1945 >    distributed twist assumes a nearly constant value in both subdomains,
1946 >    and a succession of supercoiling deformations occurs as superhelical
1947 >    stress is increasingly partitioned to writhe. The magnitude of writhe
1948 >    surpasses that of twist before also leveling off when the structure
1949 >    reaches mechanical equilibrium with the torsional load. Superhelicity
1950 >    is simultaneously right handed in one subdomain and left handed
1951 >    in the other, as predicted by the #transcription-induced##twin-supercoiled-domain#
1952 >    model [L. F. Liu and J. C. Wang, Proc. Natl. Acad. Sci. U.S.A. 84,
1953 >    7024 (1987)]. The properties of the chain at the onset of writhing
1954 >    agree well with predictions from theory, and the generated stress
1955 >    is ample for driving secondary structural transitions in physiological
1956 >    DNA. (C) 2004 American Institute of Physics.},
1957    annote = {861ZF Times Cited:3 Cited References Count:34},
1958    issn = {0021-9606},
1959    uri = {<Go to ISI>://000224456500064},
# Line 1972 | Line 1962 | Encoding: GBK
1962   @ARTICLE{Naess2001,
1963    author = {S. N. Naess and H. M. Adland and A. Mikkelsen and A. Elgsaeter},
1964    title = {Brownian dynamics simulation of rigid bodies and segmented polymer
1965 <        chains. Use of Cartesian rotation vectors as the generalized coordinates
1966 <        describing angular orientations},
1965 >    chains. Use of Cartesian rotation vectors as the generalized coordinates
1966 >    describing angular orientations},
1967    journal = {Physica A},
1968    year = {2001},
1969    volume = {294},
# Line 1981 | Line 1971 | Encoding: GBK
1971    number = {3-4},
1972    month = {May 15},
1973    abstract = {The three Eulerian angles constitute the classical choice of generalized
1974 <        coordinates used to describe the three degrees of rotational freedom
1975 <        of a rigid body, but it has long been known that this choice yields
1976 <        singular equations of motion. The latter is also true when Eulerian
1977 <        angles are used in Brownian dynamics analyses of the angular orientation
1978 <        of single rigid bodies and segmented polymer chains. Starting from
1979 <        kinetic theory we here show that by instead employing the three
1980 <        components of Cartesian rotation vectors as the generalized coordinates
1981 <        describing angular orientation, no singularity appears in the configuration
1982 <        space diffusion equation and the associated Brownian dynamics algorithm.
1983 <        The suitability of Cartesian rotation vectors in Brownian dynamics
1984 <        simulations of segmented polymer chains with spring-like or ball-socket
1985 <        joints is discussed. (C) 2001 Elsevier Science B.V. All rights reserved.},
1974 >    coordinates used to describe the three degrees of rotational freedom
1975 >    of a rigid body, but it has long been known that this choice yields
1976 >    singular equations of motion. The latter is also true when Eulerian
1977 >    angles are used in Brownian dynamics analyses of the angular orientation
1978 >    of single rigid bodies and segmented polymer chains. Starting from
1979 >    kinetic theory we here show that by instead employing the three
1980 >    components of Cartesian rotation vectors as the generalized coordinates
1981 >    describing angular orientation, no singularity appears in the configuration
1982 >    space diffusion equation and the associated Brownian dynamics algorithm.
1983 >    The suitability of Cartesian rotation vectors in Brownian dynamics
1984 >    simulations of segmented polymer chains with spring-like or ball-socket
1985 >    joints is discussed. (C) 2001 Elsevier Science B.V. All rights reserved.},
1986    annote = {433TA Times Cited:7 Cited References Count:19},
1987    issn = {0378-4371},
1988    uri = {<Go to ISI>://000168774800005},
# Line 2001 | Line 1991 | Encoding: GBK
1991   @ARTICLE{Niori1996,
1992    author = {T. Niori and T. Sekine and J. Watanabe and T. Furukawa and H. Takezoe},
1993    title = {Distinct ferroelectric smectic liquid crystals consisting of banana
1994 <        shaped achiral molecules},
1994 >    shaped achiral molecules},
1995    journal = {Journal of Materials Chemistry},
1996    year = {1996},
1997    volume = {6},
# Line 2009 | Line 1999 | Encoding: GBK
1999    number = {7},
2000    month = {Jul},
2001    abstract = {The synthesis of a banana-shaped molecule is reported and it is found
2002 <        that the smectic phase which it forms is biaxial with the molecules
2003 <        packed in the best,direction into a layer. Because of this characteristic
2004 <        packing, spontaneous polarization appears parallel to the layer
2005 <        and switches on reversal of an applied electric field. This is the
2006 <        first obvious example of ferroelectricity in an achiral smectic
2007 <        phase and is ascribed to the C-2v symmetry of the molecular packing.},
2002 >    that the smectic phase which it forms is biaxial with the molecules
2003 >    packed in the best,direction into a layer. Because of this characteristic
2004 >    packing, spontaneous polarization appears parallel to the layer
2005 >    and switches on reversal of an applied electric field. This is the
2006 >    first obvious example of ferroelectricity in an achiral smectic
2007 >    phase and is ascribed to the C-2v symmetry of the molecular packing.},
2008    annote = {Ux855 Times Cited:447 Cited References Count:18},
2009    issn = {0959-9428},
2010    uri = {<Go to ISI>://A1996UX85500025},
# Line 2030 | Line 2020 | Encoding: GBK
2020    number = {5},
2021    month = {may},
2022    abstract = {We Studied the structural changes of bilayer vesicles induced by mechanical
2023 <        forces using a Brownian dynamics simulation. Two nanoparticles,
2024 <        which interact repulsively with amphiphilic molecules, are put inside
2025 <        a vesicle. The position of one nanoparticle is fixed, and the other
2026 <        is moved by a constant force as in optical-trapping experiments.
2027 <        First, the pulled vesicle stretches into a pear or tube shape. Then
2028 <        the inner monolayer in the tube-shaped region is deformed, and a
2029 <        cylindrical structure is formed between two vesicles. After stretching
2030 <        the cylindrical region, fission occurs near the moved vesicle. Soon
2031 <        after this the cylindrical region shrinks. The trapping force similar
2032 <        to 100 pN is needed to induce the formation of the cylindrical structure
2033 <        and fission.},
2023 >    forces using a Brownian dynamics simulation. Two nanoparticles,
2024 >    which interact repulsively with amphiphilic molecules, are put inside
2025 >    a vesicle. The position of one nanoparticle is fixed, and the other
2026 >    is moved by a constant force as in optical-trapping experiments.
2027 >    First, the pulled vesicle stretches into a pear or tube shape. Then
2028 >    the inner monolayer in the tube-shaped region is deformed, and a
2029 >    cylindrical structure is formed between two vesicles. After stretching
2030 >    the cylindrical region, fission occurs near the moved vesicle. Soon
2031 >    after this the cylindrical region shrinks. The trapping force similar
2032 >    to 100 pN is needed to induce the formation of the cylindrical structure
2033 >    and fission.},
2034    annote = {Part 1 568PX Times Cited:5 Cited References Count:39},
2035    issn = {1063-651X},
2036    uri = {<Go to ISI>://000176552300084},
# Line 2056 | Line 2046 | Encoding: GBK
2046    number = {20},
2047    month = {Nov 22},
2048    abstract = {We studied the fusion dynamics of vesicles using a Brownian dynamics
2049 <        simulation. Amphiphilic molecules spontaneously form vesicles with
2050 <        a bilayer structure. Two vesicles come into contact and form a stalk
2051 <        intermediate, in which a necklike structure only connects the outer
2052 <        monolayers, as predicted by the stalk hypothesis. We have found
2053 <        a new pathway of pore opening from stalks at high temperature: the
2054 <        elliptic stalk bends and contact between the ends of the arc-shaped
2055 <        stalk leads to pore opening. On the other hand, we have clarified
2056 <        that the pore-opening process at low temperature agrees with the
2057 <        modified stalk model: a pore is induced by contact between the inner
2058 <        monolayers inside the stalk. (C) 2001 American Institute of Physics.},
2049 >    simulation. Amphiphilic molecules spontaneously form vesicles with
2050 >    a bilayer structure. Two vesicles come into contact and form a stalk
2051 >    intermediate, in which a necklike structure only connects the outer
2052 >    monolayers, as predicted by the stalk hypothesis. We have found
2053 >    a new pathway of pore opening from stalks at high temperature: the
2054 >    elliptic stalk bends and contact between the ends of the arc-shaped
2055 >    stalk leads to pore opening. On the other hand, we have clarified
2056 >    that the pore-opening process at low temperature agrees with the
2057 >    modified stalk model: a pore is induced by contact between the inner
2058 >    monolayers inside the stalk. (C) 2001 American Institute of Physics.},
2059    annote = {491UW Times Cited:48 Cited References Count:25},
2060    issn = {0021-9606},
2061    uri = {<Go to ISI>://000172129300049},
# Line 2082 | Line 2072 | Encoding: GBK
2072   @ARTICLE{Omelyan1998,
2073    author = {I. P. Omelyan},
2074    title = {On the numerical integration of motion for rigid polyatomics: The
2075 <        modified quaternion approach},
2075 >    modified quaternion approach},
2076    journal = {Computers in Physics},
2077    year = {1998},
2078    volume = {12},
# Line 2090 | Line 2080 | Encoding: GBK
2080    number = {1},
2081    month = {Jan-Feb},
2082    abstract = {A revised version of the quaternion approach for numerical integration
2083 <        of the equations of motion for rigid polyatomic molecules is proposed.
2084 <        The modified approach is based on a formulation of the quaternion
2085 <        dynamics with constraints. This allows one to resolve the rigidity
2086 <        problem rigorously using constraint forces. It is shown that the
2087 <        procedure for preservation of molecular rigidity can be realized
2088 <        particularly simply within the Verlet algorithm in velocity form.
2089 <        We demonstrate that the method presented leads to an improved numerical
2090 <        stability with respect to the usual quaternion rescaling scheme
2091 <        and it is roughly as good as the cumbersome atomic-constraint technique.
2092 <        (C) 1998 American Institute of Physics.},
2083 >    of the equations of motion for rigid polyatomic molecules is proposed.
2084 >    The modified approach is based on a formulation of the quaternion
2085 >    dynamics with constraints. This allows one to resolve the rigidity
2086 >    problem rigorously using constraint forces. It is shown that the
2087 >    procedure for preservation of molecular rigidity can be realized
2088 >    particularly simply within the Verlet algorithm in velocity form.
2089 >    We demonstrate that the method presented leads to an improved numerical
2090 >    stability with respect to the usual quaternion rescaling scheme
2091 >    and it is roughly as good as the cumbersome atomic-constraint technique.
2092 >    (C) 1998 American Institute of Physics.},
2093    annote = {Yx279 Times Cited:12 Cited References Count:28},
2094    issn = {0894-1866},
2095    uri = {<Go to ISI>://000072024300025},
# Line 2108 | Line 2098 | Encoding: GBK
2098   @ARTICLE{Omelyan1998a,
2099    author = {I. P. Omelyan},
2100    title = {Algorithm for numerical integration of the rigid-body equations of
2101 <        motion},
2101 >    motion},
2102    journal = {Physical Review E},
2103    year = {1998},
2104    volume = {58},
# Line 2116 | Line 2106 | Encoding: GBK
2106    number = {1},
2107    month = {Jul},
2108    abstract = {An algorithm for numerical integration of the rigid-body equations
2109 <        of motion is proposed. The algorithm uses the leapfrog scheme and
2110 <        the quantities involved are angular velocities and orientational
2111 <        variables that can be expressed in terms of either principal axes
2112 <        or quaternions. Due to specific features of the algorithm, orthonormality
2113 <        and unit norms of the orientational variables are integrals of motion,
2114 <        despite an approximate character of the produced trajectories. It
2115 <        is shown that the method presented appears to be the most efficient
2116 <        among all such algorithms known.},
2109 >    of motion is proposed. The algorithm uses the leapfrog scheme and
2110 >    the quantities involved are angular velocities and orientational
2111 >    variables that can be expressed in terms of either principal axes
2112 >    or quaternions. Due to specific features of the algorithm, orthonormality
2113 >    and unit norms of the orientational variables are integrals of motion,
2114 >    despite an approximate character of the produced trajectories. It
2115 >    is shown that the method presented appears to be the most efficient
2116 >    among all such algorithms known.},
2117    annote = {101XL Times Cited:8 Cited References Count:22},
2118    issn = {1063-651X},
2119    uri = {<Go to ISI>://000074893400151},
# Line 2132 | Line 2122 | Encoding: GBK
2122   @ARTICLE{Orlandi2006,
2123    author = {S. Orlandi and R. Berardi and J. Steltzer and C. Zannoni},
2124    title = {A Monte Carlo study of the mesophases formed by polar bent-shaped
2125 <        molecules},
2125 >    molecules},
2126    journal = {Journal of Chemical Physics},
2127    year = {2006},
2128    volume = {124},
# Line 2140 | Line 2130 | Encoding: GBK
2130    number = {12},
2131    month = {Mar 28},
2132    abstract = {Liquid crystal phases formed by bent-shaped (or #banana#) molecules
2133 <        are currently of great interest. Here we investigate by Monte Carlo
2134 <        computer simulations the phases formed by rigid banana molecules
2135 <        modeled combining three Gay-Berne sites and containing either one
2136 <        central or two lateral and transversal dipoles. We show that changing
2137 <        the dipole position and orientation has a profound effect on the
2138 <        mesophase stability and molecular organization. In particular, we
2139 <        find a uniaxial nematic phase only for off-center dipolar models
2140 <        and tilted phases only for the one with terminal dipoles. (c) 2006
2141 <        American Institute of Physics.},
2133 >    are currently of great interest. Here we investigate by Monte Carlo
2134 >    computer simulations the phases formed by rigid banana molecules
2135 >    modeled combining three Gay-Berne sites and containing either one
2136 >    central or two lateral and transversal dipoles. We show that changing
2137 >    the dipole position and orientation has a profound effect on the
2138 >    mesophase stability and molecular organization. In particular, we
2139 >    find a uniaxial nematic phase only for off-center dipolar models
2140 >    and tilted phases only for the one with terminal dipoles. (c) 2006
2141 >    American Institute of Physics.},
2142    annote = {028CP Times Cited:0 Cited References Count:42},
2143    issn = {0021-9606},
2144    uri = {<Go to ISI>://000236464000072},
# Line 2164 | Line 2154 | Encoding: GBK
2154    number = {6},
2155    month = {Nov},
2156    abstract = {Continuous, explicit Runge-Kutta methods with the minimal number of
2157 <        stages are considered. These methods are continuously differentiable
2158 <        if and only if one of the stages is the FSAL evaluation. A characterization
2159 <        of a subclass of these methods is developed for orders 3, 4, and
2160 <        5. It is shown how the free parameters of these methods can be used
2161 <        either to minimize the continuous truncation error coefficients
2162 <        or to maximize the stability region. As a representative for these
2163 <        methods the fifth-order method with minimized error coefficients
2164 <        is chosen, supplied with an error estimation method, and analysed
2165 <        by using the DETEST software. The results are compared with a similar
2166 <        implementation of the Dormand-Prince 5(4) pair with interpolant,
2167 <        showing a significant advantage in the new method for the chosen
2168 <        problems.},
2157 >    stages are considered. These methods are continuously differentiable
2158 >    if and only if one of the stages is the FSAL evaluation. A characterization
2159 >    of a subclass of these methods is developed for orders 3, 4, and
2160 >    5. It is shown how the free parameters of these methods can be used
2161 >    either to minimize the continuous truncation error coefficients
2162 >    or to maximize the stability region. As a representative for these
2163 >    methods the fifth-order method with minimized error coefficients
2164 >    is chosen, supplied with an error estimation method, and analysed
2165 >    by using the DETEST software. The results are compared with a similar
2166 >    implementation of the Dormand-Prince 5(4) pair with interpolant,
2167 >    showing a significant advantage in the new method for the chosen
2168 >    problems.},
2169    annote = {Ju936 Times Cited:25 Cited References Count:20},
2170    issn = {0196-5204},
2171    uri = {<Go to ISI>://A1992JU93600013},
# Line 2184 | Line 2174 | Encoding: GBK
2174   @ARTICLE{Palacios1998,
2175    author = {J. L. Garcia-Palacios and F. J. Lazaro},
2176    title = {Langevin-dynamics study of the dynamical properties of small magnetic
2177 <        particles},
2177 >    particles},
2178    journal = {Physical Review B},
2179    year = {1998},
2180    volume = {58},
# Line 2192 | Line 2182 | Encoding: GBK
2182    number = {22},
2183    month = {Dec 1},
2184    abstract = {The stochastic Landau-Lifshitz-Gilbert equation of motion for a classical
2185 <        magnetic moment is numerically solved (properly observing the customary
2186 <        interpretation of it as a Stratonovich stochastic differential equation),
2187 <        in order to study the dynamics of magnetic nanoparticles. The corresponding
2188 <        Langevin-dynamics approach allows for the study of the fluctuating
2189 <        trajectories of individual magnetic moments, where we have encountered
2190 <        remarkable phenomena in the overbarrier rotation process, such as
2191 <        crossing-back or multiple crossing of the potential barrier, rooted
2192 <        in the gyromagnetic nature of the system. Concerning averaged quantities,
2193 <        we study the linear dynamic response of the archetypal ensemble
2194 <        of noninteracting classical magnetic moments with axially symmetric
2195 <        magnetic anisotropy. The results are compared with different analytical
2196 <        expressions used to model the relaxation of nanoparticle ensembles,
2197 <        assessing their accuracy. It has been found that, among a number
2198 <        of heuristic expressions for the linear dynamic susceptibility,
2199 <        only the simple formula proposed by Shliomis and Stepanov matches
2200 <        the coarse features of the susceptibility reasonably. By comparing
2201 <        the numerical results with the asymptotic formula of Storonkin {Sov.
2202 <        Phys. Crystallogr. 30, 489 (1985) [Kristallografiya 30, 841 (1985)]},
2203 <        the effects of the intra-potential-well relaxation modes on the
2204 <        low-temperature longitudinal dynamic response have been assessed,
2205 <        showing their relatively small reflection in the susceptibility
2206 <        curves but their dramatic influence on the phase shifts. Comparison
2207 <        of the numerical results with the exact zero-damping expression
2208 <        for the transverse susceptibility by Garanin, Ishchenko, and Panina
2209 <        {Theor. Math. Phys. (USSR) 82, 169 (1990) [Teor. Mat. Fit. 82, 242
2210 <        (1990)]}, reveals a sizable contribution of the spread of the precession
2211 <        frequencies of the magnetic moment in the anisotropy field to the
2212 <        dynamic response at intermediate-to-high temperatures. [S0163-1829
2213 <        (98)00446-9].},
2185 >    magnetic moment is numerically solved (properly observing the customary
2186 >    interpretation of it as a Stratonovich stochastic differential equation),
2187 >    in order to study the dynamics of magnetic nanoparticles. The corresponding
2188 >    Langevin-dynamics approach allows for the study of the fluctuating
2189 >    trajectories of individual magnetic moments, where we have encountered
2190 >    remarkable phenomena in the overbarrier rotation process, such as
2191 >    crossing-back or multiple crossing of the potential barrier, rooted
2192 >    in the gyromagnetic nature of the system. Concerning averaged quantities,
2193 >    we study the linear dynamic response of the archetypal ensemble
2194 >    of noninteracting classical magnetic moments with axially symmetric
2195 >    magnetic anisotropy. The results are compared with different analytical
2196 >    expressions used to model the relaxation of nanoparticle ensembles,
2197 >    assessing their accuracy. It has been found that, among a number
2198 >    of heuristic expressions for the linear dynamic susceptibility,
2199 >    only the simple formula proposed by Shliomis and Stepanov matches
2200 >    the coarse features of the susceptibility reasonably. By comparing
2201 >    the numerical results with the asymptotic formula of Storonkin {Sov.
2202 >    Phys. Crystallogr. 30, 489 (1985) [Kristallografiya 30, 841 (1985)]},
2203 >    the effects of the intra-potential-well relaxation modes on the
2204 >    low-temperature longitudinal dynamic response have been assessed,
2205 >    showing their relatively small reflection in the susceptibility
2206 >    curves but their dramatic influence on the phase shifts. Comparison
2207 >    of the numerical results with the exact zero-damping expression
2208 >    for the transverse susceptibility by Garanin, Ishchenko, and Panina
2209 >    {Theor. Math. Phys. (USSR) 82, 169 (1990) [Teor. Mat. Fit. 82, 242
2210 >    (1990)]}, reveals a sizable contribution of the spread of the precession
2211 >    frequencies of the magnetic moment in the anisotropy field to the
2212 >    dynamic response at intermediate-to-high temperatures. [S0163-1829
2213 >    (98)00446-9].},
2214    annote = {146XW Times Cited:66 Cited References Count:45},
2215    issn = {0163-1829},
2216    uri = {<Go to ISI>://000077460000052},
# Line 2257 | Line 2247 | Encoding: GBK
2247   @ARTICLE{Perram1985,
2248    author = {J. W. Perram and M. S. Wertheim},
2249    title = {Statistical-Mechanics of Hard Ellipsoids .1. Overlap Algorithm and
2250 <        the Contact Function},
2250 >    the Contact Function},
2251    journal = {Journal of Computational Physics},
2252    year = {1985},
2253    volume = {58},
# Line 2280 | Line 2270 | Encoding: GBK
2270   @ARTICLE{Perrin1936,
2271    author = {F. Perrin},
2272    title = {Mouvement brownien d'un ellipsoid(II). Rotation libre et depolarisation
2273 <        des fluorescences. Translation et diffusion de moleculese ellipsoidales},
2273 >    des fluorescences. Translation et diffusion de moleculese ellipsoidales},
2274    journal = {J. Phys. Radium},
2275    year = {1936},
2276    volume = {7},
# Line 2290 | Line 2280 | Encoding: GBK
2280   @ARTICLE{Perrin1934,
2281    author = {F. Perrin},
2282    title = {Mouvement brownien d'un ellipsoid(I). Dispersion dielectrique pour
2283 <        des molecules ellipsoidales},
2283 >    des molecules ellipsoidales},
2284    journal = {J. Phys. Radium},
2285    year = {1934},
2286    volume = {5},
# Line 2307 | Line 2297 | Encoding: GBK
2297    number = {1},
2298    month = {Sep},
2299    abstract = {X-ray diffraction data taken at high instrumental resolution were
2300 <        obtained for EPC and DMPC under various osmotic pressures, primarily
2301 <        at T = 30 degrees C. The headgroup thickness D-HH was obtained from
2302 <        relative electron density profiles. By using volumetric results
2303 <        and by comparing to gel phase DPPC we obtain areas A(EPC)(F) = 69.4
2304 <        +/- 1.1 Angstrom(2) and A(DMPC)(F) = 59.7 +/- 0.2 Angstrom(2). The
2305 <        analysis also gives estimates for the areal compressibility K-A.
2306 <        The A(F) results lead to other structural results regarding membrane
2307 <        thickness and associated waters. Using the recently determined absolute
2308 <        electrons density profile of DPPC, the AF results also lead to absolute
2309 <        electron density profiles and absolute continuous transforms \F(q)\
2310 <        for EPC and DMPC, Limited measurements of temperature dependence
2311 <        show directly that fluctuations increase with increasing temperature
2312 <        and that a small decrease in bending modulus K-c accounts for the
2313 <        increased water spacing reported by Simon et al. (1995) Biophys.
2314 <        J. 69, 1473-1483. (C) 1998 Elsevier Science Ireland Ltd. All rights
2315 <        reserved.},
2300 >    obtained for EPC and DMPC under various osmotic pressures, primarily
2301 >    at T = 30 degrees C. The headgroup thickness D-HH was obtained from
2302 >    relative electron density profiles. By using volumetric results
2303 >    and by comparing to gel phase DPPC we obtain areas A(EPC)(F) = 69.4
2304 >    +/- 1.1 Angstrom(2) and A(DMPC)(F) = 59.7 +/- 0.2 Angstrom(2). The
2305 >    analysis also gives estimates for the areal compressibility K-A.
2306 >    The A(F) results lead to other structural results regarding membrane
2307 >    thickness and associated waters. Using the recently determined absolute
2308 >    electrons density profile of DPPC, the AF results also lead to absolute
2309 >    electron density profiles and absolute continuous transforms \F(q)\
2310 >    for EPC and DMPC, Limited measurements of temperature dependence
2311 >    show directly that fluctuations increase with increasing temperature
2312 >    and that a small decrease in bending modulus K-c accounts for the
2313 >    increased water spacing reported by Simon et al. (1995) Biophys.
2314 >    J. 69, 1473-1483. (C) 1998 Elsevier Science Ireland Ltd. All rights
2315 >    reserved.},
2316    annote = {130AT Times Cited:98 Cited References Count:39},
2317    issn = {0009-3084},
2318    uri = {<Go to ISI>://000076497600007},
# Line 2331 | Line 2321 | Encoding: GBK
2321   @ARTICLE{Powles1973,
2322    author = {J.~G. Powles},
2323    title = {A general ellipsoid can not always serve as a modle for the rotational
2324 <        diffusion properties of arbitrary shaped rigid molecules},
2324 >    diffusion properties of arbitrary shaped rigid molecules},
2325    journal = {Advan. Phys.},
2326    year = {1973},
2327    volume = {22},
# Line 2341 | Line 2331 | Encoding: GBK
2331   @ARTICLE{Recio2004,
2332    author = {J. Fernandez-Recio and M. Totrov and R. Abagyan},
2333    title = {Identification of protein-protein interaction sites from docking
2334 <        energy landscapes},
2334 >    energy landscapes},
2335    journal = {Journal of Molecular Biology},
2336    year = {2004},
2337    volume = {335},
# Line 2349 | Line 2339 | Encoding: GBK
2339    number = {3},
2340    month = {Jan 16},
2341    abstract = {Protein recognition is one of the most challenging and intriguing
2342 <        problems in structural biology. Despite all the available structural,
2343 <        sequence and biophysical information about protein-protein complexes,
2344 <        the physico-chemical patterns, if any, that make a protein surface
2345 <        likely to be involved in protein-protein interactions, remain elusive.
2346 <        Here, we apply protein docking simulations and analysis of the interaction
2347 <        energy landscapes to identify protein-protein interaction sites.
2348 <        The new protocol for global docking based on multi-start global
2349 <        energy optimization of an allatom model of the ligand, with detailed
2350 <        receptor potentials and atomic solvation parameters optimized in
2351 <        a training set of 24 complexes, explores the conformational space
2352 <        around the whole receptor without restrictions. The ensembles of
2353 <        the rigid-body docking solutions generated by the simulations were
2354 <        subsequently used to project the docking energy landscapes onto
2355 <        the protein surfaces. We found that highly populated low-energy
2356 <        regions consistently corresponded to actual binding sites. The procedure
2357 <        was validated on a test set of 21 known protein-protein complexes
2358 <        not used in the training set. As much as 81% of the predicted high-propensity
2359 <        patch residues were located correctly in the native interfaces.
2360 <        This approach can guide the design of mutations on the surfaces
2361 <        of proteins, provide geometrical details of a possible interaction,
2362 <        and help to annotate protein surfaces in structural proteomics.
2363 <        (C) 2003 Elsevier Ltd. All rights reserved.},
2342 >    problems in structural biology. Despite all the available structural,
2343 >    sequence and biophysical information about protein-protein complexes,
2344 >    the physico-chemical patterns, if any, that make a protein surface
2345 >    likely to be involved in protein-protein interactions, remain elusive.
2346 >    Here, we apply protein docking simulations and analysis of the interaction
2347 >    energy landscapes to identify protein-protein interaction sites.
2348 >    The new protocol for global docking based on multi-start global
2349 >    energy optimization of an allatom model of the ligand, with detailed
2350 >    receptor potentials and atomic solvation parameters optimized in
2351 >    a training set of 24 complexes, explores the conformational space
2352 >    around the whole receptor without restrictions. The ensembles of
2353 >    the rigid-body docking solutions generated by the simulations were
2354 >    subsequently used to project the docking energy landscapes onto
2355 >    the protein surfaces. We found that highly populated low-energy
2356 >    regions consistently corresponded to actual binding sites. The procedure
2357 >    was validated on a test set of 21 known protein-protein complexes
2358 >    not used in the training set. As much as 81% of the predicted high-propensity
2359 >    patch residues were located correctly in the native interfaces.
2360 >    This approach can guide the design of mutations on the surfaces
2361 >    of proteins, provide geometrical details of a possible interaction,
2362 >    and help to annotate protein surfaces in structural proteomics.
2363 >    (C) 2003 Elsevier Ltd. All rights reserved.},
2364    annote = {763GQ Times Cited:21 Cited References Count:59},
2365    issn = {0022-2836},
2366    uri = {<Go to ISI>://000188066900016},
# Line 2379 | Line 2369 | Encoding: GBK
2369   @ARTICLE{Reddy2006,
2370    author = {R. A. Reddy and C. Tschierske},
2371    title = {Bent-core liquid crystals: polar order, superstructural chirality
2372 <        and spontaneous desymmetrisation in soft matter systems},
2372 >    and spontaneous desymmetrisation in soft matter systems},
2373    journal = {Journal of Materials Chemistry},
2374    year = {2006},
2375    volume = {16},
2376    pages = {907-961},
2377    number = {10},
2378    abstract = {An overview on the recent developments in the field of liquid crystalline
2379 <        bent-core molecules (so-called banana liquid crystals) is given.
2380 <        After some basic issues, dealing with general aspects of the systematisation
2381 <        of the mesophases, development of polar order and chirality in this
2382 <        class of LC systems and explaining some general structure-property
2383 <        relationships, we focus on fascinating new developments in this
2384 <        field, such as modulated, undulated and columnar phases, so-called
2385 <        B7 phases, phase biaxiality, ferroelectric and antiferroelectric
2386 <        polar order in smectic and columnar phases, amplification and switching
2387 <        of chirality and the spontaneous formation of superstructural and
2388 <        supramolecular chirality.},
2379 >    bent-core molecules (so-called banana liquid crystals) is given.
2380 >    After some basic issues, dealing with general aspects of the systematisation
2381 >    of the mesophases, development of polar order and chirality in this
2382 >    class of LC systems and explaining some general structure-property
2383 >    relationships, we focus on fascinating new developments in this
2384 >    field, such as modulated, undulated and columnar phases, so-called
2385 >    B7 phases, phase biaxiality, ferroelectric and antiferroelectric
2386 >    polar order in smectic and columnar phases, amplification and switching
2387 >    of chirality and the spontaneous formation of superstructural and
2388 >    supramolecular chirality.},
2389    annote = {021NS Times Cited:2 Cited References Count:316},
2390    issn = {0959-9428},
2391    uri = {<Go to ISI>://000235990500001},
# Line 2411 | Line 2401 | Encoding: GBK
2401    number = {5},
2402    month = {Sep 8},
2403    abstract = {Backward error analysis has become an important tool for understanding
2404 <        the long time behavior of numerical integration methods. This is
2405 <        true in particular for the integration of Hamiltonian systems where
2406 <        backward error analysis can be used to show that a symplectic method
2407 <        will conserve energy over exponentially long periods of time. Such
2408 <        results are typically based on two aspects of backward error analysis:
2409 <        (i) It can be shown that the modified vector fields have some qualitative
2410 <        properties which they share with the given problem and (ii) an estimate
2411 <        is given for the difference between the best interpolating vector
2412 <        field and the numerical method. These aspects have been investigated
2413 <        recently, for example, by Benettin and Giorgilli in [J. Statist.
2414 <        Phys., 74 (1994), pp. 1117-1143], by Hairer in [Ann. Numer. Math.,
2415 <        1 (1994), pp. 107-132], and by Hairer and Lubich in [Numer. Math.,
2416 <        76 (1997), pp. 441-462]. In this paper we aim at providing a unifying
2417 <        framework and a simplification of the existing results and corresponding
2418 <        proofs. Our approach to backward error analysis is based on a simple
2419 <        recursive definition of the modified vector fields that does not
2420 <        require explicit Taylor series expansion of the numerical method
2421 <        and the corresponding flow maps as in the above-cited works. As
2422 <        an application we discuss the long time integration of chaotic Hamiltonian
2423 <        systems and the approximation of time averages along numerically
2424 <        computed trajectories.},
2404 >    the long time behavior of numerical integration methods. This is
2405 >    true in particular for the integration of Hamiltonian systems where
2406 >    backward error analysis can be used to show that a symplectic method
2407 >    will conserve energy over exponentially long periods of time. Such
2408 >    results are typically based on two aspects of backward error analysis:
2409 >    (i) It can be shown that the modified vector fields have some qualitative
2410 >    properties which they share with the given problem and (ii) an estimate
2411 >    is given for the difference between the best interpolating vector
2412 >    field and the numerical method. These aspects have been investigated
2413 >    recently, for example, by Benettin and Giorgilli in [J. Statist.
2414 >    Phys., 74 (1994), pp. 1117-1143], by Hairer in [Ann. Numer. Math.,
2415 >    1 (1994), pp. 107-132], and by Hairer and Lubich in [Numer. Math.,
2416 >    76 (1997), pp. 441-462]. In this paper we aim at providing a unifying
2417 >    framework and a simplification of the existing results and corresponding
2418 >    proofs. Our approach to backward error analysis is based on a simple
2419 >    recursive definition of the modified vector fields that does not
2420 >    require explicit Taylor series expansion of the numerical method
2421 >    and the corresponding flow maps as in the above-cited works. As
2422 >    an application we discuss the long time integration of chaotic Hamiltonian
2423 >    systems and the approximation of time averages along numerically
2424 >    computed trajectories.},
2425    annote = {237HV Times Cited:43 Cited References Count:41},
2426    issn = {0036-1429},
2427    uri = {<Go to ISI>://000082650600010},
# Line 2446 | Line 2436 | Encoding: GBK
2436    pages = {5093-5098},
2437    number = {48},
2438    abstract = {The recent literature in the field of liquid crystals shows that banana-shaped
2439 <        mesogenic materials represent a bewitching and stimulating field
2440 <        of research that is interesting both academically and in terms of
2441 <        applications. Numerous topics are open to investigation in this
2442 <        area because of the rich phenomenology and new possibilities that
2443 <        these materials offer. The principal concepts in this area are reviewed
2444 <        along with recent results. In addition, new directions to stimulate
2445 <        further research activities are highlighted.},
2439 >    mesogenic materials represent a bewitching and stimulating field
2440 >    of research that is interesting both academically and in terms of
2441 >    applications. Numerous topics are open to investigation in this
2442 >    area because of the rich phenomenology and new possibilities that
2443 >    these materials offer. The principal concepts in this area are reviewed
2444 >    along with recent results. In addition, new directions to stimulate
2445 >    further research activities are highlighted.},
2446    annote = {990XA Times Cited:3 Cited References Count:72},
2447    issn = {0959-9428},
2448    uri = {<Go to ISI>://000233775500001},
# Line 2461 | Line 2451 | Encoding: GBK
2451   @ARTICLE{Roy2005,
2452    author = {A. Roy and N. V. Madhusudana},
2453    title = {A frustrated packing model for the B-6-B-1-SmAP(A) sequence of phases
2454 <        in banana shaped molecules},
2454 >    in banana shaped molecules},
2455    journal = {European Physical Journal E},
2456    year = {2005},
2457    volume = {18},
# Line 2469 | Line 2459 | Encoding: GBK
2459    number = {3},
2460    month = {Nov},
2461    abstract = {A vast majority of compounds with bent core or banana shaped molecules
2462 <        exhibit the phase sequence B-6-B-1-B-2 as the chain length is increased
2463 <        in a homologous series. The B-6 phase has an intercalated fluid
2464 <        lamellar structure with a layer spacing of half the molecular length.
2465 <        The B-1 phase has a two dimensionally periodic rectangular columnar
2466 <        structure. The B-2 phase has a monolayer fluid lamellar structure
2467 <        with molecules tilted with respect to the layer normal. Neglecting
2468 <        the tilt order of the molecules in the B-2 phase, we have developed
2469 <        a frustrated packing model to describe this phase sequence qualitatively.
2470 <        The model has some analogy with that of the frustrated smectics
2471 <        exhibited by highly polar rod like molecules.},
2462 >    exhibit the phase sequence B-6-B-1-B-2 as the chain length is increased
2463 >    in a homologous series. The B-6 phase has an intercalated fluid
2464 >    lamellar structure with a layer spacing of half the molecular length.
2465 >    The B-1 phase has a two dimensionally periodic rectangular columnar
2466 >    structure. The B-2 phase has a monolayer fluid lamellar structure
2467 >    with molecules tilted with respect to the layer normal. Neglecting
2468 >    the tilt order of the molecules in the B-2 phase, we have developed
2469 >    a frustrated packing model to describe this phase sequence qualitatively.
2470 >    The model has some analogy with that of the frustrated smectics
2471 >    exhibited by highly polar rod like molecules.},
2472    annote = {985FW Times Cited:0 Cited References Count:30},
2473    issn = {1292-8941},
2474    uri = {<Go to ISI>://000233363300002},
# Line 2487 | Line 2477 | Encoding: GBK
2477   @ARTICLE{Ryckaert1977,
2478    author = {J. P. Ryckaert and G. Ciccotti and H. J. C. Berendsen},
2479    title = {Numerical-Integration of Cartesian Equations of Motion of a System
2480 <        with Constraints - Molecular-Dynamics of N-Alkanes},
2480 >    with Constraints - Molecular-Dynamics of N-Alkanes},
2481    journal = {Journal of Computational Physics},
2482    year = {1977},
2483    volume = {23},
# Line 2501 | Line 2491 | Encoding: GBK
2491   @ARTICLE{Sagui1999,
2492    author = {C. Sagui and T. A. Darden},
2493    title = {Molecular dynamics simulations of biomolecules: Long-range electrostatic
2494 <        effects},
2494 >    effects},
2495    journal = {Annual Review of Biophysics and Biomolecular Structure},
2496    year = {1999},
2497    volume = {28},
2498    pages = {155-179},
2499    abstract = {Current computer simulations of biomolecules typically make use of
2500 <        classical molecular dynamics methods, as a very large number (tens
2501 <        to hundreds of thousands) of atoms are involved over timescales
2502 <        of many nanoseconds. The methodology for treating short-range bonded
2503 <        and van der Waals interactions has matured. However, long-range
2504 <        electrostatic interactions still represent a bottleneck in simulations.
2505 <        In this article, we introduce the basic issues for an accurate representation
2506 <        of the relevant electrostatic interactions. In spite of the huge
2507 <        computational time demanded by most biomolecular systems, it is
2508 <        no longer necessary to resort to uncontrolled approximations such
2509 <        as the use of cutoffs. In particular, we discuss the Ewald summation
2510 <        methods, the fast particle mesh methods, and the fast multipole
2511 <        methods. We also review recent efforts to understand the role of
2512 <        boundary conditions in systems with long-range interactions, and
2513 <        conclude with a short perspective on future trends.},
2500 >    classical molecular dynamics methods, as a very large number (tens
2501 >    to hundreds of thousands) of atoms are involved over timescales
2502 >    of many nanoseconds. The methodology for treating short-range bonded
2503 >    and van der Waals interactions has matured. However, long-range
2504 >    electrostatic interactions still represent a bottleneck in simulations.
2505 >    In this article, we introduce the basic issues for an accurate representation
2506 >    of the relevant electrostatic interactions. In spite of the huge
2507 >    computational time demanded by most biomolecular systems, it is
2508 >    no longer necessary to resort to uncontrolled approximations such
2509 >    as the use of cutoffs. In particular, we discuss the Ewald summation
2510 >    methods, the fast particle mesh methods, and the fast multipole
2511 >    methods. We also review recent efforts to understand the role of
2512 >    boundary conditions in systems with long-range interactions, and
2513 >    conclude with a short perspective on future trends.},
2514    annote = {213KJ Times Cited:126 Cited References Count:73},
2515    issn = {1056-8700},
2516    uri = {<Go to ISI>://000081271400008},
# Line 2529 | Line 2519 | Encoding: GBK
2519   @ARTICLE{Sandu1999,
2520    author = {A. Sandu and T. Schlick},
2521    title = {Masking resonance artifacts in force-splitting methods for biomolecular
2522 <        simulations by extrapolative Langevin dynamics},
2522 >    simulations by extrapolative Langevin dynamics},
2523    journal = {Journal of Computational Physics},
2524    year = {1999},
2525    volume = {151},
# Line 2537 | Line 2527 | Encoding: GBK
2527    number = {1},
2528    month = {May 1},
2529    abstract = {Numerical resonance artifacts have become recognized recently as a
2530 <        limiting factor to increasing the timestep in multiple-timestep
2531 <        (MTS) biomolecular dynamics simulations. At certain timesteps correlated
2532 <        to internal motions (e.g., 5 fs, around half the period of the fastest
2533 <        bond stretch, T-min), visible inaccuracies or instabilities can
2534 <        occur. Impulse-MTS schemes are vulnerable to these resonance errors
2535 <        since large energy pulses are introduced to the governing dynamics
2536 <        equations when the slow forces are evaluated. We recently showed
2537 <        that such resonance artifacts can be masked significantly by applying
2538 <        extrapolative splitting to stochastic dynamics. Theoretical and
2539 <        numerical analyses of force-splitting integrators based on the Verlet
2540 <        discretization are reported here for linear models to explain these
2541 <        observations and to suggest how to construct effective integrators
2542 <        for biomolecular dynamics that balance stability with accuracy.
2543 <        Analyses for Newtonian dynamics demonstrate the severe resonance
2544 <        patterns of the Impulse splitting, with this severity worsening
2545 <        with the outer timestep. Delta t: Constant Extrapolation is generally
2546 <        unstable, but the disturbances do not grow with Delta t. Thus. the
2547 <        stochastic extrapolative combination can counteract generic instabilities
2548 <        and largely alleviate resonances with a sufficiently strong Langevin
2549 <        heat-bath coupling (gamma), estimates for which are derived here
2550 <        based on the fastest and slowest motion periods. These resonance
2551 <        results generally hold for nonlinear test systems: a water tetramer
2552 <        and solvated protein. Proposed related approaches such as Extrapolation/Correction
2553 <        and Midpoint Extrapolation work better than Constant Extrapolation
2554 <        only for timesteps less than T-min/2. An effective extrapolative
2555 <        stochastic approach for biomolecules that balances long-timestep
2556 <        stability with good accuracy for the fast subsystem is then applied
2557 <        to a biomolecule using a three-class partitioning: the medium forces
2558 <        are treated by Midpoint Extrapolation via position Verlet, and the
2559 <        slow forces are incorporated by Constant Extrapolation. The resulting
2560 <        algorithm (LN) performs well on a solvated protein system in terms
2561 <        of thermodynamic properties and yields an order of magnitude speedup
2562 <        with respect to single-timestep Langevin trajectories. Computed
2563 <        spectral density functions also show how the Newtonian modes can
2564 <        be approximated by using a small gamma in the range Of 5-20 ps(-1).
2565 <        (C) 1999 Academic Press.},
2530 >    limiting factor to increasing the timestep in multiple-timestep
2531 >    (MTS) biomolecular dynamics simulations. At certain timesteps correlated
2532 >    to internal motions (e.g., 5 fs, around half the period of the fastest
2533 >    bond stretch, T-min), visible inaccuracies or instabilities can
2534 >    occur. Impulse-MTS schemes are vulnerable to these resonance errors
2535 >    since large energy pulses are introduced to the governing dynamics
2536 >    equations when the slow forces are evaluated. We recently showed
2537 >    that such resonance artifacts can be masked significantly by applying
2538 >    extrapolative splitting to stochastic dynamics. Theoretical and
2539 >    numerical analyses of force-splitting integrators based on the Verlet
2540 >    discretization are reported here for linear models to explain these
2541 >    observations and to suggest how to construct effective integrators
2542 >    for biomolecular dynamics that balance stability with accuracy.
2543 >    Analyses for Newtonian dynamics demonstrate the severe resonance
2544 >    patterns of the Impulse splitting, with this severity worsening
2545 >    with the outer timestep. Delta t: Constant Extrapolation is generally
2546 >    unstable, but the disturbances do not grow with Delta t. Thus. the
2547 >    stochastic extrapolative combination can counteract generic instabilities
2548 >    and largely alleviate resonances with a sufficiently strong Langevin
2549 >    heat-bath coupling (gamma), estimates for which are derived here
2550 >    based on the fastest and slowest motion periods. These resonance
2551 >    results generally hold for nonlinear test systems: a water tetramer
2552 >    and solvated protein. Proposed related approaches such as Extrapolation/Correction
2553 >    and Midpoint Extrapolation work better than Constant Extrapolation
2554 >    only for timesteps less than T-min/2. An effective extrapolative
2555 >    stochastic approach for biomolecules that balances long-timestep
2556 >    stability with good accuracy for the fast subsystem is then applied
2557 >    to a biomolecule using a three-class partitioning: the medium forces
2558 >    are treated by Midpoint Extrapolation via position Verlet, and the
2559 >    slow forces are incorporated by Constant Extrapolation. The resulting
2560 >    algorithm (LN) performs well on a solvated protein system in terms
2561 >    of thermodynamic properties and yields an order of magnitude speedup
2562 >    with respect to single-timestep Langevin trajectories. Computed
2563 >    spectral density functions also show how the Newtonian modes can
2564 >    be approximated by using a small gamma in the range Of 5-20 ps(-1).
2565 >    (C) 1999 Academic Press.},
2566    annote = {194FM Times Cited:14 Cited References Count:32},
2567    issn = {0021-9991},
2568    uri = {<Go to ISI>://000080181500004},
# Line 2581 | Line 2571 | Encoding: GBK
2571   @ARTICLE{Satoh1996,
2572    author = {K. Satoh and S. Mita and S. Kondo},
2573    title = {Monte Carlo simulations using the dipolar Gay-Berne model: Effect
2574 <        of terminal dipole moment on mesophase formation},
2574 >    of terminal dipole moment on mesophase formation},
2575    journal = {Chemical Physics Letters},
2576    year = {1996},
2577    volume = {255},
# Line 2589 | Line 2579 | Encoding: GBK
2579    number = {1-3},
2580    month = {Jun 7},
2581    abstract = {The effects of dipole-dipole interaction on mesophase formation are
2582 <        investigated with a Monte Carlo simulation using the dipolar Gay-Berne
2583 <        potential. It is shown that the dipole moment at the end of a molecule
2584 <        causes a shift in the nematic-isotropic transition toward higher
2585 <        temperature and a spread of the temperature range of the nematic
2586 <        phase and that layer structures with various interdigitations are
2587 <        formed in the smectic phase.},
2582 >    investigated with a Monte Carlo simulation using the dipolar Gay-Berne
2583 >    potential. It is shown that the dipole moment at the end of a molecule
2584 >    causes a shift in the nematic-isotropic transition toward higher
2585 >    temperature and a spread of the temperature range of the nematic
2586 >    phase and that layer structures with various interdigitations are
2587 >    formed in the smectic phase.},
2588    annote = {Uq975 Times Cited:32 Cited References Count:33},
2589    issn = {0009-2614},
2590    uri = {<Go to ISI>://A1996UQ97500017},
# Line 2603 | Line 2593 | Encoding: GBK
2593   @ARTICLE{Shen2002,
2594    author = {M. Y. Shen and K. F. Freed},
2595    title = {Long time dynamics of met-enkephalin: Comparison of explicit and
2596 <        implicit solvent models},
2596 >    implicit solvent models},
2597    journal = {Biophysical Journal},
2598    year = {2002},
2599    volume = {82},
# Line 2611 | Line 2601 | Encoding: GBK
2601    number = {4},
2602    month = {Apr},
2603    abstract = {Met-enkephalin is one of the smallest opiate peptides. Yet, its dynamical
2604 <        structure and receptor docking mechanism are still not well understood.
2605 <        The conformational dynamics of this neuron peptide in liquid water
2606 <        are studied here by using all-atom molecular dynamics (MID) and
2607 <        implicit water Langevin dynamics (LD) simulations with AMBER potential
2608 <        functions and the three-site transferable intermolecular potential
2609 <        (TIP3P) model for water. To achieve the same simulation length in
2610 <        physical time, the full MID simulations require 200 times as much
2611 <        CPU time as the implicit water LID simulations. The solvent hydrophobicity
2612 <        and dielectric behavior are treated in the implicit solvent LD simulations
2613 <        by using a macroscopic solvation potential, a single dielectric
2614 <        constant, and atomic friction coefficients computed using the accessible
2615 <        surface area method with the TIP3P model water viscosity as determined
2616 <        here from MID simulations for pure TIP3P water. Both the local and
2617 <        the global dynamics obtained from the implicit solvent LD simulations
2618 <        agree very well with those from the explicit solvent MD simulations.
2619 <        The simulations provide insights into the conformational restrictions
2620 <        that are associated with the bioactivity of the opiate peptide dermorphin
2621 <        for the delta-receptor.},
2604 >    structure and receptor docking mechanism are still not well understood.
2605 >    The conformational dynamics of this neuron peptide in liquid water
2606 >    are studied here by using all-atom molecular dynamics (MID) and
2607 >    implicit water Langevin dynamics (LD) simulations with AMBER potential
2608 >    functions and the three-site transferable intermolecular potential
2609 >    (TIP3P) model for water. To achieve the same simulation length in
2610 >    physical time, the full MID simulations require 200 times as much
2611 >    CPU time as the implicit water LID simulations. The solvent hydrophobicity
2612 >    and dielectric behavior are treated in the implicit solvent LD simulations
2613 >    by using a macroscopic solvation potential, a single dielectric
2614 >    constant, and atomic friction coefficients computed using the accessible
2615 >    surface area method with the TIP3P model water viscosity as determined
2616 >    here from MID simulations for pure TIP3P water. Both the local and
2617 >    the global dynamics obtained from the implicit solvent LD simulations
2618 >    agree very well with those from the explicit solvent MD simulations.
2619 >    The simulations provide insights into the conformational restrictions
2620 >    that are associated with the bioactivity of the opiate peptide dermorphin
2621 >    for the delta-receptor.},
2622    annote = {540MH Times Cited:36 Cited References Count:45},
2623    issn = {0006-3495},
2624    uri = {<Go to ISI>://000174932400010},
# Line 2651 | Line 2641 | Encoding: GBK
2641   @ARTICLE{Shimada1993,
2642    author = {J. Shimada and H. Kaneko and T. Takada},
2643    title = {Efficient Calculations of Coulombic Interactions in Biomolecular
2644 <        Simulations with Periodic Boundary-Conditions},
2644 >    Simulations with Periodic Boundary-Conditions},
2645    journal = {Journal of Computational Chemistry},
2646    year = {1993},
2647    volume = {14},
# Line 2659 | Line 2649 | Encoding: GBK
2649    number = {7},
2650    month = {Jul},
2651    abstract = {To make improved treatments of electrostatic interactions in biomacromolecular
2652 <        simulations, two possibilities are considered. The first is the
2653 <        famous particle-particle and particle-mesh (PPPM) method developed
2654 <        by Hockney and Eastwood, and the second is a new one developed here
2655 <        in their spirit but by the use of the multipole expansion technique
2656 <        suggested by Ladd. It is then numerically found that the new PPPM
2657 <        method gives more accurate results for a two-particle system at
2658 <        small separation of particles. Preliminary numerical examination
2659 <        of the various computational methods for a single configuration
2660 <        of a model BPTI-water system containing about 24,000 particles indicates
2661 <        that both of the PPPM methods give far more accurate values with
2662 <        reasonable computational cost than do the conventional truncation
2663 <        methods. It is concluded the two PPPM methods are nearly comparable
2664 <        in overall performance for the many-particle systems, although the
2665 <        first method has the drawback that the accuracy in the total electrostatic
2666 <        energy is not high for configurations of charged particles randomly
2667 <        generated.},
2652 >    simulations, two possibilities are considered. The first is the
2653 >    famous particle-particle and particle-mesh (PPPM) method developed
2654 >    by Hockney and Eastwood, and the second is a new one developed here
2655 >    in their spirit but by the use of the multipole expansion technique
2656 >    suggested by Ladd. It is then numerically found that the new PPPM
2657 >    method gives more accurate results for a two-particle system at
2658 >    small separation of particles. Preliminary numerical examination
2659 >    of the various computational methods for a single configuration
2660 >    of a model BPTI-water system containing about 24,000 particles indicates
2661 >    that both of the PPPM methods give far more accurate values with
2662 >    reasonable computational cost than do the conventional truncation
2663 >    methods. It is concluded the two PPPM methods are nearly comparable
2664 >    in overall performance for the many-particle systems, although the
2665 >    first method has the drawback that the accuracy in the total electrostatic
2666 >    energy is not high for configurations of charged particles randomly
2667 >    generated.},
2668    annote = {Lh164 Times Cited:27 Cited References Count:47},
2669    issn = {0192-8651},
2670    uri = {<Go to ISI>://A1993LH16400011},
# Line 2690 | Line 2680 | Encoding: GBK
2680    number = {24},
2681    month = {Dec 20},
2682    abstract = {The best simple method for Newtonian molecular dynamics is indisputably
2683 <        the leapfrog Stormer-Verlet method. The appropriate generalization
2684 <        to simple Langevin dynamics is unclear. An analysis is presented
2685 <        comparing an 'impulse method' (kick; fluctuate; kick), the 1982
2686 <        method of van Gunsteren and Berendsen, and the Brunger-Brooks-Karplus
2687 <        (BBK) method. It is shown how the impulse method and the van Gunsteren-Berendsen
2688 <        methods can be implemented as efficiently as the BBK method. Other
2689 <        considerations suggest that the impulse method is the best basic
2690 <        method for simple Langevin dynamics, with the van Gunsteren-Berendsen
2691 <        method a close contender.},
2683 >    the leapfrog Stormer-Verlet method. The appropriate generalization
2684 >    to simple Langevin dynamics is unclear. An analysis is presented
2685 >    comparing an 'impulse method' (kick; fluctuate; kick), the 1982
2686 >    method of van Gunsteren and Berendsen, and the Brunger-Brooks-Karplus
2687 >    (BBK) method. It is shown how the impulse method and the van Gunsteren-Berendsen
2688 >    methods can be implemented as efficiently as the BBK method. Other
2689 >    considerations suggest that the impulse method is the best basic
2690 >    method for simple Langevin dynamics, with the van Gunsteren-Berendsen
2691 >    method a close contender.},
2692    annote = {633RX Times Cited:8 Cited References Count:22},
2693    issn = {0026-8976},
2694    uri = {<Go to ISI>://000180297200014},
# Line 2707 | Line 2697 | Encoding: GBK
2697   @ARTICLE{Skeel1997,
2698    author = {R. D. Skeel and G. H. Zhang and T. Schlick},
2699    title = {A family of symplectic integrators: Stability, accuracy, and molecular
2700 <        dynamics applications},
2700 >    dynamics applications},
2701    journal = {Siam Journal on Scientific Computing},
2702    year = {1997},
2703    volume = {18},
# Line 2715 | Line 2705 | Encoding: GBK
2705    number = {1},
2706    month = {Jan},
2707    abstract = {The following integration methods for special second-order ordinary
2708 <        differential equations are studied: leapfrog, implicit midpoint,
2709 <        trapezoid, Stormer-Verlet, and Cowell-Numerov. We show that all
2710 <        are members, or equivalent to members, of a one-parameter family
2711 <        of schemes. Some methods have more than one common form, and we
2712 <        discuss a systematic enumeration of these forms. We also present
2713 <        a stability and accuracy analysis based on the idea of ''modified
2714 <        equations'' and a proof of symplecticness. It follows that Cowell-Numerov
2715 <        and ''LIM2'' (a method proposed by Zhang and Schlick) are symplectic.
2716 <        A different interpretation of the values used by these integrators
2717 <        leads to higher accuracy and better energy conservation. Hence,
2718 <        we suggest that the straightforward analysis of energy conservation
2719 <        is misleading.},
2708 >    differential equations are studied: leapfrog, implicit midpoint,
2709 >    trapezoid, Stormer-Verlet, and Cowell-Numerov. We show that all
2710 >    are members, or equivalent to members, of a one-parameter family
2711 >    of schemes. Some methods have more than one common form, and we
2712 >    discuss a systematic enumeration of these forms. We also present
2713 >    a stability and accuracy analysis based on the idea of ''modified
2714 >    equations'' and a proof of symplecticness. It follows that Cowell-Numerov
2715 >    and ''LIM2'' (a method proposed by Zhang and Schlick) are symplectic.
2716 >    A different interpretation of the values used by these integrators
2717 >    leads to higher accuracy and better energy conservation. Hence,
2718 >    we suggest that the straightforward analysis of energy conservation
2719 >    is misleading.},
2720    annote = {We981 Times Cited:30 Cited References Count:35},
2721    issn = {1064-8275},
2722    uri = {<Go to ISI>://A1997WE98100012},
# Line 2734 | Line 2724 | Encoding: GBK
2724  
2725   @ARTICLE{Tao2005,
2726    author = {Y. G. Tao and W. K. {den Otter} and J. T. Padding and J. K. G. Dhont
2727 <        and W. J. Briels},
2727 >    and W. J. Briels},
2728    title = {Brownian dynamics simulations of the self- and collective rotational
2729 <        diffusion coefficients of rigid long thin rods},
2729 >    diffusion coefficients of rigid long thin rods},
2730    journal = {Journal of Chemical Physics},
2731    year = {2005},
2732    volume = {122},
# Line 2744 | Line 2734 | Encoding: GBK
2734    number = {24},
2735    month = {Jun 22},
2736    abstract = {Recently a microscopic theory for the dynamics of suspensions of long
2737 <        thin rigid rods was presented, confirming and expanding the well-known
2738 <        theory by Doi and Edwards [The Theory of Polymer Dynamics (Clarendon,
2739 <        Oxford, 1986)] and Kuzuu [J. Phys. Soc. Jpn. 52, 3486 (1983)]. Here
2740 <        this theory is put to the test by comparing it against computer
2741 <        simulations. A Brownian dynamics simulation program was developed
2742 <        to follow the dynamics of the rods, with a length over a diameter
2743 <        ratio of 60, on the Smoluchowski time scale. The model accounts
2744 <        for excluded volume interactions between rods, but neglects hydrodynamic
2745 <        interactions. The self-rotational diffusion coefficients D-r(phi)
2746 <        of the rods were calculated by standard methods and by a new, more
2747 <        efficient method based on calculating average restoring torques.
2748 <        Collective decay of orientational order was calculated by means
2749 <        of equilibrium and nonequilibrium simulations. Our results show
2750 <        that, for the currently accessible volume fractions, the decay times
2751 <        in both cases are virtually identical. Moreover, the observed decay
2752 <        of diffusion coefficients with volume fraction is much quicker than
2753 <        predicted by the theory, which is attributed to an oversimplification
2754 <        of dynamic correlations in the theory. (c) 2005 American Institute
2755 <        of Physics.},
2737 >    thin rigid rods was presented, confirming and expanding the well-known
2738 >    theory by Doi and Edwards [The Theory of Polymer Dynamics (Clarendon,
2739 >    Oxford, 1986)] and Kuzuu [J. Phys. Soc. Jpn. 52, 3486 (1983)]. Here
2740 >    this theory is put to the test by comparing it against computer
2741 >    simulations. A Brownian dynamics simulation program was developed
2742 >    to follow the dynamics of the rods, with a length over a diameter
2743 >    ratio of 60, on the Smoluchowski time scale. The model accounts
2744 >    for excluded volume interactions between rods, but neglects hydrodynamic
2745 >    interactions. The self-rotational diffusion coefficients D-r(phi)
2746 >    of the rods were calculated by standard methods and by a new, more
2747 >    efficient method based on calculating average restoring torques.
2748 >    Collective decay of orientational order was calculated by means
2749 >    of equilibrium and nonequilibrium simulations. Our results show
2750 >    that, for the currently accessible volume fractions, the decay times
2751 >    in both cases are virtually identical. Moreover, the observed decay
2752 >    of diffusion coefficients with volume fraction is much quicker than
2753 >    predicted by the theory, which is attributed to an oversimplification
2754 >    of dynamic correlations in the theory. (c) 2005 American Institute
2755 >    of Physics.},
2756    annote = {943DN Times Cited:3 Cited References Count:26},
2757    issn = {0021-9606},
2758    uri = {<Go to ISI>://000230332400077},
# Line 2781 | Line 2771 | Encoding: GBK
2771   @ARTICLE{Tu1995,
2772    author = {K. Tu and D. J. Tobias and M. L. Klein},
2773    title = {Constant pressure and temperature molecular dynamics simulation of
2774 <        a fully hydrated liquid crystal phase dipalmitoylphosphatidylcholine
2775 <        bilayer},
2774 >    a fully hydrated liquid crystal phase dipalmitoylphosphatidylcholine
2775 >    bilayer},
2776    journal = {Biophysical Journal},
2777    year = {1995},
2778    volume = {69},
# Line 2790 | Line 2780 | Encoding: GBK
2780    number = {6},
2781    month = {Dec},
2782    abstract = {We report a constant pressure and temperature molecular dynamics simulation
2783 <        of a fully hydrated liquid crystal (L(alpha) phase bilayer of dipalmitoylphosphatidylcholine
2784 <        at 50 degrees C and 28 water molecules/lipid. We have shown that
2785 <        the bilayer is stable throughout the 1550-ps simulation and have
2786 <        demonstrated convergence of the system dimensions. Several important
2787 <        aspects of the bilayer structure have been investigated and compared
2788 <        favorably with experimental results. For example, the average positions
2789 <        of specific carbon atoms along the bilayer normal agree well with
2790 <        neutron diffraction data, and the electron density profile is in
2791 <        accord with x-ray diffraction results. The hydrocarbon chain deuterium
2792 <        order parameters agree reasonably well with NMR results for the
2793 <        middles of the chains, but the simulation predicts too much order
2794 <        at the chain ends. In spite of the deviations in the order parameters,
2795 <        the hydrocarbon chain packing density appears to be essentially
2796 <        correct, inasmuch as the area/lipid and bilayer thickness are in
2797 <        agreement with the most refined experimental estimates. The deuterium
2798 <        order parameters for the glycerol and choline groups, as well as
2799 <        the phosphorus chemical shift anisotropy, are in qualitative agreement
2800 <        with those extracted from NMR measurements.},
2783 >    of a fully hydrated liquid crystal (L(alpha) phase bilayer of dipalmitoylphosphatidylcholine
2784 >    at 50 degrees C and 28 water molecules/lipid. We have shown that
2785 >    the bilayer is stable throughout the 1550-ps simulation and have
2786 >    demonstrated convergence of the system dimensions. Several important
2787 >    aspects of the bilayer structure have been investigated and compared
2788 >    favorably with experimental results. For example, the average positions
2789 >    of specific carbon atoms along the bilayer normal agree well with
2790 >    neutron diffraction data, and the electron density profile is in
2791 >    accord with x-ray diffraction results. The hydrocarbon chain deuterium
2792 >    order parameters agree reasonably well with NMR results for the
2793 >    middles of the chains, but the simulation predicts too much order
2794 >    at the chain ends. In spite of the deviations in the order parameters,
2795 >    the hydrocarbon chain packing density appears to be essentially
2796 >    correct, inasmuch as the area/lipid and bilayer thickness are in
2797 >    agreement with the most refined experimental estimates. The deuterium
2798 >    order parameters for the glycerol and choline groups, as well as
2799 >    the phosphorus chemical shift anisotropy, are in qualitative agreement
2800 >    with those extracted from NMR measurements.},
2801    annote = {Tv018 Times Cited:108 Cited References Count:34},
2802    issn = {0006-3495},
2803    uri = {<Go to ISI>://A1995TV01800037},
# Line 2823 | Line 2813 | Encoding: GBK
2813    number = {3},
2814    month = {Aug 1},
2815    abstract = {The Trotter factorization of the Liouville propagator is used to generate
2816 <        new reversible molecular dynamics integrators. This strategy is
2817 <        applied to derive reversible reference system propagator algorithms
2818 <        (RESPA) that greatly accelerate simulations of systems with a separation
2819 <        of time scales or with long range forces. The new algorithms have
2820 <        all of the advantages of previous RESPA integrators but are reversible,
2821 <        and more stable than those methods. These methods are applied to
2822 <        a set of paradigmatic systems and are shown to be superior to earlier
2823 <        methods. It is shown how the new RESPA methods are related to predictor-corrector
2824 <        integrators. Finally, we show how these methods can be used to accelerate
2825 <        the integration of the equations of motion of systems with Nose
2826 <        thermostats.},
2816 >    new reversible molecular dynamics integrators. This strategy is
2817 >    applied to derive reversible reference system propagator algorithms
2818 >    (RESPA) that greatly accelerate simulations of systems with a separation
2819 >    of time scales or with long range forces. The new algorithms have
2820 >    all of the advantages of previous RESPA integrators but are reversible,
2821 >    and more stable than those methods. These methods are applied to
2822 >    a set of paradigmatic systems and are shown to be superior to earlier
2823 >    methods. It is shown how the new RESPA methods are related to predictor-corrector
2824 >    integrators. Finally, we show how these methods can be used to accelerate
2825 >    the integration of the equations of motion of systems with Nose
2826 >    thermostats.},
2827    annote = {Je891 Times Cited:680 Cited References Count:19},
2828    issn = {0021-9606},
2829    uri = {<Go to ISI>://A1992JE89100044},
# Line 2850 | Line 2840 | Encoding: GBK
2840   @ARTICLE{Wegener1979,
2841    author = {W.~A. Wegener, V.~J. Koester and R.~M. Dowben},
2842    title = {A general ellipsoid can not always serve as a modle for the rotational
2843 <        diffusion properties of arbitrary shaped rigid molecules},
2843 >    diffusion properties of arbitrary shaped rigid molecules},
2844    journal = {Proc. Natl. Acad. Sci.},
2845    year = {1979},
2846    volume = {76},
# Line 2861 | Line 2851 | Encoding: GBK
2851   @ARTICLE{Withers2003,
2852    author = {I. M. Withers},
2853    title = {Effects of longitudinal quadrupoles on the phase behavior of a Gay-Berne
2854 <        fluid},
2854 >    fluid},
2855    journal = {Journal of Chemical Physics},
2856    year = {2003},
2857    volume = {119},
# Line 2869 | Line 2859 | Encoding: GBK
2859    number = {19},
2860    month = {Nov 15},
2861    abstract = {The effects of longitudinal quadrupole moments on the formation of
2862 <        liquid crystalline phases are studied by means of constant NPT Monte
2863 <        Carlo simulation methods. The popular Gay-Berne model mesogen is
2864 <        used as the reference fluid, which displays the phase sequences
2865 <        isotropic-smectic A-smectic B and isotropic-smectic B at high (T*=2.0)
2866 <        and low (T*=1.5) temperatures, respectively. With increasing quadrupole
2867 <        magnitude the smectic phases are observed to be stabilized with
2868 <        respect to the isotropic liquid, while the smectic B is destabilized
2869 <        with respect to the smectic A. At the lower temperature, a sufficiently
2870 <        large quadrupole magnitude results in the injection of the smectic
2871 <        A phase into the phase sequence and the replacement of the smectic
2872 <        B phase by the tilted smectic J phase. The nematic phase is also
2873 <        injected into the phase sequence at both temperatures considered,
2874 <        and ultimately for sufficiently large quadrupole magnitudes no coherent
2875 <        layered structures were observed. The stabilization of the smectic
2876 <        A phase supports the commonly held belief that, while the inclusion
2877 <        of polar groups is not a prerequisite for the formation of the smectic
2878 <        A phase, quadrupolar interactions help to increase the temperature
2879 <        and pressure range for which the smectic A phase is observed. The
2880 <        quality of the layered structure is worsened with increasing quadrupole
2881 <        magnitude. This behavior, along with the injection of the nematic
2882 <        phase into the phase sequence, indicate that the general tendency
2883 <        of the quadrupolar interactions is to destabilize the layered structure.
2884 <        A pressure dependence upon the smectic layer spacing is observed.
2885 <        This behavior is in much closer agreement with experimental findings
2886 <        than has been observed previously for nonpolar Gay-Berne and hard
2887 <        spherocylinder models. (C) 2003 American Institute of Physics.},
2862 >    liquid crystalline phases are studied by means of constant NPT Monte
2863 >    Carlo simulation methods. The popular Gay-Berne model mesogen is
2864 >    used as the reference fluid, which displays the phase sequences
2865 >    isotropic-smectic A-smectic B and isotropic-smectic B at high (T*=2.0)
2866 >    and low (T*=1.5) temperatures, respectively. With increasing quadrupole
2867 >    magnitude the smectic phases are observed to be stabilized with
2868 >    respect to the isotropic liquid, while the smectic B is destabilized
2869 >    with respect to the smectic A. At the lower temperature, a sufficiently
2870 >    large quadrupole magnitude results in the injection of the smectic
2871 >    A phase into the phase sequence and the replacement of the smectic
2872 >    B phase by the tilted smectic J phase. The nematic phase is also
2873 >    injected into the phase sequence at both temperatures considered,
2874 >    and ultimately for sufficiently large quadrupole magnitudes no coherent
2875 >    layered structures were observed. The stabilization of the smectic
2876 >    A phase supports the commonly held belief that, while the inclusion
2877 >    of polar groups is not a prerequisite for the formation of the smectic
2878 >    A phase, quadrupolar interactions help to increase the temperature
2879 >    and pressure range for which the smectic A phase is observed. The
2880 >    quality of the layered structure is worsened with increasing quadrupole
2881 >    magnitude. This behavior, along with the injection of the nematic
2882 >    phase into the phase sequence, indicate that the general tendency
2883 >    of the quadrupolar interactions is to destabilize the layered structure.
2884 >    A pressure dependence upon the smectic layer spacing is observed.
2885 >    This behavior is in much closer agreement with experimental findings
2886 >    than has been observed previously for nonpolar Gay-Berne and hard
2887 >    spherocylinder models. (C) 2003 American Institute of Physics.},
2888    annote = {738EF Times Cited:3 Cited References Count:43},
2889    issn = {0021-9606},
2890    uri = {<Go to ISI>://000186273200027},
# Line 2903 | Line 2893 | Encoding: GBK
2893   @ARTICLE{Wolf1999,
2894    author = {D. Wolf and P. Keblinski and S. R. Phillpot and J. Eggebrecht},
2895    title = {Exact method for the simulation of Coulombic systems by spherically
2896 <        truncated, pairwise r(-1) summation},
2896 >    truncated, pairwise r(-1) summation},
2897    journal = {Journal of Chemical Physics},
2898    year = {1999},
2899    volume = {110},
# Line 2911 | Line 2901 | Encoding: GBK
2901    number = {17},
2902    month = {May 1},
2903    abstract = {Based on a recent result showing that the net Coulomb potential in
2904 <        condensed ionic systems is rather short ranged, an exact and physically
2905 <        transparent method permitting the evaluation of the Coulomb potential
2906 <        by direct summation over the r(-1) Coulomb pair potential is presented.
2907 <        The key observation is that the problems encountered in determining
2908 <        the Coulomb energy by pairwise, spherically truncated r(-1) summation
2909 <        are a direct consequence of the fact that the system summed over
2910 <        is practically never neutral. A simple method is developed that
2911 <        achieves charge neutralization wherever the r(-1) pair potential
2912 <        is truncated. This enables the extraction of the Coulomb energy,
2913 <        forces, and stresses from a spherically truncated, usually charged
2914 <        environment in a manner that is independent of the grouping of the
2915 <        pair terms. The close connection of our approach with the Ewald
2916 <        method is demonstrated and exploited, providing an efficient method
2917 <        for the simulation of even highly disordered ionic systems by direct,
2918 <        pairwise r(-1) summation with spherical truncation at rather short
2919 <        range, i.e., a method which fully exploits the short-ranged nature
2920 <        of the interactions in ionic systems. The method is validated by
2921 <        simulations of crystals, liquids, and interfacial systems, such
2922 <        as free surfaces and grain boundaries. (C) 1999 American Institute
2923 <        of Physics. [S0021-9606(99)51517-1].},
2904 >    condensed ionic systems is rather short ranged, an exact and physically
2905 >    transparent method permitting the evaluation of the Coulomb potential
2906 >    by direct summation over the r(-1) Coulomb pair potential is presented.
2907 >    The key observation is that the problems encountered in determining
2908 >    the Coulomb energy by pairwise, spherically truncated r(-1) summation
2909 >    are a direct consequence of the fact that the system summed over
2910 >    is practically never neutral. A simple method is developed that
2911 >    achieves charge neutralization wherever the r(-1) pair potential
2912 >    is truncated. This enables the extraction of the Coulomb energy,
2913 >    forces, and stresses from a spherically truncated, usually charged
2914 >    environment in a manner that is independent of the grouping of the
2915 >    pair terms. The close connection of our approach with the Ewald
2916 >    method is demonstrated and exploited, providing an efficient method
2917 >    for the simulation of even highly disordered ionic systems by direct,
2918 >    pairwise r(-1) summation with spherical truncation at rather short
2919 >    range, i.e., a method which fully exploits the short-ranged nature
2920 >    of the interactions in ionic systems. The method is validated by
2921 >    simulations of crystals, liquids, and interfacial systems, such
2922 >    as free surfaces and grain boundaries. (C) 1999 American Institute
2923 >    of Physics. [S0021-9606(99)51517-1].},
2924    annote = {189PD Times Cited:70 Cited References Count:34},
2925    issn = {0021-9606},
2926    uri = {<Go to ISI>://000079913000008},
# Line 2949 | Line 2939 | Encoding: GBK
2939    issn = {0375-9601},
2940    uri = {<Go to ISI>://A1990EJ79800009},
2941   }
2952

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines