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

Comparing trunk/nivsRnemd/nivsRnemd.bib (file contents):
Revision 3582 by skuang, Thu Apr 8 21:23:55 2010 UTC vs.
Revision 3583 by gezelter, Tue Apr 13 19:59:50 2010 UTC

# Line 2 | Line 2
2   %% http://bibdesk.sourceforge.net/
3  
4  
5 < %% Created for Shenyu Kuang at 2010-04-08 17:18:44 -0400
5 > %% Created for Dan Gezelter at 2010-04-13 09:12:19 -0400
6  
7  
8   %% Saved with string encoding Unicode (UTF-8)
# Line 10 | Line 10
10  
11  
12   @article{ISI:000207079300006,
13 <        Abstract = {{Non-equilibrium Molecular Dynamics Simulation methods have been used to
14 <   study the ability of Embedded Atom Method models of the metals copper
15 <   and gold to reproduce the equilibrium and non-equilibrium behavior of
16 <   metals at a stationary and at a moving solid/liquid interface. The
17 <   equilibrium solid/vapor interface was shown to display a simple
18 <   termination of the bulk until the temperature of the solid reaches
19 <   approximate to 90\% of the bulk melting point. At and above such
20 <   temperatures the systems exhibit a surface disodering known as surface
21 <   melting. Non-equilibrium simulations emulating the action of a
22 <   picosecond laser on the metal were performed to determine the regrowth
23 <   velocity. For copper, the action of a 20 ps laser with an absorbed
24 <   energy of 2-5 mJ/cm(2) produced a regrowth velocity of 83-100 m/s, in
25 <   reasonable agreement with the value obtained by experiment (>60 m/s).
26 <   For gold, similar conditions produced a slower regrowth velocity of 63
27 <   m/s at an absorbed energy of 5 mJ/cm(2). This is almost a factor of two
28 <   too low in comparison to experiment (>100 m/s). The regrowth velocities
29 <   of the metals seems unexpectedly close to experiment considering that
30 <   the free-electron contribution is ignored in the Embeeded Atom Method
31 <   models used.}},
32 <        Address = {{4 PARK SQUARE, MILTON PARK, ABINGDON OX14 4RN, OXON, ENGLAND}},
33 <        Affiliation = {{Clancy, P (Reprint Author), Cornell Univ, Sch Chem Engn, Ithaca, NY 14853 USA. {[}Richardson, Clifton F.; Clancy, Paulette] Cornell Univ, Sch Chem Engn, Ithaca, NY 14853 USA.}},
34 <        Author = {Richardson, Clifton F. and Clancy, Paulette},
35 <        Date-Added = {2010-04-07 11:24:36 -0400},
36 <        Date-Modified = {2010-04-07 11:24:36 -0400},
37 <        Doc-Delivery-Number = {{V04SY}},
38 <        Issn = {{0892-7022}},
39 <        Journal = {{MOLECULAR SIMULATION}},
40 <        Journal-Iso = {{Mol. Simul.}},
41 <        Keywords = {{Non-equilibrium computer simulation; molecular dynamics; crystal growth; Embedded Atom Method models of metals}},
42 <        Language = {{English}},
43 <        Number = {{5-6}},
44 <        Number-Of-Cited-References = {{36}},
45 <        Pages = {{335-355}},
46 <        Publisher = {{TAYLOR \& FRANCIS LTD}},
47 <        Subject-Category = {{Chemistry, Physical; Physics, Atomic, Molecular \& Chemical}},
48 <        Times-Cited = {{7}},
49 <        Title = {{PICOSECOND LASER PROCESSING OF COPPER AND GOLD: A COMPUTER SIMULATION STUDY}},
50 <        Type = {{Article}},
51 <        Unique-Id = {{ISI:000207079300006}},
52 <        Volume = {{7}},
53 <        Year = {{1991}}}
13 >  Abstract =     {Non-equilibrium Molecular Dynamics Simulation
14 >                  methods have been used to study the ability of
15 >                  Embedded Atom Method models of the metals copper and
16 >                  gold to reproduce the equilibrium and
17 >                  non-equilibrium behavior of metals at a stationary
18 >                  and at a moving solid/liquid interface. The
19 >                  equilibrium solid/vapor interface was shown to
20 >                  display a simple termination of the bulk until the
21 >                  temperature of the solid reaches approximate to 90\%
22 >                  of the bulk melting point. At and above such
23 >                  temperatures the systems exhibit a surface
24 >                  disodering known as surface melting. Non-equilibrium
25 >                  simulations emulating the action of a picosecond
26 >                  laser on the metal were performed to determine the
27 >                  regrowth velocity. For copper, the action of a 20 ps
28 >                  laser with an absorbed energy of 2-5 mJ/cm(2)
29 >                  produced a regrowth velocity of 83-100 m/s, in
30 >                  reasonable agreement with the value obtained by
31 >                  experiment (>60 m/s). For gold, similar conditions
32 >                  produced a slower regrowth velocity of 63 m/s at an
33 >                  absorbed energy of 5 mJ/cm(2). This is almost a
34 >                  factor of two too low in comparison to experiment
35 >                  (>100 m/s). The regrowth velocities of the metals
36 >                  seems unexpectedly close to experiment considering
37 >                  that the free-electron contribution is ignored in
38 >                  the Embeeded Atom Method models used.},
39 >  Address =      {4 PARK SQUARE, MILTON PARK, ABINGDON OX14 4RN,
40 >                  OXON, ENGLAND},
41 >  Affiliation =  {Clancy, P (Reprint Author), Cornell Univ, Sch Chem
42 >                  Engn, Ithaca, NY 14853 USA. {[}Richardson, Clifton
43 >                  F.; Clancy, Paulette] Cornell Univ, Sch Chem Engn,
44 >                  Ithaca, NY 14853 USA.},
45 >  Author =       {Richardson, Clifton F. and Clancy, Paulette},
46 >  Date-Added =   {2010-04-07 11:24:36 -0400},
47 >  Date-Modified ={2010-04-07 11:24:36 -0400},
48 >  Doc-Delivery-Number ={V04SY},
49 >  Issn =         {0892-7022},
50 >  Journal =      {MOLECULAR SIMULATION},
51 >  Journal-Iso =  {Mol. Simul.},
52 >  Keywords =     {Non-equilibrium computer simulation; molecular
53 >                  dynamics; crystal growth; Embedded Atom Method
54 >                  models of metals},
55 >  Language =     {English},
56 >  Number =       {5-6},
57 >  Number-Of-Cited-References ={36},
58 >  Pages =        {335-355},
59 >  Publisher =    {TAYLOR \& FRANCIS LTD},
60 >  Subject-Category ={Chemistry, Physical; Physics, Atomic, Molecular
61 >                  \& Chemical},
62 >  Times-Cited =  {7},
63 >  Title =        {PICOSECOND LASER PROCESSING OF COPPER AND GOLD: A
64 >                  COMPUTER SIMULATION STUDY},
65 >  Type =         {Article},
66 >  Unique-Id =    {ISI:000207079300006},
67 >  Volume =       {7},
68 >  Year =         {1991}
69 > }
70  
71   @article{ISI:000167766600035,
72 <        Abstract = {{Molecular dynamics simulations are used to investigate the separation
73 <   of water films adjacent to a hot metal surface. The simulations clearly
74 <   show that the water layers nearest the surface overheat and undergo
75 <   explosive boiling. For thick films, the expansion of the vaporized
76 <   molecules near the surface forces the outer water layers to move away
77 <   from the surface. These results are of interest for mass spectrometry
78 <   of biological molecules, steam cleaning of surfaces, and medical
79 <   procedures.}},
80 <        Address = {{1155 16TH ST, NW, WASHINGTON, DC 20036 USA}},
81 <        Affiliation = {{Garrison, BJ (Reprint Author), Penn State Univ, Dept Chem, University Pk, PA 16802 USA. Penn State Univ, Dept Chem, University Pk, PA 16802 USA. Penn State Univ, Inst Mat Res, University Pk, PA 16802 USA. Univ Virginia, Dept Mat Sci \& Engn, Charlottesville, VA 22903 USA.}},
82 <        Author = {Dou, YS and Zhigilei, LV and Winograd, N and Garrison, BJ},
83 <        Date-Added = {2010-03-11 15:32:14 -0500},
84 <        Date-Modified = {2010-03-11 15:32:14 -0500},
85 <        Doc-Delivery-Number = {{416ED}},
86 <        Issn = {{1089-5639}},
87 <        Journal = {{JOURNAL OF PHYSICAL CHEMISTRY A}},
88 <        Journal-Iso = {{J. Phys. Chem. A}},
89 <        Keywords-Plus = {{MOLECULAR-DYNAMICS SIMULATIONS; ASSISTED LASER-DESORPTION; FROZEN AQUEOUS-SOLUTIONS; COMPUTER-SIMULATION; ORGANIC-SOLIDS; VELOCITY DISTRIBUTIONS; PARTICLE BOMBARDMENT; MASS-SPECTROMETRY; PHASE EXPLOSION; LIQUID WATER}},
90 <        Language = {{English}},
91 <        Month = {{MAR 29}},
92 <        Number = {{12}},
93 <        Number-Of-Cited-References = {{65}},
94 <        Pages = {{2748-2755}},
95 <        Publisher = {{AMER CHEMICAL SOC}},
96 <        Subject-Category = {{Chemistry, Physical; Physics, Atomic, Molecular \& Chemical}},
97 <        Times-Cited = {{66}},
98 <        Title = {{Explosive boiling of water films adjacent to heated surfaces: A microscopic description}},
99 <        Type = {{Article}},
100 <        Unique-Id = {{ISI:000167766600035}},
101 <        Volume = {{105}},
102 <        Year = {{2001}}}
72 >  Abstract =     {Molecular dynamics simulations are used to
73 >                  investigate the separation of water films adjacent
74 >                  to a hot metal surface. The simulations clearly show
75 >                  that the water layers nearest the surface overheat
76 >                  and undergo explosive boiling. For thick films, the
77 >                  expansion of the vaporized molecules near the
78 >                  surface forces the outer water layers to move away
79 >                  from the surface. These results are of interest for
80 >                  mass spectrometry of biological molecules, steam
81 >                  cleaning of surfaces, and medical procedures.},
82 >  Address =      {1155 16TH ST, NW, WASHINGTON, DC 20036 USA},
83 >  Affiliation =  {Garrison, BJ (Reprint Author), Penn State Univ,
84 >                  Dept Chem, University Pk, PA 16802 USA. Penn State
85 >                  Univ, Dept Chem, University Pk, PA 16802 USA. Penn
86 >                  State Univ, Inst Mat Res, University Pk, PA 16802
87 >                  USA. Univ Virginia, Dept Mat Sci \& Engn,
88 >                  Charlottesville, VA 22903 USA.},
89 >  Author =       {Dou, YS and Zhigilei, LV and Winograd, N and
90 >                  Garrison, BJ},
91 >  Date-Added =   {2010-03-11 15:32:14 -0500},
92 >  Date-Modified ={2010-03-11 15:32:14 -0500},
93 >  Doc-Delivery-Number ={416ED},
94 >  Issn =         {1089-5639},
95 >  Journal =      {J. Phys. Chem. A},
96 >  Journal-Iso =  {J. Phys. Chem. A},
97 >  Keywords-Plus ={MOLECULAR-DYNAMICS SIMULATIONS; ASSISTED
98 >                  LASER-DESORPTION; FROZEN AQUEOUS-SOLUTIONS;
99 >                  COMPUTER-SIMULATION; ORGANIC-SOLIDS; VELOCITY
100 >                  DISTRIBUTIONS; PARTICLE BOMBARDMENT;
101 >                  MASS-SPECTROMETRY; PHASE EXPLOSION; LIQUID WATER},
102 >  Language =     {English},
103 >  Month =        {MAR 29},
104 >  Number =       {12},
105 >  Number-Of-Cited-References ={65},
106 >  Pages =        {2748-2755},
107 >  Publisher =    {AMER CHEMICAL SOC},
108 >  Subject-Category ={Chemistry, Physical; Physics, Atomic, Molecular
109 >                  \& Chemical},
110 >  Times-Cited =  {66},
111 >  Title =        {Explosive boiling of water films adjacent to heated
112 >                  surfaces: A microscopic description},
113 >  Type =         {Article},
114 >  Unique-Id =    {ISI:000167766600035},
115 >  Volume =       {105},
116 >  Year =         {2001}
117 > }
118  
119   @article{ISI:000273472300004,
120 <        Abstract = {{The reverse nonequilibrium molecular dynamics (RNEMD) method calculates
121 <   the shear viscosity of a fluid by imposing a nonphysical exchange of
122 <   momentum and measuring the resulting shear velocity gradient. In this
123 <   study we investigate the range of momentum flux values over which RNEMD
124 <   yields usable (linear) velocity gradients. We find that nonlinear
125 <   velocity profiles result primarily from gradients in fluid temperature
126 <   and density. The temperature gradient results from conversion of heat
127 <   into bulk kinetic energy, which is transformed back into heat elsewhere
128 <   via viscous heating. An expression is derived to predict the
129 <   temperature profile resulting from a specified momentum flux for a
130 <   given fluid and simulation cell. Although primarily bounded above, we
131 <   also describe milder low-flux limitations. RNEMD results for a
132 <   Lennard-Jones fluid agree with equilibrium molecular dynamics and
133 <   conventional nonequilibrium molecular dynamics calculations at low
134 <   shear, but RNEMD underpredicts viscosity relative to conventional NEMD
135 <   at high shear.}},
136 <        Address = {{CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501 USA}},
137 <        Affiliation = {{Tenney, CM (Reprint Author), Univ Notre Dame, Dept Chem \& Biomol Engn, 182 Fitzpatrick Hall, Notre Dame, IN 46556 USA. {[}Tenney, Craig M.; Maginn, Edward J.] Univ Notre Dame, Dept Chem \& Biomol Engn, Notre Dame, IN 46556 USA.}},
138 <        Article-Number = {{014103}},
139 <        Author = {Tenney, Craig M. and Maginn, Edward J.},
140 <        Author-Email = {{ed@nd.edu}},
141 <        Date-Added = {2010-03-09 13:08:41 -0500},
142 <        Date-Modified = {2010-03-09 13:08:41 -0500},
143 <        Doc-Delivery-Number = {{542DQ}},
144 <        Doi = {{10.1063/1.3276454}},
145 <        Funding-Acknowledgement = {{U.S. Department of Energy {[}DE-FG36-08G088020]}},
146 <        Funding-Text = {{Support for this work was provided by the U.S. Department of Energy (Grant No. DE-FG36-08G088020)}},
147 <        Issn = {{0021-9606}},
148 <        Journal = {{JOURNAL OF CHEMICAL PHYSICS}},
149 <        Journal-Iso = {{J. Chem. Phys.}},
150 <        Keywords = {{Lennard-Jones potential; molecular dynamics method; Navier-Stokes equations; viscosity}},
151 <        Keywords-Plus = {{CURRENT AUTOCORRELATION-FUNCTION; IONIC LIQUID; SIMULATIONS; TEMPERATURE}},
152 <        Language = {{English}},
153 <        Month = {{JAN 7}},
154 <        Number = {{1}},
155 <        Number-Of-Cited-References = {{20}},
156 <        Publisher = {{AMER INST PHYSICS}},
157 <        Subject-Category = {{Physics, Atomic, Molecular \& Chemical}},
158 <        Times-Cited = {{0}},
159 <        Title = {{Limitations and recommendations for the calculation of shear viscosity using reverse nonequilibrium molecular dynamics}},
160 <        Type = {{Article}},
161 <        Unique-Id = {{ISI:000273472300004}},
162 <        Volume = {{132}},
163 <        Year = {{2010}},
164 <        Bdsk-Url-1 = {http://dx.doi.org/10.1063/1.3276454%7D}}
120 >  Abstract =     {The reverse nonequilibrium molecular dynamics
121 >                  (RNEMD) method calculates the shear viscosity of a
122 >                  fluid by imposing a nonphysical exchange of momentum
123 >                  and measuring the resulting shear velocity
124 >                  gradient. In this study we investigate the range of
125 >                  momentum flux values over which RNEMD yields usable
126 >                  (linear) velocity gradients. We find that nonlinear
127 >                  velocity profiles result primarily from gradients in
128 >                  fluid temperature and density. The temperature
129 >                  gradient results from conversion of heat into bulk
130 >                  kinetic energy, which is transformed back into heat
131 >                  elsewhere via viscous heating. An expression is
132 >                  derived to predict the temperature profile resulting
133 >                  from a specified momentum flux for a given fluid and
134 >                  simulation cell. Although primarily bounded above,
135 >                  we also describe milder low-flux limitations. RNEMD
136 >                  results for a Lennard-Jones fluid agree with
137 >                  equilibrium molecular dynamics and conventional
138 >                  nonequilibrium molecular dynamics calculations at
139 >                  low shear, but RNEMD underpredicts viscosity
140 >                  relative to conventional NEMD at high shear.},
141 >  Address =      {CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON
142 >                  QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501
143 >                  USA},
144 >  Affiliation =  {Tenney, CM (Reprint Author), Univ Notre Dame, Dept
145 >                  Chem \& Biomol Engn, 182 Fitzpatrick Hall, Notre
146 >                  Dame, IN 46556 USA. {[}Tenney, Craig M.; Maginn,
147 >                  Edward J.] Univ Notre Dame, Dept Chem \& Biomol
148 >                  Engn, Notre Dame, IN 46556 USA.},
149 >  Article-Number ={014103},
150 >  Author =       {Tenney, Craig M. and Maginn, Edward J.},
151 >  Author-Email = {ed@nd.edu},
152 >  Date-Added =   {2010-03-09 13:08:41 -0500},
153 >  Date-Modified ={2010-03-09 13:08:41 -0500},
154 >  Doc-Delivery-Number ={542DQ},
155 >  Doi =          {10.1063/1.3276454},
156 >  Funding-Acknowledgement ={U.S. Department of Energy
157 >                  {[}DE-FG36-08G088020]},
158 >  Funding-Text = {Support for this work was provided by the
159 >                  U.S. Department of Energy (Grant
160 >                  No. DE-FG36-08G088020)},
161 >  Issn =         {0021-9606},
162 >  Journal =      {J. Chem. Phys.},
163 >  Journal-Iso =  {J. Chem. Phys.},
164 >  Keywords =     {Lennard-Jones potential; molecular dynamics method;
165 >                  Navier-Stokes equations; viscosity},
166 >  Keywords-Plus ={CURRENT AUTOCORRELATION-FUNCTION; IONIC LIQUID;
167 >                  SIMULATIONS; TEMPERATURE},
168 >  Language =     {English},
169 >  Month =        {JAN 7},
170 >  Number =       {1},
171 >  Number-Of-Cited-References ={20},
172 >  Publisher =    {AMER INST PHYSICS},
173 >  Subject-Category ={Physics, Atomic, Molecular \& Chemical},
174 >  Times-Cited =  {0},
175 >  Title =        {Limitations and recommendations for the calculation
176 >                  of shear viscosity using reverse nonequilibrium
177 >                  molecular dynamics},
178 >  Type =         {Article},
179 >  Unique-Id =    {ISI:000273472300004},
180 >  Volume =       {132},
181 >  Year =         {2010},
182 >  Bdsk-Url-1 =   {http://dx.doi.org/10.1063/1.3276454}
183 > }
184  
185   @article{Clancy:1992,
186 <        Abstract = {{The regrowth velocity of a crystal from a melt depends on contributions
187 <   from the thermal conductivity, heat gradient, and latent heat.  The
188 <   relative contributions of these terms to the regrowth velocity of the
189 <   pure metals copper and gold during liquid-phase epitaxy are evaluated.
190 <   These results are used to explain how results from previous
191 <   nonequilibrium molecular-dynamics simulations using classical
192 <   potentials are able to predict regrowth velocities that are close to
193 <   the experimental values.  Results from equilibrium molecular dynamics
194 <   showing the nature of the solid-vapor interface of an
195 <   embedded-atom-method-modeled Cu57Ni43 alloy at a temperature
196 <   corresponding to 62\% of the melting point are presented.  The regrowth
197 <   of this alloy following a simulation of a laser-processing experiment
198 <   is also given, with use of nonequilibrium molecular-dynamics
199 <   techniques.  The thermal conductivity and temperature gradient in the
200 <   simulation of the alloy are compared to those for the pure metals.}},
201 <        Address = {{ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844 USA}},
202 <        Affiliation = {{CORNELL UNIV,SCH CHEM ENGN,ITHACA,NY 14853.}},
203 <        Author = {RICHARDSON, CF and CLANCY, P},
204 <        Date-Added = {2010-01-12 16:17:33 -0500},
205 <        Date-Modified = {2010-04-08 17:18:25 -0400},
206 <        Doc-Delivery-Number = {{HX378}},
207 <        Issn = {{0163-1829}},
208 <        Journal = {{PHYSICAL REVIEW B}},
209 <        Journal-Iso = {{Phys. Rev. B}},
210 <        Keywords-Plus = {{SURFACE SEGREGATION; MOLECULAR-DYNAMICS; TRANSITION-METALS; SOLIDIFICATION; GROWTH; CU; NI}},
211 <        Language = {{English}},
212 <        Month = {{JUN 1}},
213 <        Number = {{21}},
214 <        Number-Of-Cited-References = {{24}},
215 <        Pages = {{12260-12268}},
216 <        Publisher = {{AMERICAN PHYSICAL SOC}},
217 <        Subject-Category = {{Physics, Condensed Matter}},
218 <        Times-Cited = {{11}},
219 <        Title = {{CONTRIBUTION OF THERMAL-CONDUCTIVITY TO THE CRYSTAL-REGROWTH VELOCITY OF EMBEDDED-ATOM-METHOD-MODELED METALS AND METAL-ALLOYS}},
220 <        Type = {{Article}},
221 <        Unique-Id = {{ISI:A1992HX37800010}},
222 <        Volume = {{45}},
223 <        Year = {{1992}}}
186 >  Abstract =     {The regrowth velocity of a crystal from a melt
187 >                  depends on contributions from the thermal
188 >                  conductivity, heat gradient, and latent heat. The
189 >                  relative contributions of these terms to the
190 >                  regrowth velocity of the pure metals copper and gold
191 >                  during liquid-phase epitaxy are evaluated. These
192 >                  results are used to explain how results from
193 >                  previous nonequilibrium molecular-dynamics
194 >                  simulations using classical potentials are able to
195 >                  predict regrowth velocities that are close to the
196 >                  experimental values. Results from equilibrium
197 >                  molecular dynamics showing the nature of the
198 >                  solid-vapor interface of an
199 >                  embedded-atom-method-modeled Cu57Ni43 alloy at a
200 >                  temperature corresponding to 62\% of the melting
201 >                  point are presented. The regrowth of this alloy
202 >                  following a simulation of a laser-processing
203 >                  experiment is also given, with use of nonequilibrium
204 >                  molecular-dynamics techniques. The thermal
205 >                  conductivity and temperature gradient in the
206 >                  simulation of the alloy are compared to those for
207 >                  the pure metals.},
208 >  Address =      {ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844
209 >                  USA},
210 >  Affiliation =  {CORNELL UNIV,SCH CHEM ENGN,ITHACA,NY 14853.},
211 >  Author =       {Richardson, C.~F. and Clancy, P},
212 >  Date-Added =   {2010-01-12 16:17:33 -0500},
213 >  Date-Modified ={2010-04-08 17:18:25 -0400},
214 >  Doc-Delivery-Number ={HX378},
215 >  Issn =         {0163-1829},
216 >  Journal =      {Phys. Rev. B},
217 >  Journal-Iso =  {Phys. Rev. B},
218 >  Keywords-Plus ={SURFACE SEGREGATION; MOLECULAR-DYNAMICS;
219 >                  TRANSITION-METALS; SOLIDIFICATION; GROWTH; CU; NI},
220 >  Language =     {English},
221 >  Month =        {JUN 1},
222 >  Number =       {21},
223 >  Number-Of-Cited-References ={24},
224 >  Pages =        {12260-12268},
225 >  Publisher =    {AMERICAN PHYSICAL SOC},
226 >  Subject-Category ={Physics, Condensed Matter},
227 >  Times-Cited =  {11},
228 >  Title =        {CONTRIBUTION OF THERMAL-CONDUCTIVITY TO THE
229 >                  CRYSTAL-REGROWTH VELOCITY OF
230 >                  EMBEDDED-ATOM-METHOD-MODELED METALS AND
231 >                  METAL-ALLOYS},
232 >  Type =         {Article},
233 >  Unique-Id =    {ISI:A1992HX37800010},
234 >  Volume =       {45},
235 >  Year =         {1992}
236 > }
237  
238   @article{ISI:000090151400044,
239 <        Abstract = {{We have applied a new nonequilibrium molecular dynamics (NEMD) method
240 <   {[}F. Muller-Plathe, J. Chem. Phys. 106, 6082 (1997)] previously
241 <   applied to monatomic Lennard-Jones fluids in the determination of the
242 <   thermal conductivity of molecular fluids. The method was modified in
243 <   order to be applicable to systems with holonomic constraints. Because
244 <   the method involves imposing a known heat flux it is particularly
245 <   attractive for systems involving long-range and many-body interactions
246 <   where calculation of the microscopic heat flux is difficult. The
247 <   predicted thermal conductivities of liquid n-butane and water using the
248 <   imposed-flux NEMD method were found to be in a good agreement with
249 <   previous simulations and experiment. (C) 2000 American Institute of
250 <   Physics. {[}S0021-9606(00)50841-1].}},
251 <        Address = {{2 HUNTINGTON QUADRANGLE, STE 1NO1, MELVILLE, NY 11747-4501 USA}},
252 <        Affiliation = {{Bedrov, D (Reprint Author), Univ Utah, Dept Chem \& Fuels Engn, 122 S Cent Campus Dr,Rm 304, Salt Lake City, UT 84112 USA. Univ Utah, Dept Chem \& Fuels Engn, Salt Lake City, UT 84112 USA. Univ Utah, Dept Mat Sci \& Engn, Salt Lake City, UT 84112 USA.}},
253 <        Author = {Bedrov, D and Smith, GD},
254 <        Date-Added = {2009-11-05 18:21:18 -0500},
255 <        Date-Modified = {2009-11-05 18:21:18 -0500},
256 <        Doc-Delivery-Number = {{369BF}},
257 <        Issn = {{0021-9606}},
258 <        Journal = {{JOURNAL OF CHEMICAL PHYSICS}},
259 <        Journal-Iso = {{J. Chem. Phys.}},
260 <        Keywords-Plus = {{EFFECTIVE PAIR POTENTIALS; TRANSPORT-PROPERTIES; CANONICAL ENSEMBLE; NORMAL-BUTANE; ALGORITHMS; SHAKE; WATER}},
261 <        Language = {{English}},
262 <        Month = {{NOV 8}},
263 <        Number = {{18}},
264 <        Number-Of-Cited-References = {{26}},
265 <        Pages = {{8080-8084}},
266 <        Publisher = {{AMER INST PHYSICS}},
267 <        Subject-Category = {{Physics, Atomic, Molecular \& Chemical}},
268 <        Times-Cited = {{23}},
269 <        Title = {{Thermal conductivity of molecular fluids from molecular dynamics simulations: Application of a new imposed-flux method}},
270 <        Type = {{Article}},
271 <        Unique-Id = {{ISI:000090151400044}},
272 <        Volume = {{113}},
273 <        Year = {{2000}}}
239 >  Abstract =     {We have applied a new nonequilibrium molecular
240 >                  dynamics (NEMD) method {[}F. Muller-Plathe,
241 >                  J. Chem. Phys. 106, 6082 (1997)] previously applied
242 >                  to monatomic Lennard-Jones fluids in the
243 >                  determination of the thermal conductivity of
244 >                  molecular fluids. The method was modified in order
245 >                  to be applicable to systems with holonomic
246 >                  constraints. Because the method involves imposing a
247 >                  known heat flux it is particularly attractive for
248 >                  systems involving long-range and many-body
249 >                  interactions where calculation of the microscopic
250 >                  heat flux is difficult. The predicted thermal
251 >                  conductivities of liquid n-butane and water using
252 >                  the imposed-flux NEMD method were found to be in a
253 >                  good agreement with previous simulations and
254 >                  experiment. (C) 2000 American Institute of
255 >                  Physics. {[}S0021-9606(00)50841-1].},
256 >  Address =      {2 HUNTINGTON QUADRANGLE, STE 1NO1, MELVILLE, NY
257 >                  11747-4501 USA},
258 >  Affiliation =  {Bedrov, D (Reprint Author), Univ Utah, Dept Chem \&
259 >                  Fuels Engn, 122 S Cent Campus Dr,Rm 304, Salt Lake
260 >                  City, UT 84112 USA. Univ Utah, Dept Chem \& Fuels
261 >                  Engn, Salt Lake City, UT 84112 USA. Univ Utah, Dept
262 >                  Mat Sci \& Engn, Salt Lake City, UT 84112 USA.},
263 >  Author =       {Bedrov, D and Smith, GD},
264 >  Date-Added =   {2009-11-05 18:21:18 -0500},
265 >  Date-Modified ={2009-11-05 18:21:18 -0500},
266 >  Doc-Delivery-Number ={369BF},
267 >  Issn =         {0021-9606},
268 >  Journal =      {J. Chem. Phys.},
269 >  Journal-Iso =  {J. Chem. Phys.},
270 >  Keywords-Plus ={EFFECTIVE PAIR POTENTIALS; TRANSPORT-PROPERTIES;
271 >                  CANONICAL ENSEMBLE; NORMAL-BUTANE; ALGORITHMS;
272 >                  SHAKE; WATER},
273 >  Language =     {English},
274 >  Month =        {NOV 8},
275 >  Number =       {18},
276 >  Number-Of-Cited-References ={26},
277 >  Pages =        {8080-8084},
278 >  Publisher =    {AMER INST PHYSICS},
279 >  Subject-Category ={Physics, Atomic, Molecular \& Chemical},
280 >  Times-Cited =  {23},
281 >  Title =        {Thermal conductivity of molecular fluids from
282 >                  molecular dynamics simulations: Application of a new
283 >                  imposed-flux method},
284 >  Type =         {Article},
285 >  Unique-Id =    {ISI:000090151400044},
286 >  Volume =       {113},
287 >  Year =         {2000}
288 > }
289  
290   @article{ISI:000231042800044,
291 <        Abstract = {{The reverse nonequilibrium molecular dynamics method for thermal
292 <   conductivities is adapted to the investigation of molecular fluids. The
293 <   method generates a heat flux through the system by suitably exchanging
294 <   velocities of particles located in different regions. From the
295 <   resulting temperature gradient, the thermal conductivity is then
296 <   calculated. Different variants of the algorithm and their combinations
297 <   with other system parameters are tested: exchange of atomic velocities
298 <   versus exchange of molecular center-of-mass velocities, different
299 <   exchange frequencies, molecular models with bond constraints versus
300 <   models with flexible bonds, united-atom versus all-atom models, and
301 <   presence versus absence of a thermostat. To help establish the range of
302 <   applicability, the algorithm is tested on different models of benzene,
303 <   cyclohexane, water, and n-hexane. We find that the algorithm is robust
304 <   and that the calculated thermal conductivities are insensitive to
305 <   variations in its control parameters. The force field, in contrast, has
306 <   a major influence on the value of the thermal conductivity. While
307 <   calculated and experimental thermal conductivities fall into the same
308 <   order of magnitude, in most cases the calculated values are
309 <   systematically larger. United-atom force fields seem to do better than
310 <   all-atom force fields, possibly because they remove high-frequency
311 <   degrees of freedom from the simulation, which, in nature, are
312 <   quantum-mechanical oscillators in their ground state and do not
313 <   contribute to heat conduction.}},
314 <        Address = {{1155 16TH ST, NW, WASHINGTON, DC 20036 USA}},
315 <        Affiliation = {{Zhang, MM (Reprint Author), Int Univ Bremen, POB 750 561, D-28725 Bremen, Germany. Int Univ Bremen, D-28725 Bremen, Germany. Banco Cent Brasil, Desup, Diesp, BR-01310922 Sao Paulo, Brazil.}},
316 <        Author = {Zhang, MM and Lussetti, E and de Souza, LES and Muller-Plathe, F},
317 <        Date-Added = {2009-11-05 18:17:33 -0500},
318 <        Date-Modified = {2009-11-05 18:17:33 -0500},
319 <        Doc-Delivery-Number = {{952YQ}},
320 <        Doi = {{10.1021/jp0512255}},
321 <        Issn = {{1520-6106}},
322 <        Journal = {{JOURNAL OF PHYSICAL CHEMISTRY B}},
323 <        Journal-Iso = {{J. Phys. Chem. B}},
324 <        Keywords-Plus = {{LENNARD-JONES LIQUIDS; TRANSPORT-COEFFICIENTS; SWOLLEN POLYMERS; SHEAR VISCOSITY; MODEL SYSTEMS; SIMULATION; BENZENE; FLUIDS; POTENTIALS; DIFFUSION}},
325 <        Language = {{English}},
326 <        Month = {{AUG 11}},
327 <        Number = {{31}},
328 <        Number-Of-Cited-References = {{42}},
329 <        Pages = {{15060-15067}},
330 <        Publisher = {{AMER CHEMICAL SOC}},
331 <        Subject-Category = {{Chemistry, Physical}},
332 <        Times-Cited = {{17}},
333 <        Title = {{Thermal conductivities of molecular liquids by reverse nonequilibrium molecular dynamics}},
334 <        Type = {{Article}},
335 <        Unique-Id = {{ISI:000231042800044}},
336 <        Volume = {{109}},
337 <        Year = {{2005}},
338 <        Bdsk-Url-1 = {http://dx.doi.org/10.1021/jp0512255%7D}}
291 >  Abstract =     {The reverse nonequilibrium molecular dynamics
292 >                  method for thermal conductivities is adapted to the
293 >                  investigation of molecular fluids. The method
294 >                  generates a heat flux through the system by suitably
295 >                  exchanging velocities of particles located in
296 >                  different regions. From the resulting temperature
297 >                  gradient, the thermal conductivity is then
298 >                  calculated. Different variants of the algorithm and
299 >                  their combinations with other system parameters are
300 >                  tested: exchange of atomic velocities versus
301 >                  exchange of molecular center-of-mass velocities,
302 >                  different exchange frequencies, molecular models
303 >                  with bond constraints versus models with flexible
304 >                  bonds, united-atom versus all-atom models, and
305 >                  presence versus absence of a thermostat. To help
306 >                  establish the range of applicability, the algorithm
307 >                  is tested on different models of benzene,
308 >                  cyclohexane, water, and n-hexane. We find that the
309 >                  algorithm is robust and that the calculated thermal
310 >                  conductivities are insensitive to variations in its
311 >                  control parameters. The force field, in contrast,
312 >                  has a major influence on the value of the thermal
313 >                  conductivity. While calculated and experimental
314 >                  thermal conductivities fall into the same order of
315 >                  magnitude, in most cases the calculated values are
316 >                  systematically larger. United-atom force fields seem
317 >                  to do better than all-atom force fields, possibly
318 >                  because they remove high-frequency degrees of
319 >                  freedom from the simulation, which, in nature, are
320 >                  quantum-mechanical oscillators in their ground state
321 >                  and do not contribute to heat conduction.},
322 >  Address =      {1155 16TH ST, NW, WASHINGTON, DC 20036 USA},
323 >  Affiliation =  {Zhang, MM (Reprint Author), Int Univ Bremen, POB
324 >                  750 561, D-28725 Bremen, Germany. Int Univ Bremen,
325 >                  D-28725 Bremen, Germany. Banco Cent Brasil, Desup,
326 >                  Diesp, BR-01310922 Sao Paulo, Brazil.},
327 >  Author =       {Zhang, MM and Lussetti, E and de Souza, LES and
328 >                  M\"{u}ller-Plathe, F},
329 >  Date-Added =   {2009-11-05 18:17:33 -0500},
330 >  Date-Modified ={2009-11-05 18:17:33 -0500},
331 >  Doc-Delivery-Number ={952YQ},
332 >  Doi =          {10.1021/jp0512255},
333 >  Issn =         {1520-6106},
334 >  Journal =      {J. Phys. Chem. B},
335 >  Journal-Iso =  {J. Phys. Chem. B},
336 >  Keywords-Plus ={LENNARD-JONES LIQUIDS; TRANSPORT-COEFFICIENTS;
337 >                  SWOLLEN POLYMERS; SHEAR VISCOSITY; MODEL SYSTEMS;
338 >                  SIMULATION; BENZENE; FLUIDS; POTENTIALS; DIFFUSION},
339 >  Language =     {English},
340 >  Month =        {AUG 11},
341 >  Number =       {31},
342 >  Number-Of-Cited-References ={42},
343 >  Pages =        {15060-15067},
344 >  Publisher =    {AMER CHEMICAL SOC},
345 >  Subject-Category ={Chemistry, Physical},
346 >  Times-Cited =  {17},
347 >  Title =        {Thermal conductivities of molecular liquids by
348 >                  reverse nonequilibrium molecular dynamics},
349 >  Type =         {Article},
350 >  Unique-Id =    {ISI:000231042800044},
351 >  Volume =       {109},
352 >  Year =         {2005},
353 >  Bdsk-Url-1 =   {http://dx.doi.org/10.1021/jp0512255%7D}
354 > }
355  
356   @article{ISI:A1997YC32200056,
357 <        Abstract = {{Equilibrium molecular dynamics simulations have been carried out in the
358 <   microcanonical ensemble at 300 and 255 K on the extended simple point
359 <   charge (SPC/E) model of water {[}Berendsen et al., J. Phys. Chem. 91,
360 <   6269 (1987)]. In addition to a number of static and dynamic properties,
361 <   thermal conductivity lambda has been calculated via Green-Kubo
362 <   integration of the heat current time correlation functions (CF's) in
363 <   the atomic and molecular formalism, at wave number k=0. The calculated
364 <   values (0.67 +/- 0.04 W/mK at 300 K and 0.52 +/- 0.03 W/mK at 255 K)
365 <   are in good agreement with the experimental data (0.61 W/mK at 300 K
366 <   and 0.49 W/mK at 255 K). A negative long-time tail of the heat current
367 <   CF, more apparent at 255 K, is responsible for the anomalous decrease
368 <   of lambda with temperature. An analysis of the dynamical modes
369 <   contributing to lambda has shown that its value is due to two
370 <   low-frequency exponential-like modes, a faster collisional mode, with
371 <   positive contribution, and a slower one, which determines the negative
372 <   long-time tail. A comparison of the molecular and atomic spectra of the
373 <   heat current CF has suggested that higher-frequency modes should not
374 <   contribute to lambda in this temperature range. Generalized thermal
375 <   diffusivity D-T(k) decreases as a function of k, after an initial minor
376 <   increase at k = k(min). The k dependence of the generalized
377 <   thermodynamic properties has been calculated in the atomic and
378 <   molecular formalisms. The observed differences have been traced back to
379 <   intramolecular or intermolecular rotational effects and related to the
380 <   partial structure functions. Finally, from the results we calculated it
381 <   appears that the SPC/E model gives results in better agreement with
382 <   experimental data than the transferable intermolecular potential with
383 <   four points TIP4P water model {[}Jorgensen et al., J. Chem. Phys. 79,
384 <   926 (1983)], with a larger improvement for, e.g., diffusion,
385 <   viscosities, and dielectric properties and a smaller one for thermal
386 <   conductivity. The SPC/E model shares, to a smaller extent, the
387 <   insufficient slowing down of dynamics at low temperature already found
388 <   for the TIP4P water model.}},
389 <        Address = {{ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844 USA}},
390 <        Affiliation = {{UNIV PISA,DIPARTIMENTO CHIM \& CHIM IND,I-56126 PISA,ITALY. CNR,IST FIS ATOM \& MOL,I-56127 PISA,ITALY.}},
391 <        Author = {Bertolini, D and Tani, A},
392 <        Date-Added = {2009-10-30 15:41:21 -0400},
393 <        Date-Modified = {2009-10-30 15:41:21 -0400},
394 <        Doc-Delivery-Number = {{YC322}},
395 <        Issn = {{1063-651X}},
396 <        Journal = {{PHYSICAL REVIEW E}},
397 <        Journal-Iso = {{Phys. Rev. E}},
398 <        Keywords-Plus = {{TIME-CORRELATION-FUNCTIONS; LENNARD-JONES LIQUID; TRANSPORT-PROPERTIES; SUPERCOOLED WATER; DENSITY; SIMULATIONS; RELAXATION; VELOCITY; ELECTRON; FLUIDS}},
399 <        Language = {{English}},
400 <        Month = {{OCT}},
401 <        Number = {{4}},
402 <        Number-Of-Cited-References = {{35}},
403 <        Pages = {{4135-4151}},
404 <        Publisher = {{AMERICAN PHYSICAL SOC}},
405 <        Subject-Category = {{Physics, Fluids \& Plasmas; Physics, Mathematical}},
406 <        Times-Cited = {{18}},
407 <        Title = {{Thermal conductivity of water: Molecular dynamics and generalized hydrodynamics results}},
408 <        Type = {{Article}},
409 <        Unique-Id = {{ISI:A1997YC32200056}},
410 <        Volume = {{56}},
411 <        Year = {{1997}}}
357 >  Abstract =     {Equilibrium molecular dynamics simulations have
358 >                  been carried out in the microcanonical ensemble at
359 >                  300 and 255 K on the extended simple point charge
360 >                  (SPC/E) model of water {[}Berendsen et al.,
361 >                  J. Phys. Chem. 91, 6269 (1987)]. In addition to a
362 >                  number of static and dynamic properties, thermal
363 >                  conductivity lambda has been calculated via
364 >                  Green-Kubo integration of the heat current time
365 >                  correlation functions (CF's) in the atomic and
366 >                  molecular formalism, at wave number k=0. The
367 >                  calculated values (0.67 +/- 0.04 W/mK at 300 K and
368 >                  0.52 +/- 0.03 W/mK at 255 K) are in good agreement
369 >                  with the experimental data (0.61 W/mK at 300 K and
370 >                  0.49 W/mK at 255 K). A negative long-time tail of
371 >                  the heat current CF, more apparent at 255 K, is
372 >                  responsible for the anomalous decrease of lambda
373 >                  with temperature. An analysis of the dynamical modes
374 >                  contributing to lambda has shown that its value is
375 >                  due to two low-frequency exponential-like modes, a
376 >                  faster collisional mode, with positive contribution,
377 >                  and a slower one, which determines the negative
378 >                  long-time tail. A comparison of the molecular and
379 >                  atomic spectra of the heat current CF has suggested
380 >                  that higher-frequency modes should not contribute to
381 >                  lambda in this temperature range. Generalized
382 >                  thermal diffusivity D-T(k) decreases as a function
383 >                  of k, after an initial minor increase at k =
384 >                  k(min). The k dependence of the generalized
385 >                  thermodynamic properties has been calculated in the
386 >                  atomic and molecular formalisms. The observed
387 >                  differences have been traced back to intramolecular
388 >                  or intermolecular rotational effects and related to
389 >                  the partial structure functions. Finally, from the
390 >                  results we calculated it appears that the SPC/E
391 >                  model gives results in better agreement with
392 >                  experimental data than the transferable
393 >                  intermolecular potential with four points TIP4P
394 >                  water model {[}Jorgensen et al., J. Chem. Phys. 79,
395 >                  926 (1983)], with a larger improvement for, e.g.,
396 >                  diffusion, viscosities, and dielectric properties
397 >                  and a smaller one for thermal conductivity. The
398 >                  SPC/E model shares, to a smaller extent, the
399 >                  insufficient slowing down of dynamics at low
400 >                  temperature already found for the TIP4P water
401 >                  model.},
402 >  Address =      {ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844
403 >                  USA},
404 >  Affiliation =  {UNIV PISA,DIPARTIMENTO CHIM \& CHIM IND,I-56126
405 >                  PISA,ITALY. CNR,IST FIS ATOM \& MOL,I-56127
406 >                  PISA,ITALY.},
407 >  Author =       {Bertolini, D and Tani, A},
408 >  Date-Added =   {2009-10-30 15:41:21 -0400},
409 >  Date-Modified ={2009-10-30 15:41:21 -0400},
410 >  Doc-Delivery-Number ={YC322},
411 >  Issn =         {1063-651X},
412 >  Journal =      {Phys. Rev. E},
413 >  Journal-Iso =  {Phys. Rev. E},
414 >  Keywords-Plus ={TIME-CORRELATION-FUNCTIONS; LENNARD-JONES LIQUID;
415 >                  TRANSPORT-PROPERTIES; SUPERCOOLED WATER; DENSITY;
416 >                  SIMULATIONS; RELAXATION; VELOCITY; ELECTRON;
417 >                  FLUIDS},
418 >  Language =     {English},
419 >  Month =        {OCT},
420 >  Number =       {4},
421 >  Number-Of-Cited-References ={35},
422 >  Pages =        {4135-4151},
423 >  Publisher =    {AMERICAN PHYSICAL SOC},
424 >  Subject-Category ={Physics, Fluids \& Plasmas; Physics,
425 >                  Mathematical},
426 >  Times-Cited =  {18},
427 >  Title =        {Thermal conductivity of water: Molecular dynamics
428 >                  and generalized hydrodynamics results},
429 >  Type =         {Article},
430 >  Unique-Id =    {ISI:A1997YC32200056},
431 >  Volume =       {56},
432 >  Year =         {1997}
433 > }
434  
435   @article{Meineke:2005gd,
436 <        Abstract = {OOPSE is a new molecular dynamics simulation program that is capable of efficiently integrating equations of motion for atom types with orientational degrees of freedom (e.g. "sticky" atoms and point dipoles). Transition metals can also be simulated using the embedded atom method (EAM) potential included in the code. Parallel simulations are carried out using the force-based decomposition method. Simulations are specified using a very simple C-based meta-data language. A number of advanced integrators are included, and the basic integrator for orientational dynamics provides substantial improvements over older quaternion-based schemes. (C) 2004 Wiley Periodicals, Inc.},
437 <        Address = {111 RIVER ST, HOBOKEN, NJ 07030 USA},
438 <        Author = {Meineke, MA and Vardeman, CF and Lin, T and Fennell, CJ and Gezelter, JD},
439 <        Date-Added = {2009-10-01 18:43:03 -0400},
440 <        Date-Modified = {2009-10-01 18:43:03 -0400},
441 <        Doi = {DOI 10.1002/jcc.20161},
442 <        Isi = {000226558200006},
443 <        Isi-Recid = {142688207},
444 <        Isi-Ref-Recids = {67885400 50663994 64190493 93668415 46699855 89992422 57614458 49016001 61447131 111114169 68770425 52728075 102422498 66381878 32391149 134477335 53221357 9929643 59492217 69681001 99223832 142688208 94600872 91658572 54857943 117365867 69323123 49588888 109970172 101670714 142688209 121603296 94652379 96449138 99938010 112825758 114905670 86802042 121339042 104794914 82674909 72096791 93668384 90513335 142688210 23060767 63731466 109033408 76303716 31384453 97861662 71842426 130707771 125809946 66381889 99676497},
445 <        Journal = {Journal of Computational Chemistry},
446 <        Keywords = {OOPSE; molecular dynamics},
447 <        Month = feb,
448 <        Number = {3},
449 <        Pages = {252-271},
450 <        Publisher = {JOHN WILEY \& SONS INC},
451 <        Times-Cited = {9},
452 <        Title = {OOPSE: An object-oriented parallel simulation engine for molecular dynamics},
453 <        Volume = {26},
454 <        Year = {2005},
455 <        Bdsk-Url-1 = {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000226558200006}}
436 >  Abstract =     {OOPSE is a new molecular dynamics simulation program
437 >                  that is capable of efficiently integrating equations
438 >                  of motion for atom types with orientational degrees
439 >                  of freedom (e.g. #sticky# atoms and point
440 >                  dipoles). Transition metals can also be simulated
441 >                  using the embedded atom method (EAM) potential
442 >                  included in the code. Parallel simulations are
443 >                  carried out using the force-based decomposition
444 >                  method. Simulations are specified using a very
445 >                  simple C-based meta-data language. A number of
446 >                  advanced integrators are included, and the basic
447 >                  integrator for orientational dynamics provides
448 >                  substantial improvements over older quaternion-based
449 >                  schemes.},
450 >  Address =      {111 RIVER ST, HOBOKEN, NJ 07030 USA},
451 >  Author =       {Meineke, M. A. and Vardeman, C. F. and Lin, T and Fennell,
452 >                  CJ and Gezelter, J. D.},
453 >  Date-Added =   {2009-10-01 18:43:03 -0400},
454 >  Date-Modified ={2010-04-13 09:11:16 -0400},
455 >  Doi =          {DOI 10.1002/jcc.20161},
456 >  Isi =          {000226558200006},
457 >  Isi-Recid =    {142688207},
458 >  Isi-Ref-Recids ={67885400 50663994 64190493 93668415 46699855
459 >                  89992422 57614458 49016001 61447131 111114169
460 >                  68770425 52728075 102422498 66381878 32391149
461 >                  134477335 53221357 9929643 59492217 69681001
462 >                  99223832 142688208 94600872 91658572 54857943
463 >                  117365867 69323123 49588888 109970172 101670714
464 >                  142688209 121603296 94652379 96449138 99938010
465 >                  112825758 114905670 86802042 121339042 104794914
466 >                  82674909 72096791 93668384 90513335 142688210
467 >                  23060767 63731466 109033408 76303716 31384453
468 >                  97861662 71842426 130707771 125809946 66381889
469 >                  99676497},
470 >  Journal =      {J. Comp. Chem.},
471 >  Keywords =     {OOPSE; molecular dynamics},
472 >  Month =        feb,
473 >  Number =       {3},
474 >  Pages =        {252-271},
475 >  Publisher =    {JOHN WILEY \& SONS INC},
476 >  Times-Cited =  {9},
477 >  Title =        {OOPSE: An object-oriented parallel simulation engine
478 >                  for molecular dynamics},
479 >  Volume =       {26},
480 >  Year =         {2005},
481 >  Bdsk-Url-1 =
482 >                  {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000226558200006},
483 >  Bdsk-Url-2 =   {http://dx.doi.org/10.1002/jcc.20161}
484 > }
485  
486   @article{ISI:000080382700030,
487 <        Abstract = {A nonequilibrium method for calculating the shear viscosity is
488 <   presented. It reverses the cause-and-effect picture customarily used in
489 <   nonequilibrium molecular dynamics: the effect, the momentum flux or
490 <   stress, is imposed, whereas the cause, the velocity gradient or shear
491 <   rate, is obtained from the simulation. It differs from other
492 <   Norton-ensemble methods by the way in which the steady-state momentum
493 <   flux is maintained. This method involves a simple exchange of particle
494 <   momenta, which is easy to implement. Moreover, it can be made to
495 <   conserve the total energy as well as the total linear momentum, so no
496 <   coupling to an external temperature bath is needed. The resulting raw
497 <   data, the velocity profile, is a robust and rapidly converging
498 <   property. The method is tested on the Lennard-Jones fluid near its
499 <   triple point. It yields a viscosity of 3.2-3.3, in Lennard-Jones
500 <   reduced units, in agreement with literature results.
501 <   {[}S1063-651X(99)03105-0].},
502 <        Address = {ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844 USA},
503 <        Affiliation = {Muller-Plathe, F (Reprint Author), Max Planck Inst Polymerforsch, Ackermannweg 10, D-55128 Mainz, Germany. Max Planck Inst Polymerforsch, D-55128 Mainz, Germany.},
504 <        Author = {Muller-Plathe, F},
505 <        Date-Added = {2009-10-01 14:07:30 -0400},
506 <        Date-Modified = {2009-10-01 14:07:30 -0400},
507 <        Doc-Delivery-Number = {197TX},
508 <        Issn = {1063-651X},
509 <        Journal = {PHYSICAL REVIEW E},
510 <        Journal-Iso = {Phys. Rev. E},
511 <        Language = {English},
512 <        Month = {MAY},
513 <        Number = {5, Part A},
514 <        Number-Of-Cited-References = {17},
515 <        Pages = {4894-4898},
516 <        Publisher = {AMERICAN PHYSICAL SOC},
517 <        Subject-Category = {Physics, Fluids \& Plasmas; Physics, Mathematical},
518 <        Times-Cited = {57},
519 <        Title = {Reversing the perturbation in nonequilibrium molecular dynamics: An easy way to calculate the shear viscosity of fluids},
520 <        Type = {Article},
521 <        Unique-Id = {ISI:000080382700030},
522 <        Volume = {59},
523 <        Year = {1999}}
487 >  Abstract =     {A nonequilibrium method for calculating the shear
488 >                  viscosity is presented. It reverses the
489 >                  cause-and-effect picture customarily used in
490 >                  nonequilibrium molecular dynamics: the effect, the
491 >                  momentum flux or stress, is imposed, whereas the
492 >                  cause, the velocity gradient or shear rate, is
493 >                  obtained from the simulation. It differs from other
494 >                  Norton-ensemble methods by the way in which the
495 >                  steady-state momentum flux is maintained. This
496 >                  method involves a simple exchange of particle
497 >                  momenta, which is easy to implement. Moreover, it
498 >                  can be made to conserve the total energy as well as
499 >                  the total linear momentum, so no coupling to an
500 >                  external temperature bath is needed. The resulting
501 >                  raw data, the velocity profile, is a robust and
502 >                  rapidly converging property. The method is tested on
503 >                  the Lennard-Jones fluid near its triple point. It
504 >                  yields a viscosity of 3.2-3.3, in Lennard-Jones
505 >                  reduced units, in agreement with literature
506 >                  results. {[}S1063-651X(99)03105-0].},
507 >  Address =      {ONE PHYSICS ELLIPSE, COLLEGE PK, MD 20740-3844 USA},
508 >  Affiliation =  {Muller-Plathe, F (Reprint Author), Max Planck Inst
509 >                  Polymerforsch, Ackermannweg 10, D-55128 Mainz,
510 >                  Germany. Max Planck Inst Polymerforsch, D-55128
511 >                  Mainz, Germany.},
512 >  Author =       {M\"{u}ller-Plathe, F},
513 >  Date-Added =   {2009-10-01 14:07:30 -0400},
514 >  Date-Modified ={2009-10-01 14:07:30 -0400},
515 >  Doc-Delivery-Number ={197TX},
516 >  Issn =         {1063-651X},
517 >  Journal =      {Phys. Rev. E},
518 >  Journal-Iso =  {Phys. Rev. E},
519 >  Language =     {English},
520 >  Month =        {MAY},
521 >  Number =       {5, Part A},
522 >  Number-Of-Cited-References ={17},
523 >  Pages =        {4894-4898},
524 >  Publisher =    {AMERICAN PHYSICAL SOC},
525 >  Subject-Category ={Physics, Fluids \& Plasmas; Physics,
526 >                  Mathematical},
527 >  Times-Cited =  {57},
528 >  Title =        {Reversing the perturbation in nonequilibrium
529 >                  molecular dynamics: An easy way to calculate the
530 >                  shear viscosity of fluids},
531 >  Type =         {Article},
532 >  Unique-Id =    {ISI:000080382700030},
533 >  Volume =       {59},
534 >  Year =         {1999}
535 > }
536  
537   @article{ISI:000246190100032,
538 <        Abstract = {Atomistic simulations are conducted to examine the dependence of the
539 <   viscosity of 1-ethyl-3-methylimidazolium
540 <   bis(trifluoromethanesulfonyl)imide on temperature and water content. A
541 <   nonequilibrium molecular dynamics procedure is utilized along with an
542 <   established fixed charge force field. It is found that the simulations
543 <   quantitatively capture the temperature dependence of the viscosity as
544 <   well as the drop in viscosity that occurs with increasing water
545 <   content. Using mixture viscosity models, we show that the relative drop
546 <   in viscosity with water content is actually less than that that would
547 <   be predicted for an ideal system. This finding is at odds with the
548 <   popular notion that small amounts of water cause an unusually large
549 <   drop in the viscosity of ionic liquids. The simulations suggest that,
550 <   due to preferential association of water with anions and the formation
551 <   of water clusters, the excess molar volume is negative. This means that
552 <   dissolved water is actually less effective at lowering the viscosity of
553 <   these mixtures when compared to a solute obeying ideal mixing behavior.
554 <   The use of a nonequilibrium simulation technique enables diffusive
555 <   behavior to be observed on the time scale of the simulations, and
556 <   standard equilibrium molecular dynamics resulted in sub-diffusive
557 <   behavior even over 2 ns of simulation time.},
558 <        Address = {1155 16TH ST, NW, WASHINGTON, DC 20036 USA},
559 <        Affiliation = {Maginn, EJ (Reprint Author), Univ Notre Dame, Dept Chem \& Biomol Engn, 182 Fitzpatrick Hall, Notre Dame, IN 46556 USA. Univ Notre Dame, Dept Chem \& Biomol Engn, Notre Dame, IN 46556 USA.},
560 <        Author = {Kelkar, Manish S. and Maginn, Edward J.},
561 <        Author-Email = {ed@nd.edu},
562 <        Date-Added = {2009-09-29 17:07:17 -0400},
563 <        Date-Modified = {2009-09-29 17:07:17 -0400},
564 <        Doc-Delivery-Number = {163VA},
565 <        Doi = {10.1021/jp0686893},
566 <        Issn = {1520-6106},
567 <        Journal = {JOURNAL OF PHYSICAL CHEMISTRY B},
568 <        Journal-Iso = {J. Phys. Chem. B},
569 <        Keywords-Plus = {MOLECULAR-DYNAMICS SIMULATION; MOMENTUM IMPULSE RELAXATION; FORCE-FIELD; TRANSPORT-PROPERTIES; PHYSICAL-PROPERTIES; SIMPLE FLUID; CHLORIDE; MODEL; SALTS; ARCHITECTURE},
570 <        Language = {English},
571 <        Month = {MAY 10},
572 <        Number = {18},
573 <        Number-Of-Cited-References = {57},
574 <        Pages = {4867-4876},
575 <        Publisher = {AMER CHEMICAL SOC},
576 <        Subject-Category = {Chemistry, Physical},
577 <        Times-Cited = {35},
578 <        Title = {Effect of temperature and water content on the shear viscosity of the ionic liquid 1-ethyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide as studied by atomistic simulations},
579 <        Type = {Article},
580 <        Unique-Id = {ISI:000246190100032},
581 <        Volume = {111},
582 <        Year = {2007},
583 <        Bdsk-Url-1 = {http://dx.doi.org/10.1021/jp0686893%7D}}
584 <
585 < @article{MullerPlathe:1997xw,
586 <        Abstract = {A nonequilibrium molecular dynamics method for calculating the thermal conductivity is presented. It reverses the usual cause and effect picture. The ''effect,'' the heat flux, is imposed on the system and the ''cause,'' the temperature gradient is obtained from the simulation. Besides being very simple to implement, the scheme offers several advantages such as compatibility with periodic boundary conditions, conservation of total energy and total linear momentum, and the sampling of a rapidly converging quantity (temperature gradient) rather than a slowly converging one (heat flux). The scheme is tested on the Lennard-Jones fluid. (C) 1997 American Institute of Physics.},
587 <        Address = {WOODBURY},
588 <        Author = {MullerPlathe, F.},
589 <        Cited-Reference-Count = {13},
590 <        Date = {APR 8},
591 <        Date-Added = {2009-09-21 16:51:21 -0400},
592 <        Date-Modified = {2009-09-21 16:51:21 -0400},
593 <        Document-Type = {Article},
594 <        Isi = {ISI:A1997WR62000032},
595 <        Isi-Document-Delivery-Number = {WR620},
596 <        Iso-Source-Abbreviation = {J. Chem. Phys.},
597 <        Issn = {0021-9606},
598 <        Journal = {JOURNAL OF CHEMICAL PHYSICS},
599 <        Language = {English},
600 <        Month = {Apr},
601 <        Number = {14},
602 <        Page-Count = {4},
603 <        Pages = {6082--6085},
447 <        Publication-Type = {J},
448 <        Publisher = {AMER INST PHYSICS},
449 <        Publisher-Address = {CIRCULATION FULFILLMENT DIV, 500 SUNNYSIDE BLVD, WOODBURY, NY 11797-2999},
450 <        Reprint-Address = {MullerPlathe, F, MAX PLANCK INST POLYMER RES, D-55128 MAINZ, GERMANY.},
451 <        Source = {J CHEM PHYS},
452 <        Subject-Category = {Physics, Atomic, Molecular & Chemical},
453 <        Times-Cited = {106},
454 <        Title = {A simple nonequilibrium molecular dynamics method for calculating the thermal conductivity},
455 <        Volume = {106},
456 <        Year = {1997}}
538 >  Abstract =     {Atomistic simulations are conducted to examine the
539 >                  dependence of the viscosity of
540 >                  1-ethyl-3-methylimidazolium
541 >                  bis(trifluoromethanesulfonyl)imide on temperature
542 >                  and water content. A nonequilibrium molecular
543 >                  dynamics procedure is utilized along with an
544 >                  established fixed charge force field. It is found
545 >                  that the simulations quantitatively capture the
546 >                  temperature dependence of the viscosity as well as
547 >                  the drop in viscosity that occurs with increasing
548 >                  water content. Using mixture viscosity models, we
549 >                  show that the relative drop in viscosity with water
550 >                  content is actually less than that that would be
551 >                  predicted for an ideal system. This finding is at
552 >                  odds with the popular notion that small amounts of
553 >                  water cause an unusually large drop in the viscosity
554 >                  of ionic liquids. The simulations suggest that, due
555 >                  to preferential association of water with anions and
556 >                  the formation of water clusters, the excess molar
557 >                  volume is negative. This means that dissolved water
558 >                  is actually less effective at lowering the viscosity
559 >                  of these mixtures when compared to a solute obeying
560 >                  ideal mixing behavior. The use of a nonequilibrium
561 >                  simulation technique enables diffusive behavior to
562 >                  be observed on the time scale of the simulations,
563 >                  and standard equilibrium molecular dynamics resulted
564 >                  in sub-diffusive behavior even over 2 ns of
565 >                  simulation time.},
566 >  Address =      {1155 16TH ST, NW, WASHINGTON, DC 20036 USA},
567 >  Affiliation =  {Maginn, EJ (Reprint Author), Univ Notre Dame, Dept
568 >                  Chem \& Biomol Engn, 182 Fitzpatrick Hall, Notre
569 >                  Dame, IN 46556 USA. Univ Notre Dame, Dept Chem \&
570 >                  Biomol Engn, Notre Dame, IN 46556 USA.},
571 >  Author =       {Kelkar, Manish S. and Maginn, Edward J.},
572 >  Author-Email = {ed@nd.edu},
573 >  Date-Added =   {2009-09-29 17:07:17 -0400},
574 >  Date-Modified ={2009-09-29 17:07:17 -0400},
575 >  Doc-Delivery-Number ={163VA},
576 >  Doi =          {10.1021/jp0686893},
577 >  Issn =         {1520-6106},
578 >  Journal =      {J. Phys. Chem. B},
579 >  Journal-Iso =  {J. Phys. Chem. B},
580 >  Keywords-Plus ={MOLECULAR-DYNAMICS SIMULATION; MOMENTUM IMPULSE
581 >                  RELAXATION; FORCE-FIELD; TRANSPORT-PROPERTIES;
582 >                  PHYSICAL-PROPERTIES; SIMPLE FLUID; CHLORIDE; MODEL;
583 >                  SALTS; ARCHITECTURE},
584 >  Language =     {English},
585 >  Month =        {MAY 10},
586 >  Number =       {18},
587 >  Number-Of-Cited-References ={57},
588 >  Pages =        {4867-4876},
589 >  Publisher =    {AMER CHEMICAL SOC},
590 >  Subject-Category ={Chemistry, Physical},
591 >  Times-Cited =  {35},
592 >  Title =        {Effect of temperature and water content on the shear
593 >                  viscosity of the ionic liquid
594 >                  1-ethyl-3-methylimidazolium
595 >                  bis(trifluoromethanesulfonyl)imide as studied by
596 >                  atomistic simulations},
597 >  Type =         {Article},
598 >  Unique-Id =    {ISI:000246190100032},
599 >  Volume =       {111},
600 >  Year =         {2007},
601 >  Bdsk-Url-1 =   {http://dx.doi.org/10.1021/jp0686893%7D},
602 >  Bdsk-Url-2 =   {http://dx.doi.org/10.1021/jp0686893}
603 > }
604  
605 + @article{MullerPlathe:1997xw,
606 +  Abstract =     {A nonequilibrium molecular dynamics method for
607 +                  calculating the thermal conductivity is
608 +                  presented. It reverses the usual cause and effect
609 +                  picture. The ''effect,'' the heat flux, is imposed
610 +                  on the system and the ''cause,'' the temperature
611 +                  gradient is obtained from the simulation. Besides
612 +                  being very simple to implement, the scheme offers
613 +                  several advantages such as compatibility with
614 +                  periodic boundary conditions, conservation of total
615 +                  energy and total linear momentum, and the sampling
616 +                  of a rapidly converging quantity (temperature
617 +                  gradient) rather than a slowly converging one (heat
618 +                  flux). The scheme is tested on the Lennard-Jones
619 +                  fluid. (C) 1997 American Institute of Physics.},
620 +  Address =      {WOODBURY},
621 +  Author =       {M\"{u}ller-Plathe, F.},
622 +  Cited-Reference-Count ={13},
623 +  Date =         {APR 8},
624 +  Date-Added =   {2009-09-21 16:51:21 -0400},
625 +  Date-Modified ={2009-09-21 16:51:21 -0400},
626 +  Document-Type ={Article},
627 +  Isi =          {ISI:A1997WR62000032},
628 +  Isi-Document-Delivery-Number ={WR620},
629 +  Iso-Source-Abbreviation ={J. Chem. Phys.},
630 +  Issn =         {0021-9606},
631 +  Journal =      {J. Chem. Phys.},
632 +  Language =     {English},
633 +  Month =        {Apr},
634 +  Number =       {14},
635 +  Page-Count =   {4},
636 +  Pages =        {6082--6085},
637 +  Publication-Type ={J},
638 +  Publisher =    {AMER INST PHYSICS},
639 +  Publisher-Address ={CIRCULATION FULFILLMENT DIV, 500 SUNNYSIDE BLVD,
640 +                  WOODBURY, NY 11797-2999},
641 +  Reprint-Address ={MullerPlathe, F, MAX PLANCK INST POLYMER RES,
642 +                  D-55128 MAINZ, GERMANY.},
643 +  Source =       {J CHEM PHYS},
644 +  Subject-Category ={Physics, Atomic, Molecular & Chemical},
645 +  Times-Cited =  {106},
646 +  Title =        {A simple nonequilibrium molecular dynamics method
647 +                  for calculating the thermal conductivity},
648 +  Volume =       {106},
649 +  Year =         {1997}
650 + }
651 +
652   @article{Muller-Plathe:1999ek,
653 <        Abstract = {A novel non-equilibrium method for calculating transport coefficients is presented. It reverses the experimental cause-and-effect picture, e.g. for the calculation of viscosities: the effect, the momentum flux or stress, is imposed, whereas the cause, the velocity gradient or shear rates, is obtained from the simulation. It differs from other Norton-ensemble methods by the way, in which the steady-state fluxes are maintained. This method involves a simple exchange of particle momenta, which is easy to implement and to analyse. Moreover, it can be made to conserve the total energy as well as the total linear momentum, so no thermostatting is needed. The resulting raw data are robust and rapidly converging. The method is tested on the calculation of the shear viscosity, the thermal conductivity and the Soret coefficient (thermal diffusion) for the Lennard-Jones (LJ) fluid near its triple point. Possible applications to other transport coefficients and more complicated systems are discussed. (C) 1999 Elsevier Science Ltd. All rights reserved.},
654 <        Address = {THE BOULEVARD, LANGFORD LANE, KIDLINGTON, OXFORD OX5 1GB, OXON, ENGLAND},
655 <        Author = {Muller-Plathe, F and Reith, D},
656 <        Date-Added = {2009-09-21 16:47:07 -0400},
657 <        Date-Modified = {2009-09-21 16:47:07 -0400},
658 <        Isi = {000082266500004},
659 <        Isi-Recid = {111564960},
660 <        Isi-Ref-Recids = {64516210 89773595 53816621 60134000 94875498 60964023 90228608 85968509 86405859 63979644 108048497 87560156 577165 103281654 111564961 83735333 99953572 88476740 110174781 111564963 6599000 75892253},
661 <        Journal = {Computational and Theoretical Polymer Science},
662 <        Keywords = {viscosity; Ludwig-Soret effect; thermal conductivity; Onsager coefficents; non-equilibrium molecular dynamics},
663 <        Number = {3-4},
664 <        Pages = {203-209},
665 <        Publisher = {ELSEVIER SCI LTD},
666 <        Times-Cited = {15},
667 <        Title = {Cause and effect reversed in non-equilibrium molecular dynamics: an easy route to transport coefficients},
668 <        Volume = {9},
669 <        Year = {1999},
670 <        Bdsk-Url-1 = {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000082266500004}}
653 >  Abstract =     {A novel non-equilibrium method for calculating
654 >                  transport coefficients is presented. It reverses the
655 >                  experimental cause-and-effect picture, e.g. for the
656 >                  calculation of viscosities: the effect, the momentum
657 >                  flux or stress, is imposed, whereas the cause, the
658 >                  velocity gradient or shear rates, is obtained from
659 >                  the simulation. It differs from other
660 >                  Norton-ensemble methods by the way, in which the
661 >                  steady-state fluxes are maintained. This method
662 >                  involves a simple exchange of particle momenta,
663 >                  which is easy to implement and to analyse. Moreover,
664 >                  it can be made to conserve the total energy as well
665 >                  as the total linear momentum, so no thermostatting
666 >                  is needed. The resulting raw data are robust and
667 >                  rapidly converging. The method is tested on the
668 >                  calculation of the shear viscosity, the thermal
669 >                  conductivity and the Soret coefficient (thermal
670 >                  diffusion) for the Lennard-Jones (LJ) fluid near its
671 >                  triple point. Possible applications to other
672 >                  transport coefficients and more complicated systems
673 >                  are discussed. (C) 1999 Elsevier Science Ltd. All
674 >                  rights reserved.},
675 >  Address =      {THE BOULEVARD, LANGFORD LANE, KIDLINGTON, OXFORD OX5
676 >                  1GB, OXON, ENGLAND},
677 >  Author =       {M\"{u}ller-Plathe, F and Reith, D},
678 >  Date-Added =   {2009-09-21 16:47:07 -0400},
679 >  Date-Modified ={2009-09-21 16:47:07 -0400},
680 >  Isi =          {000082266500004},
681 >  Isi-Recid =    {111564960},
682 >  Isi-Ref-Recids ={64516210 89773595 53816621 60134000 94875498
683 >                  60964023 90228608 85968509 86405859 63979644
684 >                  108048497 87560156 577165 103281654 111564961
685 >                  83735333 99953572 88476740 110174781 111564963
686 >                  6599000 75892253},
687 >  Journal =      {Computational and Theoretical Polymer Science},
688 >  Keywords =     {viscosity; Ludwig-Soret effect; thermal
689 >                  conductivity; Onsager coefficents; non-equilibrium
690 >                  molecular dynamics},
691 >  Number =       {3-4},
692 >  Pages =        {203-209},
693 >  Publisher =    {ELSEVIER SCI LTD},
694 >  Times-Cited =  {15},
695 >  Title =        {Cause and effect reversed in non-equilibrium
696 >                  molecular dynamics: an easy route to transport
697 >                  coefficients},
698 >  Volume =       {9},
699 >  Year =         {1999},
700 >  Bdsk-Url-1 =
701 >                  {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000082266500004}
702 > }
703  
704   @article{Viscardy:2007lq,
705 <        Abstract = {The thermal conductivity is calculated with the Helfand-moment method in the Lennard-Jones fluid near the triple point. The Helfand moment of thermal conductivity is here derived for molecular dynamics with periodic boundary conditions. Thermal conductivity is given by a generalized Einstein relation with this Helfand moment. The authors compute thermal conductivity by this new method and compare it with their own values obtained by the standard Green-Kubo method. The agreement is excellent. (C) 2007 American Institute of Physics.},
706 <        Address = {CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501 USA},
707 <        Author = {Viscardy, S. and Servantie, J. and Gaspard, P.},
708 <        Date-Added = {2009-09-21 16:37:20 -0400},
709 <        Date-Modified = {2009-09-21 16:37:20 -0400},
710 <        Doi = {DOI 10.1063/1.2724821},
711 <        Isi = {000246453900035},
712 <        Isi-Recid = {156192451},
713 <        Isi-Ref-Recids = {18794442 84473620 156192452 41891249 90040203 110174972 59859940 47256160 105716249 91804339 93329429 95967319 6199670 1785176 105872066 6325196 65361295 71941152 4307928 23120502 54053395 149068110 4811016 99953572 59859908 132156782 156192449},
714 <        Journal = {Journal of Chemical Physics},
715 <        Month = may,
716 <        Number = {18},
717 <        Publisher = {AMER INST PHYSICS},
718 <        Times-Cited = {3},
719 <        Title = {Transport and Helfand moments in the Lennard-Jones fluid. II. Thermal conductivity},
720 <        Volume = {126},
721 <        Year = {2007},
722 <        Bdsk-Url-1 = {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000246453900035}}
705 >  Abstract =     {The thermal conductivity is calculated with the
706 >                  Helfand-moment method in the Lennard-Jones fluid
707 >                  near the triple point. The Helfand moment of thermal
708 >                  conductivity is here derived for molecular dynamics
709 >                  with periodic boundary conditions. Thermal
710 >                  conductivity is given by a generalized Einstein
711 >                  relation with this Helfand moment. The authors
712 >                  compute thermal conductivity by this new method and
713 >                  compare it with their own values obtained by the
714 >                  standard Green-Kubo method. The agreement is
715 >                  excellent. (C) 2007 American Institute of Physics.},
716 >  Address =      {CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON
717 >                  QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501
718 >                  USA},
719 >  Author =       {Viscardy, S. and Servantie, J. and Gaspard, P.},
720 >  Date-Added =   {2009-09-21 16:37:20 -0400},
721 >  Date-Modified ={2009-09-21 16:37:20 -0400},
722 >  Doi =          {DOI 10.1063/1.2724821},
723 >  Isi =          {000246453900035},
724 >  Isi-Recid =    {156192451},
725 >  Isi-Ref-Recids ={18794442 84473620 156192452 41891249 90040203
726 >                  110174972 59859940 47256160 105716249 91804339
727 >                  93329429 95967319 6199670 1785176 105872066 6325196
728 >                  65361295 71941152 4307928 23120502 54053395
729 >                  149068110 4811016 99953572 59859908 132156782
730 >                  156192449},
731 >  Journal =      {J. Chem. Phys.},
732 >  Month =        may,
733 >  Number =       {18},
734 >  Publisher =    {AMER INST PHYSICS},
735 >  Times-Cited =  {3},
736 >  Title =        {Transport and Helfand moments in the Lennard-Jones
737 >                  fluid. II. Thermal conductivity},
738 >  Volume =       {126},
739 >  Year =         {2007},
740 >  Bdsk-Url-1 =
741 >                  {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000246453900035},
742 >  Bdsk-Url-2 =   {http://dx.doi.org/10.1063/1.2724821}
743 > }
744  
745   @article{Viscardy:2007bh,
746 <        Abstract = {The authors propose a new method, the Helfand-moment method, to compute the shear viscosity by equilibrium molecular dynamics in periodic systems. In this method, the shear viscosity is written as an Einstein-type relation in terms of the variance of the so-called Helfand moment. This quantity is modified in order to satisfy systems with periodic boundary conditions usually considered in molecular dynamics. They calculate the shear viscosity in the Lennard-Jones fluid near the triple point thanks to this new technique. They show that the results of the Helfand-moment method are in excellent agreement with the results of the standard Green-Kubo method. (C) 2007 American Institute of Physics.},
747 <        Address = {CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501 USA},
748 <        Author = {Viscardy, S. and Servantie, J. and Gaspard, P.},
749 <        Date-Added = {2009-09-21 16:37:19 -0400},
750 <        Date-Modified = {2009-09-21 16:37:19 -0400},
751 <        Doi = {DOI 10.1063/1.2724820},
752 <        Isi = {000246453900034},
753 <        Isi-Recid = {156192449},
754 <        Isi-Ref-Recids = {18794442 89109900 84473620 86837966 26564374 23367140 83161139 75750220 90040203 110174972 5885 67722779 91461489 42484251 77907850 93329429 95967319 105716249 6199670 1785176 105872066 6325196 129596740 120782555 51131244 65361295 41141868 4307928 21555860 23120502 563068 120721875 142813985 135942402 4811016 86224873 57621419 85506488 89860062 44796632 51381285 132156779 156192450 132156782 156192451},
755 <        Journal = {Journal of Chemical Physics},
756 <        Month = may,
757 <        Number = {18},
758 <        Publisher = {AMER INST PHYSICS},
759 <        Times-Cited = {1},
760 <        Title = {Transport and Helfand moments in the Lennard-Jones fluid. I. Shear viscosity},
761 <        Volume = {126},
762 <        Year = {2007},
763 <        Bdsk-Url-1 = {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000246453900034}}
746 >  Abstract =     {The authors propose a new method, the Helfand-moment
747 >                  method, to compute the shear viscosity by
748 >                  equilibrium molecular dynamics in periodic
749 >                  systems. In this method, the shear viscosity is
750 >                  written as an Einstein-type relation in terms of the
751 >                  variance of the so-called Helfand moment. This
752 >                  quantity is modified in order to satisfy systems
753 >                  with periodic boundary conditions usually considered
754 >                  in molecular dynamics. They calculate the shear
755 >                  viscosity in the Lennard-Jones fluid near the triple
756 >                  point thanks to this new technique. They show that
757 >                  the results of the Helfand-moment method are in
758 >                  excellent agreement with the results of the standard
759 >                  Green-Kubo method. (C) 2007 American Institute of
760 >                  Physics.},
761 >  Address =      {CIRCULATION \& FULFILLMENT DIV, 2 HUNTINGTON
762 >                  QUADRANGLE, STE 1 N O 1, MELVILLE, NY 11747-4501
763 >                  USA},
764 >  Author =       {Viscardy, S. and Servantie, J. and Gaspard, P.},
765 >  Date-Added =   {2009-09-21 16:37:19 -0400},
766 >  Date-Modified ={2009-09-21 16:37:19 -0400},
767 >  Doi =          {DOI 10.1063/1.2724820},
768 >  Isi =          {000246453900034},
769 >  Isi-Recid =    {156192449},
770 >  Isi-Ref-Recids ={18794442 89109900 84473620 86837966 26564374
771 >                  23367140 83161139 75750220 90040203 110174972 5885
772 >                  67722779 91461489 42484251 77907850 93329429
773 >                  95967319 105716249 6199670 1785176 105872066 6325196
774 >                  129596740 120782555 51131244 65361295 41141868
775 >                  4307928 21555860 23120502 563068 120721875 142813985
776 >                  135942402 4811016 86224873 57621419 85506488
777 >                  89860062 44796632 51381285 132156779 156192450
778 >                  132156782 156192451},
779 >  Journal =      {J. Chem. Phys.},
780 >  Month =        may,
781 >  Number =       {18},
782 >  Publisher =    {AMER INST PHYSICS},
783 >  Times-Cited =  {1},
784 >  Title =        {Transport and Helfand moments in the Lennard-Jones
785 >                  fluid. I. Shear viscosity},
786 >  Volume =       {126},
787 >  Year =         {2007},
788 >  Bdsk-Url-1 =
789 >                  {http://gateway.isiknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=Alerting&SrcApp=Alerting&DestApp=WOS&DestLinkType=FullRecord;KeyUT=000246453900034},
790 >  Bdsk-Url-2 =   {http://dx.doi.org/10.1063/1.2724820}
791 > }
792 >

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines