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

Comparing trunk/chainLength/GoldThiolsPaper.tex (file contents):
Revision 3849 by kstocke1, Fri Dec 21 19:07:54 2012 UTC vs.
Revision 3850 by gezelter, Fri Dec 21 19:29:36 2012 UTC

# Line 77 | Line 77 | thermal-transfer fluids, although careful experiments
77  
78   Metallic particles have also been proposed for use in highly efficient
79   thermal-transfer fluids, although careful experiments by Eapen {\it et
80 <  al.}  have shown that metal-particle-based ``nanofluids'' have
81 < thermal conductivities that match Maxwell
82 < predictions.\cite{Eapen:2007th} The likely cause of previously
83 < reported non-Maxwell
80 >  al.}  have shown that metal-particle-based nanofluids have thermal
81 > conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
82 > likely cause of previously reported non-Maxwell
83   behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
84   is percolation networks of nanoparticles exchanging energy via the
85 < solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
86 < picture of particle-ligand and particle-solvent interactions in order
87 < to understand the thermal behavior of complex fluids. To date, there
88 < have been few reported values (either from theory or experiment) for
89 < $G$ for ligand-protected nanoparticles embedded in liquids, and there
90 < is a significant gap in knowledge about how chemically distinct
91 < ligands or protecting groups will affect heat transport from the
92 < particles.
85 > solvent,\cite{Eapen:2007mw} so it is important to get a detailed
86 > molecular picture of particle-ligand and ligand-solvent interactions
87 > in order to understand the thermal behavior of complex fluids. To
88 > date, there have been few reported values (either from theory or
89 > experiment) for $G$ for ligand-protected nanoparticles embedded in
90 > liquids, and there is a significant gap in knowledge about how
91 > chemically distinct ligands or protecting groups will affect heat
92 > transport from the particles.
93  
94   Experimentally, the thermal properties of various nanostructured
95   interfaces have been investigated by a number of groups. Cahill and
96 < coworkers studied nanoscale thermal transport from metal
97 < nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxide
98 < interfaces, and hydrophilic and hydrophobic interfaces between water
99 < and solids with different self-assembled
96 > coworkers studied thermal transport from metal nanoparticle/fluid
97 > interfaces, epitaxial TiN/single crystal oxide interfaces, and
98 > hydrophilic and hydrophobic interfaces between water and solids with
99 > different self-assembled
100   monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
101   Wang {\it et al.} studied heat transport through long-chain
102   hydrocarbon monolayers on gold substrate at the individual molecular
# Line 108 | Line 107 | suggested that specific ligands (capping agents) could
107   interface resistance of glass-embedded metal
108   nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
109   normally considered barriers for heat transport, Alper {\it et al.}
110 < suggested that specific ligands (capping agents) could completely
110 > have suggested that specific ligands (capping agents) could completely
111   eliminate this barrier
112   ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
113  
# Line 116 | Line 115 | calculations suggest an explanation for the very large
115   non-equilibrium molecular dynamics (RNEMD) to calculate the
116   interfacial thermal conductance at a metal / organic solvent interface
117   that had been chemically protected by butanethiolate groups.  Our
118 < calculations suggest an explanation for the very large thermal
118 > calculations suggested an explanation for the very large thermal
119   conductivity at alkanethiol-capped metal surfaces when compared with
120   bare metal/solvent interfaces.  Specifically, the chemical bond
121   between the metal and the ligand introduces a vibrational overlap that
# Line 124 | Line 123 | One notable result of our previous work was the observ
123   the vibrational spectra (metal to ligand, ligand to solvent) provides
124   a mechanism for rapid thermal transport across the interface.
125  
126 < One notable result of our previous work was the observation of
127 < non-monotonic dependence of the thermal conductance on the coverage of
128 < a metal surface by a chemical protecting group.  Our explanation for
129 < this behavior was that gaps in surface coverage allowed solvent to
130 < penetrate close to the capping molecules that had been heated by the
131 < metal surface, to absorb thermal energy from these molecules, and then
132 < diffuse away.  The effect of surface coverage was relatively difficult
133 < to study as the individual protecting groups have lateral mobility on
135 < the surface and can aggregate to form domains on the timescale of the
136 < simulation.  
126 > A notable result of the previous simulations was the non-monotonic
127 > dependence of $G$ on the fractional coverage of the metal surface by
128 > the chemical protecting group.  Gaps in surface coverage allow the
129 > solvent molecules come into direct contact with ligands that had been
130 > heated by the metal surface, absorb thermal energy from the ligands,
131 > and then diffuse away.  Quantifying the role of surface coverage is
132 > difficult as the ligands have lateral mobility on the surface and can
133 > aggregate to form domains on the timescale of the simulation.
134  
135 < To prevent lateral mobility of the surface ligands, the current work
136 < involves mixed-chain monolayers in which the length mismatch between
137 < long and short chains is sufficient to accomodate solvent
138 < molecules. These complete (but mixed-chain) surfaces are significantly
139 < less prone to surface domain formation, and allow us to further
140 < investigate the mechanism of heat transport to the solvent.  A thermal
141 < flux is created using velocity shearing and scaling reverse
142 < non-equilibrium molecular dynamics (VSS-RNEMD), and the resulting
143 < temperature profiles are analyzed to yield information about the
144 < interfacial thermal conductance.
135 > To isolate this effect without worrying about lateral mobility of the
136 > surface ligands, the current work involves mixed-chain-length
137 > monolayers in which the length mismatch between long and short chains
138 > is sufficient to accomodate solvent molecules. These completely
139 > covered (but mixed-chain) surfaces are significantly less prone to
140 > surface domain formation, and allow us to further investigate the
141 > mechanism of heat transport to the solvent.  A thermal flux is created
142 > using velocity shearing and scaling reverse non-equilibrium molecular
143 > dynamics (VSS-RNEMD), and the resulting temperature profiles are
144 > analyzed to yield information about the interfacial thermal
145 > conductance.
146  
147  
148   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 171 | Line 169 | box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuan
169   {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
170   a different approach in that an unphysical {\it flux} is imposed
171   between different regions or ``slabs'' of the simulation
172 < box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
172 > box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
173   system responds by developing a temperature or momentum {\it gradient}
174   between the two regions. Since the amount of the applied flux is known
175   exactly, and the measurement of a gradient is generally less
# Line 207 | Line 205 | the two slabs.\cite{2012MolPh.110..691K} This method p
205  
206   The most useful alternative RNEMD approach developed so far utilizes a
207   series of simultaneous velocity shearing and scaling exchanges between
208 < the two slabs.\cite{2012MolPh.110..691K} This method provides a set of
208 > the two slabs.\cite{Kuang2012} This method provides a set of
209   conservation constraints while simultaneously creating a desired flux
210   between the two slabs.  Satisfying the constraint equations ensures
211   that the new configurations are sampled from the same NVE ensemble.
# Line 286 | Line 284 | a single 1 ns simulation.\cite{2012MolPh.110..691K}
284   kept to a minimum.  This ability to generate simultaneous thermal and
285   shear fluxes has been previously utilized to map out the shear
286   viscosity of SPC/E water over a wide range of temperatures (90~K) with
287 < a single 1 ns simulation.\cite{2012MolPh.110..691K}
287 > a single 1 ns simulation.\cite{Kuang2012}
288  
289   \begin{figure}
290    \includegraphics[width=\linewidth]{figures/rnemd}
# Line 348 | Line 346 | the gold/thiolate interface and the thiolate/solvent i
346  
347   In the particular case we are studying here, there are two interfaces
348   involved in the transfer of heat from the gold slab to the solvent:
349 < the gold/thiolate interface and the thiolate/solvent interface. We
349 > the metal/thiolate interface and the thiolate/solvent interface. We
350   could treat the temperature on each side of an interface as discrete,
351   making the interfacial conductance the inverse of the Kaptiza
352   resistance, or $G = \frac{1}{R_k}$. To model the total conductance
# Line 383 | Line 381 | simulations. Force field parameters are needed for int
381   Our simulations include a number of chemically distinct components.
382   Figure \ref{fig:structures} demonstrates the sites defined for both
383   the {\it n}-hexane and alkanethiolate ligands present in our
384 < simulations. Force field parameters are needed for interactions both
387 < between the same type of particles and between particles of different
388 < species.
384 > simulations.
385  
386   \begin{figure}
387    \includegraphics[width=\linewidth]{figures/structures}
# Line 412 | Line 408 | interactions, Lennard-Jones potentials are used.  We h
408   sites are located at the carbon centers for alkyl groups. Bonding
409   interactions, including bond stretches and bends and torsions, were
410   used for intra-molecular sites closer than 3 bonds. For non-bonded
411 < interactions, Lennard-Jones potentials are used.  We have previously
411 > interactions, Lennard-Jones potentials were used.  We have previously
412   utilized both united atom (UA) and all-atom (AA) force fields for
413 < thermal conductivity in previous work,\cite{} and since the united
413 > thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
414   atom force fields cannot populate the high-frequency modes that are
415   present in AA force fields, they appear to work better for modeling
416   thermal conductivity.  The TraPPE-UA model for alkanes is known to
417   predict a slightly lower boiling point than experimental values. This
418   is one of the reasons we used a lower average temperature (200K) for
419 < our simulations.
419 > our simulations.
420  
421   The TraPPE-UA force field includes parameters for thiol
422   molecules\cite{TraPPE-UA.thiols} which were used for the
# Line 445 | Line 441 | We have implemented the VSS-RNEMD algorithm in OpenMD,
441   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
442   \subsection{Simulation Protocol}
443  
444 < We have implemented the VSS-RNEMD algorithm in OpenMD, our
445 < group molecular dynamics code. A 1188 atom gold slab was
446 < equilibrated at 1 atm and 200 K. The periodic box was then expanded
447 < in the $z$-direction to expose two Au(111) faces on either side of the 11-atom thick slab.
444 > We have implemented the VSS-RNEMD algorithm in OpenMD, our group
445 > molecular dynamics code.\cite{openmd} An 1188 atom gold slab was
446 > equilibrated at 1 atm and 200 K.  The periodic box was then expanded
447 > in the $z$-direction to expose two Au(111) faces on either side of the
448 > 11-atom thick slab.
449  
450 < A full monolayer of thiolates (1/3 the number of surface gold atoms) was placed on three-fold hollow sites on the gold interfaces. To efficiently test the effect of thiolate binding sites on the thermal conductance, all systems had one gold interface with thiolates placed only on fcc hollow sites and the other interface with thiolates only on hcp hollow sites. To test the effect of thiolate chain length on interfacial thermal conductance, full coverages of five chain lengths were tested: butanethiolate, hexanethiolate, octanethiolate, decanethiolate, and dodecanethiolate. To test the effect of mixed chain lengths, full coverages of butanethiolate/decanethiolate and butanethiolate/dodecanethiolate mixtures were created in short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and 87.5/12.5. The short and long chains were placed on the surface hollow sites in a random configuration.
450 > A full monolayer of thiolates (1/3 the number of surface gold atoms)
451 > was placed on three-fold hollow sites on the gold interfaces. The
452 > effect of thiolate binding sites on the thermal conductance was tested
453 > by placing thiolates at both fcc and hcp hollow sites.  No appreciable
454 > difference in the temperature profile was noted due to the location of
455 > thiolate binding.  
456  
457 < The simulation box $z$-dimension was set to roughly double the length of the gold/thiolate block. Hexane solvent molecules were placed in the vacant portion of the box using the packmol algorithm. Hexane, a straight chain flexible alkane, is very structurally similar to the thiolate alkane tails; previous work has shown that UA models of hexane and butanethiolate have a high degree of vibrational overlap.\cite{Kuang2011} This overlap should provide a mechanism for thermal energy transfer from the thiolates to the solvent.
457 > To test the role of thiolate chain length on interfacial thermal
458 > conductance, full coverages of each of five chain lengths were tested:
459 > butanethiolate (C$_4$), hexanethiolate (C$_6$), octanethiolate
460 > (C$_8$), decanethiolate (C$_{10}$), and dodecanethiolate
461 > (C$_{12}$). To test the effect of mixed chain lengths, full coverages
462 > of C$_4$ / C$_{10}$ and C$_4$ / C$_{12}$ mixtures were created in
463 > short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and
464 > 87.5/12.5. The short and long chains were placed on the surface hollow
465 > sites in a random configuration.
466  
467 < The system was equilibrated to 220 K in the NVT ensemble, allowing the thiolates and solvent to warm gradually. Pressure correction to 1 atm was done in an NPT ensemble that allowed expansion or contraction only in the z direction, so as not to disrupt the crystalline structure of the gold. The diagonal elements of the pressure tensor were monitored during the pressure correction step. If the xx and/or yy elements had a mean above zero throughout the simulation -- indicating residual pressure in the plane of the gold slab -- an additional short NPT equilibration step was performed allowing all box dimensions to change. Once the pressure was stable at 1 atm, a final NVT simulation was performed. All systems were equilibrated in the microcanonical (NVE) ensemble before proceeding with the VSS-RNEMD step.
467 > The simulation box $z$-dimension was set to roughly double the length
468 > of the gold/thiolate block. Hexane solvent molecules were placed in
469 > the vacant portion of the box using the packmol algorithm.  
470  
471 < A kinetic energy flux was applied using VSS-RNEMD in the NVE ensemble. The total simulation time was 3 nanoseconds, with velocity scaling occurring every 10 femtoseconds. The hot slab was centered in the gold and the cold slab was placed in the center of the solvent region. The average temperature was 220 K, with a temperature difference between the hot and cold slabs of approximately 30 K. The average temperature and kinetic energy flux were carefully selected with two considerations in mind: 1) if the cold bin gets too cold (below ~180 K) the solvent may freeze or undergo a glassy transition, and 2) due to the deep sulfur-gold potential well, sulfur atoms routinely embed into the gold slab, particularly at temperatures above 250 K. Simulation conditions were chosen to avoid both of these undesirable situations. A reversed flux direction resulted in frozen long chain thiolates and solvent too near its boiling point.
471 > The system was equilibrated to 220 K in the canonical (NVT) ensemble,
472 > allowing the thiolates and solvent to warm gradually. Pressure
473 > correction to 1 atm was done using an isobaric-isothermal (NPT)
474 > integrator that allowed expansion or contraction only in the $z$
475 > direction, maintaining the crystalline structure of the gold as close
476 > to the bulk result as possible.  The diagonal elements of the pressure
477 > tensor were monitored during the pressure equilibration stage.  If the
478 > $xx$ and/or $yy$ elements had a mean above zero throughout the
479 > simulation -- indicating residual surface tension in the plane of the
480 > gold slab -- an additional short NPT equilibration step was performed
481 > allowing all box dimensions to change.  Once the pressure was stable
482 > at 1 atm, a final equilibration stage was performed at constant
483 > temperature. All systems were equilibrated in the microcanonical (NVE)
484 > ensemble before proceeding with the VSS-RNEMD and data collection
485 > stages.
486  
487 < A temperature profile of the system was created by dividing the box into $\sim$ 3 \AA \, bins along the z axis and recording the average temperature of each bin.
487 > A kinetic energy flux was applied using VSS-RNEMD. The total
488 > simulation time was 3 nanoseconds, with velocity scaling occurring
489 > every 10 femtoseconds.  The hot slab was centered in the gold and the
490 > cold slab was placed in the center of the solvent region.  The entire
491 > system has a (time-averaged) temperature of 220 K, with a temperature
492 > difference between the hot and cold slabs of approximately 30 K.  The
493 > average temperature and kinetic energy flux were selected to prevent
494 > solvent freezing (or glass formation) and to not allow the thiolates
495 > to bury in the gold slab.  The Au-S interaction has a deep potential
496 > energy well, which allows sulfur atoms to embed into the gold slab at
497 > temperatures above 250 K.  Simulation conditions were chosen to avoid
498 > both of these undesirable situations.
499  
500 + Temperature profiles of the system were created by dividing the box
501 + into $\sim$ 3 \AA \, bins along the $z$ axis and recording the average
502 + temperature of each bin.
503  
504 +
505                  
506   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
507   %                          **RESULTS**
508   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%                        
509   \section{Results}
510  
511 + The solvent, hexane, a straight chain flexible alkane, is structurally
512 + similar to the thiolate alkane tails, and previous work has shown that
513 + UA models of hexane and butanethiolate have a high degree of
514 + vibrational overlap.\cite{kuang:AuThl} This overlap provides a
515 + mechanism for thermal energy transfer from the thiolates to the
516 + solvent.
517  
518   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
519   % CHAIN LENGTH

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines