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

Comparing trunk/matt_papers/canidacy_paper/canidacy_paper.tex (file contents):
Revision 108 by mmeineke, Wed Sep 11 15:58:41 2002 UTC vs.
Revision 110 by mmeineke, Mon Sep 16 22:10:09 2002 UTC

# Line 36 | Line 36 | atoms,and the head group of the lipid will typically c
36  
37   Simulations of phospholipid bilayers are, by necessity, quite
38   complex. The lipid molecules are large molecules containing many
39 < atoms,and the head group of the lipid will typically contain charge
39 > atoms, and the head group of the lipid will typically contain charge
40   separated ions which set up a large dipole within the molecule. Adding
41   to the complexity are the number of water molecules needed to properly
42 < solvate the lipid bilayer. Because of these factors, many current
43 < simulations are limited in both length and time scale due to to the
44 < sheer number of calculations performed at every time step and the
45 < lifetime of the researcher. A typical
42 > solvate the lipid bilayer, typically 25 water molecules for every
43 > lipid molecule. Because of these factors, many current simulations are
44 > limited in both length and time scale due to to the sheer number of
45 > calculations performed at every time step and the lifetime of the
46 > researcher. A typical
47   simulation\cite{saiz02,lindahl00,venable00,Marrink01} will have around
48   64 phospholipids forming a bilayer approximately 40~$\mbox{\AA}$ by
49   50~$\mbox{\AA}$ with roughly 25 waters for every lipid. This means
50   there are on the order of 8,000 atoms needed to simulate these systems
51 < and the trajectories in turn are integrated for times up to 10 ns.
51 > and the trajectories are integrated for times up to 10 ns.
52  
53   These limitations make it difficult to study certain biologically
54   interesting phenomena that don't fit within the short time and length
# Line 60 | Line 61 | these numbers are reasonable.
61   roughly 25 waters for every lipid to fully solvate the bilayer. With
62   the large number of atoms involved in a simulation of this magnitude,
63   steps \emph{must} be taken to simplify the system to the point where
64 < these numbers are reasonable.
64 > the numbers of atoms becomes reasonable.
65  
66   Another system of interest would be drug molecule diffusion through
67 < the membrane. Due to the fluid like properties of a lipid membrane,
67 > the membrane. Due to the fluid-like properties of a lipid membrane,
68   not all diffusion takes place at membrane channels. It is of interest
69   to study certain molecules that may incorporate themselves directly
70   into the membrane. These molecules may then have an appreciable
71   waiting time (on the order of nanoseconds) within the
72   bilayer. Simulation of such a long time scale again requires
73   simplification of the system in order to lower the number of
74 < calculations needed at each time step.
74 > calculations needed at each time step or to increase the length of
75 > each time step.
76  
77  
78   \section{Methodology}
# Line 79 | Line 81 | computational cost of the system. This is done by a co
81  
82   The length scale simplifications are aimed at increasing the number of
83   molecules that can be simulated without drastically increasing the
84 < computational cost of the system. This is done by a combination of
85 < substituting less expensive interactions for expensive ones and
86 < decreasing the number of interaction sites per molecule. Namely,
87 < point charge distributions are replaced with dipoles, and unified atoms are
88 < used in place of water, phospholipid head groups, and alkyl groups.
84 > computational cost of the simulation. This is done through a
85 > combination of substituting less expensive interactions for expensive
86 > ones and decreasing the number of interaction sites per
87 > molecule. Namely, point charge distributions are replaced with
88 > dipoles, and unified atoms are used in place of water, phospholipid
89 > head groups, and alkyl groups.
90  
91   The replacement of charge distributions with dipoles allows us to
92 < replace an interaction that has a relatively long range,
93 < $(\frac{1}{r})$, for the coulomb potential, with that of a relatively
94 < short range, $(\frac{1}{r^{3}})$, for dipole - dipole
95 < potentials. Combined with a computational simplification algorithm
96 < such as a Verlet neighbor list,\cite{allen87:csl} this should give
97 < computational scaling by $N$. This is in comparison to the Ewald
98 < sum\cite{leach01:mm} needed to compute the coulomb interactions, which
99 < scales at best by $N \ln N$.
92 > replace an interaction that has a relatively long range ($\frac{1}{r}$
93 > for the coulomb potential) with that of a relatively short range
94 > ($\frac{1}{r^{3}}$ for dipole - dipole potentials). Combined with
95 > Verlet neighbor lists,\cite{allen87:csl} this should result in an
96 > algorithm wich scales linearly with increasing system size. This is in
97 > comparison to the Ewald sum\cite{leach01:mm} needed to compute
98 > periodic replicas of the coulomb interactions, which scales at best by
99 > $N \ln N$.
100  
101   The second step taken to simplify the number of calculations is to
102   incorporate unified models for groups of atoms. In the case of water,
103   we use the soft sticky dipole (SSD) model developed by
104 < Ichiye\cite{Liu96} (Section~\ref{sec:ssdModel}). For the phospholipids, a
105 < unified head atom with a dipole will replace the atoms in the head
106 < group, while unified $\text{CH}_2$ and $\text{CH}_3$ atoms will
107 < replace the alkyl units in the tails (Section~\ref{sec:lipidModel}).
104 > Ichiye\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md}
105 > (Section~\ref{sec:ssdModel}). For the phospholipids, a unified head
106 > atom with a dipole will replace the atoms in the head group, while
107 > unified $\text{CH}_2$ and $\text{CH}_3$ atoms will replace the alkyl
108 > units in the tails (Section~\ref{sec:lipidModel}).
109  
110   The time scale simplifications are introduced so that we can take
111   longer time steps. By increasing the size of the time steps taken by
112 < the simulation, we are able to integrate the simulation trajectory
113 < with fewer calculations. However, care must be taken that any
112 > the simulation, we are able to integrate a given length of time using
113 > fewer calculations. However, care must be taken that any
114   simplifications used, still conserve the total energy of the
115   simulation. In practice, this means taking steps small enough to
116   resolve all motion in the system without accidently moving an object
# Line 124 | Line 128 | conserve energy when bonds lengths are allowed to osci
128   \subsection{The Soft Sticky Water Model}
129   \label{sec:ssdModel}
130  
131 < %\begin{floatingfigure}{55mm}
132 < %\includegraphics[width=45mm]{ssd.epsi}
133 < %\caption{The SSD model with the oxygen and hydrogen atoms drawn in for reference. \vspace{5mm}}
134 < %\label{fig:ssdModel}
135 < %\end{floatingfigure}
131 > \begin{figure}
132 > \begin{center}
133 > \includegraphics[width=50mm]{ssd.epsi}
134 > \caption{The SSD model with the oxygen and hydrogen atoms drawn in for reference.}
135 > \end{center}
136 > \label{fig:ssdModel}
137 > \end{figure}
138  
139   The water model used in our simulations is a modified soft
140   Stockmayer-sphere model.\cite{stevens95} Like the Stockmayer-sphere, the SSD
141 < model\cite{Liu96} consists of a Lennard-Jones interaction site and a
141 > model consists of a Lennard-Jones interaction site and a
142   dipole both located at the water's center of mass (Figure
143   \ref{fig:ssdModel}). However, the SSD model extends this by adding a
144   tetrahedral potential to correct for hydrogen bonding.
145  
146 < The SSD water potential is then given by the following equation:
146 > The SSD water potential for a pair of water molecules is then given by
147 > the following equation:
148   \begin{equation}
149   V_{\text{SSD}} = V_{\text{LJ}}(r_{i\!j}) + V_{\text{dp}}(\mathbf{r}_{i\!j},
150          \boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})
# Line 145 | Line 152 | $V_{\text{LJ}}$ is the Lennard-Jones potential:
152          \boldsymbol{\Omega}_{j})
153   \label{eq:ssdTotPot}
154   \end{equation}
155 < $V_{\text{LJ}}$ is the Lennard-Jones potential:
155 > where $\mathbf{r}_{ij}$ is the vector between molecules $i$ and $j$,
156 > and $\boldsymbol{\Omega}$ is the orientation of molecule $i$ or $j$
157 > respectively. $V_{\text{LJ}}$ is the Lennard-Jones potential:
158   \begin{equation}
159   V_{\text{LJ}} =
160          4\epsilon_{ij} \biggl[
# Line 154 | Line 163 | where $r_{ij}$ is the distance between two $ij$ pairs,
163          \biggr]
164   \label{eq:lennardJonesPot}
165   \end{equation}
166 < where $r_{ij}$ is the distance between two $ij$ pairs, $\sigma_{ij}$
166 > here $\sigma_{ij}$
167   scales the length of the interaction, and $\epsilon_{ij}$ scales the
168   energy of the potential. For SSD, $\sigma_{\text{SSD}} = 3.051 \mbox{
169   \AA}$ and $\epsilon_{\text{SSD}} = 0.152\text{ kcal/mol}$.
# Line 169 | Line 178 | where $\mathbf{r}_{ij}$ is the vector between $i$ and
178                  {r^{5}_{ij}} \biggr]
179   \label{eq:dipolePot}
180   \end{equation}
181 < where $\mathbf{r}_{ij}$ is the vector between $i$ and $j$,
182 < $\boldsymbol{\Omega}$ is the orientation of the species, and
183 < $\boldsymbol{\mu}$ is the dipole vector. The SSD model specifies a dipole
184 < magnitude of 2.35~D for water.
181 > where $\boldsymbol{\mu}_i$ is the dipole vector of molecule $i$,
182 > $\boldsymbol{\mu}_i$ takes its orientation from
183 > $\boldsymbol{\Omega}_i$. The SSD model specifies a dipole magnitude of
184 > 2.35~D for water.
185  
186 < The hydrogen bonding of the model is governed by the $V_{\text{sp}}$
186 > The hydrogen bonding is modeled by the $V_{\text{sp}}$
187   term of the potential. Its form is as follows:
188   \begin{equation}
189   V_{\text{sp}}(\mathbf{r}_{i\!j},\boldsymbol{\Omega}_{i},
# Line 208 | Line 217 | polar coordinates of the position of sphere $j$ in the
217   \label{eq:spCorrection}
218   \end{equation}
219   The angles $\theta_{ij}$ and $\phi_{ij}$ are defined by the spherical
220 < polar coordinates of the position of sphere $j$ in the reference frame
221 < fixed on sphere $i$ with the z-axis aligned with the dipole moment.
220 > coordinates of the position of molecule $j$ in the reference frame
221 > fixed on molecule $i$ with the z-axis aligned with the dipole moment.
222   The correction
223   $w^{x}_{ij}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},\boldsymbol{\Omega}_{j})$
224   is needed because
# Line 217 | Line 226 | $s(r_{ij})$ that scales smoothly between 0 and 1. It i
226   vanishes when $\theta_{ij}$ is $0^\circ$ or $180^\circ$.
227  
228   Finally, the sticky potential is scaled by a cutoff function,
229 < $s(r_{ij})$ that scales smoothly between 0 and 1. It is represented
229 > $s(r_{ij})$, that scales smoothly between 0 and 1. It is represented
230   by:
231   \begin{equation}
232   s(r_{ij}) =
# Line 237 | Line 246 | molecules. Therefore, it's predominant interaction is
246   being equal to either 3.35~$\mbox{\AA}$ for $s(r_{ij})$ or
247   4.0~$\mbox{\AA}$ for $s'(r_{ij})$, the sticky potential is only active
248   over an extremely short range, and then only with other SSD
249 < molecules. Therefore, it's predominant interaction is through it's
250 < point dipole and Lennard-Jones sphere.
249 > molecules. Therefore, it's predominant interaction is through the
250 > point dipole and the Lennard-Jones sphere. Of these, only the dipole
251 > interaction can be considered ``long-range''.
252  
253   \subsection{The Phospholipid Model}
254   \label{sec:lipidModel}
255  
256 < %\begin{floatingfigure}{90mm}
257 < %\includegraphics[angle=-90,width=80mm]{lipidModel.epsi}
258 < %\caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ is the bend angle, $\mu$ is the dipole moment of the head group, and n is the chain length. \vspace{5mm}}
259 < %\label{fig:lipidModel}
260 < %\end{floatingfigure}
256 > \begin{figure}
257 > \begin{center}
258 > \includegraphics[angle=-90,width=80mm]{lipidModel.epsi}
259 > \caption{A representation of the lipid model. $\phi$ is the torsion angle, $\theta$ is the bend angle, $\mu$ is the dipole moment of the head group, and n is the chain length. \vspace{5mm}}
260 > \end{center}
261 > \label{fig:lipidModel}
262 > \end{figure}
263  
264   The lipid molecules in our simulations are unified atom models. Figure
265 < \ref{fig:lipidModel} shows a template drawing for one of our
265 > \ref{fig:lipidModel} shows a schematic for one of our
266   lipids. The Head group of the phospholipid is replaced by a single
267   Lennard-Jones sphere with a freely oriented dipole placed at it's
268 < center. The magnitude of it's dipole moment is 20.6 D. The tail atoms
269 < are unified $\text{CH}_2$ and $\text{CH}_3$ atoms and are also modeled
270 < as Lennard-Jones spheres. The total potential for the lipid is
271 < represented by Equation \ref{eq:lipidModelPot}.
268 > center. The magnitude of the dipole moment is 20.6 D, chosen to match
269 > that of DPPC\cite{Cevc87}. The tail atoms are unified $\text{CH}_2$
270 > and $\text{CH}_3$ atoms and are also modeled as Lennard-Jones
271 > spheres. The total potential for the lipid is represented by Equation
272 > \ref{eq:lipidModelPot}.
273  
274   \begin{equation}
275 < V_{\mbox{lipid}} = \overbrace{%
276 <        V_{\text{bend}}(\theta_{ijk}) +%
277 <        V_{\text{tors.}}(\phi_{ijkl})}^{bonded}
278 <        + \overbrace{%
279 <        V_{\text{LJ}}(r_{i\!j}) +
280 <        V_{\text{dp}}(r_{i\!j},\Omega_{i},\Omega_{j})%
268 <        }^{non-bonded}
275 > V_{\text{lipid}} =
276 >        \sum_{i}V_{i}^{\text{internal}}
277 >        + \sum_i \sum_{j>i} \sum_{\text{$\alpha$ in $i$}}
278 >        \sum_{\text{$\beta$ in $j$}}
279 >        V_{\text{LJ}}(r_{\alpha_{i}\beta_{j}})
280 >        +\sum_i\sum_{j>i}V_{\text{dp}}(r_{1_i,1_j},\Omega_{1_i},\Omega_{1_j})
281   \label{eq:lipidModelPot}
282   \end{equation}
283 + where,
284 + \begin{equation}
285 + V_{i}^{\text{internal}} =
286 +        \sum_{\text{bends}}V_{\text{bend}}(\theta_{\alpha\beta\gamma})
287 +        + \sum_{\text{torsions}}V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta})
288 +        + \sum_{\alpha} \sum_{\beta>\alpha}V_{\text{LJ}}(r_{\alpha \beta})
289 + \label{eq:lipidModelPotInternal}
290 + \end{equation}
291  
292   The non-bonded interactions, $V_{\text{LJ}}$ and $V_{\text{dp}}$, are
293   the Lennard-Jones and dipole-dipole interactions respectively. For the
294 < non-bonded potentials, only the bend and the torsional potentials are
294 > bonded potentials, only the bend and the torsional potentials are
295   calculated. The bond potential is not calculated, and the bond lengths
296   are constrained via RATTLE.\cite{leach01:mm} The bend potential is of
297   the form:
298   \begin{equation}
299 < V_{\text{bend}}(\theta_{ijk}) = k_{\theta}\frac{(\theta_{ijk} - \theta_0)^2}{2}
299 > V_{\text{bend}}(\theta_{\alpha\beta\gamma})
300 >        = k_{\theta}\frac{(\theta_{\alpha\beta\gamma} - \theta_0)^2}{2}
301   \label{eq:bendPot}
302   \end{equation}
303   Where $k_{\theta}$ sets the stiffness of the bend potential, and $\theta_0$
304   sets the equilibrium bend angle. The torsion potential was given by:
305   \begin{equation}
306 < V_{\text{tors.}}(\phi_{ijkl})= c_1[1+\cos\phi_{ijkl}]
307 <        + c_2 [1 - \cos(2\phi_{ijkl})] + c_3[1 + \cos(3\phi_{ijkl})]
306 > V_{\text{tors.}}(\phi_{\alpha\beta\gamma\zeta})
307 >        = c_1 [1+\cos\phi_{\alpha\beta\gamma\zeta}]
308 >        + c_2 [1 - \cos(2\phi_{\alpha\beta\gamma\zeta})]
309 >        + c_3 [1 + \cos(3\phi_{\alpha\beta\gamma\zeta})]
310   \label{eq:torsPot}
311   \end{equation}
312   All parameters for bonded and non-bonded potentials in the tail atoms
# Line 300 | Line 323 | Our first simulation is an array of 25 single chained
323   \subsection{Starting Configuration and Parameters}
324   \label{sec:5x5Start}
325  
326 < Our first simulation is an array of 25 single chained lipids in a sea
326 > \begin{figure}
327 > \begin{center}
328 > \includegraphics[width=70mm]{5x5-initial.eps}
329 > \caption{The starting configuration of the 25 lipid system. A box is drawn around the periodic image.}
330 > \end{center}
331 > \label{fig:5x5Start}
332 > \end{figure}
333 >
334 > \begin{figure}
335 > \begin{center}
336 > \includegraphics[width=70mm]{5x5-6.27ns.eps}
337 > \caption{The 25 lipid system at 6.27~ns}
338 > \end{center}
339 > \label{fig:5x5Final}
340 > \end{figure}
341 >
342 > Our first simulation is an array of 25 single chain lipids in a sea
343   of water (Figure \ref{fig:5x5Start}). The total number of water
344   molecules is 1386, giving a final of water concentration of 70\%
345   wt. The simulation box measures 34.5~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$
# Line 312 | Line 351 | Figure \ref{fig:5x5Final} shows a snapshot of the syst
351   \label{sec:5x5Results}
352  
353   Figure \ref{fig:5x5Final} shows a snapshot of the system at
354 < 3.6~ns. Note that the system has spontaneously self assembled into a
354 > 6.27~ns. Note that the system has spontaneously self assembled into a
355   bilayer. Discussion of the length scales of the bilayer will follow in
356   this section. However, it is interesting to note a key qualitative
357   property of the system revealed by this snapshot, the tail chains are
# Line 341 | Line 380 | g_{\gamma}(r) = foobar
380   For the species containing dipoles, a second pair wise distribution
381   function was used, $g_{\gamma}(r)$. It is of the form:
382   \begin{equation}
383 < g_{\gamma}(r) = foobar
383 > g_{\gamma}(r) = \langle \sum_i \sum_{j>i}
384 >        (\cos \gamma_{ij}) \delta(| \mathbf{r} - \mathbf{r}_{ij}|) \rangle
385   \label{eq:gammaofr}
386   \end{equation}
387   Where $\gamma_{ij}$ is the angle between the dipole of atom $j$ with
# Line 355 | Line 395 | means that the two neighbors on the same monolayer hav
395   for the Head groups of the lipids. The first peak in the $g(r)$ at
396   4.03~$\mbox{\AA}$ is the nearest neighbor separation of the heads of
397   two lipids. This corresponds to a maximum in the $g_{\gamma}(r)$ which
398 < means that the two neighbors on the same monolayer have their dipoles
399 < aligned. The broad peak at 6.5~$\mbox{\AA}$ is the inter-bilayer
398 > means that the two neighbors on the same leaf have their dipoles
399 > aligned. The broad peak at 6.5~$\mbox{\AA}$ is the inter-surface
400   spacing. Here, there is a corresponding anti-alignment in the angular
401   correlation. This means that although the dipoles are aligned on the
402   same monolayer, the dipoles will orient themselves to be anti-aligned
# Line 386 | Line 426 | The second simulation consists of 50 single chained li
426   \subsection{Starting Configuration and Parameters}
427   \label{sec:r50Start}
428  
429 + \begin{figure}
430 + \begin{center}
431 + \includegraphics[width=70mm]{r50-initial.eps}
432 + \caption{The starting configuration of the 50 lipid system.}
433 + \end{center}
434 + \label{fig:r50Start}
435 + \end{figure}
436 +
437 + \begin{figure}
438 + \begin{center}
439 + \includegraphics[width=70mm]{r50-2.21ns.eps}
440 + \caption{The 50 lipid system at 2.21~ns}
441 + \end{center}
442 + \label{fig:r50Final}
443 + \end{figure}
444 +
445   The second simulation consists of 50 single chained lipid molecules
446   embedded in a sea of 1384 SSD waters (54\% wt.). The lipids in this
447   simulation were started with random orientation and location (Figure
448 < \ref{fig:r50Start} ) The simulation box measured 34.5~$\mbox{\AA}$ x
449 < 39.4~$\mbox{\AA}$ x 39.4~$\mbox{\AA}$ with periodic boundary conditions
448 > \ref{fig:r50Start} ) The simulation box measured 26.6~$\mbox{\AA}$ x
449 > 26.6~$\mbox{\AA}$ x 108.4~$\mbox{\AA}$ with periodic boundary conditions
450   imposed. The simulation was run in the NVE ensemble with an average
451   temperature of 300~K.
452  
# Line 406 | Line 462 | the high degree of similarity between the correlation
462   Figures \ref{fig:r50HHCorr}, \ref{fig:r50CCg}, and \ref{fig:r50} are
463   the same correlation functions for the random 50 simulation as for the
464   previous simulation of 25 lipids. What is most interesting to note, is
465 < the high degree of similarity between the correlation functions for
466 < each system. Even though the 25 lipid simulation formed a bilayer and
467 < the random 50 simulation is still in the micelle stage, both have a
468 < inter surface spacing of 6.5~$\mbox{\AA}$ with the same characteristic
469 < anti-alignment between surfaces. Not as surprising, is the consistency
470 < of the closest packing statistics between systems. Namely, a head-head
471 < closest approach distance of 4~$\mbox{\AA}$, and similar findings for
472 < the chain-chain and head-water distributions as in the 25 lipid
473 < system.
465 > the high degree of similarity between the correlation functions
466 > between systems. Even though the 25 lipid simulation formed a bilayer
467 > and the random 50 simulation is still in the micelle stage, both have
468 > an inter-surface spacing of 6.5~$\mbox{\AA}$ with the same
469 > characteristic anti-alignment between surfaces. Not as surprising, is
470 > the consistency of the closest packing statistics between
471 > systems. Namely, a head-head closest approach distance of
472 > 4~$\mbox{\AA}$, and similar findings for the chain-chain and
473 > head-water distributions as in the 25 lipid system.
474  
475   \section{Future Directions}
476  
477   Current simulations indicate that our model is a feasible one, however
478 < improvements will need to be made to allow the system to simulate an
479 < isobaric-isothermal ensemble. This will allow the system to relax to
480 < an equilibrium configuration at room temperature and pressure allowing
481 < us to compare our model to experimental results. Also, we plan to
482 < parallelize the code for an even greater speedup. This will allow us
483 < to simulate the size systems needed to examine phenomena such as the
484 < ripple phase and drug molecule diffusion
478 > improvements will need to be made to allow the system to be simulated
479 > in the isobaric-isothermal ensemble. This will relax the system to an
480 > equilibrium configuration at room temperature and pressure allowing us
481 > to compare our model to experimental results. Also, we are in the
482 > process of parallelizeing the code for an even greater speedup. This
483 > will allow us to simulate the size systems needed to examine phenomena
484 > such as the ripple phase and drug molecule diffusion
485  
486 < Once the work has completed on the simulation engine, we would then
487 < like to use it to explore phase diagram for our model. By
486 > Once the work has been completed on the simulation engine, we will
487 > then use it to explore the phase diagram for our model. By
488   characterizing how our model parameters affect the bilayer properties,
489 < we hope to tailor our model to more closely match real biological
490 < molecules. With this information, we then hope to incorporate
489 > we will tailor our model to more closely match real biological
490 > molecules. With this information, we will then incorporate
491   biologically relevant molecules into the system and observe their
492   transport properties across the membrane.
493  
# Line 440 | Line 496 | Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, a
496   I would like to thank Dr. J.Daniel Gezelter for his guidance on this
497   project. I would also like to acknowledge the following for their help
498   and discussions during this project: Christopher Fennell, Charles
499 < Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, and Dan Combest.
499 > Vardeman, Teng Lin, Megan Sprague, Patrick Conforti, and Dan
500 > Combest. Funding for this project came from the National Science
501 > Foundation.
502  
503   \pagebreak
504   \bibliographystyle{achemso}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines