ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/mattDisertation/lipid.tex
Revision: 1083
Committed: Fri Mar 5 00:07:19 2004 UTC (20 years, 6 months ago) by mmeineke
Content type: application/x-tex
File size: 19344 byte(s)
Log Message:
added several figures to the lipid paper, changed the ndthesis class around a little. some small fixes in oopse.

File Contents

# User Rev Content
1 mmeineke 971
2    
3     \chapter{\label{chapt:lipid}Phospholipid Simulations}
4    
5     \section{\label{lipidSec:Intro}Introduction}
6    
7 mmeineke 1083 In the past 10 years, increasing computer speeds have allowed for the
8     atomistic simulation of phospholipid bilayers for increasingly
9     relevant lenghths of time. These simulations have ranged from
10     simulation of the gel phase ($L_{\beta}$) of
11 mmeineke 1061 dipalmitoylphosphatidylcholine (DPPC),\cite{lindahl00} to the
12 mmeineke 1001 spontaneous aggregation of DPPC molecules into fluid phase
13 mmeineke 1061 ($L_{\alpha}$) bilayers.\cite{marrink01} With the exception of a few
14     ambitious
15     simulations,\cite{marrink01:undulation,marrink:2002,lindahl00} most
16 mmeineke 1083 investigations are limited to a range of 64 to 256
17 mmeineke 1061 phospholipids.\cite{lindahl00,sum:2003,venable00,gomez:2003,smondyrev:1999,marrink01}
18 mmeineke 1083 The expense of the force calculations involved when performing these
19     simulations limits the system size. To properly hydrate a bilayer, one
20 mmeineke 1001 typically needs 25 water molecules for every lipid, bringing the total
21     number of atoms simulated to roughly 8,000 for a system of 64 DPPC
22 mmeineke 1061 molecules. Added to the difficulty is the electrostatic nature of the
23 mmeineke 1083 phospholipid head groups and water, requiring either the
24     computationally expensive Ewald sum or the faster, particle mesh Ewald
25     sum.\cite{nina:2002,norberg:2000,patra:2003} These factors all limit
26     the system size and time scales of bilayer simulations.
27 mmeineke 1001
28     Unfortunately, much of biological interest happens on time and length
29 mmeineke 1083 scales well beyond the range of current simulation technology. One
30     such example is the observance of a ripple phase
31     ($P_{\beta^{\prime}}$) between the $L_{\beta}$ and $L_{\alpha}$ phases
32     of certain phospholipid bilayers.\cite{katsaras00,sengupta00} These
33     ripples are shown to have periodicity on the order of
34     100-200~$\mbox{\AA}$. A simulation on this length scale would have
35     approximately 1,300 lipid molecules with an additional 25 water
36     molecules per lipid to fully solvate the bilayer. A simulation of this
37     size is impractical with current atomistic models.
38 mmeineke 1001
39 mmeineke 1083 The time and length scale limitations are most striking in transport
40     phenomena. Due to the fluid-like properties of a lipid membrane, not
41     all diffusion across the membrane happens at pores. Some molecules of
42     interest may incorporate themselves directly into the membrane. Once
43     here, they may possess an appreciable waiting time (on the order of
44     10's to 100's of nanoseconds) within the bilayer. Such long simulation
45     times are nearly impossible to obtain when integrating the system with
46     atomistic detail.
47 mmeineke 1001
48 mmeineke 1083 To address these issues, several schemes have been proposed. One
49 mmeineke 1061 approach by Goetz and Liposky\cite{goetz98} is to model the entire
50 mmeineke 1001 system as Lennard-Jones spheres. Phospholipids are represented by
51     chains of beads with the top most beads identified as the head
52     atoms. Polar and non-polar interactions are mimicked through
53     attractive and soft-repulsive potentials respectively. A similar
54 mmeineke 1061 model proposed by Marrinck \emph{et. al}.\cite{marrink04}~uses a
55 mmeineke 1001 similar technique for modeling polar and non-polar interactions with
56     Lennard-Jones spheres. However, they also include charges on the head
57     group spheres to mimic the electrostatic interactions of the
58     bilayer. While the solvent spheres are kept charge-neutral and
59     interact with the bilayer solely through an attractive Lennard-Jones
60     potential.
61    
62     The model used in this investigation adds more information to the
63 mmeineke 1026 interactions than the previous two models, while still balancing the
64     need for simplifications over atomistic detail. The model uses
65     Lennard-Jones spheres for the head and tail groups of the
66 mmeineke 1061 phospholipids, allowing for the ability to scale the parameters to
67 mmeineke 1026 reflect various sized chain configurations while keeping the number of
68     interactions small. What sets this model apart, however, is the use
69 mmeineke 1061 of dipoles to represent the electrostatic nature of the
70 mmeineke 1026 phospholipids. The dipole electrostatic interaction is shorter range
71 mmeineke 1083 than Coulombic ($\frac{1}{r^3}$ versus $\frac{1}{r}$), and eliminates
72     the need for a costly Ewald sum.
73 mmeineke 1001
74 mmeineke 1026 Another key feature of this model, is the use of a dipolar water model
75 mmeineke 1061 to represent the solvent. The soft sticky dipole ({\sc ssd}) water
76     \cite{liu96:new_model} relies on the dipole for long range electrostatic
77     effects, but also contains a short range correction for hydrogen
78     bonding. In this way the systems in this research mimic the entropic
79     contribution to the hydrophobic effect due to hydrogen-bond network
80 mmeineke 1083 deformation around a non-polar entity, \emph{i.e.}~the phospholipid
81     molecules.
82 mmeineke 1026
83     The following is an outline of this chapter.
84 mmeineke 1083 Sec.~\ref{lipidSec:Methods} is an introduction to the lipid model used
85     in these simulations, as well as clarification about the water model
86     and integration techniques. The various simulations explored in this
87     research are outlined in
88     Sec.~\ref{lipidSec:ExpSetup}. Sec.~\ref{lipidSec:resultsDis} gives a
89     summary and interpretation of the results. Finally, the conclusions
90     of this chapter are presented in Sec.~\ref{lipidSec:Conclusion}.
91 mmeineke 1026
92 mmeineke 971 \section{\label{lipidSec:Methods}Methods}
93    
94 mmeineke 1026 \subsection{\label{lipidSec:lipidModel}The Lipid Model}
95    
96 mmeineke 971 \begin{figure}
97 mmeineke 1061 \centering
98     \includegraphics[width=\linewidth]{twoChainFig.eps}
99     \caption[The two chained lipid model]{Schematic diagram of the double chain phospholipid model. The head group (in red) has a point dipole, $\boldsymbol{\mu}$, located at its center of mass. The two chains are eight methylene groups in length.}
100 mmeineke 971 \label{lipidFig:lipidModel}
101     \end{figure}
102    
103     The phospholipid model used in these simulations is based on the
104     design illustrated in Fig.~\ref{lipidFig:lipidModel}. The head group
105     of the phospholipid is replaced by a single Lennard-Jones sphere of
106 mmeineke 1061 diameter $\sigma_{\text{head}}$, with $\epsilon_{\text{head}}$ scaling
107     the well depth of its van der Walls interaction. This sphere also
108     contains a single dipole of magnitude $|\boldsymbol{\mu}|$, where
109     $|\boldsymbol{\mu}|$ can be varied to mimic the charge separation of a
110 mmeineke 971 given phospholipid head group. The atoms of the tail region are
111 mmeineke 1083 modeled by beads representing multiple methyl groups. They are free
112     of partial charges or dipoles, and contain only Lennard-Jones
113     interaction sites at their centers of mass. As with the head groups,
114     their potentials can be scaled by $\sigma_{\text{tail}}$ and
115     $\epsilon_{\text{tail}}$.
116 mmeineke 971
117     The long range interactions between lipids are given by the following:
118     \begin{equation}
119 mmeineke 1061 V_{\text{LJ}}(r_{ij}) =
120     4\epsilon_{ij} \biggl[
121     \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{12}
122     - \biggl(\frac{\sigma_{ij}}{r_{ij}}\biggr)^{6}
123     \biggr]
124 mmeineke 971 \label{lipidEq:LJpot}
125     \end{equation}
126     and
127     \begin{equation}
128 mmeineke 1061 V_{\text{dipole}}(\mathbf{r}_{ij},\boldsymbol{\Omega}_{i},
129     \boldsymbol{\Omega}_{j}) = \frac{|\mu_i||\mu_j|}{4\pi\epsilon_{0}r_{ij}^{3}} \biggl[
130     \boldsymbol{\hat{u}}_{i} \cdot \boldsymbol{\hat{u}}_{j}
131     -
132 mmeineke 1083 3(\boldsymbol{\hat{u}}_i \cdot \mathbf{\hat{r}}_{ij}) %
133     (\boldsymbol{\hat{u}}_j \cdot \mathbf{\hat{r}}_{ij} \biggr]
134 mmeineke 971 \label{lipidEq:dipolePot}
135     \end{equation}
136     Where $V_{\text{LJ}}$ is the Lennard-Jones potential and
137     $V_{\text{dipole}}$ is the dipole-dipole potential. As previously
138     stated, $\sigma_{ij}$ and $\epsilon_{ij}$ are the Lennard-Jones
139     parameters which scale the length and depth of the interaction
140     respectively, and $r_{ij}$ is the distance between beads $i$ and $j$.
141     In $V_{\text{dipole}}$, $\mathbf{r}_{ij}$ is the vector starting at
142 mmeineke 1083 bead $i$ and pointing towards bead $j$. Vectors $\mathbf{\Omega}_i$
143 mmeineke 971 and $\mathbf{\Omega}_j$ are the orientational degrees of freedom for
144     beads $i$ and $j$. $|\mu_i|$ is the magnitude of the dipole moment of
145     $i$, and $\boldsymbol{\hat{u}}_i$ is the standard unit orientation
146 mmeineke 1083 vector rotated with euler angles: $\boldsymbol{\Omega}_i$.
147 mmeineke 971
148 mmeineke 1061 The model also allows for the bonded interactions bends, and torsions.
149     The bond between two beads on a chain is of fixed length, and is
150     maintained according to the {\sc rattle} algorithm.\cite{andersen83}
151 mmeineke 971 The bends are subject to a harmonic potential:
152     \begin{equation}
153 mmeineke 1061 V_{\text{bend}}(\theta) = k_{\theta}( \theta - \theta_0 )^2
154 mmeineke 971 \label{lipidEq:bendPot}
155     \end{equation}
156 mmeineke 1061 where $k_{\theta}$ scales the strength of the harmonic well, and
157     $\theta$ is the angle between bond vectors
158     (Fig.~\ref{lipidFig:lipidModel}). In addition, we have placed a
159     ``ghost'' bend on the phospholipid head. The ghost bend adds a
160     potential to keep the dipole pointed along the bilayer surface, where
161 mmeineke 1083 $\theta$ is now the angle the dipole makes with respect to the {\sc
162     head}-$\text{{\sc ch}}_2$ bond vector
163     (Fig.~\ref{lipidFig:ghostBend}). The torsion potential is given by:
164 mmeineke 971 \begin{equation}
165 mmeineke 1061 V_{\text{torsion}}(\phi) =
166     k_3 \cos^3 \phi + k_2 \cos^2 \phi + k_1 \cos \phi + k_0
167 mmeineke 971 \label{lipidEq:torsionPot}
168     \end{equation}
169     Here, the parameters $k_0$, $k_1$, $k_2$, and $k_3$ fit the cosine
170     power series to the desired torsion potential surface, and $\phi$ is
171 mmeineke 1061 the angle the two end atoms have rotated about the middle bond
172     (Fig.:\ref{lipidFig:lipidModel}). Long range interactions such as the
173     Lennard-Jones potential are excluded for atom pairs involved in the
174     same bond, bend, or torsion. However, internal interactions not
175 mmeineke 971 directly involved in a bonded pair are calculated.
176    
177 mmeineke 1083 \begin{figure}
178     \centering
179     \includegraphics[width=\linewidth]{ghostBendFig.eps}
180     \caption[Depiction of the ``ghost'' bend]{The ``ghost'' bend is a bend potential added to constrain the motion of the dipole on the {\sc head} group. The potential follows Eq.~\ref{lipidEq:bendPot} where $\theta$ is now the angle that the dipole makes with the {\sc head}-$\text{{\sc ch}}_2$ bond vector.}
181     \label{lipidFig:ghostBend}
182     \end{figure}
183    
184 mmeineke 1026 All simulations presented here use a two chained lipid as pictured in
185 mmeineke 1061 Fig.~\ref{lipidFig:lipidModel}. The chains are both eight beads long,
186 mmeineke 1026 and their mass and Lennard Jones parameters are summarized in
187     Table~\ref{lipidTable:tcLJParams}. The magnitude of the dipole moment
188     for the head bead is 10.6~Debye, and the bend and torsion parameters
189 mmeineke 1061 are summarized in Table~\ref{lipidTable:tcBendParams} and
190     \ref{lipidTable:tcTorsionParams}.
191 mmeineke 971
192 mmeineke 1061 \begin{table}
193     \caption{The Lennard Jones Parameters for the two chain phospholipids.}
194     \label{lipidTable:tcLJParams}
195     \begin{center}
196     \begin{tabular}{|l|c|c|c|}
197     \hline
198     & mass (amu) & $\sigma$($\mbox{\AA}$) & $\epsilon$ (kcal/mol) \\ \hline
199     {\sc head} & 72 & 4.0 & 0.185 \\ \hline
200     {\sc ch}\cite{Siepmann1998} & 13.02 & 4.0 & 0.0189 \\ \hline
201     $\text{{\sc ch}}_2$\cite{Siepmann1998} & 14.03 & 3.95 & 0.18 \\ \hline
202     $\text{{\sc ch}}_3$\cite{Siepmann1998} & 15.04 & 3.75 & 0.25 \\ \hline
203     {\sc ssd}\cite{liu96:new_model} & 18.03 & 3.051 & 0.152 \\ \hline
204     \end{tabular}
205     \end{center}
206     \end{table}
207 mmeineke 1026
208 mmeineke 1061 \begin{table}
209     \caption[Bend Parameters for the two chain phospholipids]{Bend Parameters for the two chain phospholipids. All alkane parameters are based off of those from TraPPE.\cite{Siepmann1998}}
210     \label{lipidTable:tcBendParams}
211     \begin{center}
212     \begin{tabular}{|l|c|c|}
213     \hline
214     & $k_{\theta}$ ( kcal/($\text{mol deg}^2$) ) & $\theta_0$ ( deg ) \\ \hline
215     {\sc ghost}-{\sc head}-$\text{{\sc ch}}_2$ & 0.00177 & 129.78 \\ \hline
216     $x$-{\sc ch}-$y$ & 58.84 & 112.0 \\ \hline
217     $x$-$\text{{\sc ch}}_2$-$y$ & 58.84 & 114.0 \\ \hline
218     \end{tabular}
219     \end{center}
220     \end{table}
221    
222     \begin{table}
223     \caption[Torsion Parameters for the two chain phospholipids]{Torsion Parameters for the two chain phospholipids. Alkane parameters based on TraPPE.\cite{Siepmann1998}}
224     \label{lipidTable:tcTorsionParams}
225     \begin{center}
226     \begin{tabular}{|l|c|c|c|c|}
227     \hline
228     All are in kcal/mol $\rightarrow$ & $k_3$ & $k_2$ & $k_1$ & $k_0$ \\ \hline
229     $x$-{\sc ch}-$y$-$z$ & 3.3254 & -0.4215 & -1.686 & 1.1661 \\ \hline
230     $x$-$\text{{\sc ch}}_2$-$\text{{\sc ch}}_2$-$y$ & 5.9602 & -0.568 & -3.802 & 2.1586 \\ \hline
231     \end{tabular}
232     \end{center}
233     \end{table}
234    
235    
236     \section{\label{lipidSec:furtherMethod}Further Methodology}
237    
238 mmeineke 1026 As mentioned previously, the water model used throughout these
239 mmeineke 1061 simulations was the {\sc ssd} model of
240     Ichiye.\cite{liu96:new_model,liu96:monte_carlo,chandra99:ssd_md} A
241     discussion of the model can be found in Sec.~\ref{oopseSec:SSD}. As
242     for the integration of the equations of motion, all simulations were
243     performed in an orthorhombic periodic box with a thermostat on
244     velocities, and an independent barostat on each Cartesian axis $x$,
245     $y$, and $z$. This is the $\text{NPT}_{xyz}$. ensemble described in
246     Sec.~\ref{oopseSec:Ensembles}.
247 mmeineke 1026
248    
249     \subsection{\label{lipidSec:ExpSetup}Experimental Setup}
250    
251     Two main starting configuration classes were used in this research:
252     random and ordered bilayers. The ordered bilayer starting
253     configurations were all started from an equilibrated bilayer at
254     300~K. The original configuration for the first 300~K run was
255     assembled by placing the phospholipids centers of mass on a planar
256     hexagonal lattice. The lipids were oriented with their long axis
257     perpendicular to the plane. The second leaf simply mirrored the first
258     leaf, and the appropriate number of waters were then added above and
259     below the bilayer.
260    
261     The random configurations took more work to generate. To begin, a
262     test lipid was placed in a simulation box already containing water at
263     the intended density. The waters were then tested for overlap with
264     the lipid using a 5.0~$\mbox{\AA}$ buffer distance. This gave an
265     estimate for the number of waters each lipid would displace in a
266     simulation box. A target number of waters was then defined which
267     included the number of waters each lipid would displace, the number of
268 mmeineke 1083 waters desired to solvate each lipid, and a factor to pad the
269     initial box with a few extra water molecules.
270 mmeineke 1026
271     Next, a cubic simulation box was created that contained at least the
272     target number of waters in an FCC lattice (the lattice was for ease of
273     placement). What followed was a RSA simulation similar to those of
274     Chapt.~\ref{chapt:RSA}. The lipids were sequentially given a random
275     position and orientation within the box. If a lipid's position caused
276 mmeineke 1083 atomic overlap with any previously placed lipid, its position and
277     orientation were rejected, and a new random placement site was
278 mmeineke 1026 attempted. The RSA simulation proceeded until all phospholipids had
279 mmeineke 1083 been adsorbed. After placement of all lipid molecules, water
280     molecules with locations that overlapped with the atomic coordinates
281     of the lipids were removed.
282 mmeineke 1026
283 mmeineke 1083 Finally, water molecules were removed at random until the desired
284     number of waters per lipid was reached. The typical low final density
285     for these initial configurations was not a problem, as the box shrinks
286     to an appropriate size within the first 50~ps of a simulation in the
287     $\text{NPT}_{xyz}$ ensemble.
288 mmeineke 1026
289 mmeineke 1083 \subsection{\label{lipidSec:Configs}Configurations}
290 mmeineke 1026
291 mmeineke 1083 The first class of simulations were started from ordered
292     bilayers. They were all configurations consisting of 60 lipid
293     molecules with 30 lipids on each leaf, and were hydrated with 1620
294     {\sc ssd} molecules. The original configuration was assembled
295     according to Sec.~\ref{lipidSec:ExpSetup} and simulated for a length
296     of 10~ns at 300~K. The other temperature runs were started from a
297     frame 7~ns into the 300~K simulation. Their temperatures were reset
298     with the thermostating algorithm in the $\text{NPT}_{xyz}$
299     integrator. All of the temperature variants were also run for 10~ns,
300     with only the last 5~ns being used for accumulation of statistics.
301    
302     The second class of simulations were two configurations started from
303     randomly dispersed lipids in a ``gas'' of water. The first
304     ($\text{R}_{\text{I}}$) was a simulation containing 72 lipids with
305     1800 {\sc ssd} molecules simulated at 300~K. The second
306     ($\text{R}_{\text{II}}$) was 90 lipids with 1350 {\sc ssd} molecules
307     simulated at 350~K. Both simulations were integrated for more than
308     20~ns, and illustrate the spontaneous aggregation of the lipid model
309     into phospholipid macrostructures: $\text{R}_{\text{I}}$ into a
310     bilayer, and $\text{R}_{\text{II}}$ into a inverted rod.
311    
312     \section{\label{lipidSec:resultsDis}Results and Discussion}
313    
314     \subsection{\label{lipidSec:scd}$\text{S}_{\text{{\sc cd}}}$ order parameters}
315    
316     The $\text{S}_{\text{{\sc cd}}}$ order parameter is often reported in
317     the experimental charecterizations of phospholipids. It is obtained
318     through deuterium NMR, and measures the ordering of the carbon
319     deuterium bond in relation to the bilayer normal at various points
320     along the chains. In our model, there are no explicit hydrogens, but
321     the order parameter can be written in terms of the carbon ordering at
322     each point in the chain:\cite{egberts88}
323     \begin{equation}
324     S_{\text{{\sc cd}}} = \frac{2}{3}S_{xx} + \frac{1}{3}S_{yy}
325     \label{lipidEq:scd1}
326     \end{equation}
327     Where $S_{ij}$ is given by:
328     \begin{equation}
329     S_{ij} = \frac{1}{2}\Bigl<(3\cos\Theta_i\cos\Theta_j - \delta_{ij})\Bigr>
330     \label{lipidEq:scd2}
331     \end{equation}
332     Here, $\Theta_i$ is the angle the $i$th carbon atom frame axis makes
333     with the bilayer normal. The brackets denote an average over time and
334     molecules. The carbon atom axes are defined:
335     $\mathbf{\hat{z}}\rightarrow$ vector from $C_{n-1}$ to $C_{n+1}$;
336     $\mathbf{\hat{y}}\rightarrow$ vector that is perpindicular to $z$ and
337     in the plane through $C_{n-1}$, $C_{n}$, and $C_{n+1}$;
338     $\mathbf{\hat{x}}\rightarrow$ vector perpindicular to
339     $\mathbf{\hat{y}}$ and $\mathbf{\hat{z}}$.
340    
341     The order parameter has a range of $[1,-\frac{1}{2}]$. A value of 1
342     implies full order aligned to the bilayer axis, 0 implies full
343     disorder, and $-\frac{1}{2}$ implies full order perpindicular to the
344     bilayer axis. The {\sc cd} bond vector for carbons near the head group
345     are usually ordered perpindicular to the bilayer normal, with tails
346     farther away tending toward disorder. This makes the order paramter
347     negative for most carbons, and as such $|S_{\text{{\sc cd}}}|$ is more
348     commonly reported than $S_{\text{{\sc cd}}}$.
349    
350    
351    
352    
353     \begin{figure}
354     \centering
355     \includegraphics[width=\linewidth]{scdFig.eps}
356     \caption[$\text{S}_{\text{{\sc cd}}}$ order parameter for our model]{A comparison of $|\text{S}_{\text{{\sc cd}}}|$ between our model (blue) and DMPC\cite{petrache00} (black) near 300~K.}
357     \label{lipidFig:scdFig}
358     \end{figure}
359    
360    
361     \begin{figure}
362     \centering
363     \includegraphics[width=\linewidth]{densityProfile.eps}
364     \caption[The density profile of the lipid bilayers]{The density profile of the lipid bilayers along the bilayer normal. The black lines are the {\sc head} atoms, red lines are the {\sc ch} atoms, green lines are the $\text{{\sc ch}}_2$ atoms, blue lines are the $\text{{\sc ch}}_3$ atoms, and the magenta lines are the {\sc ssd} atoms.}
365     \label{lipidFig:densityProfile}
366     \end{figure}
367    
368    
369    
370     \begin{figure}
371     \centering
372     \includegraphics[width=\linewidth]{diffusionFig.eps}
373     \caption[The lateral difusion constants versus temperature]{The lateral diffusion constants for the bilayers as a function of temperature.}
374     \label{lipidFig:diffusionFig}
375     \end{figure}
376    
377     \begin{table}
378     \caption[Structural properties of the bilayers]{Bilayer Structural properties as a function of temperature.}
379     \begin{center}
380     \begin{tabular}{|c|c|c|c|c|}
381     \hline
382     Temperature (K) & $<L_{\perp}>$ ($\mbox{\AA}$) & %
383     $<A_{\parallel}>$ ($\mbox{\AA}^2$) & $<P_2>_{\text{Lipid}}$ & %
384     $<P_2>_{\text{{\sc head}}}$ \\ \hline
385     270 & 18.1 & 58.1 & 0.253 & 0.494 \\ \hline
386     275 & 17.2 & 56.7 & 0.295 & 0.514 \\ \hline
387     277 & 16.9 & 58.0 & 0.301 & 0.541 \\ \hline
388     280 & 17.4 & 58.0 & 0.274 & 0.488 \\ \hline
389     285 & 16.9 & 57.6 & 0.270 & 0.616 \\ \hline
390     290 & 17.0 & 57.6 & 0.263 & 0.534 \\ \hline
391     293 & 17.5 & 58.0 & 0.227 & 0.643 \\ \hline
392     300 & 16.9 & 57.6 & 0.315 & 0.536 \\ \hline
393     \end{tabular}
394     \end{center}
395     \end{table}