ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3810
Committed: Tue Dec 11 18:22:20 2012 UTC (11 years, 6 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 22518 byte(s)
Log Message:
CO parameters section, dipole moment, quadrupole moment, and experimental values obtained

File Contents

# User Rev Content
1 gezelter 3808 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4 jmichalk 3802 \usepackage{setspace}
5 gezelter 3808 \usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{tabularx}
8     \usepackage{graphicx}
9     \usepackage{multirow}
10     %\usepackage{booktabs}
11     %\usepackage{bibentry}
12     %\usepackage{mathrsfs}
13     %\usepackage[ref]{overcite}
14     \usepackage[square, comma, sort&compress]{natbib}
15     \usepackage{url}
16     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19 jmichalk 3802
20 gezelter 3808 % double space list of tables and figures
21     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22     \setlength{\abovecaptionskip}{20 pt}
23     \setlength{\belowcaptionskip}{30 pt}
24    
25     %\renewcommand\citemid{\ } % no comma in optional reference note
26     \bibpunct{[}{]}{,}{n}{}{;}
27     \bibliographystyle{achemso}
28    
29     \begin{document}
30    
31    
32 jmichalk 3802 %%
33     %Introduction
34     % Experimental observations
35     % Previous work on Pt, CO, etc.
36     %
37     %Simulation Methodology
38     % FF (fits and parameters)
39     % MD (setup, equilibration, collection)
40     %
41     % Analysis of trajectories!!!
42     %Discussion
43     % CO preferences for specific locales
44     % CO-CO interactions
45     % Differences between Au & Pt
46     % Causes of 2_layer reordering in Pt
47     %Summary
48     %%
49    
50     %Title
51 gezelter 3808 \title{Investigation of the Pt and Au 557 Surface Reconstructions
52     under a CO Atmosphere}
53     \author{Joseph R. Michalka, Patrick W. MacIntyre and J. Daniel
54     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
55     Department of Chemistry and Biochemistry,\\
56     University of Notre Dame\\
57     Notre Dame, Indiana 46556}
58 jmichalk 3802 %Date
59     \date{Dec 15, 2012}
60     %authors
61 gezelter 3808
62 jmichalk 3802 % make the title
63     \maketitle
64    
65 gezelter 3808 \begin{doublespace}
66 jmichalk 3802
67 gezelter 3808 \begin{abstract}
68 jmichalk 3802
69 gezelter 3808 \end{abstract}
70 jmichalk 3802
71 gezelter 3808 \newpage
72    
73    
74 jmichalk 3802 \section{Introduction}
75     % Importance: catalytically active metals are important
76     % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
77     % Sub: Designing catalysis is the future, and will play an important role in numerous processes (ones that are currently seen to be impractical, or at least inefficient)
78     % Theory can explore temperatures and pressures which are difficult to work with in experiments
79     % Sub: Also, easier to observe what is going on and provide reasons and explanations
80     %
81    
82 gezelter 3808 Industrial catalysts usually consist of small particles exposing
83     different atomic terminations that exhibit a high concentration of
84     step, kink sites, and vacancies at the edges of the facets. These
85 jmichalk 3810 sites are thought to be the locations of catalytic
86 gezelter 3808 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
87     significant evidence to demonstrate that solid surfaces are often
88     structurally, compositionally, and chemically {\it modified} by
89     reactants under operating conditions.\cite{Tao2008,Tao:2010,Tao2011}
90     The coupling between surface oxidation state and catalytic activity
91     for CO oxidation on Pt, for instance, is widely
92     documented.\cite{Ertl08,Hendriksen:2002} Despite the well-documented
93     role of these effects on reactivity, the ability to capture or predict
94     them in atomistic models is currently somewhat limited. While these
95     effects are perhaps unsurprising on the highly disperse, multi-faceted
96     nanoscale particles that characterize industrial catalysts, they are
97     manifest even on ordered, well-defined surfaces. The Pt(557) surface,
98     for example, exhibits substantial and reversible restructuring under
99     exposure to moderate pressures of carbon monoxide.\cite{Tao:2010}
100 jmichalk 3802
101 gezelter 3808 This work is part of an ongoing effort to understand the causes,
102     mechanisms and timescales for surface restructuring using molecular
103     simulation methods. Since the dynamics of the process is of
104     particular interest, we utilize classical molecular dynamic methods
105     with force fields that represent a compromise between chemical
106     accuracy and the computational efficiency necessary to observe the
107     process of interest.
108    
109 jmichalk 3806 High-index surfaces of catalytically active metals are an important area of exploration because they are typically more reactive than an ideal surface of the same metal. The greater number of low-coordinated surface atoms is believed responsible for this increased reactivity \cite{}. Additionally, the activity and specificity of many metals towards certain chemical processes has been shown to strongly depend on the structure of the surface \cite{}. Prior work has also shown that reaction conditions, such as high pressures and high temperatures are able to cause reconstructions of the metallic surface, either through changing the displayed surface facets or by changing the number and types of high-index sites available for reactions \cite{doi:10.1126/science.1197461,doi:10.1021/nn3015322, doi:10.1021/jp302379x}. A greater understanding of these high-index surfaces and the restructuring processes they undergo is needed as a prerequisite for more intelligent catalyst design. While current experimental work has started exploring systems at \emph{in situ} conditions, for a long time such experiments were limited to ideal surfaces in high vacuum. New techniques, such as ambient pressure XPS (AP-XPS) \cite{}, high-pressure XPS (HP-XPS) \cite{}, high-pressure STM \cite{}, environmental transmission electron microscopy (E-TEM) \cite{} and many others, are providing clearer pictures of the processes that are occurring on metal surfaces under these conditions. Nevertheless, all of these techniques still have limitations, especially in observing what is occurring at an atomic level. Theoretical models and simulations in combination with experiment have proven their worth in explaining the underlying reasons for some of these reconstructions \cite{}.
110 jmichalk 3802 \\
111 jmichalk 3806 By examining two different metal-CO systems the effect that the metal and the metal-CO interaction plays can be elucidated. Our first system is composed of platinum and CO and has been the subject of many experimental and theoretical studies primarily because of platinum's strong reactivity toward CO oxidation. The focus has primarily been on adsorption energies, preferred adsorption sites, and catalytic activities. The second system we examined is composed of gold and CO. The gold-CO interaction is much weaker than the platinum-CO interaction and it seems likely that this difference in attraction would lead to differences in any potential surface reconstructions.
112 jmichalk 3802 %It has also been a good test for new quantum methods because of the difficulty with modeling the preference CO has for the atop binding site \cite{doi:10.1021/jp002302t}.
113 jmichalk 3806 %Now that dynamic surface events are known to play a role in many catalytic systems, additional research is being done to more closely examine many systems. Recent work by Tao et al. \cite{doi:10.1126/science.1182122} shows that a high-index platinum surface undergoes surface reconstructions when exposed to a small amount of CO, $\sim$~1 torr. Unexpectedly, the reconstruction was metastable and reverted upon removal of the CO. Work by McCarthy et al. \cite{doi:10.1021/jp302379x} examined temperature programmed desorption's of CO from various platinum samples and saw that species which had large amounts of low-coordinated surface atoms, highly sputtered surfaces or small nano particles, developed a higher temperature desorption peak, suggesting that binding of CO to the platinum surface is strongly dependent on local geometry.
114 jmichalk 3802
115    
116    
117    
118    
119     \section{Simulation Methods}
120 gezelter 3808 The challenge in modeling any solid/gas interface problem is the
121     development of a sufficiently general yet computationally tractable
122     model of the chemical interactions between the surface atoms and
123     adsorbates. Since the interfaces involved are quite large (10$^3$ -
124     10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
125     molecular dynamics
126     (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
127     methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
128     mechanical potential energy surfaces remain out of reach.
129     Additionally, the ``bonds'' between metal atoms at a surface are
130     typically not well represented in terms of classical pairwise
131     interactions in the same way that bonds in a molecular material are,
132     nor are they captured by simple non-directional interactions like the
133     Coulomb potential. For this work, we have been using classical
134     molecular dynamics with potential energy surfaces that are
135     specifically tuned for transition metals. In particular, we use the
136     EAM potential for Au-Au and Pt-Pt interactions, while modeling the CO
137     using a model developed by Straub and Karplus for studying
138     photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and Pt-CO
139     cross interactions were parameterized as part of this work.
140    
141     \subsection{Metal-metal interactions}
142     Many of the potentials used for classical simulation of transition
143     metals are based on a non-pairwise additive functional of the local
144     electron density. The embedded atom method (EAM) is perhaps the best
145     known of these
146     methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
147     but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
148     the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
149     parameter sets. The glue model of Ercolessi {\it et al.} is among the
150     fastest of these density functional approaches.\cite{Ercolessi88} In
151     all of these models, atoms are conceptualized as a positively charged
152     core with a radially-decaying valence electron distribution. To
153     calculate the energy for embedding the core at a particular location,
154     the electron density due to the valence electrons at all of the other
155     atomic sites is computed at atom $i$'s location,
156     \begin{equation*}
157     \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
158     \end{equation*}
159     Here, $\rho_j(r_{ij})$ is the function that describes the distance
160     dependence of the valence electron distribution of atom $j$. The
161     contribution to the potential that comes from placing atom $i$ at that
162     location is then
163     \begin{equation*}
164     V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
165     \end{equation*}
166     where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
167     $\phi_{ij}(r_{ij})$ is an pairwise term that is meant to represent the
168     overlap of the two positively charged cores.
169 jmichalk 3807
170 gezelter 3808 The {\it modified} embedded atom method (MEAM) adds angular terms to
171     the electron density functions and an angular screening factor to the
172     pairwise interaction between two
173     atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
174     MEAM has become widely used to simulate systems in which angular
175     interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
176     metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
177     MEAM presents significant additional computational costs, however.
178 jmichalk 3807
179 gezelter 3808 The EAM, Finnis-Sinclair, MEAM, and the Quantum Sutton-Chen potentials
180     have all been widely used by the materials simulation community for
181     simulations of bulk and nanoparticle
182     properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
183     melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
184     fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
185     propagation,\cite{BECQUART:1993rg} and alloying
186     dynamics.\cite{Shibata:2002hh} All of these potentials have their
187     strengths and weaknesses. One of the strengths common to all of the
188     methods is the relatively large library of metals for which these
189     potentials have been
190     parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
191    
192 jmichalk 3802 \subsection{CO}
193 gezelter 3808 Since one explanation for the strong surface CO repulsion on metals is
194     the large linear quadrupole moment of carbon monoxide, the model
195     chosen for this molecule exhibits this property in an efficient
196     manner. We used a model first proposed by Karplus and Straub to study
197     the photodissociation of CO from myoglobin.\cite{Straub} The Straub and
198     Karplus model is a rigid three site model which places a massless M
199 jmichalk 3810 site at the center of mass along the CO bond. The geometry along with the interaction
200     parameters are reproduced in Table 1. The effective dipole moment is still
201     small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
202     to the experimental (-2.63 D~\AA\cite{}) and quantum mechanical predictions (WHAT VALUE, Coriana et al?).
203 jmichalk 3802 %CO Table
204     \begin{table}[H]
205 jmichalk 3810 \caption{Positions, $\sigma$, $\epsilon$ and charges for CO geometry and self-interactions\cite{}. Distances are in \AA~, energies are in kcal/mol, and charges are in $e$.}
206 jmichalk 3802 \centering
207 jmichalk 3810 \begin{tabular}{| c | c | ccc |}
208 jmichalk 3802 \hline
209 jmichalk 3810 \multicolumn{5}{|c|}{\textbf{Self-Interactions}}\\
210 jmichalk 3802 \hline
211 jmichalk 3810 & r & $\sigma$ & $\epsilon$ & q\\
212 jmichalk 3802 \hline
213 jmichalk 3810 \textbf{C} & 0.0 & 0.0262 & 3.83 & -0.75 \\
214     \textbf{O} & 1.13 & 0.1591 & 3.12 & -0.85 \\
215     \textbf{M} & 0.6457 & - & - & 1.6 \\
216 jmichalk 3802 \hline
217     \end{tabular}
218     \end{table}
219 gezelter 3808
220 jmichalk 3802 \subsection{Cross-Interactions}
221 gezelter 3808 The cross-interactions between the metals and the CO needed to be parameterized. Previous attempts at parameterization have used two different functional forms to model these interactions\cite{}. A LJ model was fit for the Metal-Carbon interaction and a Morse potential was parameterized for the Metal-Oxygen interaction. The parameter sets chosen, as shown in Table 2, did a suitable job at reproducing experimental adsorption energies as shown in Table 3, but more importantly, they were able to capture the binding site preference. The Pt-CO parameters show a slight preference for the atop binding site which matches the experimental observations.
222 jmichalk 3802
223 gezelter 3808 \subsection{Construction and Equilibration of 557 Metal interfaces}
224 jmichalk 3802
225 gezelter 3808 Our model systems are composed of approximately 4000 metal atoms cut along the 557 plane. The bare crystals were initially run in the Canonical ensemble at 1000K and 800K respectively for Pt and Au. The difference in temperature is necessary because of the two metals different melting points. Various amounts of CO were added to the simulation box and allowed to absorb to the metal surfaces over a short period of 100 ps. After further thermal relaxation the simulations were all run for at least 40 ns. A subset of the runs that showed interesting effects were allowed to run longer. The system
226 jmichalk 3802
227    
228 gezelter 3808 Our model systems are composed of approximately 4000 metal atoms cut along the 557 plane. This cut creates a stepped surface of 6x(111) surface plateaus separated by a single (100) atomic step height. The abundance of low-coordination atoms along the step edges acts as a suitable model for industrial catalysts which tend to have a high concentration of high-index sites. Experimental work has shown that such surfaces are notable for reconstructing upon adsorption\cite{}. Reconstructions have been seen for the Pt 557 surface that involve doubling of the step height and further formation of nano clusters with a triangular motif \cite{doi:10.1126/science.1182122}. To shed insight on whether this reconstruction is limited to the platinum surface, simulations of gold under similar conditions will also be examined. To properly observe these changes, our system size needs to be greater than the periodic phenomena we are examining. The large size and the long time scales needed precluded us from using quantum approaches. Thus, a forcefield describing the Metal-Metal, CO-CO, and CO-Metal interactions was parameterized and the simulations were run using OpenMD\cite{} an open-source molecular dynamics package.
229    
230    
231    
232    
233 jmichalk 3802 %\subsection{System}
234     %Once equilibration was reached, the systems were exposed to various sub-monolayer coverage of CO: $0, \frac{1}{10}, \frac{1}{4}, \frac{1}{3},\frac{1}{2}$. The CO was started many \AA~above the surface with random velocity and rotational velocity vectors sampling from a Gaussian distribution centered on the temperature of the equilibrated metal block. Full adsorption occurred over the period of approximately 10 ps for Pt, while the binding energy between Au and CO is smaller and led to an incomplete adsorption. The metal-metal interactions were treated using the Embedded Atom Method while the Pt-CO and Au-CO interactions were fit to experimental data and quantum calculations. The raised temperature helped shorten the length of the simulations by allowing the activation barrier of reconstruction to be more easily overcome. A few runs at lower temperatures showed the very beginnings of reconstructions, but their simulation lengths limited their usefulness.
235    
236    
237     %Table of Parameters
238     %Pt Parameter Set 9
239     %Au Parameter Set 35
240     \begin{table}[H]
241     \caption{Best fit parameters for metal-adsorbate interactions. Distances are in \AA~and energies are in kcal/mol}
242     \centering
243     \begin{tabular}{| c | cc | c | ccc |}
244     \hline
245     \multicolumn{7}{| c |}{\textbf{Cross-Interactions} }\\
246     \hline
247     & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ \\
248     \hline
249     \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
250     \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
251    
252     \hline
253     \end{tabular}
254     \end{table}
255    
256     %Table of energies
257     \begin{table}[H]
258 jmichalk 3805 \caption{Adsorption energies in eV}
259 jmichalk 3802 \centering
260     \begin{tabular}{| c | cc |}
261     \hline
262     & Calc. & Exp. \\
263     \hline
264     \textbf{Pt-CO} & -1.9 & -1.4~\cite{Kelemen}-- -1.9~\cite{Yeo} \\
265     \textbf{Au-CO} & -0.39 & -0.44~\cite{TPD_Gold_CO} \\
266     \hline
267     \end{tabular}
268     \end{table}
269    
270    
271    
272    
273    
274    
275     % Just results, leave discussion for discussion section
276     \section{Results}
277     \subsection{Diffusion}
278 jmichalk 3806 While an ideal metallic surface is unlikely to experience much surface diffusion, high-index surfaces have large numbers of low-coordinated atoms which have a much easier time overcoming the energetic barriers limiting diffusion, leading to easier surface reconstructions. Surface movement was divided between the parallel ($\parallel$) and perpendicular ($\perp$) directions relative to the step edge. We were then able to calculate diffusion constants as a function of CO coverage. As can be seen in Table 4, the presence and amount of CO directly affects the diffusion constants of surface platinum atoms. The presence of two 50\% coverage systems is to show how the diffusion process is affected by time. The majority of the systems were run for approximately 50 ns while the half monolayer system been running continuously. The lowered diffusion constant at longer run times will be examined in-depth in the discussion section.
279 jmichalk 3802
280     %Table of Diffusion Constants
281     %Add gold?M
282     \begin{table}[H]
283     \caption{Platinum diffusion constants parallel and perpendicular to the 557 step edge. As the coverage increases, the diffusion constants parallel and perpendicular to the step edge both initially increase and then decrease slightly. There were two approaches of analysis. One looking at the surface atoms that had moved more than a prescribed amount over the run time and the other looking at all surface atoms. Units are \AA\textsuperscript{2}/ns}
284     \centering
285     \begin{tabular}{| c | ccc | ccc | c |}
286     \hline
287     \textbf{System Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Atoms} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Atoms} & \textbf{Time (ns)}\\
288     \hline
289     &\multicolumn{3}{c|}{\textbf{Mobile}}&\multicolumn{3}{c|}{\textbf{Surface Atoms}} & \\
290     \hline
291     50\% & 3.74 & 0.89 & 497 & 2.05 & 0.49 & 912 & 116 \\
292     50\% & 5.81 & 1.59 & 365 & 2.41 & 0.68 & 912 & 46 \\
293     33\% & 6.73 & 2.47 & 332 & 2.51 & 0.93 & 912 & 46 \\
294     25\% & 5.38 & 2.04 & 361 & 2.18 & 0.84 & 912 & 46 \\
295     5\% & 5.54 & 0.63 & 230 & 1.44 & 0.19 & 912 & 46 \\
296     0\% & 3.53 & 0.61 & 282 & 1.11 & 0.22 & 912 & 56 \\
297     \hline
298     50\%-r & 6.91 & 2.00 & 198 & 2.23 & 0.68 & 925 & 25\\
299     0\%-r & 4.73 & 0.27 & 128 & 0.72 & 0.05 & 925 & 43\\
300     \hline
301     \end{tabular}
302     \end{table}
303    
304    
305    
306     %Discussion
307     \section{Discussion}
308 jmichalk 3806 Comparing the results from simulation to those reported previously by Tao et al. the similarities in the platinum and CO system are quite strong. As shown in figure 1, the simulated platinum system under a CO atmosphere will restructure slightly by doubling the terrace heights. The restructuring appears to occur slowly, one to two platinum atoms at a time. Looking at individual snapshots, these adatoms tend to either rise on top of the plateau or break away from the step edge and then diffuse perpendicularly to the step direction until reaching another step edge. This combination of growth and decay of the step edges appears to be in somewhat of a state of dynamic equilibrium. However, once two previously separated edges meet as shown in figure 1.B, this point tends to act as a focus or growth point for the rest of the edge to meet up, akin to that of a zipper. From the handful of cases where a double layer was formed during the simulation. Measuring from the initial appearance of a growth point, the double layer tends to be fully formed within $\sim$~35 ns.
309 jmichalk 3802
310     \subsection{Diffusion}
311     As shown in the results section, the diffusion parallel to the step edge tends to be much faster than that perpendicular to the step edge. Additionally, the coverage of CO appears to play a slight role in relative rates of diffusion, as shown in Table 4. Thus, the bottleneck of the double layer formation appears to be the initial formation of this growth point, which seems to be somewhat of a stochastic event. Once it appears, parallel diffusion, along the now slightly angled step edge, will allow for a faster formation of the double layer than if the entire process were dependent on only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the more likely a growth point is to be formed. One driving force behind this reconstruction appears to be the lowering of surface energy that occurs by doubling the terrace widths. (I'm not really proving this... I have the surface flatness to show it, but surface energy?)
312     \\
313     \\
314     %Evolution of surface
315     \begin{figure}[H]
316     \includegraphics[scale=0.5]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
317     \caption{Four snapshots at various times a) 258 ps b) 19 ns c) 31.2 ns d) 86.1 ns. Slight disruption of the surface occurs fairly quickly. However, the doubling of the layers seems to be very dependent on the initial linking of two separate step edges. The focal point in b, appears to be a growth spot for the rest of the double layer.}
318     \end{figure}
319    
320    
321    
322    
323     %Peaks!
324     \includegraphics[scale=0.25]{doublePeaks_noCO.png}
325     \section{Conclusion}
326    
327    
328 gezelter 3808 \section{Acknowledgments}
329     Support for this project was provided by the National Science
330     Foundation under grant CHE-0848243 and by the Center for Sustainable
331     Energy at Notre Dame (cSEND). Computational time was provided by the
332     Center for Research Computing (CRC) at the University of Notre Dame.
333 jmichalk 3802
334 gezelter 3808 \newpage
335     \bibliography{firstTryBibliography}
336     \end{doublespace}
337     \end{document}