1: % INJ
2:
3: %\documentclass[aps,pra,superscriptaddress,showpacs,twocolumn]{revtex4}
4: \documentclass[12pt]{elsart}
5: %\documentclass[aps,prb,superscriptaddress,showpacs,twocolumn]{revtex4}
6: %\documentclass[aps,prb,superscriptaddress,showpacs,preprint]{revtex4}
7: %\usepackage[draft]{graphicx}
8: \usepackage{graphicx}
9: \usepackage{amssymb}
10: \usepackage{amsmath}
11:
12: \newcommand{\bk}{{\bf k}}
13: \newcommand{\bq}{{\bf q}}
14: \newcommand{\bQ}{{\bf Q}}
15: \newcommand{\bp}{{\bf p}}
16: \newcommand{\bx}{{\bf x}}
17: \newcommand{\br}{{\bf r}}
18: \newcommand{\Li}{{\mathop{\rm{Li}}\nolimits}}
19: \renewcommand{\Im}{{\mathop{\rm{Im}}\nolimits\,}}
20: \renewcommand{\Re}{{\mathop{\rm{Re}}\nolimits\,}}
21: \newcommand{\sgn}{{\mathop{\rm{sgn}}\nolimits\,}}
22: \newcommand{\Tr}{{\mathop{\rm{Tr}}\nolimits\,}}
23: \newcommand{\EF}{E_{\mathrm{F}}}
24: \newcommand{\kB}{k_{\mathrm{B}}}
25: \newcommand{\kF}{k_{\mathrm{F}}}
26: \newcommand{\nF}{n_{\mathrm{F}}}
27: \newcommand{\Green}{{G}}
28: \newcommand{\dSC}{{\mathrm{dSC}}}
29: \newcommand{\dPG}{{\mathrm{dPG}}}
30: \newcommand{\dDW}{{\mathrm{dDW}}}
31: \newcommand{\Ret}{{\mathrm{R}}}
32: \newcommand{\Tau}{T_\tau}
33: \newcommand{\HF}{{\mathrm{HF}}}
34: \newcommand{\TF}{{\mathrm{TF}}}
35: \newcommand{\vW}{{\mathrm{vW}}}
36:
37: \begin{document}
38: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
39: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
40:
41: \journal{Phys. Lett. A}
42:
43: \begin{frontmatter}
44: \title{Modeling vacancies and hydrogen impurities in graphene: A molecular
45: point of view}
46:
47: \author[CTChem]{G. Forte}
48: \author[CTChem]{A. Grassi}
49: \author[CTChem]{G. M. Lombardo}
50: \author[IMM]{A. La Magna}
51: \author[CTPhys,INFN,CNISM]{G. G. N. Angilella\thanksref{corr}}
52: \author[CTPhys,CNISM]{R. Pucci}
53: \author[CTPhys]{R. Vilardi}
54:
55: \address[CTChem]{Dipartimento di Scienze Chimiche, Facolt\`a di Farmacia,
56: Universit\`a di Catania,\\
57: Viale A. Doria, 6, I-95126 Catania, Italy}
58: \address[IMM]{IMM, CNR, Catania, Italy}
59:
60: \address[CTPhys]{Dipartimento di Fisica e Astronomia, Universit\`a di
61: Catania,\\ Via S. Sofia, 64, I-95123 Catania, Italy}
62: \address[INFN]{INFN, Sez. Catania, Italy}
63: \address[CNISM]{CNISM, UdR di Catania, Italy}
64: \thanks[corr]{Corresponding author. E-mail: {\tt giuseppe.angilella@ct.infn.it}.}
65: \date{\today}
66:
67:
68: \begin{abstract}
69: We have followed a `molecular' approach to study impurity effects in graphene.
70: This is thought as the limiting case of an infinitely large cluster of benzene
71: rings. Therefore, we study several carbon clusters, with increasing size, from
72: phenalene, including three benzene rings, up to coronene~61, with 61 benzene
73: rings. The impurities considered were a chemisorbed H atom, a vacancy, and a
74: substitutional proton. We performed HF and UHF calculations using the STO-3G
75: basis set. With increasing cluster size in the absence of impurities, we find a
76: decreasing energy gap, here defined as the HOMO-LUMO difference. In the case of
77: H chemisorption or a vacancy, the gap does not decrease appreciably, whereas it
78: is substantially reduced in the case of a substitutional proton. The presence of
79: an impurity invariably induces an increase of the density of states near the
80: HOMO level. We find a zero mode only in the case of a substitutional proton. In
81: agreement with experiments, we find that both the chemisorbed H, the
82: substitutional proton, and the C atom near a vacancy acquire a magnetic moment.
83: The relevance of graphene clusters for the design of novel electronic devices is
84: also discussed.
85:
86: PACS:
87: 36.40.Cg, %Electronic and magnetic properties of clusters
88: 73.22.-f, %Electronic structure of nanoscale materials: clusters, nanoparticles, nanotubes, and nanocrystals
89: 73.20.Hb %Impurity and defect levels; energy states of adsorbed species
90: %36.40.Mr, % Spectroscopy and geometrical structure of clusters
91: %31.25.Qm, % Electron correlation calculations for polyatomic molecules
92: %33.20.Tp % Vibrational analysis
93: \end{abstract}
94:
95: \end{frontmatter}
96:
97: \section{Introduction}
98:
99: Since the pioneering works of Wallace \cite{Wallace:47}, Coulson
100: \cite{Coulson:52a}, and Semenoff \cite{Semenoff:54}, several authors have
101: studied the peculiar electronic and structural properties of graphene, a single
102: sheet of graphite, originally only considered as a prototype of the graphite's
103: surface. A breakthrough in these studies arrived with the experimental
104: realization of graphene \cite{Novoselov:04}. The increasing interest in this
105: material stems both from the analogy of its behavior with many phenomena of
106: quantum electrodynamics (QED) \cite{Katsnelson:07a}, and because of its
107: prospective applications in the fabrication of novel electronic devices. To this
108: aim, it is important to have systems with a tunable band gap. We will stress
109: that the graphene clusters (GC) here considered, even if they are characterized
110: by a zig-zag boundary, have a semiconducting behavior, at variance with zig-zag
111: graphene nanoribbons (ZGNR), which have a metallic behavior \cite{Yan:07}. The
112: gap strongly depends on the cluster size, so that it will be highly desirable to
113: have method allowing to control the size of graphene quantum dots.
114:
115: Graphene is a robust material. However, defects, such as ripples, adatoms,
116: vacancies and charges induced by the substrate, can modify its electronic
117: properties \cite{CastroNeto:08}. In the past, the study of vacancies and
118: impurities in carbon-based systems has been important for several aspects.
119:
120: (i) Graphene nanostructures are believed to be attractive structures for
121: hydrogen storage. Preliminary results in this direction have been reported in
122: the case of single-walled carbon nanotubes (SWCNT) \cite{Dillon:97}.
123:
124: (ii) It has been found \cite{Esquinazi:03} that proton irradiation of highly
125: oriented pyrolitic graphite samples triggers ferro- or ferrimagnetism. It is now
126: commonly accepted that not only vacancies, but also hydrogen chemisorbed on
127: graphene can acquire a magnetic moment.
128:
129: (iii) From an astrophysical point of view, it would be of interest to understand
130: why H$_2$ is the most abundant molecule in the interstellar medium (ISM),
131: despite the continuous dissociation of molecular hydrogen by UV radiation and
132: cosmic rays \cite{Hornekaer:06}.
133:
134: Of course, there can be relevant differences in the characterization of defects
135: in graphite, nanotubes, or graphene. For instance, if one considers the
136: screening of a charged impurity, Cheianov and Fal'ko \cite{Cheianov:06} have
137: shown that the electron density $\delta\rho$ in graphene decays as
138: $\delta\rho(r) \sim r^{-3}$ at long distances $r$ from the impurity. Such a
139: behavior should be contrasted with that derived by Lau and Kohn \cite{Lau:78}
140: for a two-dimensional (2D) electron gas, where $\delta\rho(r) \sim r^{-2}$. The
141: reason of such a difference is due to the peculiar chiral properties of the
142: electrons in graphene. It has been suggested \cite{Bena:07} that such a
143: difference can be used to distinguish between monolayer and bilayer graphene in
144: Fourier transformed scanning tunneling spectroscopy (FTSTS). Actually, impurity
145: induced quantum interferences have been used in scanning tunneling microscopy
146: (STM) \cite{Mallet:07} to identify experimentally mono- or bilayer graphene.
147: Previously, STM was used \cite{Ruffieux:00,Ruffieux:05} to study chemisorption
148: of hydrogen on the basal plane of graphite and atomic vacancy formation.
149:
150: From a theoretical point of view, the effects on the electronic structure of
151: graphene due to the presence of vacancies, local impurities, and substitutional
152: impurities have been studied by Pereira \emph{et al.} \cite{Pereira:08}, by
153: using a tight-binding Hamiltonian with a local potential $U$. These authors have
154: found general trends in the low-energy spectrum of graphene connected with
155: localized zero modes, resonances and gaps induced by the above mentioned
156: defects. However, in some cases, it is important to have a more detailed
157: description of the local interaction of the defects with graphene. For this
158: reason, in the present work, we will make recourse to molecular models,
159: which in the past have been used to study the chemisorption of hydrogen on
160: graphite surface \cite{Klose:92,Fromherz:93}. According to these calculations,
161: the graphite surface is modeled as a cluster of benzene rings, whose external
162: dangling bonds are saturated by hydrogen atoms. The most frequently used model
163: clusters are polycyclic aromatic hydrocarbons (PAH), and particularly coronene
164: (C$_{24}$H$_{12}$). More recently, this model cluster has been used by
165: Patchkovskii \emph{et al.} \cite{Patchkovskii:05} to study the interaction of
166: molecular hydrogen with graphene nanostructures. These authors have performed
167: calculations at the level of second-order M\o{}ller-Plesset (MP2) perturbation
168: theory \cite{Moeller:34}.
169:
170: In the present work, we will use such model clusters to simulate graphene, at
171: the level of \emph{ab initio} Hartree-Fock (HF) and unrestricted \emph{ab
172: initio} Hartree-Fock (UHF) theory. In some cases, also MP2 calculations will be
173: presented, but the main emphasis will be more on the trends obtained by
174: increasing the cluster size, than in the study of correlation effects. Actually,
175: Martin \emph{et al.} \cite{Martin:08} have argued that correlation effects are
176: small in graphene. A similar result was found by Siringo \cite{Siringo:84} in
177: the study of H chemisorbed on graphene \cite{note:Siringo}. Furthermore,
178: recently the Manchester group has found that the electron energy levels in
179: quantum dots in graphene follow a statistics characteristic of `Dirac
180: billiards', rather than a Poisson distribution, as is expected for an infinite
181: system \cite{Ponomarenko:08}: therefore, quantization effects are quite
182: important. It is interesting, in this context, to derive these results also
183: within a molecular approach. In the present work, we have used clusters up to 61
184: benzene rings, and we have considered the cases of a single vacancy, of a
185: chemisorbed H atom, and of a substitutional proton.
186:
187: The paper is organized as follows. In Sec.~\ref{sec:models}, we describe the
188: models used. In Sec.~\ref{sec:results} we present our results and a comparison
189: thereof with density functional theory (DFT) and other types of calculations.
190: Finally, in Sec.~\ref{sec:conclusions} we summarize and draw our conclusions.
191:
192: \section{The model clusters}
193: \label{sec:models}
194:
195: \begin{figure}[t]
196: \centering
197: \includegraphics[bb=253 21 609 343,clip,width=0.9\columnwidth]{CORO_PLUS_2.ps}
198: \caption{(Color online.) Showing coronene, C$_{24}$H$_{12}$ (gray rods),
199: coronene~19, C$_{54}$H$_{18}$ (gray and blue rods), and coronene~37,
200: C$_{96}$H$_{24}$ (gray, blue, and red rods).}
201: \label{fig:coronene2plus}
202: \end{figure}
203:
204: The molecular clusters which we have considered are: (i) phenalene (phenalenyl
205: radical); (ii) pyrene; (iii) coronene; (iv) coronene~19, \emph{viz.} coronene
206: surrounded by another series of benzene rings, for a total of 19 benzene rings;
207: (v) coronene~37, \emph{viz.} coronene surrounded by two more series of benzene
208: rings, for a total of 37 benzene rings; (vi) coronene~61. Coronene~37 is shown
209: in Fig.~\ref{fig:coronene2plus}. We believe that the larger the cluster, the
210: better it models graphene. For this reason, we have chosen to place in the
211: middle of the cluster: (a) a chemisorbed H atom; (b) a carbon vacancy; (c) a
212: proton in the substitutional position of a carbon atom.
213:
214: For clusters (i)--(iii) we have used HF calculations with a STO-3G basis set and
215: HF calculations with a 6-311G basis set \cite{Frisch:04}. In these cases we have
216: also performed MP2 calculations. All the remaining larger clusters have been
217: treated at the HF level using a STO-3G basis set. In the cases of H
218: chemisorption, carbon vacancy and proton substitution, we have performed UHF
219: calculations, because in all these cases there is one $\pi$-electron missing.
220: The cluster geometry has been optimized in all cases.
221:
222: \section{Results}
223: \label{sec:results}
224:
225: \begin{figure}[t]
226: \centering
227: %\includegraphics[width=0.9\columnwidth]{ENERGY_BASI.ps}
228: \includegraphics[width=0.9\columnwidth]{NEW_FIG_2.ps}
229: \caption{(Color online.) Showing the dependence of the cluster energy as a
230: function of the cluster under consideration, for the various methods (HF or MP2)
231: and basis sets (STO-3G or 6-311G) employed in the present work. On this scale,
232: the results of different approximations are almost indistinguishable.}
233: \label{fig:levels}
234: \end{figure}
235:
236: \begin{figure}[t]
237: \centering
238: %\includegraphics[width=0.9\columnwidth]{FIT_named.ps}
239: \includegraphics[width=0.9\columnwidth]{NEW_FIG_3.ps}
240: %\includegraphics[width=0.9\columnwidth]{GAP_BASI.ps}
241: %\caption{(Color online.) Showing the dependence of the HOMO-LUMO difference (gap
242: %energy) \emph{vs} the various clusters under consideration, for the various methods (HF or MP2)
243: %and basis sets (STO-3G or 6-311) employed in the present work.}
244: \caption{(Color online.) Showing the dependence of the HOMO-LUMO difference (gap
245: energy) \emph{vs} the number of carbon cycles for the various clusters studied
246: in the present work. The dashed line is a guide to the eye.}
247: \label{fig:HOMO-LUMO}
248: \end{figure}
249:
250: \subsection{Pure graphene clusters}
251: \label{sec:puregraphene}
252:
253: First of all, we note that the total energy as a function of the number of
254: benzene rings in the clusters (cf. Fig.~\ref{fig:levels}) depends weakly
255: (\emph{i.e.} to within 2~\%) on the basis set or on the specific approximation
256: employed (HF, MP2). This lends support to our conviction that relevant
257: conclusions can be drawn already for one of the largest cluster considered,
258: \emph{i.e.} coronene~37, within the framework of the HF approximation with the
259: STO-3G basis set.
260:
261: Secondly, we note that the gap, which in these kind of calculations is described
262: by the HOMO-LUMO difference, is decreasing by increasing the size of the
263: cluster, as shown in Fig.~\ref{fig:HOMO-LUMO}. This is in agreement with the
264: consideration that our clusters resemble more closely graphene, as the number of
265: benzene rings increases. The smallest gap which we have obtained is 5.48~eV.
266:
267: Extrapolation of our data seems to indicate that the gap would close at a
268: cluster size of $\gtrsim 1000$~benzene rings. This is in qualitative agreement
269: with the experimental finding that graphene quantum dots as large as $\sim
270: 30$~nm are characterized by a gap of $\sim 0.5$~eV \cite{Ponomarenko:08}. We
271: find a C--C nearest neighbours equilibrium distance $d_{\mathrm{C-C}} =
272: 1.42$~\AA, which is stable with respect to the different approximations and
273: cluster sizes considered here. This in striking agreement with the typical bond
274: length for $sp^2$ C--C bonds for graphene and graphite, which is 1.42~\AA.
275:
276: Our minimization procedure yields for the clusters considered in this case a
277: planar configuration, at variance with the experimental finding for graphene,
278: showing the presence of ripples \cite{Stolyarova:07,Meyer:07}. This discrepancy
279: may be related to the specific boundary conditions used in our calculations.
280: Such a result is not in contradiction with the known fact that a nonzero amount
281: of curvature is required in order to stabilize a two-dimensional crystal, as a
282: consequence of Mermin-Wagner theorem \cite{Peierls:34,Mermin:66,Mermin:68}.
283: Indeed, the systems we are considering here are not truly infinite lattices, but
284: rather finite-size clusters. Furthermore, we note that Choi \emph{et al.}
285: \cite{Choi:08}, by using a cluster with a monovacancy and 127~C atoms, and
286: performing spin-polarized density functional theory calculations, have found an
287: almost planar configuration.
288:
289:
290: \begin{figure}[t]
291: \centering
292: \includegraphics[width=0.9\columnwidth]{coronene_37_DOS.ps}\\
293: \includegraphics[width=0.9\columnwidth]{coronene_37_DOS_2.ps}
294: \caption{(Color online.) Showing the DOS for pure coronene~37 (upper panel), and
295: for coronene~37 with a chemisorbed H atom (lower panel), as a function of
296: energy.}
297: \label{fig:DOS}
298: \end{figure}
299:
300: \begin{figure}[t]
301: \centering
302: \includegraphics[width=0.8\columnwidth]{coronene_37_PS_S_15eV.ps}\\
303: \includegraphics[width=0.8\columnwidth]{coronene_37_PS_S_10eV.ps}\\
304: \includegraphics[width=0.8\columnwidth]{coronene_37_PS_S_5eV.ps}
305: \caption{(Color online.) Showing the level-spacing distribution $P(s)$ for
306: coronene~37, corresponding to different energy intervals centered around zero:
307: 15~eV (topmost panel), 10~eV (intermediate panel), 5~eV (bottom panel). It can
308: be noted that $P(s)$ looses the Poissonian behavior in favor of a Gaussian
309: behavior as the energy interval decreases.}
310: \label{fig:Ps}
311: \end{figure}
312:
313:
314: In Fig.~\ref{fig:DOS} we report the density of states (DOS) as a function of
315: energy (eV) for coronene~37. We note that (i) a nearly straight line joins the
316: peaks near the HOMO energy: such a trend mimics the linear dependence which is
317: found in the DOS near the Fermi level in the infinite system
318: \cite{CastroNeto:08}; (ii) particle-hole symmetry is absent, as is obtained in
319: tight-binding calculations with a nonzero next-nearest hopping parameter
320: $t^\prime$ \cite{CastroNeto:08}. However, here this asymmetry is not very
321: pronounced than in the case $t^\prime \neq 0$. This is probably due to the fact
322: that in our calculations we do not restrict the interaction to next-nearest
323: neighbors. We have also considered the level-spacing distribution $P(s)$, where
324: the dimensionless spacing $s$ is defined as the energy difference $\Delta E_i =
325: E_i - E_{i-1}$ between successive levels, divided by the average $\langle \Delta
326: E_i \rangle$ of the energy difference between successive levels
327: \cite{DeRaedt:08}. The quantity $P(s)$ yields the number of energy differences
328: for which $s-\Delta/2 < \Delta E_i / \langle \Delta E_i \rangle \leq s +
329: \Delta/2$, where $\Delta$ is the step of the histogram. We have found that
330: $P(s)$ strongly depends on the energy range considered around $E=0$ (cf.
331: Fig.~\ref{fig:Ps}). One can also recognize a strong dependence on the cluster
332: size, so that no definite conclusion can be achieved at the present status of
333: the calculations.
334:
335: \subsection{Clusters with a chemisorbed H atom}
336: \label{sec:chemisorbed}
337:
338: \begin{figure}[t]
339: \centering
340: \includegraphics[bb=152 20 705 341,clip,width=0.9\columnwidth]{CORO_PLUS_HF_3G_H_TOP_B_S.ps}
341: \caption{(Color online.) Showing coronene~37 with a chemisorbed H atom.}
342: \label{fig:largestHatom}
343: \end{figure}
344:
345:
346:
347: The largest cluster considered in the presence of a chemisorbed H atom is
348: reported in Fig.~\ref{fig:largestHatom}, where the H adatom is depicted with a
349: yellow color.
350: %We have indicated with A and B the carbon atoms belonging to the
351: %two different sublattices.
352: The total energy, as a function of the cluster size
353: and the approximation used, is quite similar to that reported in
354: Fig.~\ref{fig:levels}, and will not be shown here.
355:
356: We find an average equilibrium distance of 1.53~\AA{} between two in-plane
357: nearest neighbor C atoms, forming an angle $\angle$HCC$=105.3^\circ$ with the H
358: adatom, while the C atom, on top of which the H atom resides, emerges from the
359: plane of 0.71~\AA. The distance of the H atom from the underneath C atom is
360: 1.10~\AA. The distance of the C atoms surrounding the adatom is of 0.30~\AA{}
361: with respect to the planar configuration. These values are in qualitative
362: agreement with recent DFT calculations \cite{Boukhvalov:08}. We find that (i)
363: the most stable position is that with H on top of a carbon atom; (ii) the
364: underlying C atom is induced out of the plane; (iii) there are small differences
365: between the $\alpha$ and $\beta$ sets of eigenvalues positions. We will
366: therefore report only the results for the $\alpha$ set.
367:
368: An important difference found in the presence of a chemisorbed H atom, with
369: respect to the pure graphene clusters discussed in Sec.~\ref{sec:puregraphene},
370: is the trend of the gap as a function of the number of the benzene rings in the
371: cluster, which is reported in Fig.~\ref{fig:rings}. No evidence of a gap
372: decrease appears in this case, which supports the results of Pereira \emph{et
373: al.} \cite{Pereira:08} that hydrogen chemisorption on graphene induces the
374: opening of a gap. Also in agreement with such findings \cite{Pereira:08} is the
375: behavior of the DOS, which is reported a dashed blue line in Fig.~\ref{fig:DOS}.
376: One can see a clear enhancement of the DOS near the HOMO level, which Pereira
377: \emph{et al.} \cite{Pereira:08} have found as due to a resonance induced by the
378: chemisorbed hydrogen. Of course, in this case we have used UHF calculations,
379: which allowed us to estimate also the magnetic moment acquired by the H atom,
380: which for a corenene~37 cluster is $-0.33$~$\mu_{\mathrm{B}}$.
381:
382: The above mentioned results, in the framework here considered, do not depend
383: strongly on the cluster size considered, as reported in
384: Tab.~\ref{tab:moment}, where a comparison is presented between the results
385: obtained with the coronene~19 and the coronene~37 clusters, by using a STO-3G
386: basis set.
387:
388: \begin{figure}[t]
389: \centering
390: %\includegraphics[width=0.9\columnwidth]{GAP_BASI_H_TOP.ps}
391: \includegraphics[width=0.9\columnwidth]{NEW_FIG_7.ps}
392: \caption{(Color online.) Showing the dependence of the HOMO-LUMO difference (gap
393: energy) \emph{vs} the various clusters considered, in the presence
394: of a chemisorbed H atom, for the different methods and basis sets employed in
395: this work.}
396: \label{fig:rings}
397: \end{figure}
398:
399: \begin{table}[t]
400: \caption{Showing values of the total energy $E$ (Hartree), HOMO energy
401: (Hartree), LUMO energy (Hartree), HOMO-LUMO gap (eV), and magnetic moment $\mu$
402: ($\mu_{\mathrm{B}}$), for the coronene~19 and coronene~37 clusters, and the same
403: clusters with a H adatom on top, with a vacancy, with substitutional proton, and
404: with a substitutional H atom, using a STO-3G basis set.}
405: \label{tab:moment}
406: \begin{tabular}{lrrrrr}
407: \hline
408: & $E$ & HOMO & LUMO & gap & $\mu$ \\
409: \hline
410: \multicolumn{6}{l}{coronene~19} \\
411: pure & $-2030.796$ & $-0.159$ & $0.128$ & $7.810$ & --- \\
412: with H top & $-2031.701$ & $-0.252$ & $0.243$ & $13.465$ &
413: $-0.333$\\
414: with vacancy & $-1993.461$ & $-0.262$ & $0.216$ & $13.007$ & $1.838$
415: \\
416: with proton & $-1993.834$ & $-0.335$ & $-0.120$ & $5.850$ & $0.917$
417: \\
418: with H subst. & $-1994.058$ & $-0.263$ & $0.212$ & $12.925$ & $0.849$ \\
419: \hline
420: \multicolumn{6}{l}{coronene~37} \\
421: pure & $-3605.818$ & $-0.137$ & $0.101$ & $6.476$ & --- \\
422: with H top & $-3607.059$ & $-0.251$ & $0.240$ & $13.361$ &
423: $-0.334$\\
424: with vacancy & $-3568.808$ & $-0.259$ & $0.215$ & $12.898$ & $1.838$
425: \\
426: with proton & $-3569.186$ & $-0.319$ & $-0.114$ & $5.578$ & $0.833$ \\
427: with H subst. & $-3569.401$ & $-0.261$ & $0.212$ & $12.871$ & $0.851$ \\
428: \hline
429: \end{tabular}
430: \end{table}
431:
432:
433:
434: \subsection{Clusters with a vacancy}
435: \label{sec:vacancy}
436:
437: \begin{figure}[t]
438: \centering
439: \includegraphics[width=0.9\columnwidth]{coro_plus_2_VACANCY_stick.ps}
440: \caption{(Color online.) Showing coronene~37 in the presence of a vacancy. Number
441: 68 indicates the position of the C atom around the vacancy referred to in the
442: main text.}
443: \label{fig:vacancy}
444: \end{figure}
445:
446:
447: \begin{figure}[t]
448: \centering
449: \includegraphics[width=0.9\columnwidth]{coronene_37_H_DOS.ps}\\
450: \includegraphics[width=0.9\columnwidth]{coronene_37_PROTON_DOS.ps}
451: \caption{Upper panel: Showing DOS of coronene~37 with a vacancy.
452: Lower panel: Showing DOS of coronene~37 with a substitutional proton.}
453: \label{fig:DOSvac}
454: \end{figure}
455:
456:
457: In Fig.~\ref{fig:vacancy} we report the clusters considered in the presence of a
458: vacancy. The same considerations made above for the total energy per number of
459: benzene rings in Sec.~\ref{sec:puregraphene} and \ref{sec:chemisorbed} above
460: apply to this case as well. The most stable electronic configuration is the
461: triplet state, and the gap has a similar trend and similar values as those
462: obtained for H chemisorption (cf. Tab.~\ref{tab:moment}). From
463: Fig.~\ref{fig:DOSvac}, one can see that also in this case the DOS presents an
464: increase near the HOMO level with respect to the case in the absence of the
465: vacancy. The main difference with respect to the case with H on top, is relative
466: to the magnetic moment acquired by a carbon atom near the vacancy (see atom
467: labelled C$^{68}$ for coronene~37 in Fig.~\ref{fig:vacancy}). The value obtained
468: in this case is $\approx 1.84$~$\mu_{\mathrm{B}}$, which presents just small
469: changes in going from coronene~19 to coronene~37, or by using the $\alpha$ or
470: $\beta$ sets of eigenvalues. The fact that the magnetic moment is smaller when
471: we consider H chemisorption instead of a vacancy is in qualitative agreement
472: with the results of Yaziev and Helm \cite{Yaziev:07}. These authors have found
473: $1$~$\mu_{\mathrm{B}}$ per hydrogen chemisorption defect, and
474: $1.12-1.53$~$\mu_{\mathrm{B}}$ per vacancy defect, depending on the defect
475: concentration. A value of $1.52$~$\mu_{\mathrm{B}}$ for a monovacancy is
476: reported by Choi \emph{et al.} \cite{Choi:08}. From this comparison, we can
477: deduce that the magnetic moments are sensitive to the model calculations
478: employed, but our results are in overall agreement with the most recent works.
479:
480: Fig.~\ref{fig:vacdist} and Tab.~\ref{tab:vacdist} report the distances between
481: the various C~atoms around the vacancy.
482:
483:
484: \begin{figure}[t]
485: \centering
486: \includegraphics[width=0.9\columnwidth]{VACANCY_DISTANCE.ps}
487: \caption{Showing C atoms around the vacancy. The distances between the various
488: atoms are reported in Tab.~\ref{tab:vacdist}.}
489: \label{fig:vacdist}
490: \end{figure}
491:
492:
493: \begin{table}[t]
494: \caption{Distances (in \AA{}) between the various C~atoms $(i,j)$ around the vacancy, as
495: listed in Fig.~\ref{fig:vacdist}.}
496: \label{tab:vacdist}
497: \begin{tabular}{ccc}
498: \hline
499: $i$ & $j$ & $d_{ij}$ \\
500: \hline
501: 68 & 69 & 1.38\\
502: 68 & 70 & 1.38\\
503: 68 & 71 & 2.39\\
504: 68 & 66 & 2.39\\
505: 68 & 17 & 2.77\\
506: 68 & 6 & 2.77\\
507: 68 & 1 & 2.63\\
508: 68 & 5 & 2.63\\
509: 68 & 80 & 2.41\\
510: 68 & 82 & 2.41\\
511: \hline
512: \end{tabular}
513: \end{table}
514:
515:
516: \subsection{Clusters with a substitutional proton}
517:
518: Last, we have considered the same clusters as in Fig.~\ref{fig:coronene2plus},
519: but now with a proton substituting a carbon atom at the cluster center. Such a
520: configuration has been realized by considering a de-electronated hydrogen atom.
521: The same considerations for the total energy apply also to this system, as
522: discussed in the presvious sections. The distances between the proton and the
523: nearest neighbor C atoms are 1.03~\AA{} for the C$^{68}$ atom, and 1.83~\AA{}
524: for the other two next nearest neighbor atoms. The cluster turns out to be flat.
525: The DOS presents a peak near zero energy (cf. Fig.~\ref{fig:DOSvac}), as found
526: by several authors when considering the effect of impurities in graphene
527: \cite{Pereira:08,Hu:08}. We note that, for a substitutional proton, the gap is
528: more than a half smaller with respect to the cases considered in
529: Sections~\ref{sec:chemisorbed} and \ref{sec:vacancy} (Cf.
530: Tab.~\ref{tab:moment}). It is even smaller with respect to a cluster with a
531: substitutional H, \emph{i.e.} a system presenting charge neutrality, which is
532: also reported in Tab.~\ref{tab:moment}. We argue that the gap is a sensitive
533: quantity with respect to the injections of extra charges. The magnetic moment on
534: the proton has an intermediate value with respect to the two cases considered
535: previously. It is more sensitive to the cluster size going from
536: $0.91$~$\mu_{\mathrm{B}}$ from coronene~19 to $0.83$~$\mu_{\mathrm{B}}$ for
537: coronene~37, but is similar to that obtained for the substitutional H, which is
538: $0.85$~$\mu_{\mathrm{B}}$ for both clusters (cf. Tab.~\ref{tab:moment}). The
539: present model is probably the most interesting in realtion to the experiments
540: performed by irradiating graphite with protons \cite{Esquinazi:03}.
541:
542: \section{Conclusions}
543: \label{sec:conclusions}
544:
545: In the present work we have followed a `molecular' approach to study impurity
546: effects in graphene. The impurities considered were a chemisorbed H atom, a
547: vacancy, and a a positively charged substitutional H. We modelled these systems
548: by considering clusters containing up to 37 benzene rings. We performed HF and
549: UHF calculations using the STO-3G basis set.
550:
551: The main emphasis of this work is not on correlation effects, which in some
552: cases have been treated at the MP2 level, but on the size of the cluster. From
553: the results discussed in the present work, we believe that coronene~37 is
554: sufficiently large to ensure a local behavior similar to that of bulk graphene,
555: especially near the center of the cluster. This is supported by the small
556: changes in the physical properties studied here in going from coronene~19 to
557: coronene~37. Of course, some features differ considerably from the finite
558: clusters and the infinite system. For instance, we still observe quite a large
559: gap in the density of states of finite clusters, as compared to that of
560: graphene, while we have shown some important trends, which look quite promising
561: in interpreting the general features of graphene: (i) The gap decreases by
562: increasing the cluster size; (ii) The gap does not decrease if we consider H
563: chemisorption or a vacancy; (iii) the gap is drastically reduced if we consider
564: a substitutional proton. Other important features are: (i) an increase of the
565: DOS near the HOMO level for all the impurities considered here with respect to
566: the pure system; (ii) the appearance of a zero mode in the DOS if we consider a
567: substitutional proton.
568:
569: These results are for many aspects in agreement with the theoretical work of
570: Pereira \emph{et al.} \cite{Pereira:08}, who have studied local disorder in
571: graphene by using a tight-binding approach. Such a method allowed these authors
572: to treat lattices with $4\times 10^6$ carbon atoms, thus reaching the size of
573: real samples. However, they cannot treat lattice distortions around the defect,
574: or the specific features connected with the particular impurity introduced.
575:
576: In agreement with experiments \cite{Esquinazi:03}, we have found that
577: chemisorbed H, substitutional proton, and C atom near a vacancy acquire a
578: magnetic moment. The determination of a precise value thereof is quite delicate,
579: but we believe that the trend on going from one defect to another is correct, in
580: view of the comparison of our results with the calculations of Yazyev and Helm
581: \cite{Yaziev:07}.
582:
583: We could not derive definite conclusions as regards the spacing of the levels
584: near the HOMO state, and further work is thus required in this direction. In the
585: future, it should be possible to treat even larger clusters, and these will be
586: able to treat more realistically especially graphene quantum dots. We find that
587: the gap strongly depends on the graphene cluster size, so that it is essential
588: to control the size of graphene quantum dots in order to obtain a particular gap
589: value. This is relevant for the design of novel electronic devices, as well as
590: it is important to control the impurity level.
591:
592: %\cite{Schedin:07}
593: %\cite{Wehling:08}
594: %\cite{Leenaerts:08}
595: %\cite{Nakada:96}
596: %\cite{Russo:08}
597:
598: \ack
599:
600: The authors thank Professor N. H. March for useful discussions over the general
601: area embraced by the present paper.
602:
603: \begin{small}
604: %\bibliographystyle{apsrev}
605: \bibliographystyle{mprsty}
606: \bibliography{a,b,c,d,e,f,g,h,i,j,k,l,m,n,o,p,q,r,s,t,u,v,w,x,y,z,zzproceedings,Angilella,notes}
607: \end{small}
608:
609:
610:
611:
612: \end{document}
613:
614: