0807.1953/bn.tex
1: %\documentclass[aps,prb,preprint]{revtex4}
2: \documentclass[aps, prb, twocolumn, showpacs, floatfix]{revtex4}
3: \usepackage{graphicx}
4: 
5: \begin{document}
6: 
7: \title{Electronic structures of organic molecule encapsulated BN nanotubes under transverse electric field}
8: 
9: \author{Wei He}
10: \author{Zhenyu Li}
11: \author{Jinlong Yang}
12: \thanks{Corresponding author. E-mail: jlyang@ustc.edu.cn}
13: \author{J. G. Hou}
14: 
15: \affiliation{Hefei National Laboratory for Physical Sciences at
16:      Microscale,  University of Science and Technology of
17:      China, Hefei,  Anhui 230026, China}
18: 
19: \date{\today}
20: 
21: \begin{abstract}
22: The electronic structures of boron nitride nanotubes (BNNTs) doped
23: by different organic molecules under a transverse electric field
24: were investigated via first-principles calculations. The external
25: field reduces the energy gap of BNNT, thus makes the molecular bands
26: closer to the BNNT band edges and enhances the charge transfers
27: between BNNT and molecules. The effects of the electric field
28: direction on the band structure are negligible. The electric field
29: shielding effect of BNNT to the inside organic molecules is
30: discussed. Organic molecule doping strongly modifies the optical
31: property of BNNT, and the absorption edge is red-shifted under
32: static transverse electric field.
33: \end{abstract}
34: 
35: \pacs{75.75.+a, 73.22.-f, 72.80.Rj}
36: 
37: \maketitle
38: 
39: \section{introduction}
40: 
41: Nanotube is a very important actor in nanodevice applications due to
42: its novel properties. Compared to the more widely studied carbon
43: nanotube (CNT), boron nitride nanotube (BNNT) has its own
44: distinctive properties. Besides the excellent mechanical stiffness
45: and thermal conductivity comparable to CNT, \cite{review} BNNT has a
46: unified electronic structure regardless of its diameter, chirality,
47: and the number of walls of the tube. \cite{bn1, bn2, bn3} BNNT also
48: shows pronounced resistance to oxidation, and it is stable up to 700
49: $^{\circ}$C in air, while CNT survives only below 400 $^{\circ}$C.
50: \cite{bn5} All these properties make BNNT an attractive candidate
51: for nano-electronics.
52: 
53: Pristine BNNT is a wide-gap semiconductor. For electronics
54: application, it is desirable to make $p$- or $n$-type doping to
55: BNNT. Our previous study shows that the electronic structure of BNNT
56: can be modified by organic molecule encapsulation. \cite{bnhe}
57: Electrophilic molecule introduces acceptor states in the wide gap of
58: BNNT close to the valence band edge (VBE), which makes the doped
59: system a $p$-type semiconductor. However, with typical nucleophilic
60: organic molecules, instead of shallow donor states, only deep
61: occupied molecular states are observed. There is a significant
62: electron transfer from BNNT to an electrophilic molecule, while the
63: charge transfer between a nucleophilic molecule and BNNT is
64: negligible.
65: 
66: On the other hand, previous theoretical studies \cite{field0,
67: field1, field2} showed that the band gap of CNT or BNNT can be
68: reduced and even closed by applying a transverse electric field, due
69: to the giant Stark effect (GSE). \cite{field1} The electric field
70: can mix the nearby subband states in the valence band (VB) complex
71: and conduction band (CB) complex separately, leading to an electric
72: field dependence for the band gap. This effect is more remarkable in
73: BNNTs than in CNTs, due to the reduced screening of the electric
74: field in BNNTs. The GSE was lately confirmed experimentally using
75: the bias dependent scanning tunneling microscopy and scanning
76: tunneling spectroscopy. \cite{field3}
77: 
78: Since the electronic structure of BNNT can be tuned by both organic
79: molecule encapsulation and transverse electric field, it is thus
80: interesting to see what will happen if we apply both. In this paper,
81: by performing density functional theory (DFT) calculations, we study
82: the electronic structures of organic molecule encapsulated BNNTs
83: under a transverse electric field. The electrostatic shielding of
84: BNNT to the inside molecules and the effects of organic molecule
85: doping and electric field on the optical properties of BNNT are also
86: studied.
87: 
88: Following our previous study, \cite{bnhe} several typical
89: electrophilic and nucleophilic molecules are considered for BNNT
90: encapsulation. The two electrophilic molecules studied are
91: tetracyano-p-quinodimethane (TCNQ) and
92: tetrafluorocyano-p-quinodimethane (F4TQ). Three nucleophilic
93: molecules are selected: tetrakis(dimethylamino)ethylene (TDAE),
94: anthracene (ANTR), and tetrathia-fulvalene (TTF). The (16,0) BNNT is
95: chosen as a prototype in this study. We name the organic molecules
96: capsulated BNNT as M@BN, where M is the name of the doped organic
97: molecules.
98: 
99: \begin{figure}[!hbp]
100: \includegraphics[width=8cm]{total.eps}
101: \caption{(Color online) Optimized geometrical structures of the M@BN
102: systems under a 0.5 V/\AA\ electric field. B in dark-salmon, N in
103: blue, C in grey, H in yellow, F in aqua, and S in goldenrod. }
104: \label{fig:geo}
105: \end{figure}
106: 
107: The rest of the paper is organized as follows: In Sec. II, the
108: theoretical approach and computational details are briefly
109: described. In Sec. III, the calculated band structures, charge
110: transfers between BNNT and molecules, shielding effect under
111: electric field, and optical properties are discussed. Finally, we
112: conclude in Sec. IV.
113: 
114: \section{Computational Details}
115: 
116: To investigate the geometry structure and electronic states of the
117: organic molecule encapsulated BNNT, we performed first-principles
118: spin-polarized DFT calculations. The computational details of
119: geometry optimization and band structure calculation can be found in
120: our previous study, \cite{bnhe} and we only give a brief summary
121: here. We used the projector augmented-wave (PAW) method \cite{20,21}
122: implemented in the Vienna Ab Initio Simulation Package (VASP).
123: \cite{17,18} The Perdew, Burke and Ernzerhof (PBE)
124: exchange-correlation functional was chosen. \cite{19} We note that
125: more rigorous band structure calculations are required to yield
126: correct band gaps. \cite{Baumeier07} However, PBE functional has
127: been widely used in similar systems, and it should give correct
128: trends which are interested in this study. The energy cutoff for
129: plane-waves was 400 eV. In our calculation, each BNNT was separated
130: by 10 \AA\ of vacuum, and the minimum distance between periodically
131: repeated organic molecules along the BNNT tube axis was
132: $\sim$8.7\AA. A 4$\times$1$\times$1 $\Gamma$-centered Monhkorst-Pack
133: $k$-point grid was used for Brillouin zone sampling.
134: 
135: 
136: The way to handle electric field in VASP is adding an artificial
137: dipole sheet in the middle of the vacuum part in the periodic
138: supercell. \cite{vaspdipole1, vaspdipole2} In our study, the
139: direction of the electric field was chosen to be perpendicular to
140: the nanotube's axis (along the $y$ direction), with a magnitude of
141: 0.20 or 0.50 V/\AA. As discussed later, the selected fields are
142: strong enough to make significant change of the electronic
143: structure, and weak enough to avoid artificial field emission.
144: 
145: 
146: \begin{figure}[!hbp]
147: \includegraphics[width=8cm]{E0.5vacuum.eps}
148: \caption{Electrostatic potential (in volt) across the BNNT in the
149: M@BN systems under 0.5 V/\AA\ external electric field. The top of
150: the highest occupied bands are marked by dash lines. The numbers on
151: the top of the curves are the effective electric field (in V/\AA),
152: calculated by the slope of the line connecting the left
153: high-potential point and the right low-potential point in vacuum.}
154: \label{fig:potential}
155: \end{figure}
156: 
157: All linear optical properties, such as absorption coefficient and
158: refractive index, can be obtained as functions of dielectric
159: function. The real part of the dielectric function can be obtained
160: from its imaginary part by Kramer-Kronig transformation, and the
161: imaginary part $\epsilon_2$ can be calculated based on the
162: independent-particle approximation, \cite{Adolph0108} i.e.,
163: \begin{equation}
164:  \epsilon_2(\omega)=\frac{4\pi}{\Omega\omega^2}\sum_{i,j}\sum_\mathbf{k}
165:  W_\mathbf{k}|P_{ij}|^2\delta(E_{\mathbf{k}j}-E_{\mathbf{k}i}-\omega)
166: \end{equation}
167: where $\Omega$ is the unit cell volume, and $\omega$ is the photon
168: energy. Summations for $i$ and $j$ are performed for valence and
169: conduction bands, respectively. $P_{ij}$ denotes the dipole
170: transition matrix elements obtained from the self-consistent band
171: structure calculations and $W_\mathbf{k}$ is the $\mathbf{k}$ point
172: weighting. The same method has been applied to h-BN, and the results
173: are in reasonably good agreement with experiments. \cite{Guo0502}
174: The number of bands used for dielectric function calculations are
175: the sum of the electron number and the ion number. The number of
176: occupied bands is half of the number of electron, and the number of
177: unoccupied bands are more than 380 for all the systems studied in
178: this paper. The optical polarization in dielectric function
179: calculations is perpendicular to the nanotube axis. In this way, for
180: some parts of the tube wall, the optical polarization is nearly
181: perpendicular to the BN layer, while for the other parts of the tube
182: wall, it is roughly parallel to the BN layer.
183: 
184: \section{RESULTS AND DISCUSSION}
185: 
186: \subsection{Geometric and band structures}
187: 
188: As reported in our previous study, \cite{bnhe} after the organic
189: molecules are encapsulated into BNNT, their geometries changes
190: little. This suggests the weak interaction between BNNT and the
191: encapsulated molecules, due to the large distances between them. The
192: separations between neighboring molecules and between molecule and
193: BNNT are larger than additions of the corresponding atomic van der
194: Waals radius. Optimizations under the transverse electric field gave
195: similar structure as in the zero-field case. The optimized
196: geometrical structures under a 0.5 V/\AA\ electric field are shown
197: in Fig. \ref{fig:geo}. The relaxed geometries of the M@BN systems
198: before and after applying electric field change little. The largest
199: change of bond length, comparing to the geometry under zero field,
200: is smaller than 0.01 \AA. Consistently, the direction of the
201: transverse field does not affect the geometric and electronic
202: structures. Our test calculation gave almost identical band
203: structures for F4TQ@BN under 0.5 V/\AA\ transverse electric field
204: along two different directions. For simplicity, in the rest part of
205: the paper, all electronic structure calculations were done based on
206: the geometrical structures obtained without electric field.
207: 
208: 
209: 
210: Before calculating the electronic structure of M@BN with transverse
211: electric field applied, two important issues are needed to be
212: discussed. One is that the vacuum region should be broad enough to
213: avoid the overlap between the charge density and the artificial
214: dipole sheet used to generate electric field. The width of the
215: vacuum layer between two neighboring BNNTs in our calculation is 10
216: \AA, and the work function of (16,0) BNNT is 6.36 eV. The charge
217: density of BNNT roughly decay as $\exp(-\sqrt{\phi}r)$, where $\phi$
218: is the work function in eV and $r$ is the distance from the BNNT in
219: \AA. \cite{vaspdipole3} The charge density thus decays one magnitude
220: every 0.91 \AA. The dipole sheet is 5 \AA\ far from the BNNT, and
221: there is thus only negligible charge density there. The other issue
222: is that the vacuum region should not be too broad so as to attract
223: electrons to vacuum (field emission). In Fig. \ref{fig:potential},
224: we plot the electrostatic potential of the M@BN system under a 0.5
225: V/\AA\ electric field, the strongest electric field used in our
226: calculations. We can see that the lowest potential in vacuum region
227: is still higher than the edge of the highest occupied band (HOBand),
228: which confirms that the artificial field emission has not been
229: introduced in this study.
230: 
231: \begin{figure}[!hbp]
232: \includegraphics[width=8cm]{BNNTBand.eps}
233: \includegraphics[width=8cm]{charge.eps}
234: \caption{Band structures (from $\Gamma$ to X) of the pristine BNNT
235: under (a) 0.0, (b) 0.2, and (c) 0.5 V/\AA\ electric fields. Fermi
236: levels are marked by dashed lines. Density distribution for HOBand,
237: LUBand, and LUBand+1 in pristine BNNT under (d-f) 0.0 and (g-i) 0.5
238: V/\AA\ electric fields. The electric field direction is from up to
239: down.} \label{fig:bandbn}
240: \end{figure}
241: 
242: We first calculate the band structure of the pristine BNNT under the
243: 0.2 and 0.5 V/\AA\ electric fields, and the results are shown in
244: Fig. \ref{fig:bandbn}. The band gap decreases from 4.45 to 3.73 and
245: 1.98 eV, when the field increases from 0 to 0.2 and 0.5 V/\AA,
246: respectively. Our result is consistent with the previous theoretical
247: study on similar systems. \cite{field1} Such band gap decrease is a
248: result of the extensive mixing among the subband states within the
249: valence and conduction band complex under electric filed. The band
250: mixing is also reflected by the band-resolved charge density
251: distributions, as shown in Fig. \ref{fig:bandbn}, where the density
252: of the HOBand, the lowest unoccupied band (LUBand), and LUBand+1
253: under 0.0 and 0.5 V/\AA\ field are plotted. The orbital density move
254: along the electric field for occupied states, and against the field
255: direction for unoccupied states. From the density distributions, the
256: LUBand state can be identified as a nearly free electron (NFE)
257: state. \cite{field1}
258: 
259: \begin{figure}[!hbp]
260: \includegraphics[width=8cm]{E0.5bandcompare.eps}
261: \caption{Band structures (from $\Gamma$ to X) of the pristine BNNT
262: and the M@BN systems under 0.5 V/\AA\ electric field. Fermi levels
263: are marked by dashed lines. } \label{fig:band0.5}
264: \end{figure}
265: 
266: The band structures of the M@BN systems under 0.2 and 0.5 V/\AA\
267: electric field are then calculated. Similar to results under zero
268: electric field, \cite{bnhe} the bands are either mainly from
269: molecules or mainly from BNNT. The molecular bands are much flatter
270: than the BNNT bands. Since there is no qualitative difference
271: between the band structure of the M@BN systems under 0.2 and 0.5
272: V/\AA\ electric fields, we will focus on the 0.5 V/\AA\ case (Fig.
273: \ref{fig:band0.5}). For the two electrophilic molecules (TCNQ and
274: F4TQ), the lowest unoccupied molecular bands become very close to
275: the BNNT VBE, which means good $p$-type doping. It also suggests a
276: considerable charge transfer between the molecules and the BNNT. We
277: define the gap between the lowest unoccupied molecular band and the
278: BNNT VBE as $E_p$, and the gap between the highest occupied
279: molecular band and the BNNT conduction band edge (CBE) as $E_n$.
280: Their values are listed in Table \ref{tbl:eg}. For the nucleophilic
281: organic molecules TDAE and TTF, $E_n$ drops from 2.30 and 2.85 eV at
282: zero field \cite{bnhe} to 1.08 and 1.54 eV, respectively. However,
283: it is still too large to form good $n$-type doping. For ANTR@BN, the
284: highest occupied molecular band is pushed into the BNNT valence band
285: manifest, and there is only an unoccupied molecular band close to
286: the BNNT CBE. Therefore, neither $n$-type or $p$-type doping is
287: introduced to BNNT by ANTR encapsulation.
288: 
289: \begin{table}[b]
290: \caption{Energy gaps $E_p$, $E_n$, and $E_g^{BN}$ of the pristine
291: BNNT and the M@BN systems under electric field $\mathcal{E}$. The
292: energy gaps are in eV, and the electric field $\mathcal{E}$ is in
293: V/\AA. } \label{tbl:eg}
294: \begin{tabular}{cccccccc}
295:  \hline\hline
296:  Gap & $\mathcal{E}$ & BN & TCNQ & F4TQ & TDAE & TTF & ANTR   \\
297:  \hline
298: $E_p$      & 0.0 &  --- & 0.15 & 0.09 & 4.37 & 3.55 & 2.96 \\
299:            & 0.2 &  --- & 0.17 & 0.11 &  --- & 3.25 & 2.65 \\
300:            & 0.5 &  --- & 0.09 & 0.05 &  --- &  --- & 1.93 \\
301: $E_n$      & 0.0 &  --- &  --- &  --- & 2.30 & 2.85 & 3.71 \\
302:            & 0.2 &  --- &  --- &  --- & 1.86 & 2.49 & 3.40 \\
303:            & 0.5 &  --- &  --- &  --- & 1.08 & 1.54 &  --- \\
304: $E_g^{BN}$ & 0.0 & 4.45 & 4.45 & 4.48 & 4.45 & 4.46 & 4.43 \\
305:            & 0.2 & 3.73 & 4.00 & 4.18 & 3.63 & 3.79 & 3.81 \\
306:            & 0.5 & 1.98 & 2.79 & 2.91 & 2.31 & 2.14 & 2.18 \\
307:  \hline\hline
308: \end{tabular}
309: \end{table}
310: 
311: To consider the molecular effect on the intrinsic electronic
312: structure of BNNT under electric field, we define the energy gap
313: between the CBE and VBE of BNNT as $E_g^{BN}$. As shown in Table
314: \ref{tbl:eg}, without external field, it is not very sensitive to
315: the type of molecules. However, $E_g^{BN}$ for TCNQ@BN and F4TQ@BN
316: under electric field is much larger than other systems under the
317: same field. This can be easily understood by their electronic
318: structures. Under strong external field, the energy gaps of TCNQ@BN
319: and F4TQ@BN become very small, which means enhanced metallicity and
320: stronger electrostatic screening. With stronger screening, the
321: reduction of electrostatic potential along the direction of electric
322: field is slower (see Fig. \ref{fig:potential}), therefore,
323: $E_g^{BN}$ reduction compared to pristine BNNT and other M@BN
324: systems is smaller. Similar behavior has been observed in our study
325: on defective BNNTs under an electric field. \cite{bnhu}
326: 
327: 
328: 
329: \subsection{Charge transfer}
330: 
331: An important property of the molecule encapsulated system is the
332: charge transfer between the molecules and BNNT. We calculate the
333: charge transfer using the method described in our previous work,
334: \cite{bnhe} originally proposed by Lu et al..\cite{charge1} Briefly
335: speaking, the boundary between the organic molecule and the BNNT is
336: determined by the maximum/minimum position of the
337: cylinder-integrated differential electron density curve, and the
338: corresponding maximum/minimum value is the charge transfer value.
339: The differential electron density is defined as the difference
340: between electron densities of the M@BN system under electric field
341: and the sum of charge densities of BNNT and molecules without
342: electric field. The range of the radial position $R_b$ of boundary
343: between molecule and BNNT can be estimated by the density
344: distribution of the BNNT bands. For electrophilic molecules, $R_b$
345: should be inside the region where the BNNT HOBand has a high density
346: ($R_b<6$ \AA), while for nucleophilic molecules, $R_b$ should be
347: inside region where the BNNT LUBand has a high density ($R_b<3.5$
348: \AA). \cite{bnhe}
349: 
350: \begin{figure}[!hbp]
351: \includegraphics[width=8cm]{E0.5charge.eps}
352: \caption{(Color online) The cylinder-integrated differential
353: electron density curves for the the pristine BNNT and the M@BN
354: systems under a 0.5 V/A electric field. The blue vertical lines
355: indicate the radial boundary between molecules and the BNNT, in the
356: viewpoint of charge transfer. The black vertical line in the first
357: panel marks the radius of the BNNT.} \label{fig:E0.5charge}
358: \end{figure}
359: 
360: For pristine BNNT, the differential electron density is just the
361: density change before and after applying electric field. As shown in
362: Fig. \ref{fig:E0.5charge}, the electron density increases inside the
363: nanotube wall, while it decrease outside the wall. Of course, the
364: absolute value of this change is very small ($\sim$0.01 electron).
365: For the electrophilic organic molecules TCNQ and F4TQ, under 0.5
366: V/\AA\ electric field, there are 0.28 and 0.38 electrons per
367: molecule transfer from BNNT, about twice as the value without field
368: (0.10 and 0.24 electron). \cite{bnhe} This is due to the closer
369: molecular state in energy to VBE. The boundary radius is 4.3 \AA,
370: which is basically the midpoint between the molecule and the BNNT as
371: in the no field case. For the nucleophilic organic molecules TDAE,
372: TTF and ANTR, the charge transfers form molecules to BN also
373: increase, which is 0.006, 0.011, and 0.008 electron per molecule,
374: respectively (0.004, 0.008, and 0.005 electrons without external
375: field). Such small electron transfers also indicate that the doped
376: system is not a good $n$-type semiconductor.
377: %, which is mainly contributed by the NFE states ({\it empty states})
378: 
379: 
380: \subsection{Shielding effect}
381: 
382: It is interesting to study the effects of BNNT to the geometry and
383: electronic structure of the inside molecules under an electric
384: field. In Fig. \ref{fig:dos}, we compare the molecular partial
385: densities of states (PDOS) with and without external field for two
386: representative molecules: an electrophilic one (F4TQ) and a
387: nucleophilic one (TTF). The calculated PDOS shows different
388: behaviors for these two types of molecules. For the electrophilic
389: organic molecules F4TQ (Fig. \ref{fig:dos}a), there is an nearly
390: rigid upshift of the molecular PDOS after applying the electric
391: field. This is due to the electron transfer from BNNT to the
392: molecules. For the nucleophilic organic molecules TTF (Fig.
393: \ref{fig:dos}c), the molecular PDOS almost does not change after
394: applying the electric field, which is consistent with the little
395: charge transfers between BNNT and the molecules.
396: 
397: \begin{figure}[!hbp]
398: \includegraphics[width=8cm]{dos.eps}
399: \caption{(Color online) Molecular PDOS of (a) F4TQ@BN and (c) TTF@BN
400: under 0.0 (solid) and 0.5 (dashed) V/\AA\ electric field. DOS of
401: individual (b) F4TQ and (d) TTF molecules under 0.0 (solid) and 0.5
402: (dashed) V/\AA\ electric field.} \label{fig:dos}
403: \end{figure}
404: 
405: We also check the shielding effect by applying electric field to
406: individual molecules without a BNNT encapsulation. First, we relax
407: the geometry of the individual molecules under the 0.5 V/\AA\
408: electric field. Contrasting to the encapsulated case, we found a
409: significant geometry relaxation after applying the electric field.
410: Then, we fix the molecular geometry to the one as inside BNNT, and
411: calculate the electronic structure under zero and finite electric
412: field. The effect of electric field to the molecular electronic
413: structures for individual molecules is different from that for
414: molecules encapsulated by BNNT (see Fig. \ref{fig:dos}). For F4TQ
415: molecule, the DOS energy shift caused by electric field is smaller
416: than that for PDOS of F4TQ@BN. However, we can see a notable change
417: of the overall DOS shape after applying the electric field. For TTF,
418: electric field shifts the DOS of individual molecules, but it does
419: not change the molecular PDOS. In summary, our results indicates,
420: although not good as CNT, the BNNT still provides a significant
421: electrostatic shielding for the inside molecules: the geometry of
422: the inside molecules and the electronic structures of nucleophilic
423: molecules are almost not affected by an external field under the
424: protection of BNNT.
425: 
426: \subsection{Optical properties}
427: 
428: The calculated imaginary part of the dielectric function of the
429: pristine BNNT without external electric field is shown in the inset
430: of Fig. \ref{fig:optics}a. A much denser $k$-point sampling is
431: required to obtain smooth curves. \cite{Guo0502} However, the main
432: features of $\epsilon_2$ are already presented in our results. And
433: our calculated $\epsilon_2$ of prinstine BNNT agrees well with the
434: previous theoretical result on the zigzag BNNTs. \cite{Guo0502}
435: Mainly, there are three peaks at about 5.6, 10.6, and 13.5 eV,
436: respectively. The strong peak around 5.6 eV corresponds to the
437: $\pi\rightarrow\pi^*$ interband transitions. The two other peaks
438: higher in energy and weaker in intensity are associated with
439: interband transitions also involving the $\sigma$ bands. We note
440: that the peak around 10.6 eV does not appear in our test calculation
441: with the optical polarization along the tube axis. By optical
442: selection rule analysis, \cite{Guo0502} the 10.6 eV peak should
443: correspond to $\pi\rightarrow\sigma^*$ and $\sigma\rightarrow\pi^*$
444: transitions, and the 13.5 peak mainly contributed by
445: $\sigma\rightarrow\sigma^*$ transitions. In Fig. \ref{fig:optics}a,
446: we also plot $\epsilon_2$ of the pristine BNNT under transverse
447: electric field. Similar to a previous study on pristine BNNT,
448: \cite{field5} the whole imaginary dielectric function is not
449: strongly affected by the static field, but the absorption edge is
450: red-shifted.
451: 
452: \begin{figure}[!hbp]
453: \includegraphics[width=8cm]{opticsper.eps}
454: \caption{(Color online) The imaginary part ($\epsilon_2$) of the
455: theoretical dielectric function of the BNNT and the M@BN systems,
456: with (red) and without (blue) the external static electric field.
457: Inset: $\epsilon_2$ plotted in a larger energy scale. }
458: \label{fig:optics}
459: \end{figure}
460: 
461: 
462: With molecule intercalation, all the three peaks are still clearly
463: shown in the M@BN system. However, new features appear in the
464: low-energy region from 0 to 5 eV. These peaks below the BNNT
465: absorption edge ($\sim$4.5 eV) are clearly from transitions between
466: molecular bands and the BNNT bands or between molecular bands
467: themselves. For the two electrophilic molecules, there are peaks
468: close to the zero energy. This is consistent with the very narrow
469: band gap calculated for these two systems. There is no significant
470: peaks below 4.5 eV for TDAE@BN, and the new peaks for TTF@BN and
471: ANTR@BN are close to the BNNT absorption edge. Our results indicate
472: that the optical property of BNNT can be strongly modified by
473: different organic molecule encapsulation.
474: 
475: Next, we study the case where both molecule intercalation and
476: electric field exist. The theoretical imaginary part of the
477: dielectric functions of the molecule doped BNNT under a 0.5 V/\AA\
478: transverse static electric field are also presented in Fig.
479: \ref{fig:optics}. The three peaks from BNNT are still not strongly
480: affected by the static field. However, if we look into the details
481: below 5 eV, we can see three different behaviors. For transitions
482: between valence band and conductance bands of BNNT, the electric
483: field GSE broadens the band manifests, thus broadens the absorption
484: peaks too. This broadening effect red-shifts the absorption edge of
485: BNNT as shown in Fig. \ref{fig:optics}d. For transitions between
486: molecular bands, due to electrostatic shielding of BNNT, the
487: electric field does not strongly affect the peaks, as shown in Fig.
488: \ref{fig:optics}e and \ref{fig:optics}f. For transitions between
489: BNNT bands and molecular bands, the electric field affects both the
490: amplitude of the peaks and the peak positions, as shown in Fig.
491: \ref{fig:optics}b and \ref{fig:optics}c.
492: 
493: 
494: \section{conclusion}
495: 
496: In summary, we calculate the electronic structure and optical
497: properties of BNNT under transverse electric field with different
498: molecule intercalations. When applying transverse electric field,
499: the band gap of BNNT decrease, while the band structures of organic
500: molecules change little, or only with a rigid shift. After applying
501: electric field, there is more charge transfer between molecules and
502: nanotube in the $p$-type semiconductors TCNQ@BN and F4TQ@BN. BNNT
503: provides relatively good electrostatic shielding for the geometry
504: relaxation of the inside organic molecular chain, and for the
505: electronic structure of nucleophilic molecules. Organic molecule
506: doping strongly modifies the optical properties of the composite
507: M@BN systems. The absorption edge is slightly red-shifted under
508: static electric field.
509: 
510: 
511: \begin{acknowledgements}
512: The postprocessing routine downloaded from the homepage of Prof.
513: Furthmuller has been used in the optical property calculations. This
514: work is partially supported by the National Natural Science
515: Foundation of China (50721091, 20533030, 50731160010), by National
516: Key Basic Research Program under Grant No.2006CB922004, by the
517: USTC-HP HPC project, and by the SCCAS and Shanghai Supercomputer
518: Center.
519: \end{acknowledgements}
520: 
521: \begin{thebibliography}{99}
522: 
523: \bibitem{review} D. Golberg, Y. Bando, C. C. Tang, and C. Y. Zhi, Adv. Mater. {\bf 19}, 2413(2007).
524: 
525: \bibitem{bn1} A. Rubio, J. L. Corkill, and M. L. Cohen, Phys. Rev. B  {\bf 49}, 5081 (1994).
526: 
527: \bibitem{bn2} N. G. Chopra, R. J. Luyken, K. Cherrey, V. H. Crespi, M. L. Cohen, S. G. Louie, and A. Zettl, Science {\bf 269}, 966 (1995).
528: 
529: \bibitem{bn3} X. Blase, A. Rubio, S. G. Louie, and M. L. Cohen, Europhy. Lett. {\bf 28}, 335 (1994).
530: 
531: %\bibitem{bn4} M.Terauchi, M.Tanaka, K.Suzuki, A.Ogino, and K.Kimura, Chem. Phys. Lett. {\bf 324}, 359 (2000).
532: 
533: \bibitem{bn5} Y. Chen, J. Zou, S. J. Campbell, and G. L. Caer, Appl. Phys. Lett. {\bf 84}, 2430 (2004).
534: 
535: \bibitem{bnhe} W. He, Z. Li, J. Yang, and J. G. Hou, J. Chem. Phys., 128, 164701 (2008).
536: 
537: \bibitem{field0} J. O'Keeffe, C. Wei, and K. Cho, Appl. Phys. Lett. {\bf 80}, 676 (2002).
538: 
539: \bibitem{field1} K. H. Khoo, M. S. C. Mazzoni, and S. G. Louie, Phys, Rev. B {\bf 69}, 201401 (2004)
540: 
541: \bibitem{field2} C. W. Chen, M. H. Lee, adn S. J. Clark, Nanotechnology {\bf 15}, 1837 (2004)
542: 
543: \bibitem{field3} M. Ishigami, J. D. Sau, S. Aloni, M. L. Cohen, and A. Zettl, Phys, Rev. Lett. {\bf 94}, 056804 (2005)
544: 
545: \bibitem{20} G. Kresse, D. Joubert, Phys. Rev. B {\bf 59}, 1758 (1999).
546: 
547: \bibitem{21} P. E. Blochl, Phys. Rev. {\bf 50}, 17953 (1994).
548: 
549: \bibitem{17} G. Kresse, J. Furthm¨šller, Phys. Rev. B {\bf 64}, 11169 (1996).
550: 
551: \bibitem{18} G. Kresse, J. Furthmuller, Comput. Mater. Sci. {\bf 6}, 15 (1996).
552: 
553: \bibitem{19} J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. {\bf 77}, 3865 (1996).
554: 
555: \bibitem{Baumeier07} B. Baumeier, P. Kruger, and J. Pollmann, Phys.
556: Rev. B {\bf 76}, 085407 (2007).
557: 
558: \bibitem{vaspdipole1} J. Neugebauer and M. Scheffler, Phys. Rev. B 46, 16067 (1992).
559: 
560: \bibitem{vaspdipole2} G. Kresse and J. Furthmuller, VASP Guide.
561: 
562: \bibitem{Adolph0108} B. Adolph, J. Furthmuller, and F. Bechsted,
563: Phys. Rev. B {\bf 63}, 125108 (2001)
564: 
565: \bibitem{Guo0502} G. Y. Guo and J. C. Lin, Phys. Rev. B {\bf 71},
566: 165402 (2005)
567: 
568: \bibitem{vaspdipole3}P. J. Feibelman, Phys. Rev. B 64, 125403(2001).
569: 
570: \bibitem{bnhu} S. Hu, Z. Li, X. C. Zeng, and J. Yang, J. Phys. Chem.
571: C, ASAP article (2008)
572: 
573: \bibitem{charge1} X. Lu, M. Grobis, K. H. Khoo, S. G. Louie, and M. F. Crommie, Phys. Rev. B 70, 115418 (2004).
574: 
575: %\bibitem{shield1} E. N. Brothers, G. E. Scuseria, and K. N. Kudin, J Chem Phys {\bf 124}, 041101
576: %(2006)
577: 
578: %\bibitem{shield3} L. Wang, J. Lu, L. Lai, W. Song, M. Ni, Z. Gao, and W. N. Mei, J Phys Chem C {\bf 111}, 3285
579: %(2007)
580: 
581: %\bibitem{shield2} B. Kozinsky and N. Marzari, Phys Rev Lett {\bf 96}, 166801
582: %(2006)
583: 
584: \bibitem{field5} C. W. Chen, M. H. Lee, and Y. T. Lin, Appl. Phys. Lett. {\bf 89}, 223105 (2006).
585: 
586: %\bibitem{organic0} T. Takenobu, T. Takano, M. Shiraishi, Y. Murakami, M. Ata, H. Kataura, Y. Achiba, and Y. Iwasa, Nat. Mater. {\bf 2}, 683 (2003).
587: 
588: %\bibitem{organic1} J. Lu, S. Nagase, D. Yu, H. Ye, R. Han, Z. Gao, S. Zhang, and L. Peng, Phys. Rev. Lett.{\bf 93}, 116804 (2004).
589: 
590: %\bibitem{organic2} W. Liang, J. Yang, and J. Sun, Appl. Phys. Lett. {\bf 86}, 223113 (2005).
591: 
592: %\bibitem{Yamashiro0310} A. Yamashiro, Y. Shimoi, K. Harigaya, and K.
593: %Wakabayashi, Phys. Rev. B \textbf{68}, 193410 (2003)
594: 
595: \end{thebibliography}
596: 
597: \end{document}
598: