You are on page 1of 13

P1: IZO

Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Structural Chemistry, Vol. 14, No. 5, October 2003 (°


C 2003)

Density Functional Tight-Binding Studies


of Carbon Nanotube Structures

Zenaida Peralta-Inga,1,2 Sylke Boyd,1 Jane S. Murray,1


Charles J. O’Connor,1,2 and Peter Politzer1–3

Received March 23, 2002; accepted April 8, 2002

A density functional tight-binding self-consistent charge approach has been used to study the struc-
tures and elastic properties of nine model carbon nanotubes of different helicities and diameters
between 5.5 and 10.8 A. The systems contain from 112 to 268 atoms and were optimized under
periodic boundary conditions in the axial direction. Both the carbon networks and the overall tube di-
mensions were optimized. Most of the C C bond lengths are slightly lengthened relative to graphene
(two-dimensional graphite); the others remain essentially the same or are shorter. There is overall
a longitudinal compression of the tube. The strain energy per atom, relative to graphene, varies in-
versely with the square of the tube radius. The Young’s moduli decrease with increasing radius but do
not depend upon chirality. The Poisson ratios are nearly constant. The consequences of removing an
electron from each system were also investigated. In most instances, the tube dimensions were little
affected; in only a few cases is there a change in length or radius (positive or negative) as large as
0.10%. The Young’s moduli remain the same as for the neutral systems, but the Poisson ratios tend
to increase for metals and semimetals and to decrease for semiconductors.

KEY WORDS: Carbon nanotubes; structural properties; density functional tight-binding calculations
(DFTB-SCC).

INTRODUCTION A single-walled carbon nanotube can be viewed as


the result of rolling up a two-dimensional layer of graphite
Carbon nanotubes, first reported by Iijima [1], con- (“graphene”) [3, 5, 8], producing a tubular structure. The
tinue to be of considerable and increasing interest to the direction in which this rolling up is performed is specified
condensed-matter and materials research communities. by a vector c, defined in terms of the lattice vectors a1 and
One indication of this is the frequency with which the a2 (Fig. 1),
field has been reviewed in recent years [2–8]. Nanotubes
present themselves in a variety of structural forms, which c = na1 + ma2 ≡ (n, m) (1)
give rise to remarkable mechanical and electronic proper-
ties. For example, they have high levels of tensile strength and the nanotube is commonly labeled by means of the
[9–12], yet are very elastic and can withstand considerable integer indices (n, m). The vector c links points that are
(reversible) bending, twisting, and deformation [10, 13– superimposed in forming the tube, which, therefore, has
15]. Depending upon the details of their structures, they circumference |c|; its axis is along T, which, together with
range from metallic to semiconducting in nature [16–24]. c, defines the unit cell of the structure. Each set of indices
denotes a tube of different diameter and helicity; Fig. 1
shows the characteristic vectors c for some examples. In
1 Department of Chemistry, University of New Orleans, New Orleans, general, the systems (n, m) are chiral; the only excep-
Louisiana 70148. tions are (n, n), denoted “armchair,” and (n, 0), “zigzag.”
2 Advanced Materials Research Institute, University of New Orleans,

New Orleans, Louisiana 70148.


(Because of the hexagonal symmetry of graphene, only
3 To whom correspondence should be addressed; email: ppolitze@ the cases 0 < |m| < n need be considered; thus (0, n) is
uno.edu equivalent to (n, 0).) The chiral angle θ is that between

431
1040-0400/03/1000-0431/0 °
C 2003 Plenum Publishing Corporation
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

432 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

carbon nanotube electrodes, both of them elongate, but to


different extents, producing a curvature in the electrode
system. It was suggested that this could be the basis for
an electromechanical actuator that would provide a very
high work density.
In the present work, we have used a computational
procedure to investigate the structures and several prop-
erties of a group of carbon nanotubes with various he-
licities. Specifically, we present geometries, strain ener-
gies, Young’s moduli, and Poisson ratios. The size of
the systems studied is limited primarily by computational
resources. We have to limit ourselves to high-symmetry
low-index tubes. However, the properties investigated are
found to depend more on the tube radius than on the chi-
rality. The variation of the mechanical properties with re-
spect to increasing tube radius allows conclusions about
larger radii, even extrapolation to graphite. We have also
examined the effects of removing an electron, to impose
Fig. 1. Diagram showing scheme used to describe nanotubes. The indices an overall positive charge on the system. The results are
(n, m) would indicate the tube with axis parallel to T that is formed by
rolling up the graphene sheet in the direction of the vector c = na1 +
inconclusive with respect to geometric and mechanical
ma2 ≡ (n, m), where a1 and a2 are the lattice vectors. The magnitude of properties, since we are working at the very limit of the
c would be the circumference of the tube; T and c define the unit cell. For method used. However, we find them interesting enough,
the vector c in this figure, (n, m) = (4, 2). For illustrative purposes, the for an initial approach, to include in the discussion.
vectors characterizing three other nanotubes—(4, 0), (3, 3), and (0, 3)—
are also shown.

PROCEDURE
a1 and c; its range is given by 0◦ ≤ |θ| ≤ 30◦ , where the We used a density functional tight-binding approach
two extrema correspond to the achiral zigzag (θ = 0◦ ) (DFTB), with a minimal valence basis set of Slater-type
and armchair (θ = 30◦ ) structures. The chiral angle can orbitals [31, 32]. It involves a second-order expansion of
be evaluated for each (n, m) via Eq. (2) [3, 5]: the Kohn–Sham total energy, with respect to fluctuations
√ in the electronic density, and a self-consistent redistribu-
m 3
tan θ = (2) tion of atomic charges (SCC). Our choice of method was
2n + m
guided by computational efficiency, availability, and the
Single-walled nanotubes frequently assemble into possibility of modeling periodic systems. It is particularly
bundles of parallel tubes (“ropes”). Another structural pos- relevant that the DFTB technique has been applied suc-
sibility is multiwalled systems, which are composed of cessfully to a variety of carbon systems (among others),
several coaxial tubes. including amorphous state [28], C60 [29], and diamond
It was predicted theoretically [16–19] and confirmed [30]. The DFTB-SCC procedure allows the treatment of
experimentally [20–22, 24] that single-walled carbon nan- systems with about 200 atoms under periodic boundary
otubes can be metallic, semimetallic, or semiconducting, conditions. The time investment for one energy calcula-
depending upon their helicities and diameters. Those hav- tion on a Compaq XP1000 for a tube with 224 atoms is
ing n = m are metallic, those for which (n − m)/3 is a about 1.5 min.
nonzero integer are semimetallic (i.e., small-gap semi- The limitations of DFTB lie in its neglect of exchange
conductors), and the rest are larger-gap semiconductors. and correlation effects, as well as its restriction to the 0
This raises the interesting possibility of creating metal- point of the Brillouin zone. The latter factor affects our
semiconductor junctions by joining tubes with different understanding of the electronic band structure of the nan-
helicities [25, 26]. otubes, since the semiconducting/metallic character stems
Baughman et al. [27] suggested that bundles of from π–π ∗ interactions at the K points with respect to
single-walled nanotubes can act as nano actuators, due graphene. A thorough discussion of this point from ge-
to their large strain response to external charging. They ometry considerations can be found in Odom et al. [8].
showed that when a voltage is introduced between two Although the shape of the valence band in the electronic
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 433

density of states is, in general, well described, the HOMO– would require at least 632 atoms. Such systems are in-
LUMO gap is usually overestimated. This renders the accessible to our computational resources.
method unsuitable for the study of negatively charged A conjugate gradient technique, which includes a
tubes. We did, however, examine systems from which one minimum residual force as the cutoff criterion, was used to
electron had been removed. optimize the carbon network geometry. This was done in
Each model tube was produced by rolling up a several stages, in which the system was expanded or com-
graphene sheet [3, 5]. Periodic boundary conditions were pressed by up to 10% along the tube axis; the resulting
applied in the direction of the tube axis, thus effectively energies were fit to a parabolic curve (Fig. 2). The final
providing for infinite length and avoiding the need to sat- minimum total energy then corresponds to having opti-
isfy unfilled valencies at the ends. In some cases, as for the mized the overall tube dimensions as well as the carbon
highly symmetric armchair and zigzag tubes, we used an network.
integer multiple of the basic periodicity length to achieve This procedure provides us with not only the energy
real-space dimensions sufficient to diminish interference of the optimum structure but also with considerable in-
with the periodic boundary conditions. Other tubes, e.g. formation about the dynamic behavior of the tube and al-
(7, 3), possess a periodicity length of 75.45 A, which lows us to calculate two of its elastic properties: Young’s

Fig. 2. Top, energy per atom relative to graphene vs. tube length for the (7, 1) model. The minimum
indicates the strain energy per carbon atom, as well as the equilibrium length L 0 . Bottom, tube radius
vs. tube length for the (7, 1) model. The equilibrium radius R0 is indicated.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

434 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

Table I. Properties of Neutral Model Systemsa

(n, m) Type Radius, |c|/(2π ) Chiral angle Model length Number of carbons

(7, 0) Zigzag 2.73 0.0 17.1 112


(7, 1) Chiral 2.95 6.6 21.4 152
(7, 2) Chiral 3.19 12.2 34.8 268
(7, 4) Chiral 3.76 21.1 13.7 124
(7, 7) Armchair 4.73 30.0 10.0 168
(8, 0) Zigzag 3.12 0.0 17.1 128
(8, 2) Chiral 3.58 10.9 13.0 112
(8, 4) Chiral 4.13 19.1 22.5 224
(8, 8) Armchair 5.41 30.0 10.0 192
a Distances, Angstroms; angles, degrees.

modulus, Y , which is a measure of longitudinal elasticity, 1R/R0


σ =− (4)
and the Poisson ratio, σ , which relates transverse to lon- 1L/L 0
gitudinal strain, in response to a given stress. These are
1L and 1R are the changes in the tube’s equilibrium
given by Eqs. (3) and (4) [11, 33]:
length and radius, L 0 and R0 , in response to a force F
F/A applied against the cross-sectional area A. We determined
Y = (3) Y from the parabolic E vs. L relationship, as shall be
1L/L 0

Fig. 3. Structures of optimized neutral (7, m) model systems. Colors of bonds indicate relative lengths: black < gray < white.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 435

discussed in the next section. To evaluate σ , we used the were determined from the indices (n, m) and the lattice
slope of the linear R(L) function (Fig. 2). vectors using formulas given by Saito et al. [3] and by
Harris [5]; the chiral angles were calculated with Eq. (2).
RESULTS AND DISCUSSION In graphene (two-dimensional graphite), all C C
bonds are equivalent, with a DFTB-SCC optimized length
Neutral Systems of 1.426 Å(vs. the experimental 1.420 Å in graphite [34]).
In contrast, the nanotubes show a variety of C C bond
The model nanotubes that are included in this work lengths, as can be seen in Figs. 3 and 4. Most of the bonds
are listed in Table I and shown in Figs. 3 and 4. Their radii in each tube are slightly stretched, relative to graphene, but

Fig. 4. Structures of optimized neutral (8, m) model systems. Colors of bonds indicate relative lengths: black < gray < white.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

436 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

Fig. 5. Bond length distributions in optimized neutral (7, m) model systems. The optimized C C distance in graphene is 1.426 A.

some are essentially unchanged or shortened by a small orientation of the carbon sp2 hybrid orbitals with respect to
amount. The actual distributions of bond lengths, given in the tube axis. The net effect of the strain associated with
Figs. 5 and 6, show that they fall into two or three groups the tubular structure is an overall decrease in length for
for each tube. For most of the systems, the more transverse each tube, by about 0.8%, compared to simply rolled-up
bonds are lengthened, while the more longitudinal remain graphene. The radii increase in some cases; it decreases
about the same or are shorter. In three instances, however, in others.
the reverse occurs (see Figs. 3–6); these are the (7, 7), Another measure of this strain, which also reflects the
(8, 4), and (8, 8). Two of these tubes are of the armchair distortion of each carbon from sp2 hybridization, is the in-
type, while the (8, 4) is approaching it. We are currently crease in energy per atom over that in graphene. This is
investigating this indication that a key factor may be the shown for each set of indices (n, m) in Fig. 7; the strain
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 437

Fig. 6. Bond length distributions in optimized neutral (8, m) model systems. The optimized C C distance in graphene is 1.426 A.

energy decreases with increasing tube radius, in good proaches were found to yield very similar values (Fig. 8).
agreement with the prediction from continuum elastic the- The magnitudes gradually diminish as the radius of the
ory, E str = C/r 2 [3, 36], in which C depends upon the tube becomes larger. The Young’s modulus of graphene
Young’s modulus and structural properties of the tube. A has been estimated to be about 1.2 TPa [11, 39], while the
least-squares fit of our data yields C = 1.99 eV Å2 /atom, experimental value for graphite is 1.06 TPa [3, 35]. Thus,
which is quite similar to the values reported in earlier work the decrease that we observe is consistent with graphene
[35–38], ranging from 2.00 to 2.16 eV Å2 /atom. being the limiting case of a nanotube as its radius in-
We used two approaches to compute Young’s mod- creases. We find no significant dependence upon chirality.
ulus, which is a measure of tensile strength. Both involve Our results are within the rather large range encompassed
the parabolic E vs. L relationship, E = aL2 + bL + c. by previous theoretical [11, 35, 39–42] and experimental
Replacing F/1L in Eq. (3) by 2a(= d 2 E/dL2 ) and L 0 [9, 10, 43–45] studies, some of which did not find a ra-
by −b/2a (found by setting dE/dL = 0) gives Y1 = dial dependence. Sanchez-Portal et al. [35] have discussed
−b/(π R02 ). Alternatively, inserting the strain energy ex- possible reasons for discrepancies.
pression E str = C/R02 leads to Y2 = −(bE str )/πC. The The Poisson ratio describes the response of the radius
second expression uses the C parameter that reflects all to a length dilation. The values were obtained by averaging
of the tubes. However, no assumption about the effective over the expansion and compression steps in the optimiza-
cross-sectional area of the tube has to be made. The two ap- tion processes. As shown in Fig. 9, most of them are in the
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

438 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

Fig. 7. Strain energy per atom vs. tube radius for neutral systems.

neighborhood of 0.27, regardless of the radius or the chiral out, leading to the bond length distributions shown in
angle. This is consistent with Hernandez et al. [39] and Lu Figs. 10 and 11. The groups now cover wider ranges than
[42], but is higher than was found by Van Lier et al. [11], for the neutral systems, the most extreme example being
Sanchez-Portal et al. [35], and Yakobson et al. [40, 41]. the (7, 2), which has at least one bond with every length
All agree that σ is independent of the tube radius. between 1.423 and 1.447 Å. (We are studying this tube fur-
ther, by higher-level methods.) The majority of the bonds
are again slightly longer than in graphene.
Positively Charged Systems In Table II are presented the percentage changes in
tube dimensions that accompany the loss of an electron.
The positively charged model nanotubes were pro- Most of the changes are rather small and can be consid-
duced by removing one electron from the neutral analogs. ered to be within the error bars of the study. In only four
The same geometry optimization procedure was carried cases is there an increase or decrease as large as 0.10%.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 439

Fig. 8. Young’s modulus vs. tube radius, for both neutral and positively charged systems. Results for both calculations of Y are
included (see text).

Our results suggest that the elongation upon charging that the other hand, increases in some instances and decreases
was observed by Baughman et al. [27] may not be a uni- in others (Fig. 9). In general, the metallic and semimetallic
versal feature of carbon nanotubes, but rather may depend systems are in the former category, while the semiconduc-
upon their chiralities and other physical factors. They used tors are in the latter.
bundles of tubes that were believed to have individual di-
ameters of 12 to 14 Å, which is larger than any included SUMMARY
in our study (Table I).
Figure 8 shows that Young’s modulus is essentially Our DFTB-SCC analysis of a series of model car-
unaffected by the loss of an electron. The Poisson ratio, on bon nanotubes with various helicities and diameters has
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

440 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

Fig. 9. Poisson ratio vs. tube radius (top) and chiral angle (bottom). Data for both neutral and positively charged systems are
included.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 441

Fig. 10. Bond length distributions in optimized positively charged (7, m) model systems. The optimized C C distance in graphene is 1.426 A.

brought out the following points: 2. The strain energy per atom, relative to gra-
phene, diminishes with tube radius in accor-
1. Most of the optimized C C bond lengths are dance with E str = C/r 2 , where C = 1.99 eV Å2 /
slightly longer than in graphene. The orientation atom.
of the carbon sp2 orbitals with respect to the tube 3. The tensile strengths, as measured by Young’s
axis appears to be an important factor in deter- modulus, are greater than for graphene, but grad-
mining whether there is lengthening, shortening ually decrease as the radius increases. They
or no change. Overall, the strain associated with are independent of chirality. The Poisson ratios
the tubular structure causes about a 0.8% decrease are nearly constant, perhaps becoming slightly
in the length of the tube. smaller with larger radii.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

442 Peralta-Inga, Boyd, Murray, O’Connor, and Politzer

Fig. 11. Bond length distributions in optimized positively charged (8, m) model systems. The optimized C C distance in graphene is 1.426 A.

4. The loss of an electron to form a positively negative, as much as 0.10%. The Young’s moduli
charged system usually has little effect upon over- are unaffected, but the Poisson ratio tends to in-
all tube dimensions; in only a few instances is the crease for metals and semimetals and decrease for
change in length or radius, whether positive or semiconductors.

ACKNOWLEDGMENTS
Table II. Changes in Overall Dimensions in Going from Neutral to
Positively Charged Systems
We thank Dr. Kevin Boyd for the graphics software,
Percentage change Percentage change Dr. Th. Frauenheim for the DFTB-SCC code, and the ref-
(n, m) in length in radius eree for very constructive comments. We appreciate the
(7, 0) −0.095 +0.049 support provided by the Advanced Materials Research In-
(7, 1) −0.054 −0.033 stitute through DARPA Grant No. MDA 972-97-1-0003.
(7, 2) +0.099 −0.026
(7, 4) −0.018 −0.043
(7, 7) −0.064 +0.005 REFERENCES
(8, 0) +0.045 −0.105
(8, 2) −0.064 +0.018 1. Iijima, S. Nature (London) 1991, 354, 56.
(8, 4) −0.063 +0.016 2. Dresselhaus, M. S.; Dresselhaus, G.; Eklund, P. C. Science of
(8, 8) +0.110 −0.024 Fullerenes and Carbon Nanotubes; Academic Press: San Diego,
1996.
P1: IZO
Structural Chemistry (STUC) PP958-stuc-471248 November 5, 2003 14:17 Style file version Nov. 07, 2000

Density Functional Tight-Binding Studies of Carbon Nanotubes 443

3. Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Physical Properties 25. Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Phys. Rev. B 1996,
of Carbon Nanotubes; Imperial College Press: London, 1998. 53, 2044.
4. Dai, H.; Kong, J.; Zhou, C.; Franklin, N.; Tombler, T.; Cassell, A.; 26. Treboux, G.; Lapstun, P.; Silverbrook, K. J. Phys. Chem. B 1999,
Fan, S.; Chapline, M. J. Phys. Chem. B 1999, 103, 11246. 103, 1871.
5. Harris, P. J. F. Carbon Nanotubes and Related Structures; Cambridge 27. Baughman, R. H.; Cui, C.; Zakhidov, A. A.; Iqbal, Z.; Barisci, J. N.;
University Press: Cambridge, UK, 1999. Spinks, G. M.; Wallace, G. G.; Mazzoldi, A.; De Rossi, D.; Rinzler,
6. Tanaka, K.; Yamabe, T.; Fukui K. Eds., The Science and Technology A. G.; Jaschinski, O.; Roth, S.; Kertesz, M. Science 1999, 284,
of Carbon Nanotubes; Elsevier: Amsterdam, 1999. 1340.
7. Ajayan, P. M. Chem. Rev. 1999, 99, 1787. 28. Frauenheim, T.; Jungnickel, G.; Kohler, T.; Stephan, U.; J. Noncrys-
8. Odom, T. W.; Huang, J.-L.; Kim, P.; Lieber, C. M. J. Phys. Chem. B tall. Solids 1995, 182, 186.
2000, 104, 2794. 29. Porezag, D.; Pederson, M. R.; Frauenheim, T.; Kohler, T. Phys. Rev.
9. Treacy, M. M. J.; Ebbeson, T. W.; Gibson, J. M. Nature (London) B 1995, 52, 14963.
1996, 381, 678. 30. Kohler, T.; Sternberg, M.; Porezag, D.; Frauenheim, T. Phys. Stat.
10. Wong, E. W.; Sheehan, P. E.; Lieber, C. M. Science 1997, 277, 1971. Sol. A 1996, 154, 69.
11. Van Lier, G.; Van Alsenoy, C.; Van Doren, V.; Geerlings, P. Chem. 31. Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.;
Phys. Lett. 2000, 326, 181. Frauenheim, Th.; Suhai, S.; Seifert, G. Phys. Rev. B 1998, 58,
12. Yu, M.-F.; Files, B. S.; Arepalli, S.; Ruoff, R. S. Phys. Rev. Lett. 7260.
2000, 84, 5552. 32. Frauenheim, Th.; Seifert, G.; Elstner, M.; Hajnal, Z.; Jungnickel, G.;
13. Iijima, S.; Brabec, C.; Maiti, A.; Bernholc, J. J. Chem. Phys. 1996, Porezag, D.; Suhai, S.; Scholz, R. Phys. Stat. Sol. 2000, 217, 41.
104, 2089. 33. Condon, E. U.; Odishaw, H. (Eds.), Handbook of Physics; McGraw-
14. Falvo, M. R.; Clary, G. J.; Taylor, R. M., II; Chi, V.; Brooks, F. P., Hill: New York, 1958.
Jr.; Washburn, S.; Superfine, R. Nature (London) 1997, 389, 582. 34. Baskin, Y.; Meyer, L. Phys. Rev. 1955, 100, 544.
15. Chesnokov, S. A.; Nalimova, V. A.; Rinzler, A. G.; Smalley, R. E.; 35. Sanchez-Portal, D.; Artacho, E.; Soler, J. M.; Rubio, A.; Ordejon, P.
Fischer, J. E. Phys. Rev. Lett. 1999, 82, 343. Phys. Rev. B 1999, 59, 12678.
16. Mintmire, J. W.; Dunlap, B. I.; White, C. T. Phys. Rev. Lett. 1992, 36. Robertson, D. H.; Brenner, D. W.; Mintmire, J. W. Phys. Rev. B 1992,
68, 631. 45, 12592.
17. Hamada, N.; Sawada, S.; Oshiyama, A. Phys. Rev. Lett. 1992, 68, 37. Adams, G.; Sankey, O.; Page, J.; O’Keeffe, M.; Drabold, D. Science
1579. 1992, 256, 1792.
18. Saito, R.; Fujita, M.; Dresselhaus, G.; Dresselhaus, M. S. Phys. Rev. 38. Kurti, J.; Kresse, G. Kuzmany, H. Phys. Rev. B 1998, 58, R8869.
B 1992, 46, 1804; Appl. Phys. Lett. 1992, 60, 2204. 39. Hernandez, E.; Goze, C.; Bernier, P.; Rubio, A. Phys. Rev. Lett. 1998,
19. White, C. T.; Robertson, D. H.; Mintmire, J. W. Phys. Rev. B 1993, 80, 4502.
47, 5485. 40. Yakobson, B. I.; Brabec, C. J.; Bernholc, J. Phys. Rev. Lett. 1996,
20. Wildoer, J. W. G.; Venema, L. C.; Rinzler, A. G.; Smalley, R. E.; 76, 2511.
Dekker, C. Nature (London) 1998, 391, 59. 41. Yakobson, B. I.; Smalley, R. E. Am. Sci. 1997, 85, 324.
21. Odom, T. W.; Huang, J.-L.; Kim, P.; Lieber, C. M. Nature (London) 42. Lu, J. P. Phys. Rev. Lett. 1997, 79, 1297; J. Phys. Chem. Solids 1997,
1998, 391, 62. 58, 1649.
22. Wirth, I.; Eisebitt, S.; Kann, G.; Eberhardt, W. Phys. Rev. B 2000, 43. Treacy, M. M. J.; Ebbeson, T. W.; Gibson, J. M. Nature (London)
61, 5719. 1996, 381, 678.
23. Dresselhaus, M. S. Science 2001, 292, 650. 44. Wong, E. W.; Sheehan, P. E.; Lieber, C. M. Science 1997, 277, 1971.
24. Ouyang, M.; Huang, J.-L.; Cheung, C. L.; Lieber, C. M. Science 45. Yu, M.-F.; Files, B. S.; Arepalli, S.; Ruoff, R. S. Phys. Rev. Lett.
2001, 292, 702. 2000, 84, 5552.

You might also like