modulating the electronic properties along carbon …the g band features in spectrum (b) show a...

21
Modulating the Electronic Properties along Carbon Nanotubes via Tube-Substrate Interaction Jaqueline S. Soares, Ana Paula M. Barboza, Paulo T. Araujo, Newton M. Barbosa Neto, †,‡ Denise Nakabayashi, Nitzan Shadmi, § Tohar S. Yarden, § Ariel Ismach, § Noam Geblinger, § Ernesto Joselevich, § Cecilia Vilani, | Luiz G. Canc ¸ado, †,Lukas Novotny, Gene Dresselhaus, # Mildred S. Dresselhaus, Bernardo R. A. Neves, Mario S. C. Mazzoni, and Ado Jorio* ,† Departamento de Fı ´sica, Universidade Federal de Minas Gerais, Belo Horizonte, MG, 31270-901, Brazil, Instituto de Fı ´sica, Universidade Federal de Uberla ˆndia, Uberla ˆndia, MG, 38400-902, Brazil, § Department of Materials and Interfaces, Weizmann Institute of Science, Rehovot, 76100, Israel, | Divisa ˜o de Metrologia de Materiais, Instituto Nacional de Metrologia, Normalizac ¸a ˜o e Qualidade Industrial (INMETRO), Duque de Caxias, RJ, 25250-020, Brazil, Institute of Optics, University of Rochester, Rochester, New York 14627, United States, # Francis Bitter Magnet Laboratory, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States, and Department of Physics and Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States ABSTRACT We study single wall carbon nanotubes (SWNTs) deposited on quartz. Their Raman spectrum depends on the tube-substrate morphology, and in some cases, it shows that the same SWNT-on-quartz system exhibits a mixture of semiconductor and metal behavior, depending on the orientation between the tube and the substrate. We also address the problem using electric force microscopy and ab initio calculations, both showing that the electronic properties along a single SWNT are being modulated via tube-substrate interaction. KEYWORDS Carbon nanotubes, tube-substrate interaction, Raman spectroscopy, electric force microscopy, quartz S ingle wall carbon nanotubes (SWNTs) are quasi-one- dimensional structures consisting of a rolled up graphene nanoribbon. 1-4 Due to their unusually large surface-to-volume ratio, SWNTs are strongly affected by the environment. 5 Contact with a supporting substrate modifies their properties, and such interactions have been broadly studied as either a drawback or a solution for developing nanotube-based nanotechnologies. 6-23 Researchers have for example studied the interaction of SWNTs with silicon substrates 6,8,10,13-16,19 as a possible route for the integration between SWNTs and silicon-based microelectronics. Ab initio calculations for small diameter SWNTs adsorbed on unpassivated Si surfaces predict stable structures through the formation of covalent Si-C bonds. 8,13-15 Other substrates have also been studied, 11,17,20-22 with quartz becoming identified as a promising substrate for the epitaxial growth of SWNTs. 17,20-22 With the development of nanotube epit- axy combined with the controlled application of external forces, which can generate complex carbon nanotube structures, 17,20-22 the effect of nanotube-substrate interac- tion can be controlled and measured along the same physical nanotube, as reported here. Carbon nanotubes were grown by catalytic chemical vapor deposition (CVD) on miscut single-crystal quartz wafers, as previously reported. 17 The resulting vicinal R-SiO 2 (11 ¯ 01) substrate is insulating, and terminated with parallel atomic steps. 17 At the temperature of nanotube growth, the surface contains exposed unpassivated Si atoms, 24 thus promoting a strong tube-substrate interaction, especially when the nanotube lies along a step. 8,13-15 Alternatively, when the nanotube lies across the surface steps, the interac- tion is discontinuous and weaker. Nanotube epitaxy com- bined with gas flow directed growth 17,20-22 leads to the formation of carbon nanotube serpentines (see Figure 1a,b), i.e., SWNTs with parallel straight segments (labeled S 1 ,S 2 , S 3 , ..., in Figure 1b) connected by alternating U-turns (labeled U 12 ,U 23 ,U 34 , ..., in Figure 1b). The straight segments usually lie along the quartz steps, while the U-turns lie across the steps (see Figure 1b), so that the tube-substrate interaction is modulated along the tube. The main physical effects that have to be revealed and studied in this tube-substrate system are strain and charge transfer and how these effects vary when changing the orientation between the nanotube and the quartz surface steps. Strain depends not only on the tube-substrate inter- action but also on the dynamics of the serpentine formation process, which involves a competition between the tube- * To whom correspondence should be addressed, adojorio@fisica.ufmg.br. Received for review: 09/14/2010 Published on Web: 11/04/2010 pubs.acs.org/NanoLett © 2010 American Chemical Society 5043 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043–5048

Upload: others

Post on 07-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Modulating the Electronic Properties alongCarbon Nanotubes via Tube-SubstrateInteractionJaqueline S. Soares,† Ana Paula M. Barboza,† Paulo T. Araujo,† Newton M. Barbosa Neto,†,‡Denise Nakabayashi,† Nitzan Shadmi,§ Tohar S. Yarden,§ Ariel Ismach,§ Noam Geblinger,§Ernesto Joselevich,§ Cecilia Vilani,| Luiz G. Cançado,†,⊥ Lukas Novotny,⊥ Gene Dresselhaus,#Mildred S. Dresselhaus,∇ Bernardo R. A. Neves,† Mario S. C. Mazzoni,† and Ado Jorio*,†

    †Departamento de Fı́sica, Universidade Federal de Minas Gerais, Belo Horizonte, MG, 31270-901, Brazil, ‡ Institutode Fı́sica, Universidade Federal de Uberlândia, Uberlândia, MG, 38400-902, Brazil, §Department of Materials andInterfaces, Weizmann Institute of Science, Rehovot, 76100, Israel, |Divisão de Metrologia de Materiais, InstitutoNacional de Metrologia, Normalização e Qualidade Industrial (INMETRO), Duque de Caxias, RJ, 25250-020, Brazil,⊥ Institute of Optics, University of Rochester, Rochester, New York 14627, United States, #Francis Bitter MagnetLaboratory, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States, and∇Department of Physics and Department of Electrical Engineering and Computer Science, Massachusetts Institute ofTechnology, Cambridge, Massachusetts 02139, United States

    ABSTRACT We study single wall carbon nanotubes (SWNTs) deposited on quartz. Their Raman spectrum depends on thetube-substrate morphology, and in some cases, it shows that the same SWNT-on-quartz system exhibits a mixture of semiconductorand metal behavior, depending on the orientation between the tube and the substrate. We also address the problem using electricforce microscopy and ab initio calculations, both showing that the electronic properties along a single SWNT are being modulated viatube-substrate interaction.

    KEYWORDS Carbon nanotubes, tube-substrate interaction, Raman spectroscopy, electric force microscopy, quartz

    Single wall carbon nanotubes (SWNTs) are quasi-one-dimensional structures consisting of a rolled upgraphene nanoribbon.1-4 Due to their unusually largesurface-to-volume ratio, SWNTs are strongly affected by theenvironment.5 Contact with a supporting substrate modifiestheir properties, and such interactions have been broadlystudied as either a drawback or a solution for developingnanotube-based nanotechnologies.6-23 Researchers have forexample studied the interaction of SWNTs with siliconsubstrates6,8,10,13-16,19 as a possible route for the integrationbetween SWNTs and silicon-based microelectronics. Abinitio calculations for small diameter SWNTs adsorbed onunpassivated Si surfaces predict stable structures through theformation of covalent Si-C bonds.8,13-15 Other substrateshave also been studied,11,17,20-22 with quartz becomingidentified as a promising substrate for the epitaxial growthof SWNTs.17,20-22 With the development of nanotube epit-axy combined with the controlled application of externalforces, which can generate complex carbon nanotubestructures,17,20-22 the effect of nanotube-substrate interac-tion can be controlled and measured along the same physicalnanotube, as reported here.

    Carbon nanotubes were grown by catalytic chemicalvapor deposition (CVD) on miscut single-crystal quartzwafers, as previously reported.17 The resulting vicinal R-SiO2(11̄01) substrate is insulating, and terminated with parallelatomic steps.17 At the temperature of nanotube growth, thesurface contains exposed unpassivated Si atoms,24 thuspromoting a strong tube-substrate interaction, especiallywhen the nanotube lies along a step.8,13-15 Alternatively,when the nanotube lies across the surface steps, the interac-tion is discontinuous and weaker. Nanotube epitaxy com-bined with gas flow directed growth17,20-22 leads to theformation of carbon nanotube serpentines (see Figure 1a,b),i.e., SWNTs with parallel straight segments (labeled S1, S2,S3, ..., in Figure 1b) connected by alternating U-turns (labeledU12, U23, U34, ..., in Figure 1b). The straight segments usuallylie along the quartz steps, while the U-turns lie across thesteps (see Figure 1b), so that the tube-substrate interactionis modulated along the tube.

    The main physical effects that have to be revealed andstudied in this tube-substrate system are strain and chargetransfer and how these effects vary when changing theorientation between the nanotube and the quartz surfacesteps. Strain depends not only on the tube-substrate inter-action but also on the dynamics of the serpentine formationprocess, which involves a competition between the tube-

    * To whom correspondence should be addressed, [email protected] for review: 09/14/2010Published on Web: 11/04/2010

    pubs.acs.org/NanoLett

    © 2010 American Chemical Society 5043 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043–5048

  • surface interaction and the gas-flow-related drag forces.17

    Raman spectroscopy is the ideal tool for analyzing thetube-substrate interaction without disturbing the tube-surface interaction.25 Analysis of the tangential stretchingmode frequency and line shape (named the G band, at1500-1600 cm-1, see Figure 1c)25 can be used to measurestrain26-28 and doping29-31 and to distinguish betweenmetallic and semiconducting tubes,25 because of the pres-ence of a Kohn anomaly in the phonon dispersion of metallicSWNTs.30-32 In addition, frequency shifts of the dominantsecond-order mode (G′ band, at 2600-2700 cm-1) can alsobe used to differentiate between electron donor (n) andacceptor (p) doping, even at the individual single atomdoping level.33 These aspects of Raman spectroscopy areexemplified in the three representative Raman spectrashown in Figure 2. Spectra (a) and (b) were taken at differentlocations of the same nanotube, whereas spectrum (c) wastaken on a different nanotube. The G-band features inspectrum (a) are typical of a semiconducting SWNT, with twosharp Lorentzian peaks,25 G+ and G-. The G-band featuresin spectrum (c) are typical of a metallic nanotube,25 with alarge downshift and broadening of the G- peak. The G bandfeatures in spectrum (b) show a superposition of metal andsemiconducting behavior. Spectra (a) and (c) both show aG′ band with a single peak, while spectrum (b) shows asplitting of the G′ band. According to ref 33, the shifts in theG′ band features between spectra (a) and (c) can be ex-plained by doping, which can cause either an upshift or

    downshift due to doping, with respect to spectra (a) and (c).The surprise is that the spectra (a) and (b) were measuredfrom one single SWNT serpentine on quartz, the one shownin Figure 1. The difference between (a) and (b) in Figure 2 isthat, in (b) the tube lies along the surface steps, while in (a)the tube lies across the steps.

    FIGURE 1. Raman spectra along a SWNT serpentine on (11̄01) quartz. (a) Confocal image of the G-band integrated intensity (Elaser ) 1.96 eV,spatial resolution ∼500 nm). The spectral intensity is stronger in segments aligned along the steps because of the light polarization dependencefor Raman scattering (the light was polarized along the tube axis).25 Changing the light polarization with respect to the sample changes theoverall intensities but does not change the Raman lineshapes. (b) Schematics of part of the SWNT serpentine in (a) (yellow) on top of themiscut quartz. Miscut means the quartz was polished with a small angle with respect to the atomic layers, so that the substrate exhibitsatomic steps (gray lines). The straight segments are labeled by Si, with i ) 1, 2, 3, ..., numbering the segments according to the growth direction(from point 1 to 41 in (a) (see Supporting Information)). The U-turns are labeled by Ujk, which are the U-turns connecting segments Sj and Sk.(c) The G-band Raman spectra obtained at the 41 points indicated by vertical green pointers and numbered in (a). There is a blue shift andred shift of the higher frequency G+ feature (∼1590-1605 cm-1) along with the appearance and disappearance of the lower frequency G-feature (∼1540 cm-1), related to the tube-substrate morphology and interaction.

    FIGURE 2. Spectroscopic analysis of two SWNT serpentines grownon quartz. (a) and (b) come from the same serpentine in Figure 1a,differing with regard to the tube orientation with respect to thesubstrate, i.e., across (S2 in Figure 1b) vs along (S3 in Figure 1b) thesurface steps, respectively. Spectrum (a) exhibits a G band with aline shape typical of a semiconducting SWNT. Spectrum (b) exhibitsa G band with a line shape showing a mixture of lineshapes typicalof semiconducting (blue Lorentzians) and a metallic (red Lorentz-ians) SWNT behavior. Note that the G′ peak splits in (b). (c) The Gand G′ bands of a different serpentine SWNT exhibiting a G bandwith a metallic character, irrespective of tube orientation with regardto the substrate steps.

    © 2010 American Chemical Society 5044 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043-–5048

  • To understand the spectral evolution between (a) and (b)in Figure 2, we took spectra at different locations alongthe same nanotube serpentine of Figure 2a,b, as shown bythe 41 locations indicated in Figure 1a. The confocal imageshown in Figure 1a was obtained by integrating of theG-band Raman signal from the SWNT serpentine whilescanning the sample. Figure 1c shows the G-band spectraobtained from all 41 points indicated by the vertical greenpointers in Figure 1a. Analysis of the spectral features showsthat the G-band spectra look like the semiconducting profile(a) in Figure 2 when crossing the atomic steps (U-turns andstraight lines passing through points 11-14, named S2 inFigure 1b), while at segments parallel to the steps, such aspoints 15-18 (S3 in Figure 1b), the spectra look like spec-trum (b) in Figure 2. There is a variation in the higherfrequency G+ feature (∼1590-1605 cm-1), which is alsorelated to the tube-substrate morphology, as discussedbelow.

    Figure 3 shows the frequency behavior for both the G andG′ bands (ωG vs ωG′) as a method to characterize thestrain26-28 and doping29-31 induced by the tube-substrateinteraction and to understand the observation of a mixedmetal-semiconductor behavior such as found in spectrum(b) in Figure 2. As mentioned earlier, the G band in semi-conducting SWNTs is composed of two peaks; the higherfrequency G+ feature is related to the C-C stretching modealong the tube axis and is named the longitudinal optical (LO)mode, while the lower frequency G- feature is related toC-C stretching along the tube circumference and is namedthe transverse optical (TO) mode.25 Their frequencies (ω)depend sensitively on strain26-28 and doping,29-31 andwhen the tube becomes metallic, the Kohn anomaly appearsfor the LO phonon, which decreases ωGLO and broadens thepeak.30,32 In Figure 3a, ωGLO and ωGTO, which are observedat the 41 points along the SWNT serpentine depicted in

    Figure 1, are plotted as a function of the G′ band frequency(ωG′, second-order feature related to a breathing of thecarbon hexagons) observed at the same points. Pointslocated at the portions of the serpentine crossing the atomicsteps exhibit only one G′ peak, which we call G′1, and two Gband peaks characteristics of semiconducting SWNTs (seespectrum (a) in Figure 2 and also the region between the twodashed vertical lines in Figure 3a). Points where the nano-tube lies along to the steps exhibit two G′ peaks (seespectrum (b) in Figure 2), which we call G′2 and G′3, respec-tively, in Figure 3a. At these locations, the G band exhibitsfour peaks, two of which show a line shape characteristic ofa semiconducting SWNT and two of which show a line shapecharacteristic of a metallic SWNT (blue and red Lorentziansin spectrum (b) of Figure 2, respectively). The two G bandpeaks characteristic of a semiconducting tube exhibit asmooth correlation between the G′1 frequency and the G′3frequency (blue symbols in Figure 3a). The higher frequencyG peak characteristic of metallic SWNTs (TO) exhibits afrequency that correlates with that of G′2 (red symbols inFigure 3a), while the LO mode frequency is strongly red-shifted, as expected due to the Kohn anomaly.30,32 Thebehavior shown in Figure 3a suggests that the G′2 peak isrelated to a metallic SWNT, while the G′3 is related to asemiconducting SWNT. It is known that n (p) doping causesa downshift (upshift) in the G′ frequency33.

    In Figure 3b, the ωGLO frequency for semiconductingSWNTs is plotted as a function of the distance s from point#1, which is measured along the SWNT. Clearly the ωGLO isobserved to oscillate, showing maxima at the center of thestraight tube segments (labeled Si in Figures 3b and 1b) andminima at the center of the U-turns (labeled Ujk in Figures3b and 1b). Larger frequencies indicate stronger strain26-28

    and doping,29-31 thus corroborating the stronger vs weakermodulated tube-substrate interaction when the nanotube

    FIGURE 3. Strain and doping effects on the SWNT G and G′ Raman bands. (a) Relation between ωG and ωG′ for the SWNT serpentine shown inFigures 1 and 2a,b. When the tube is making the U-turn, the G′ band exhibits a single peak, named G′1, while the G band exhibits two peaks,characteristic of a semiconducting SWNT. When the tube is aligned along a substrate step, the G′ band exhibits two peaks, named G′2 and G′3,while the G band exhibits four peaks, two characteristic of a semiconducting SWNT (blue symbols) and two characteristic of a metallic SWNT(red symbols). Not all the 41 G band data sets from Figure 1 are shown here because the G′ band is absent in 12 cases. (b) The G+ frequencyof semiconducting SWNTs (LO mode) observed at the 41 points shown in Figure 1a is plotted as a function of the distance s from point #1measured along the SWNT. Si and Ujk locate the center of the straight segments and U-turns, respectively (see Figure 1b), showing that thefrequency changes are correlated with the tube-substrate morphology. The frequency uncertainty is better than (2 cm-1.

    © 2010 American Chemical Society 5045 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043-–5048

  • lies along versus across the surface steps. The modulationof ωGLO along the straight segments, on the other hand, canbe attributed to oscillations in the strain along the nanotuberesulting from the competition between the tube-surfaceinteraction and the aerodynamic drag forces during theformation of the serpentine, consistent with the “fallingspaghetti” mechanism previously proposed.17 All the maxi-mal ωGLO values correspond to the center of the straightsegments aligned along a quartz step. The fact that themisaligned straight segment S2 actually shows a minimumin ωGLO indicates that the effect of tube-surface interactionon the electronic properties of the nanotube is relativelystronger than that of strain, while the remaining effect ofthe aerodynamic drag forces should be responsible for thesmaller ωGLO maxima for the Si’s with even i. Scanningelectron microscopy shows the serpentine growth happenedfrom point 1 to 41, so that i odd indicate straight segmentsfalling to the left, while i even indicate straight segmentsfalling to the right, taking Figure 1a as a reference (checkthe “falling spaghetti” mechanism17). Therefore, the largerωGLO maxima for Si odd indicate that, during the “fallingspaghetti” process, there was a gas-flow drag componentpointing to the left. This asymmetry can be attributed to thefact that the flow was not perfectly perpendicular to the steps(see Supporting Information). All these effects are stable asa function of the sample lifetime, since this study was carriedout within 2 years time after the growth.

    To shed light in our findings, we performed first-principlescalculations for a nanotube placed on top of a crystalline SiO2substrate. Our calculations are based on pseudopotentialDensity Functional Theory (DFT)34-36 formalism, as imple-mented in the SIESTA program,37,38 which makes use of abasis set composed of pseudo atomic functions of finiterange. To fully justify this strong tube-substrate interaction,we developed different models where the SWNTs are incontact with nonpassivated surfaces, with either Si or Oexposed (see Supporting Information). One of the modeledsystems is shown in Figure 4a after geometry relaxation. Weconsider in the calculation a (19,0) carbon nanotube sitting

    on the (001) surface of a SiO2 substrate. The (19,0) has adiameter of 1.4 nm, matching the diameter of the tubeanalyzed in Figures 1-3 (see Supporting Information). Uponrelaxation, the silicon atoms in the contact region are foundto experience an upward displacement and the bottom partof the nanotube becomes flat. The deformation is the resultof a strong interaction between carbon bonding states andsurface dangling bonds, which show up in the band structureas dispersive bands crossing the Fermi level (see SupportingInformation). The yellowish clouds in Figure 4a represent aplot of the electronic density for states within energies ofup to 0.1 eV around the Fermi level. Therefore, the bandsresponsible for the gap closure are predominantly localizedspatially in the contact region along the flat surface of thenanotube. These calculations elucidate how a strong nano-tube-substrate adhesion of the nanotube along an exposedSiO2 surface rehybridizes the electronic states to producetwo tube segments (bottom and top with respect to thesubstrate) with different electronic configurations, thus ex-plaining the mixed metal-semiconducting character ob-served in the Raman spectroscopy measurements (Figures1-3). Such kind of hybrid structure requires flat crystallinesubstrates to be present, where the interaction is extendedand does not generate localized states.

    For a more theoretical quantitative analysis, parts (b) and(c) of Figures 4 show how the substrate affects the bondlengths (strain) and charge distribution (doping) around thetube circumference, respectively. Figure 4b shows that theaverage distance between carbon atoms (aC-C) increaseswhen the tube is in contact with the substrate, with ∆aC-CMax

    ∼ 1.7%. A slightly larger increase is observed for the aCCmeasured along the tube axis (open circles), suggesting astronger effect should be felt by ωGLO. Figure 4c shows thatthe average density of electrons per carbon atom alsoincreases in the tube-substrate contact region, changingfrom the expected value of 4 electrons/C (the two 1s coreelectrons are not considered in our model) to up to ∼4.05electrons/C at the surface. Increase in bond lengths shouldsoften the tangential mode frequencies, but this effect is

    FIGURE 4. Modeling the tube-substrate interaction based on pseudopotential DFT. (a) Real space representation of the electronic states atthe Fermi level of the distorted tube. The (19,0) SWNT is sitting on the (001) surface of a SiO2 substrate (see arrow) with tube axis along (100).(b) Average bond length (aC-C) along the tube axis (open circles) and along the tube circumference (filled circles), and (c) the electron percarbon atom density (G) of the SWNT deposited on a quartz substrate. In this polar representation, aC-C and G change along the tubecircumference, the values varying from the center to the edge of the circle, as quantified by the axis on the left, which applies for 90° and270°. The aC-C increases at the tube-substrate contact region (240-300°). The electron density and bond lengths change around thecircumference, and the changes exhibit maxima near the tube-substrate contact region.

    © 2010 American Chemical Society 5046 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043-–5048

  • compensated by the phonon stiffening expected by thedisappearance of the Kohn anomaly due to doping. Wecannot clearly distinguish in our experiment between thespectral changes arising from strain and those arisingfrom doping, and these effects are probably correlated toachieve structure equilibrium, as indicated by parts (b)and (c) of Figure 4.

    Although the results discussed here are based on datafrom one SWNT, we have studied nine SWNT serpentines,and significant spectral modifications clearly related to theanisotropic tube-substrate interaction could be identifiedin all of them (see Supporting Information). In all nine casesthe frequencies of the G and G′ bands change according tothe orientation between the tube axis and the crystallinequartz substrate, consistent with the results shown here. Thedetailed frequency changes vary from sample to sampleindicating a possible (n,m) dependence, where both the tubediameter and the angle between the hexagons and the SiO2substrate could play a role. The influence of tube curvatureis negligible, since the U-turns are in the order of 1 µmdiameter while the tube is 1 nm in diameter. The importanceof defects is also excluded by the complete absence of thedefect induced D band (∼1350 cm-1) in both far-field andnear-field Raman spectroscopy imaging (see SupportingInformation). The modulated doping-strain, generating mixedmetal-semiconductor Raman profiles have been observedin four different serpentines. When the tube exhibits spectracharacteristic of a metallic system (one tube measured), themixed behavior was not observed. This is supported by ourDFT calculations, since metallic tubes always have electronscrossing the Fermi level (see Supporting Information).

    For completeness, we have also employed electric forcemicroscopy (EFM) characterization of the SWNT serpentinefor the direct determination of its metallicity. Figure 5 showsthe EFM analysis of the same SWNT discussed in Figure 1.

    The nanotube appears in the image as regions of negativefrequency shifts (in blue) contrasting with the quartz sub-strate (in brown). Even though the EFM signal seems homo-geneous in Figure 5a, a more careful analysis reveals struc-ture in the line shape, as shown in the line profiles acrosstwo different regions in the nanotube presented in Figure5b. As discussed in ref 39, the different polarizabilities ofmetallic nanotubes on substrate versus semiconductingnanotubes on substrate create a distinct coupling with theelectric field emanating from the EFM tip for each case.These differences result in a “W-shaped” EFM line profileacross a semiconducting tube, while a metallic SWNT por-trays a “V-shaped” line profile (see Supporting Informationfor an schematic explaining the effect). The red and blueprofiles in Figure 5b correspond to the regions marked by thered and dark blue lines in Figure 5a, respectively. The“W-shaped” dark blue profile, typical for a region with weaksubstrate interaction, is the EFM signature for a semiconductingnanotube, whereas the “V-shaped” red profile, typical for astrong substrate interaction region, attests to a metallic behav-ior for this region of the SWNT. Therefore, the EFM resultsshown in Figure 5 corroborate the metal-semiconductingalternating behavior of our SWNT serpentine Raman spectra.

    In summary, we have measured a SWNT-substratesystem where the tube substrate interaction can be tunedby changing the tube-substrate orientation. The effects onthe electronic and vibrational properties of the SWNTs,observed using resonance Raman spectroscopy and electricforce microscopy, indicate important changes in the proper-ties depending on the tube-substrate orientation. Thesechanges are clearly related to the tube-substrate interactionresulting from the tube-substrate morphology and forma-tion dynamics. The periodic change on the tube-substrateinteraction existing in our SWNT serpentines seems togenerate a set of alternate doped-undoped tube segments,and different complex superlattices could be created throughsubstrate engineering. Interestingly, Huang and Choi21 ob-served that the length-normalized resistance of carbonnanotube serpentines increases with the number of U-turns,and such behavior was explained by the presence of defectsin the curved regions. Our finding can explain the resultsobserved in ref 21 with perfectly crystalline junctions, wherethe “defects” would be the substrate changing the tubeelectronic behavior. This is supported by our Raman spec-troscopy results with the complete absence of the disorder-induced D-band peak (∼1300 cm-1), which is normallyobserved in defective sp2 carbon materials.25

    Acknowledgment. A.J., M.S.C.M., and B.R.A.N. acknowl-edge financial support from Rede Nacional de SPM, RedeNacional de Pesquisa em Nanotubos de Carbono, Institutode Nanotecnologia (MCT-CNPq), and AFOSR/SOARD (award#FA9550-08-1-0236). L.N. acknowledges financial supportfrom the U.S. Department of Energy (Grant DE-FG02-05ER46207). E.J. acknowledges support from the IsraelScience Foundation, the US-Israel Binational Science Foun-

    FIGURE 5. Electric force microscopy (EFM) analysis of the alternatingmetal-semiconductor carbon nanotube heterojunction. (a) EFMimage of the same SWNT serpentine discussed in Figure 1 (Experi-mental conditions: V )-5 V; lift height, 50 nm; ω0 ) 27.6 kHz; k )0.60 N/m). (b) Scanning across the SWNT, the cantilever frequencyshift is different if a tube is semiconducting (“W-shaped” blue line)or metallic (“V-shaped” red line) (see Supporting Information). Theprofiles in (b) are as measured. The red and dark blue lines in (a)indicate the two regions where these profiles in (b) were acquired,which correspond to S3 and S2 in Figure 1b, respectively.

    © 2010 American Chemical Society 5047 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043-–5048

  • dation, the Kimmel Center for Nanoscale Science, and theLegrain, Djanogly, Alhadeff, and Perlman Family founda-tions. M.S.D. acknowledges NSF DMR # 07-04197.

    Supporting Information Available. This material is avail-able free of charge via the Internet at http://pubs.acs.org.

    REFERENCES AND NOTES(1) Iijima, S.; Ichihashi, T. Nature 1993, 363, 603–605.(2) Yu, D.; Liu, F. Nano Lett. 2007, 7, 3046–3050.(3) Jiao, L.; Zhang, L.; Wang, X.; Diankov, G.; Dai, H. Nature 2009,

    458, 877–880.(4) Santos, H.; Chico, L.; Brey, L. Phys. Rev. Lett. 2009, 103, 086801.(5) Jorio, A.; Dresselhaus, M. S. MRS Bull. 2007, 32, 988–993.(6) Heertel, T.; Walkup, R. E.; Avouris, Ph. Phys. Rev. B 1998, 58,

    13870–13873.(7) Mazzoni, M. S. C.; Chacham, H. Appl. Phys. Lett. 2000, 76, 1561–

    1563.(8) Orellana, W.; Miwa, R. H.; Fazzio, A. Phys. Rev. Lett. 2003, 91,

    166802.(9) Jiang, C.; Zhao, J.; Therese, H. A.; Friedrich, M.; Mews, A. J. Phys.

    Chem. B 2003, 107, 8742–8745.(10) Tsukruk, V. V.; Ko, H.; Peleshanko, S. Phys. Rev. Lett. 2004, 92,

    065502.(11) Kim, Y.-H.; Heben, M. J.; Zhang, S. B. Phys. Rev. Lett. 2004, 92,

    176102.(12) Ismach, A.; Segev, L.; Wachtel, E.; Joselevich, E. Angew. Chem.,

    Int. Ed 2004, 43, 6140–6143.(13) Miwa, R. H.; Orellana, W.; Fazzio, A. Appl. Phys. Lett. 2005, 86,

    213111.(14) Berber, S.; Oshiyama, A. Phys. Rev. Lett. 2006, 96, 105505.(15) Peng, G. W.; Huan, A. C. H.; Liu, L.; Feng, Y. P. Phys. Rev. B 2006,

    74, 235416.(16) Albrecht, P. M.; Lyding, J. W. Nanotechnology 2007, 18, 125302.(17) Geblinger, N.; Ismach, A.; Joselevich, E. Nat. Nanotechnol. 2008,

    3, 195–200.(18) Barboza, A. P. M.; Gomes, A. P.; Archanjo, B. S.; Araújo, P. T.;

    Jorio, A.; Ferlauto, A. S.; Mazzoni, M. S. C.; Chacham, H.; Neves,B. R. A. Phys. Rev. Lett. 2008, 100, 256804.

    (19) You, Y.; Yu, T.; Kasim, J.; Song, H.; Fan, X.; Ni, Z.; Cao, L.; Jiang,H.; Shen, D.; Kuo, J.; Shen, Z. Appl. Phys. Lett. 2008, 93, 103111.

    (20) Jeon, S.; Lee, C.; Tang, J.; Hone, J.; Nuckolis, C. Nano Res. 2008,1, 427–433.

    (21) Huang, J.; Choi, W. Nanotechnology 2008, 19, 505601.(22) Xiao, J.; Dunham, S.; Liu, P.; Zhang, Y.; Kocabas, C.; Moh, L.;

    Huang, Y.; Hwang, K.-C.; Lu, C.; Huang, W.; Rogers, J. A. NanoLett. 2009, 9, 4311–4319.

    (23) Steiner, M.; Freitag, M.; Tsang, J. C.; Perebeinos, V.; Bol, A. A.;Failla, A. V.; Avouris, Ph. Appl. Phys. A 2009, 96, 271–282.

    (24) Viefhaus, H.; Rossow, W. Surf. Sci. 1984, 141, 341–354.(25) Dresselhaus, M. S.; Dresselhaus, G.; Saito, R.; Jorio, A. Phys. Rep.

    2005, 409, 47–99.(26) Cronin, S. B.; Swan, A. K.; Ünlü, M. S.; Goldberg, B. B.; Dressel-

    haus, M. S.; Tinkhan, M. Phys. Rev. Lett. 2004, 93, 167401.(27) Duan, X.; Son, H.; Gao, B.; Zhang, J.; Wu, T.; Samsonidze, Ge.

    G.; Dresselhaus, M. S.; Liu, Z.; Kong, J. Nano Lett. 2007, 7, 2116–2121.

    (28) Gao, B.; Jiang, L.; Ling, X.; Zhang, J.; Liu, Z. J. Phys. Chem. C 2008,112, 20123–20125.

    (29) Das, A.; Sood, A. K.; Govindaraj, A.; Marco Saitta, A.; Lazzeri, M.;Mauri, F.; Rao, C. N. R.; Liu, Z.; Kong, J. Phys. Rev. Lett. 2007, 99,136803.

    (30) Dubay, O.; Kresse, G.; Kuzmany, H. Phys. Rev. Lett. 2002, 88,235506.

    (31) Tsang, J. C.; Freitag, M.; Perebeinos, V.; Liu, J.; Avouris, Ph. Nat.Nanotechnol. 2007, 2, 725–730.

    (32) Piscanec, S.; Lazzeri, M.; Robertson, J.; Ferrari, A. C.; Mauri, F.Phys. Rev. B 2007, 75, 035427.

    (33) Maciel, I. O.; Anderson, N.; Pimenta, M. A.; Hartschuh, A.; Qian,H.; Terrones, M.; Terrones, H.; Campos-Delgado, J.; Rao, A.;Novotny, L.; Jorio, A. Nat. Mater. 2008, 7, 878–883.

    (34) Kohn, W.; Sham, L. J. Phys. Rev. 1965, 140, A1133–A1138.(35) Kleinman, L.; Bylander, D. M. Phys Rev. Lett. 1982, 48, 1425–

    1428.(36) Troullier, N.; Martins, J. L. Phys. Rev. B 1991, 43, 1993–1996.(37) Ordejón, P.; Artacho, E.; Soler, J. M. Phys. rev. B 1996, 53,

    R10441–R10444.(38) Soler, J. M.; Artacho, E.; Gale, J. D.; Garcia, A.; Junquera, J.;

    Ordejón, P.; Sánchez-Portal, D. J. Phys.: Condens. Matter 2002,14, 2745–2779.

    (39) Barboza, A. P. M.; Gomes, A. P.; Chacham, H.; Neves, B. R. A.Carbon 2010, 48, 3287–3292.

    © 2010 American Chemical Society 5048 DOI: 10.1021/nl103245q | Nano Lett. 2010, 10, 5043-–5048

  • 1

    Modulating the electronic properties along carbon

    nanotubes via tube-substrate interaction

    Jaqueline S. Soares,† Ana Paula M. Barboza,

    † Paulo T. Araujo,

    † Newton M. Barbosa Neto,

    †, Denise

    Nakabayashi,† Nitzan Shadmi,

    § Tohar S. Yarden,

    § Ariel Ismach,

    § Noam Geblinger,

    § Ernesto Joselevich,

    §

    Cecilia Vilani, Luiz G. Cançado,

    †,# Lukas Novotny,

    # Gene Dresselhaus,

    ¥ Mildred S. Dresselhaus,

    £

    Bernardo R. A. Neves,† Mario S. C. Mazzoni,

    † Ado Jorio*

    ,†

    Supporting Information

    Materials and Methods

    Sample preparation. Carbon nanotubes were grown by catalytic chemical vapor deposition (CVD) on

    top of one-side polished single-crystal quartz wafers (Sawyer Research Products, Inc.), as discussed in

    ref. (1). The substrates were cleaned by sonication in acetone. Parallel stripes of amorphous SiO2 (Kurt

    J.Lesker, 99.99 %) were laid down in the surface step direction by photolithography and electron-beam

    evaporation and deposition of 0.5 nm thick evaporated layer of Fe, the growth catalyst, (Kurt J.Lesker,

    99.95 %) was evaporated on the amorphous SiO2 stripes. A lift-off was done by weak sonication in

    acetone and the Fe particles were oxidized in air. The samples were placed inside the middle of a quartz

    tube at a defined angle α between the surface steps and the gas flow. The tube was placed in a furnace,

    and CVD was carried out at 900 °C and 1 atm for 60 min from a mixture of 60 % Ar, 40 % H2, 0.4 %

    C2H4 at flow rate of 500 sccm (cm3/min).

  • 2

    Raman spectroscopy. Confocal Raman measurements were performed on an inverted optical

    microscope with the addition of an x,y-stage for raster-scanning the samples. Light from a He-Ne laser

    (632.8 nm) is reflected by means of a beam splitter and then focused onto the surface of the sample

    using an oil-objective with 60 magnification. Having obtained a tight focal spot at the sample surface

    and using the x,y-scan stage to raster scan the sample, Raman scattered light is collected by the same

    microscope objective and recorded using either a single-photon counting avalanche photodiode (APD)

    or a spectrograph with a charged-coupled device (CCD).

    Near-field Raman spectroscopy. Near-field Raman spectroscopy from carbon nanotube serpentine

    samples were obtained using the same system described above, by approaching a gold tip to the sample

    surface2. The near field is obtained through the tip-enhanced Raman scattering (TERS) effect, with a

    typical tip diameter of 20-40nm and typical sample-tip distance of ~2nm, as described in details in3.

    Atomic Force Microscopy. Detailed atomic force microscopy (AFM) measurements were performed

    by JPK Nanowizard (JPK Instruments AG) atomic force microscope operating in tapping mode. A

    cantilever commercially available (Mikromasch 35 NSC/ALBS - nominal constant spring – k ~ 40 N/m)

    was employed for AFM images acquisition.

    Electric Force Microscopy. Electric force microscopy (EFM) and joint atomic force microscopy

    (AFM) measurements were performed by scanning probe microscopy (SPM) (Nanoscope IV

    MultiMode SPM, from Veeco Instruments). AuCr-covered silicon cantilevers with nominal spring

    constant k ~ 0.3 to 0.6 N/m, nominal radius of curvature R ~ 30 nm and resonant frequency 0 ~ 20 to

  • 3

    40 kHz were employed throughout this work for AFM (intermittent contact modes) and EFM

    characterization.

    Scanning Electron Microscopy. Scanning Electron Microscopy (SEM) images were taken using

    field-emission SEM instruments Supra 55VP FEG LEO and Ultra 55 Zeiss, Oberkochen, Germany, at

    low working voltages of 0.5-2.0 kV.

    Supporting Text

    First-principles calculations. While theoretical and experimental works have been developed for

    studying the effect of strain4-6

    and doping7,8

    on GLO

    and GTO

    , these published models cannot be

    directly used here to quantify the effects reported in this study. The published models address strain and

    doping levels that are much smaller in magnitude than the effects expected for the carbon nanotube-

    SiO2 substrate interaction9-12

    . This is made evident from the observed shifts in the present work, as

    compared with previous observations4-8,13

    , and confirmed by our own calculations. First-principles

    calculations were performed within the Pseudopotential Density Functional Theory (DFT)14-16

    formalism as implemented in the SIESTA program17,18

    , which makes use of a basis set composed of

    pseudo atomic functions of finite range. The Generalized Gradient Approximation (GGA) within the

    Perdew-Burke-Erzerhof (PBE) parametrization19

    was employed for the exchange-correlation functional.

    In our geometric models, the nanotube is placed on top of a SiO2 slab saturated by H atoms at its

    bottom part. Periodic boundary conditions are used with the vertical lattice vector large enough to

    prevent interactions between the system and its repetitions. We consider distinct arrangements for the

    contact region between the nanotube and the slab. We mention, for instance, a contact region with either

    Si or O atoms exposed with dangling bonds and an intermediate geometry in which only some of the

  • 4

    surface O atoms are non-passivated. Outside the contact region, all Si atoms make bonds with –OH

    groups. These geometries are meant to mimic conditions in which a strong adhesion of the nanotube

    may take place. Although we can not pinpoint which specific mechanism is responsible for the adhesion

    (covalent bonds between exposed O or Si atoms with the nanotube or a charge transfer between O atoms

    and the nanotube), the idea is that the existence of a periodic array of interaction sites (this is the

    importance of the crystalline quartz) may cause the appearance of dispersive bands crossing the Fermi

    level. These may be defective bands (dispersive because of the periodicity), if they originate from

    covalent bonds, or simply a shift of the Fermi level, if charge is transferred from the nanotube to surface

    atoms (without making additional bonds). The former case is illustrated in Fig. S1, which shows the

    band structure of an originally semiconductor (19,0) nanotube interacting with a SiO2 slab having Si

    dangling bonds, as described in our paper (see Fig. 4 on the main text). Here the system is periodic in

    both the [100] and [010] directions normal to (100). The lateral lattice vector has a length of 2.6 nm,

    which is enough to avoid interactions between the nanotube and its periodic repetitions. The ten oxygen

    atoms at the bottom side of the surface are held fixed to mimic the structural effect of the missing part of

    the substrate, and they are saturated by hydrogen atoms.

    When the nanotube is placed across the surface steps, a strong interaction is supposed to exist only

    locally, at the steps themselves. The consequence in the band structure would be a set of non-dispersive

    defective bands, and this is consistent with preliminary efforts we made to calculate such structures.

    Therefore, the general picture is a significant interaction for tubes aligned along the steps, which may

    modulate the electronic properties of the nanotube (such as a semiconductor-metal transition evidenced

    by the presence of dispersive bands crossing the Fermi level in Fig. S1), and localized interactions in the

    steps, which act as isolated defects in the band structure.

  • 5

    Figure S1: Band structure of the system discussed in Fig. 4 of the main paper, i.e. a (19,0) carbon

    nanotube interacting with a SiO2 slab. The Fermi level is set to zero.

    Imaging a SWNT serpentine by Scanning Electron Microscopy and Atomic Force Microscopy –

    The tube growth direction. We have used scanning electron microscopy (SEM) and atomic force

    microscopy (AFM) in order to fully characterize the SWNT serpentines. Figure S2 shows the SEM and

    AFM images of the serpentine discussed in our paper. The SEM image shows the markers used to locate

    the serpentines and the amorphous silicon, where the catalystic particles are deposited, from where the

    carbon nanotubes grow. The black arrow indicates the serpentine location that is discussed in detail in

    our paper. Since the tube grows from the large white/gray track in (a), where the catalyst particles are

    placed, by image magnification we know that the growth direction in Fig. (b) occurred from the bottom-

    right to the top-left part of the tube displayed. Notice the growth is generally perpendicular to the

    amorphous Si track, but exactly at the point we measure, there is a shift to the side (to the right in panel

  • 6

    (a) and to the left in panel (b) – the figures are mirror symmetric), probably related to fluctuations in the

    gas flow direction near the lithographic markers.

    Figure S2: (a) SEM and (b) AFM images of the SWNT serpentine on top of the quartz substrate. The

    black arrow in the SEM image shows exactly the point where the AFM image in (b) was taken.

  • 7

    Figure S3: (a) Confocal Raman image corresponding to the G band intensity of a carbon nanotube

    serpentine. (b) Near-field optical Raman image corresponding to the G band intensity recorded in the

    boxed area in panel (a). (c) Topographic image obtained simultaneously with the near-field image

    shown in panel (b). Panels (d) and (e) show two hyper-spectral Raman images corresponding to the G

    mode and RBM intensities, respectively, obtained from the boxed area in panel (b).

    Imaging a SWNT serpentine by Near-field Raman Spectroscopy – The tube crystallinity. Near-

    field scanning optical (NSOM) Raman spectroscopy measurements were obtained from 6 SWNT

    serpentines to check local effects related by the presence of defects in the tube structure. One example is

    shown in Fig.S3. The spectra quality indicates an unusually high crystallinity of the samples, with no

    local emission characteristics of defects, usually identified by the presence of a Raman peak in the

    1300-1400 cm-1

    spectral range, known as the D band (see also ref. 2), or a localized variation in the

    intensity of allowed Raman peaks (G and RBM)20,21

    . Huang and Choi22

    observed that the length-

    normalized resistance of carbon nanotube serpentines increase with the number of U-turns, and such

    behavior was explained by the presence of defects in the curved regions. Our metal-semiconductor

    junctions can explain the results observed in ref. (22) with perfectly crystalline junctions, where the

    “defects” would be the substrate changing the tube electronic behavior. This is supported by our Raman

    spectroscopy results with the complete absence of the disorder-induced D-band peak (~1300 cm-1

    ),

    which is normally observed in defective sp2 carbon materials

    23. Notice this serpentine has an RBM, but

    as discussed in the next topic of this supporting online material, an RBM based (n,m) assignment is not

    yet clear for our SWNTs on quartz. The high crystallinity reported here was also observed for SWNT

    serpentines where no RBM emission was observed.

  • 8

    Comment about D-band mode. Clearly our SWNTs are distinct by the complete absence of the

    defect induced D-band. Considering the work from different collaborators, we know by now studies on

    41 SWNT serpentines grown by the group of Ernesto Joselevich, and none of these serpentines

    exhibited a D-band, while all of them show important changes on G and G´ bands. In some cases the

    samples are dirty (see Figure S2b), but still, the presence of impurities on top of the SWNT serpentines

    do not cause a D-band in their Raman spectra. You can even press these SWNT serpentines with AFM

    tips while acquiring the Raman spectra, and usually no D band is seen. We do see a very small D peak

    in areas where we clearly caused some damage with the AFM tip (such studies have been performed in

    our laboratory and will be reported in the near future).

    Considerations about the radial breathing mode. The authors are aware that the RBM is a very

    useful signature for SWNTs, including for assigning the metal vs. semiconducting character. However,

    we understand that the observation of the RBM, in the present study, is neither a proof of isolated tubes

    (since another non-resonant SWNT could be bundling with the measured SWNT), nor a proof of the

    metallic vs. semiconducting character of the tube, since the substrate-induced distortions may be

    significant, i.e. significant frequency changes or RBM quenching are expected. From the 9 serpentines

    we measured, 6 of them have RBM modes, and the RBMs, involving atomic displacements normal to

    the plane of the substrate, are always unchanged along the whole serpentines. From the 4 SWNTs

    exhibiting the metal-semiconducting hybrid behavior, only one of them has an observable RBM, but the

    RBM-based (n,m) analysis, or even the metal vs. semiconducting assignment, is not clear from a

    Kataura plot based analysis24

    . The same holds for the other 5 cases where no metal-semiconducting

    hybrid behavior is observed. Here we base our metal vs. semiconducting assignment on the G band

    analysis. The physics behind this peak is now well developed and surpasses the structural information

    delivered by the RBM with respect to metallic vs. semiconducting character. Nevertheless, we are now

  • 9

    developing a Kataura plot based RBM analysis for tubes on the quartz substrate, and strong deviations

    from the established RBM frequencies and optical transition energies seem to apply.

    Comment about the measurements in other SWNT serpentines. As mentioned in the main text, we

    measured 9 SWNT serpentines and significant tube-substrate interaction could be identified in all of

    them. The formation of alternating metal-semiconductor behavior has been observed in 4 cases. In 4

    other serpentines, we measured a semiconductor SWNT behavior all over the tubes. We also measured

    one SWNT serpentine with a metallic character.

    Figure S4 shows the G band Raman spectra of two SWNT serpentines, one with a semiconducting

    character, shown in Fig. S4a, and another with a metallic character, shown in Fig. S4b. Both G band

    spectra change when changing location along the serpentine. The G+ peak for semiconducting case

    shifts to higher frequencies when the tube is along the substrate steps (Fig. S4a top) and the G- peak for

    the metallic tube shifts to lower frequencies when the tube is along the substrate steps (Fig. S4b

    bottom). These changes show the tube-substrate interaction and modulation in other SWNT serpentines,

    and we find the spectra taken on different serpentines to be qualitatively consistent with observations in

    the paper.

  • 10

    Figure S4: Spectroscopic analysis of two SWNT serpentines grown on quartz. Spectrum (a) exhibits a

    G band with a lineshape typical of a semiconducting SWNT. Spectrum (b) exhibits a G band with a

    lineshape typical of a metallic SWNT (see text).

    Direct determination of the metallic/semiconducting character of a SWNT via Electric Force

    Microscopy (EFM) imaging. The determination of metallic vs. semiconducting character of a SWNT

    using EFM has been demonstrated and reported recently.27

    Here we provide a schematic for a clear

    understanding of the effect (see Fig.S5). The EFM measures the dielectric response of the whole

    sample, i.e. SWNT plus substrate. In general, when the tip approaches the dielectric material, there is a

    decrease in . Then two effects happen in our measurement, as depicted in Fig.S5: (1) decreases

    when the tip approaches the tube, reaches a minimum when at the top of the tube, and increases back

    when the tip departs from the tube (see column a in Fig.S5); (2) shows a slight increase during the

    constant-height scan, because the EFM tip retracts when crossing the tube (see column b in Fig.S5). The

    change in surface-tip distance causes a lowering in the tip-substrate interaction. The overall result is a

    sum of these two effects (see column c in Fig.S5). In the case of semiconducting SWNT (blue), when

    the EFM tip is really going on top of the tube, the tip retraction due to change in surface height causes a

    lowering in the tip-substrate interaction, giving rise to the “W shaped” in the semiconducting SWNTs.

    This effect is indeed present in the metallic SWNT (red), but not observable due to the different

    (stronger and sharper) dielectric response, thus keeping a slightly distorter “V shaped” EFM profile.

  • 11

    Figure S5: The EFM profiles for semiconducting and metallic SWNTs. In column (a), the effect of the

    EFM tip crossing a semiconductor (blue) and a metallic (red) SWNTs on are shown. The lineshapes

    are different due to the differences in dielectric response. In column (b) the effect of the EFM tip

    retracting from the substrate on are shown. Here the effect is the same for both semiconductors and

    metals. In column (c) the two effects are combined. The plots shown here were built using the equations

    in ref. (27).

    Besides the case study in the main paper, we also have SWNT serpentines that are only metallic or

    only semiconducting in our samples. Here we show the EFM and G band Raman measurements

  • 12

    performed in SWNT serpentines that are only metallic and only semiconducting (see Figures S6 and S7,

    respectively). The EFM results are self consistent and consistent with the Raman spectroscopy results.

    Figure S6: Analysis of a metallic SWNT serpentine grown on quartz. (a) EFM image of the metallic

    SWNT serpentine. The red and dark blue lines indicate the two regions where the Raman spectra in (b)

    and the profiles in (c) and (d) were acquired. (b) Raman spectra of the metallic SWNT serpentine in

    the straight segment (blue) and at the edge of the U turn (red). (c) and (d) EFM scan across the metallic

    SWNT at the edge of the U turn (c) and in the straight segment (d). The “V” shaped line profile

    shows the EFM signal characteristic of metallic tubes.

  • 13

    Figure S7: Analysis of a semiconducting SWNT serpentine grown on quartz. (a) EFM image of the

    semiconducting SWNT serpentine. The red and dark blue lines indicate the two regions where the

    Raman spectra in (b) and the profiles in (c) and (d) were acquired. (b) Raman spectra of the

    semiconducting SWNT serpentine in the straight segment (blue) and at the edge of the U turn (red). (c)

    and (d) EFM scan across the semiconducting SWNT at the edge of the U turn (c) and in the straight

    segment (d). The “W” shaped line profile shows the EFM signal characteristic of semiconducting

    tubes.

    References

    (1) Geblinger, N.; Ismach, A.; Joselevich, E. Nature Nanotech. 3, 195-200 (2008).

    (2) Cancado, L. G.; Jorio, A.; Ismach, A.; Joselevich, E.; Hartschuh, A.; Novotny, L. Phys. Rev. Lett.

    103, 186101 (2009).

    (3) Cancado, L. G.; Hartschuh, A.; Novotny, L. J. Raman Spec. 40, 1420-1426 (2009).

  • 14

    (4) Cronin, S. B.; Swan, A. K.; Ünlü, M. S.; Goldberg, B. B.; Dresselhaus, M. S.; Tinkhan, M. Phys.

    Rev. Lett. 93, 167401 (2004).

    (5) Duan, X.; Son, H.; Gao, B.; Zhang, J.; Wu, T.; Samsonidze, Ge. G.; Dresselhaus, M. S.; Liu, Z.;

    Kong, J. Nano Lett. 7, 2116-2121 (2007).

    (6) Gao, B.; Jiang, L.; Ling, X.; Zhang, J.; Liu, Z. J. Phys. Chem. C 112, 20123-20125 (2008).

    (7) Das, A.; Sood, A. K.; Govindaraj, A.; Marco Saitta, A.; Lazzeri, M.; Mauri, F.; Rao, C. N. R.

    Phys. Rev. Lett. 99, 136803 (2007).

    (8) Tsang, J. C.; Freitag, M.; Perebeinos, V.; Liu, J.; Avouris, PH. Nature Nanotech. 2, 725-730

    (2007).

    (9) Orellana, W.; Miwa, R. H.; Fazzio, A. Phys. Rev. Lett. 91, 166802 (2003).

    (10) Miwa, R. H.; Orellana, W.; Fazzio, A. Appl. Phys. Lett. 86, 213111 (2005).

    (11) Berber, S.; Oshiyama, A. Phys. Rev. Lett. 96, 105505 (2006).

    (12) Peng, G. W.; Huan, A. C. H.; Liu, L.; Feng, Y. P. Phys. Rev. B 74, 235416 (2006).

    (13) Steiner, M.; Freitag, M.; Tsang, J. C.; Perebeinos, V.; Bol, A. A.; Failla, A. V.; Avouris, PH.

    Appl. Phys. A 96, 271-282 (2009).

    (14) Kohn, W.; Sham, L. J. Phys. Rev. 140, A1133- A1138 (1965).

    (15) Kleinman, L.; Bylander, D. M. Phys. Rev. Lett. 48, 1425-1428 (1982).

    (16) Troullier, N.; Martins, J. L. Phys. Rev. B 43, 1993-1996 (1991).

    (17) Ordejón, P.; Artacho, E.; Soler, J. M. Phys. Rev. B 53, R10441- R10444 (1996).

    (18) Soler, J. M.; Artacho, E.; Gale, J. D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. J.

    Phys.: Condens. Matter 14, 2745-2779 (2002).

  • 15

    (19) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 77, 3865-3868 (1996).

    (20) Anderson, N.; Hartschuh, A.; Cronin, S.; Novotny, L. J. Am. Chem. Soc. 127, 2533-2537 (2005).

    (21) Maciel, I. O.; Anderson, N.; Pimenta, M. A.; Hartschuh, A.; Qian, H.; Terrones, M.; Terrones,

    H.; Campos-Delgado, J.; Rao, A. M.; Novotny, L.; Jorio, A. Nature Mat. 7, 878-883 (2008).

    (22) Huang, J.; Choi, W. Nanotechnology 19, 505601 (2008).

    (23) Dresselhaus, M. S.; Dresselhaus, G.; Saito, R.; Jorio, A. Phys. Rep. 409, 47-99 (2005).

    (24) Araujo, P. T.; Pesce, P. B. C.; Dresselhaus, M. S.; Sato, K.; Saito, R.; Jorio, A. Physica E 42,

    1251-1261 (2010).

    (25) Bonnell, D. A. Scanning Probe Microscopy and Spectroscopy (Wiley-VCH, New York, 2001).

    (26) Barboza, A. P. M.; Gomes, A. P.; Archanjo, B. S.; Araújo, P. T.; Jorio, A.; Ferlauto, A. S.;

    Mazzoni, M. S. C.; Chacham, H.; Neves, B. R. A. Phys. Rev. Lett. 100, 256804 (2008).

    (27) Barboza, A. P. M.; Gomes, A. P.; Chacham, H.; Neves, B. R. A. Carbon 48, 3287-3292 (2010).