Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. ·...

24
Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh 1,2† , Jesse Kinder 2† , Lihong H. Herman 1† , Sang-Yong Ju 2 , Michael A. Segal 2 , Jeffreys N. Johnson 2 , Garnet K.-L. Chan 2 and Jiwoong Park 2,3 * Although metallic nanostructures are useful for nanoscale optics, all of their key optical properties are determined by their geometry. This makes it difficult to adjust these properties independently, and can restrict applications. Here we use the absolute intensity of Rayleigh scattering to show that single-walled carbon nanotubes can form ideal optical wires. The spatial distribution of the radiation scattered by the nanotubes is determined by their shape, but the intensity and spectrum of the scattered radiation are determined by exciton dynamics, quantum-dot-like optical resonances and other intrinsic properties. Moreover, the nanotubes display a uniform peak optical conductivity of 8 e 2 /h, which we derive using an exciton model, suggesting universal behaviour similar to that observed in nanotube conductance. We further demonstrate a radiative coupling between two distant nanotubes, with potential applications in metamaterials and optical antennas. D eveloping novel optical nanostructures with well-defined optical spectra and spatial scattering mode is an exciting scientific challenge, with possible applications in nanoscale optics, photovoltaics and bioimaging 1–6 . One approach to manipu- lating electromagnetic waves and localized energy at optical frequencies is to develop optical nanostructures with desirable spectral responses and near-field optical coupling. Metal nanostruc- tures are frequently used in nanoscale optical devices. However, in most of these plasmonic structures, all key radiation properties, such as spatial modes, wavelength-dependence and scattering inten- sity, are determined by the geometry of the nanostructure, a restric- tion that potentially limits the accuracy and design flexibility of these devices 1,7 . A possibly more flexible alternative is to use optical nanostructures for which the spectral response can also be controlled by the intrinsic electronic structure on the nanoscale, independent of the macroscopic shape. Single-walled carbon nano- tubes are macromolecules with wire-like physical dimensions that may be ideal candidates for such an approach, for two main reasons. First, the physical dimensions of nanotubes (lengths much larger than optical wavelengths) make them an analogue of thin conducting wires. As a result, the light scattered from nano- tubes resembles the radiation pattern of long thin wires in classical radiation theory 8 (upper panel, Fig. 1a). Second, single-walled carbon nanotubes display sharp optical resonances 9–15 in which the resonant frequency scattering and absorption strength is con- trolled by the nanoscale structure, in particular the chiral index (n, m) 9,15–18 (middle and lower panels, Fig. 1a). Combined, these properties make single-walled carbon nanotubes prototypical one- dimensional optical nanostructures, or excitonic optical wires, the spectral response of which is mainly controlled by the intrinsic quantum-mechanical properties of the material, whereas the scat- tered field can be independently determined by their shape. Furthermore, the cylindrical symmetry of nanotubes allows a quantitatively accurate description of the scattered optical field even in the near-field regime, the strength of which decreases much more slowly compared to that of a localized dipole (upper panel, Fig. 1a) 19 . Rayleigh scattering in carbon-nanotube optical wires The full characterization of single-walled carbon nanotubes as exci- tonic optical wires requires an accurate description of several key par- ameters, including the scattering pattern, spectral response, strength of the scattered optical fields and the optical conductivity s(v). For this, we measure the absolute Rayleigh scattering intensity from indi- vidual nanotubes, with detailed spatial and spectral resolution. This allows a joint investigation of their classical wire-like properties and discrete quantum optical spectra. Previous optical experiments using Raman 14 , photoluminescence 9,12 and Rayleigh 10,11,13 spectroscopy elucidated the excitonic nature of optical excitations and their relationship to structural (n, m) indices, but they present an incom- plete picture because the relative rather than absolute intensities of optical responses were measured. Furthermore, although several classical antenna-like optical properties of carbon nanotubes were described in earlier theoretical studies 8,20–23 and experiments 24–26 ,a direct measurement of the spatial pattern for optical fields scattered by individual nanotubes has not been reported to date. To this end, we used an on-chip Rayleigh imaging platform recently developed in our laboratory to investigate elastic scattering from nanotubes placed directly on a solid transparent substrate 27 . Our setup produces high signal-to-noise ratio optical images of individual nanotubes under wide-field laser illumination with dif- fraction-limited spatial resolution as a function of the excitation wavelength l. In addition, it allows optical measurements for mul- tiple nanostructures at varying distances, which is essential for studying novel optical coupling behaviour. An example is shown in Fig. 1b, where a series of scattering images taken at different wave- lengths is processed to generate a coloured map of elastic scattering intensity. This method, combined with nanotube diameters (d exp ) obtained from calibrated atomic force microscopy (AFM) measure- ments, allows accurate identification of interband transitions 1 School of Applied and Engineering Physics, Cornell University, Ithaca, New York 14853, USA, 2 Department of Chemistry and Chemical Biology, Cornell University, Ithaca, New York 14853, USA, 3 Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, New York 14853, USA; These authors contributed equally to this work. *e-mail: [email protected] ARTICLES PUBLISHED ONLINE: 19 DECEMBER 2010 | DOI: 10.1038/NNANO.2010.248 NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 1 © 2010 Macmillan Publishers Limited. All rights reserved.

Transcript of Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. ·...

Page 1: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

Single-walled carbon nanotubes as excitonicoptical wiresDaniel Y. Joh1,2†, Jesse Kinder2†, Lihong H. Herman1†, Sang-Yong Ju2, Michael A. Segal2,

Jeffreys N. Johnson2, Garnet K.-L. Chan2 and Jiwoong Park2,3*

Although metallic nanostructures are useful for nanoscale optics, all of their key optical properties are determined by theirgeometry. This makes it difficult to adjust these properties independently, and can restrict applications. Here we use theabsolute intensity of Rayleigh scattering to show that single-walled carbon nanotubes can form ideal optical wires. Thespatial distribution of the radiation scattered by the nanotubes is determined by their shape, but the intensity andspectrum of the scattered radiation are determined by exciton dynamics, quantum-dot-like optical resonances and otherintrinsic properties. Moreover, the nanotubes display a uniform peak optical conductivity of ∼8 e2/h, which we deriveusing an exciton model, suggesting universal behaviour similar to that observed in nanotube conductance. We furtherdemonstrate a radiative coupling between two distant nanotubes, with potential applications in metamaterials andoptical antennas.

Developing novel optical nanostructures with well-definedoptical spectra and spatial scattering mode is an excitingscientific challenge, with possible applications in nanoscale

optics, photovoltaics and bioimaging1–6. One approach to manipu-lating electromagnetic waves and localized energy at opticalfrequencies is to develop optical nanostructures with desirablespectral responses and near-field optical coupling. Metal nanostruc-tures are frequently used in nanoscale optical devices. However, inmost of these plasmonic structures, all key radiation properties,such as spatial modes, wavelength-dependence and scattering inten-sity, are determined by the geometry of the nanostructure, a restric-tion that potentially limits the accuracy and design flexibility ofthese devices1,7. A possibly more flexible alternative is to useoptical nanostructures for which the spectral response can also becontrolled by the intrinsic electronic structure on the nanoscale,independent of the macroscopic shape. Single-walled carbon nano-tubes are macromolecules with wire-like physical dimensions thatmay be ideal candidates for such an approach, for two mainreasons. First, the physical dimensions of nanotubes (lengthsmuch larger than optical wavelengths) make them an analogue ofthin conducting wires. As a result, the light scattered from nano-tubes resembles the radiation pattern of long thin wires in classicalradiation theory8 (upper panel, Fig. 1a). Second, single-walledcarbon nanotubes display sharp optical resonances9–15 in whichthe resonant frequency scattering and absorption strength is con-trolled by the nanoscale structure, in particular the chiral index(n, m)9,15–18 (middle and lower panels, Fig. 1a). Combined, theseproperties make single-walled carbon nanotubes prototypical one-dimensional optical nanostructures, or excitonic optical wires, thespectral response of which is mainly controlled by the intrinsicquantum-mechanical properties of the material, whereas the scat-tered field can be independently determined by their shape.Furthermore, the cylindrical symmetry of nanotubes allows aquantitatively accurate description of the scattered optical fieldeven in the near-field regime, the strength of which decreases

much more slowly compared to that of a localized dipole (upperpanel, Fig. 1a)19.

Rayleigh scattering in carbon-nanotube optical wiresThe full characterization of single-walled carbon nanotubes as exci-tonic optical wires requires an accurate description of several key par-ameters, including the scattering pattern, spectral response, strengthof the scattered optical fields and the optical conductivity s(v). Forthis, we measure the absolute Rayleigh scattering intensity from indi-vidual nanotubes, with detailed spatial and spectral resolution. Thisallows a joint investigation of their classical wire-like properties anddiscrete quantum optical spectra. Previous optical experiments usingRaman14, photoluminescence9,12 and Rayleigh10,11,13 spectroscopyelucidated the excitonic nature of optical excitations and theirrelationship to structural (n, m) indices, but they present an incom-plete picture because the relative rather than absolute intensities ofoptical responses were measured. Furthermore, although severalclassical antenna-like optical properties of carbon nanotubes weredescribed in earlier theoretical studies8,20–23 and experiments24–26, adirect measurement of the spatial pattern for optical fields scatteredby individual nanotubes has not been reported to date.

To this end, we used an on-chip Rayleigh imaging platformrecently developed in our laboratory to investigate elastic scatteringfrom nanotubes placed directly on a solid transparent substrate27.Our setup produces high signal-to-noise ratio optical images ofindividual nanotubes under wide-field laser illumination with dif-fraction-limited spatial resolution as a function of the excitationwavelength l. In addition, it allows optical measurements for mul-tiple nanostructures at varying distances, which is essential forstudying novel optical coupling behaviour. An example is shownin Fig. 1b, where a series of scattering images taken at different wave-lengths is processed to generate a coloured map of elastic scatteringintensity. This method, combined with nanotube diameters (dexp)obtained from calibrated atomic force microscopy (AFM) measure-ments, allows accurate identification of interband transitions

1School of Applied and Engineering Physics, Cornell University, Ithaca, New York 14853, USA, 2Department of Chemistry and Chemical Biology, CornellUniversity, Ithaca, New York 14853, USA, 3Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, New York 14853, USA; †These authorscontributed equally to this work. *e-mail: [email protected]

ARTICLESPUBLISHED ONLINE: 19 DECEMBER 2010 | DOI: 10.1038/NNANO.2010.248

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 1

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 2: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

(denoted by EiiS and Eii

M for semiconducting and metallic types,respectively) and structural (n, m) indices for individual nanotubes.In addition to determining the relative Rayleigh scattering intensityat different l, we were also able to measure the absolute Rayleighscattering intensity and characterize the scattering propagationmode for individual nanotubes for the first time, using a wide-field illumination geometry that excites the entire length ofaligned nanotubes with uniform laser intensity, polarization andimpinging angle (see Supplementary Information).

The observed propagation pattern of scattered light fromindividual nanotubes is described by classical scattering theoryand radiation theory, which yield the same prediction whenlNT .. l .. dNT .. .. (where lNT is the nanotube length anddNT the diameter) (Fig. 1a,c). In this regime, a nanotube behavesas a long radiating wire with a surface current I(z,v)¼pdNTs(v)E(z,v) excited by a laser, where E(z,v) is the electricfield along the nanotube axis (Fig. 1c). (Note that the same radiationequations hold for a semiconducting tube as for a metallic tube: inthe former case, the oscillating polarization a(v) implies a surfacecurrent, and the relation between the polarizability per unitsurface area and the conductivity are given by a(v)¼ is(v)/v).In our experiment, a collimated laser (beam diameter, �70 mm) isused to excite the nanotube with an electromagnetic plane waveincident at angle uex with respect to the nanotube axis, which deter-mines the direction of the scattered electromagnetic wave8,28. Whenuex¼ 908, the scattered field is predicted by classical scatteringtheory to become focused onto a rotationally symmetric disk

perpendicular to the nanotube axis. This is indeed the behaviourwe observed from our nanotube, as shown in the Fraunhoferimage measured with a charge-coupled device (CCD) placedafter the objective lens with additional optical elements, whichmeasures the spatial distribution of the scattered light (Fig. 1d; seeSupplementary Information). In contrast, if uex = 908, the scatteredradiation is focused onto a rotationally symmetric cone having thesame direction as the excitation laser (Fig. 1e). In addition, nano-tubes display a highly directed scattering pattern, as seen from thenarrow angular width (,78) of the propagation disk/cone inFig. 1d,e (measured from the width of the bright strip). We alsoobserve strong polarization dependence in the scattered light inten-sity, consistent with previous optical studies on nanotubes10 (seeSupplementary Information).

Although the spatial pattern of the scattered radiation under laserillumination is determined by the macroscopic shape of the nano-tube, its intrinsic material properties at the nanoscale determinethe scattering spectra and intensity. This leads to the possibility ofindependent control of the spectral profile and scattering pattern.The correlation between Rayleigh spectra and nanotube structure(chirality) can be seen in Fig. 1b, where different nanotubesdisplay distinct colours that are uniform along the entire length ofa nanotube, independent of its shape (straight versus curved; seeSupplementary Information), and this novel property presents amajor advantage of these excitonic optical wires over their metalcounterparts. To demonstrate that the spectral response of a nano-tube is independent of its shape, we patterned individual nanotubes

c

d

e

b Wavelength (nm)

650 550

20 μm

2 μm

450

θexLaser

θex=90°

Detector

−40.3°

+40.3°

θ det

a z

r

1|E|scatt r∞

eh

Eexθex

I

θdet CCD

OL TL

Eex

SWNT

Laser

θex ≠ 90°

θex

Laser

−40.3°

+40.3°

θ det

kz

ES22

Ener

gy

ΓX

Scat

terin

g in

tens

ity (a

.u.)

Energy

Figure 1 | Optical scattering from single-walled carbon nanotubes. a, Upper panel: schematic illustrating electromagnetic radiation from a long, straight wire.

Intensity plot corresponds to the strength of the scattered electric field (Escatt), which decays as 1/p

r. Middle panel: schematic of an electronic band diagram

for a semiconducting nanotube (where kz represents the azimuthal quantum number). An interband transition from valence to conduction band (red arrow)

resulting from optical excitation leads to the formation of an excitonic bound state that dissociates into free particles in lower sub-bands at rate GX (dashed

blue arrows). Lower panel: optical excitation and exciton formation (e–h) in a nanotube by an electromagnetic plane wave (Eex). The solid arrow indicates the

direction of propagation of the incident photons. b, Representative spatial Rayleigh image, showing more than 10 nanotubes simultaneously (in false colour,

corresponding to strongly scattered wavelengths). c, Optical setup schematic. The incident electric field Eex is linearly polarized and incident at angle uex with

respect to the nanotube axis, inducing a surface current I along the nanotube. The scattered light is focused onto a disk or cone around the nanotube axis

and is collected with a detection angle udet¼ 80.68. The Fraunhofer image is effectively obtained by removing the tube lens (TL) and placing the CCD

immediately after the objective lens (OL) to image the nanotube radiation pattern, as indicated by the dotted blue line (see Supplementary Information).

d,e, Main panels: radiation behaviour for nanotubes. Spatial Rayleigh images (upper inset) and radiation patterns from Fraunhofer imaging (lower inset) of

individual nanotubes when uex¼ 908 (d) and when uex = 90 (e).

ARTICLES NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.248

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology2

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 3: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

into shorter periodic segments with well-defined separation dis-tances (h) by cutting them using a high-intensity supercontinuumlaser beam that was focused and raster-scanned on the sampleplane (Fig. 2a; see Supplementary Information). Comparing theRayleigh spectra before and after patterning shows that nanotubesmaintain their spectral profiles even after being cut into smaller per-iodic segments (Fig. 2b). We also characterized the spatial distri-bution of the light scattered by these patterned nanotubes usingFraunhofer imaging, and found that the scattering pattern showsadditional side peaks (cones denoted by asterisks) in addition tothe central peak (disk), resembling the interference patterns of clas-sical diffraction gratings or multiple-slit apertures (Fig. 2c,d).

Uniform optical conductivity on resonanceUnlike the scattering patterns, the scattering spectra and intensity ofcarbon-nanotube optical wires are quantum-mechanical in natureand determined by the exciton dynamics through a single par-ameter, the complex conductivity s(v). Plotting the scatteringintensity as a function of excitation energy for individual single-walled carbon nanotubes reveals strong optical resonances(Fig. 3a), where the observed peak corresponds to a scattering reson-ance in which the photon energy matches the energy of an allowedexciton transition (middle and lower panels, Fig. 1a)9–11,15. By com-bining Rayleigh scattering and AFM information, we assign specifictransitions to each optical resonance27. Figure 3a demonstrates suchassignments for five representative nanotubes, all of which show asingle narrow peak at visible wavelengths, with the resonanceenergy determined by the diameter and electronic type. The relationbetween the scattering spectra (as well as the absolute scattering effi-ciency at a given wavelength) and the dynamic conductivity s(v) isgiven by equation (1), which also describes the radiation efficiencyof long conducting wires (see Supplementary Information for deri-vation)29. To measure the absolute scattering intensity of nanotubeson and off resonance, we obtained the scattering cross-section width(dscatt) from the Rayleigh intensity data after accounting for the col-lection and transmission efficiency of the optics, the quantum effi-ciency of the camera and the laser intensity at the nanotubeposition, which were all measured at the excitation wavelength(see Supplementary Information). In general, we observed thatdscatt on resonance ranges from �0.2 pm to 2.3 pm and increasesnonlinearly with nanotube diameter (Fig. 3b). Measurements ofdscatt at the resonance wavelength for 19 different nanotubes areplotted in Fig. 3c, which shows that all data points follow thescaling curve dscatt/ d NT

2 /l. This is in agreement with the classicalradiation model, which predicts

dscatt =12p

3Z2|s(v)|2d2l−1 (1)

for an infinite conducting wire (or pipe)29 that carries a surfacecurrent I(z,v)¼ pdNTs(v)E(z,v), where Z is the impedance ofthe medium (308 V for glycerol/quartz in our experiment), withthe perhaps surprising implication (discussed below) that s(v) onresonance is uniform across different single-walled carbon nano-tubes. Although there are qualitative theoretical treatments20,22,23,a quantitative measurement of the absolute value of s (v) forsingle-walled carbon nanotubes with known diameter and elec-tronic type has not been measured experimentally.

As mentioned above, the scaling behaviour in Fig. 3c suggeststhat the dynamic conductivity on resonance (|s|peak) is independentof nanotube diameter and chiral index (n, m) (see SupplementaryInformation for (n, m) assignments). Using equation (1), weextracted |s|peak for all 19 nanotube resonances in our sample.They display a similar value for |s|peak, showing a narrow distri-bution: |s|peak¼ (8.3+1.5)G0, where G0¼ e2/h is the natural unitfor sheet conductivity (inset, Fig. 3c) and is also well-known

as the d.c. conductance quantum of one-dimensional wires.We observe little difference between nanotubes with different diam-eters, or between different interband transitions (not including theE11

S transition). This unusual behaviour suggests that the opticalproperties of nanotubes might be described by a universal model,much like the d.c. conductance of single-walled nanotubes.

Exciton models have previously been shown to be important inexplaining the chiral index dependence of optical excitations9,15–

18,30. Here, we find by considering the exciton dynamics that wecan also understand the uniform dynamic conductivity on reson-ance observed above, as well as other features of the scattering.Most of the oscillator strength of an interband transition such asE22

S is concentrated in a single exciton level, leading to a singlesharp peak in absorption and scattering spectra16,18,30. The radiativelifetime of the exciton strongly affects the dynamic conduc-tivity18,30,31, with shorter lifetimes leading to broader peaks andlower peak conductivity. Excluding the lowest interband transition,the lifetime is determined primarily by the rate of dissociation of theexciton into free carriers in lower bands17.

We calculated the exciton spectrum, wavefunctions and line-widths for an interband transition, and used these to determinethe conductivity at the level of linear response theory (seeSupplementary Information). A key finding is that the excitonlinewidth GX is inversely proportional to the nanotube diameter(schematic in Fig. 1b), which underlies the uniform optical behav-iour we observe. The dissociation process associated with theexciton lifetime is found to be on the order of tens of femtoseconds,

aBefore

After

d

−10 +10

θ (deg)

h = 2.0 µm

h = 1.5 µm

Intensity (a.u.)

**

* *

0

c

θ

Inte

nsity

h

**

Norm

alized intensity (a.u.)

0.0

0.5

1.0

0.0

0.5

1.0

b

2.0 2.4eV

Figure 2 | Scattering patterns for periodically spaced nanotube segments.

a, Colour Rayleigh image before (upper panel) and after (lower panel)

cutting. Scale bar, 5 mm. b, Rayleigh spectra (normalized to the peak)

before and after patterning for the nanotube marked by the symbols in

a. c, Schematic illustrating the scattering pattern for periodically spaced

(h¼ pitch distance) linear structures. d, Upper panel: Fraunhofer image for a

nanotube with h¼ 2.0 mm after cutting. Lower panel: measured scattering

patterns for nanotubes of different pitch sizes after cutting. Asterisks denote

side bands depicted in c.

NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.248 ARTICLES

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 3

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 4: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

in good agreement with previous experimental work and theoreticalcalculations18,32. Our computed peak optical conductivities are inde-pendent of nanotube diameter, and vary little over the transitionsprobed in the experiment (inset, Fig. 3d). Our model shows a mag-nitude consistent with the experimental data (7–10G0), even thoughother effects such as trigonal warping for E11

M transitions and (n, m)-dependent electron–phonon coupling may contribute to thevariation seen in Fig. 3c. As an example, the calculated dynamicconductivity for the E33

S transition of a semiconducting nanotubeis shown in Fig. 3d, where we plot the real, imaginary and absoluteconductivity (denoted as Re[s(v)], Im[s(v)] and |s(v)|, respect-ively). The peak widths predicted by the model are also consistentwith the experimental data. Furthermore, the calculated complexconductivities s(v) provide an additional check on the experimen-tal results. Within the exciton model, the imaginary optical conduc-tivity is nearly zero on resonance and Re[s(v)] ≈ |s|peak (Fig. 3d).Consequently, the measured absolute conductivity |s|peak yields adirect estimate of the peak absorption cross-section width (dabs),given by dabs(v)¼ pdNTZ Re[s(v)]. This leads to a scaling relationfor dabs at the peak: dabs ≈ (0.3)dNT (equivalent to 2.5 × 10217 cm2

per carbon atom), consistent with recent time-resolved lumines-cence studies on individual (6, 5) nanotubes33.

Long-range optical coupling between nanotubesThe Rayleigh scattering properties of nanotubes discussed so far(directed radiation pattern, wavelength dependence and uniformpeak optical conductivity) suggest that single-walled carbon nano-tubes indeed behave as unique excitonic optical wires, and presentan ideal optical element in which the spectral response and scatteredfield can be determined accurately in both near- and far-fieldregimes. The integration of these nanoscale optical wires in anarray geometry (where novel properties often emerge) may lead tothe development of novel metamaterials and optical devices, andthis requires an understanding of their coupling behaviour. Infact, nanotubes are expected to couple strongly, because the magni-tude of the scattered (electric) field is expected to decrease as 1/

pr

near nanotubes (instead of 1/r) thanks to the wire-like shape ofthe nanotubes (Fig. 1a)19. This may result in a radiative couplingbetween two parallel nanotube wires over a relatively long distance(much longer than the typical resonant energy transfer distance ofseveral nanometres)34. Furthermore, because nanotubes are stronglyabsorptive on resonance and reflective otherwise (see Fig. 3d), thisrapid phase change of s(v) can generate a novel coupling behav-iour between neighbouring nanotubes with strong wavelengthdependence.

b

c

2.2 nm

dexp = 0.7 nm

1.3 nm

dexp (nm)0.0 0.5 1.0 1.5 2.0 2.5

0.0

0.5

1.0

0 5 10 150

5

10

Count no.

Con

duct

ivity

(e2 /h

)d

M11EM22ES22ES33ES44E

dNT (nm)1 2 3

0

5

10

|σ|peak (G0)

|σ| pe

ak (G

0)

Energy (eV)

scat

λ(n

m)

1.54 eV

1.90 eV

2.20 eV

2.35 eV

2.44 eV

1.15 nm

1.31 nm

1.46 nm

0.98 nm

1.50 nm

1.5 2.5

M11E

S33E

S33E

M11E

S22E

a

Scat

terin

g in

tens

ity (a

.u.)

|σ(ω)|Im[σ(ω)]Re[σ(ω)]

1 2 3

−5

0

5

10

eV

2 μm

Figure 3 | Frequency-dependent scattering intensity and optical conductivity of single-walled carbon nanotubes. a, Colour Rayleigh images of five

nanotubes, together with their corresponding resonant Rayleigh peaks and diameters determined by AFM (height traces shown in red). b, Normalized spatial

Rayleigh scattering intensity plots for tubes of different diameter (dexp) measured at their respective resonance energies. c,p

(dscattl) as a function of dexp,

measured for 19 nanotube resonances of various sub-band transitions. Inset: histogram of peak dynamic conductivity |s|peak in units of G0 (e2/h). d, Main

panel: calculated plots of Re[s(v)], Im[s(v)] and |s(v)| versus �v for a representative semiconducting nanotube using the exciton model. Inset: calculated

values of |s|peak for different resonances and nanotube diameters using the exciton model (same symbols as in c).

ARTICLES NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.248

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology4

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 5: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

We demonstrate such long-range coupling in an AFM image(Fig. 4a) of a pair of nanotubes that has segments running parallelover 10 mm while maintaining a separation of �300 nm. Thesenanotubes have Rayleigh peaks at 634 nm (nanotube L) and664 nm (nanotube R) with significant spectral overlap. We assignchiral indices of (18, 7) for nanotube L and (14, 5) or (11, 8) fornanotube R based on our Rayleigh scattering and AFM measure-ments. A spatial map of the Rayleigh scattering spectrum revealsunusual behaviour in the overlap region, where the position of thescattering signal shifts abruptly from one nanotube to the other,with a slight change in excitation wavelength, leaving the originallybright nanotube almost invisible (insets, Fig. 4b). This unusual

coupling behaviour is shown more clearly in the middle panel ofFig. 4c, where the Rayleigh intensity is plotted as a function of wave-length and position for the interacting region. Outside the coupledregion, however, each nanotube has a typical Rayleigh scatteringspectrum, as shown in the left and right panels (L and R, respect-ively) of Fig. 4c. Figure 4d compares the intensities from nanotubesL and R (after Gaussian deconvolution) in the coupled anduncoupled regions. A sharp transition (width, ,10 nm) was againobserved near 650 nm. Moreover, below the transition wavelength,the intensity from nanotube L is enhanced as much as twofoldat the expense of the intensity of nanotube R; above the transition,the roles are reversed.

M11ES

33E

Data

1.35 nm

1.75 nm

300 nm

Wav

elen

gth

(nm

)

Wavelength (nm)

Scat

terin

g in

tens

ity (a

.u.)

Fit

600

650

700

Position

750

2.4 μm

Max

Min

LR

SWNT R

SWNT L

664 nm

635 640 645 650 655 660

620 640 660 680

R-Uncoupled

L-Uncoupled

R-Coupled

L-Coupled

λλ

λL λR

634 nm

a b

c

dCou

pled

regi

on

L R

L RL R

1 μm

Figure 4 | Radiative coupling between distant single-walled carbon nanotubes. a, AFM image of two nanotubes with overlapping spectral resonances

centred at 634 nm (lL) and 664 nm (lR) (for nanotubes L and R, respectively). The two nanotubes run parallel to each other over a distance of 10mm.

Corresponding height traces are shown in red. b, Main panels: optical coupling (insets show spatial Rayleigh images near the coupling region, denoted by

blue lines) recorded at 634 and 664 nm. Re[s(v)] represented by solid lines and Im[s(v)] by dashed lines. Scale bar, 5 mm. c, Rayleigh scattering intensity

contour of excitation wavelength versus position for the green cross-sections marked in the inset of b for coupled (middle panel) and uncoupled (left and

right panels) regions. d, Intensity contributions for nanotubes L and R in the uncoupled and coupled regions obtained from Gaussian deconvolution (left

inset) of c. Inset: calculated values for the observed transition. Arrows indicate the changes in the scattering intensity of tube R before and after

intertube coupling.

NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.248 ARTICLES

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 5

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 6: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

Although intertube coupling and even exciton transfer betweensingle-walled carbon nanotubes have been reported in bundlednanotubes35–37, the transfer of brightness between two distanttubes as shown in Fig. 4 has not been observed previously. Adirect radiative coupling between the two nanotube optical wiresprovides a qualitative explanation for the observed long-range coup-ling behaviour (schematic shown in Fig. 4b). The driving field of onenanotube is a superposition of the applied field and the scatteredfield of the other nanotube. When two distant nanotubes in oursample are driven in phase, the scattered field of a reflective nano-tube (off resonance) may amplify the driving field of the other(on resonance), increasing its effective optical conductivity andthe radiation intensity. The scattered field of an absorptive nanotubecan reduce the driving field of the other, decreasing its conductivity.We developed a more quantitative model using classical radiationtheory and an effective mean field description of the conductivity,which accounts for the geometric phase, multiple scattering, nearfields and the phase behaviour of s(v) (Fig. 3d) (calculations aredescribed in the Supplementary Information). Our model qualitat-ively reproduces the transition (inset, Fig. 4d) observed in our exper-iment, and warrants further study of correlated local fluctuations tofully understand the mechanism responsible for the abrupt tran-sition observed. This radiative coupling might provide a meansfor studying intertube exciton energy transfer and for improvingthe peak absorption efficiency of nanotubes in parallel arrays.

ConclusionsSingle-walled carbon nanotubes are ideal molecular optical wires,providing a novel platform for investigating the coupling betweenoptical fields and electrical conductors at the nanometre scale.The agreement between our model calculations and the experimen-tal data shows that accurate predictions of the properties of thesesystems are possible. In addition, the long-range radiative couplingdemonstrated in this work provides a new design principle for chan-nelling optical energy to other nanostructures and individualmolecules in the vicinity of nanotubes, an exciting prospect withbroad impact in molecular imaging, photodetection and solarenergy harvesting.

Received 25 October 2010; accepted 12 November 2010;published online 19 December 2010

References1. Bharadwaj, P., Deutsch, B. & Novotny, L. Optical antennas. Adv. Opt. Photon. 1,

438–483 (2009).2. Schuck, P. J., Fromm, D. P., Sundaramurthy, A., Kino, G. S. & Moerner, W. E.

Improving the mismatch between light and nanoscale objects with gold bowtienanoantennas. Phys. Rev. Lett. 94, 017402 (2005).

3. Muhlschlegel, P., Eisler, H. J., Martin, O. J. F., Hecht, B. & Pohl, D. W. Resonantoptical antennas. Science 308, 1607–1609 (2005).

4. Farahani, J. N., Pohl, D. W., Eisler, H. J. & Hecht, B. Single quantum dot coupledto a scanning optical antenna: a tunable superemitter. Phys. Rev. Lett. 95,017402 (2005).

5. Alu, A. & Engheta, N. Tuning the scattering response of optical nanoantennaswith nanocircuit loads. Nature Photon. 2, 307–310 (2008).

6. Taminiau, T. H., Stefani, F. D., Segerink, F. B. & Van Hulst, N. F. Opticalantennas direct single-molecule emission. Nature Photon. 2, 234–237 (2008).

7. Kelly, K. L., Coronado, E., Zhao, L. L. & Schatz, G. C. The optical properties ofmetal nanoparticles: the influence of size, shape, and dielectric environment.J. Phys. Chem. B 107, 668–677 (2003).

8. Slepyan, G. Y., Shuba, M. V., Maksimenko, S. A. & Lakhtakia, A. Theory ofoptical scattering by achiral carbon nanotubes and their potential as opticalnanoantennas. Phys. Rev. B 73, 195416 (2006).

9. Wang, F., Dukovic, G., Brus, L. E. & Heinz, T. F. The optical resonances incarbon nanotubes arise from excitons. Science 308, 838–841 (2005).

10. Sfeir, M. et al. Probing electronic transitions in individual carbon nanotubesby Rayleigh scattering. Science 306, 1540–1543 (2004).

11. Sfeir, M. et al. Optical spectroscopy of individual single-walled carbonnanotubes of defined chiral structure. Science 312, 554–556 (2006).

12. Bachilo, S. et al. Structure-assigned optical spectra of single-walled carbonnanotubes. Science 298, 2361–2366 (2002).

13. Berciaud, S. et al. Excitons and high-order optical transitions in individualcarbon nanotubes: a Rayleigh scattering spectroscopy study. Phys. Rev. B 81,041414 (2010).

14. Dresselhaus, M., Dresselhaus, G., Saito, R. & Jorio, A. Raman spectroscopy ofcarbon nanotubes. Phys. Rep. 409, 47–99 (2005).

15. Dresselhaus, M. S., Dresselhaus, G., Saito, R. & Jorio, A. Exciton photophysics ofcarbon nanotubes. Annu. Rev. Phys. Chem. 58, 719–747 (2007).

16. Ando, T. Excitons in carbon nanotubes. J. Phys. Soc. Jpn 66, 1066–1073 (1997).17. Kane, C. L. & Mele, E. J. Ratio problem in single carbon nanotube fluorescence

spectroscopy. Phys. Rev. Lett. 90, 207401 (2003).18. Spataru, C. D., Ismail-Beigi, S., Benedict, L. X. & Louie, S. G. Quasiparticle

energies, excitonic effects and optical absorption spectra of small-diametersingle-walled carbon nanotubes. Appl. Phys. A 78, 1129–1136 (2004).

19. Jackson, J. D. Classical Electrodynamics 3rd edn (Wiley, 1998).20. Slepyan, G. Y., Maksimenko, S. A., Lakhtakia, A., Yevtushenko, O. &

Gusakov, A. V. Electrodynamics of carbon nanotubes: dynamic conductivity,impedance boundary conditions, and surface wave propagation. Phys. Rev. B60, 17136–17149 (1999).

21. Burke, P. J., Li, S. D. & Yu, Z. Quantitative theory of nanowire and nanotubeantenna performance. IEEE Trans. Nanotechnol. 5, 314–334 (2006).

22. Hanson, G. W. Fundamental transmitting properties of carbon nanotubeantennas. IEEE Trans. Antenn. Propag. 53, 3426–3435 (2005).

23. Hao, J. & Hanson, G. W. Infrared and optical properties of carbon nanotubedipole antennas. IEEE Trans. Nanotechnol. 5, 766–775 (2006).

24. Kempa, K. et al. Carbon nanotubes as optical antennae. Adv. Mater. 19,421–426 (2007).

25. Wang, Y. et al. Receiving and transmitting light-like radio waves: antenna effectin arrays of aligned carbon nanotubes. Appl. Phys. Lett. 85, 2607–2609 (2004).

26. Cubukcu, E. et al. Aligned carbon nanotubes as polarization-sensitive, molecularnear-field detectors. Proc. Natl Acad. Sci. USA 106, 2495–2499 (2009).

27. Joh, D. et al. On-chip Rayleigh imaging and spectroscopy of carbon nanotubes.Nano Lett. doi:10.1021/nl1012568 (2010).

28. Bohren, C. & Huffman, D. Absorption and Scattering of Light by Small Particles(Wiley, 1998).

29. Abbott, T. A. & Griffiths, D. J. Acceleration without radiation. Am. J. Phys. 53,1203–1211 (1985).

30. Perebeinos, V., Tersoff, J. & Avouris, P. Scaling of excitons in carbon nanotubes.Phys. Rev. Lett. 92, 257402 (2004).

31. Spataru, C. D., Ismail-Beigi, S., Capaz, R. B. & Louie, S. G. Theory and ab initiocalculation of radiative lifetime of excitons in semiconducting carbon nanotubes.Phys. Rev. Lett. 95, 247402 (2005).

32. Manzoni, C. et al. Intersubband exciton relaxation dynamics in single-walledcarbon nanotubes. Phys. Rev. Lett. 94, 207401 (2005).

33. Berciaud, S., Cognet, L. & Lounis, B. Luminescence decay and the absorptioncross section of individual single-walled carbon nanotubes. Phys. Rev. Lett. 101,077402 (2008).

34. Scholes, G. D. Long-range resonance energy transfer in molecular systems.Annu. Rev. Phys. Chem. 54, 57–87 (2003).

35. Wang, F. et al. Interactions between individual carbon nanotubes studied byRayleigh scattering spectroscopy. Phys. Rev. Lett. 96, 167401 (2006).

36. Ju, S. Y., Kopcha, W. P. & Papadimitrakopoulos, F. Brightly fluorescent single-walled carbon nanotubes via an oxygen-excluding surfactant organization.Science 323, 1319–1323 (2009).

37. Tan, P. H. et al. Photoluminescence spectroscopy of carbon nanotube bundles:evidence for exciton energy transfer. Phys. Rev. Lett. 99, 137402 (2007).

AcknowledgementsThe authors thank P.L. McEuen for useful discussions, and Y.J. Kim and R. Havener forassistance with sample fabrication and numerical modelling. This work was supported bythe National Science Foundation (NSF) through the Center for Nanoscale Systems, CornellCenter for Materials Research, Center for Chemical Innovation and an NSF CAREER grant.Additional funding was received from the David and Lucile Packard Foundation, Alfred P.Sloan Foundation, Camille and Henry Dreyfus Foundation, and the US Department ofDefense through the Air Force Office of Scientific Research. Sample fabrication wasperformed at the Cornell Nanoscale Science and Technology Facility, a NationalNanotechnology Infrastructure Network node.

Author contributionsD.J. and L.H. performed optical measurements and analysed the data. J.K. carried outtheoretical calculations. S.-Y.J. and J.J. performed Raman spectroscopy measurements.M.S. carried out nanotube synthesis. G.C. and J.P. supervised the project.

Additional informationThe authors declare no competing financial interests. Supplementary informationaccompanies this paper at www.nature.com/naturenanotechnology. Reprints andpermission information is available online at http://npg.nature.com/reprintsandpermissions/.Correspondence and requests for materials should be addressed to J.P.

ARTICLES NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.248

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology6

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 7: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

nature nanotechnology | www.nature.com/naturenanotechnology 1

1

Supplementary Online Information for

Single-Walled Carbon Nanotubes as Excitonic Optical Wires

Daniel Y. Joh1,2*, Jesse Kinder2*, Lihong H. Herman1*, Sang-Yong Ju2, Michael A.

Segal2, Jeffreys N. Johnson2, Garnet K.-L. Chan2, and Jiwoong Park2,3†

*These authors contributed equally to this work

†To whom correspondence should be addressed. Email: [email protected]

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 8: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

2 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

2

Supplementary Information Contents Experimental Apparatus for On-Chip Rayleigh Imaging

Figure S1. On-chip Rayleigh imaging setup.

Measuring Spatial Distribution of Scattered Light (Fraunhofer Imaging)

Figure S2: Schematic of SWNT Fraunhofer imaging.

Polarization Dependence

Figure S3: Polarization dependence for SWNT Rayleigh scattering.

Spectral response for bent SWNTs

Figure S4: Optical scattering from curved vs. straight regions of an isolated

SWNT.

Figure S5: Radiation directivity dependence of the observed Rayleigh scattering

intensity.

Patterning SWNTs with a Laser

Figure S6: Schematic for laser-induced cutting of SWNTs

Derivation of Equation (1)

Chiral Index Assignments for SWNTs in Figure 3c and Figure 4

Figure S7: Data points from Fig. 3C with labels

Table S1: List of diameters, resonant energies, transition and possible (n, m)

values

Measurement of scattering cross section width on SWNT resonance

Models for Dynamic Conductivity and Radiative Coupling of SWNTs

A. Peak Conductivity

B. Radiative Coupling

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 9: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 3

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

3

Experimental Apparatus for On-Chip Rayleigh Imaging

Aligned SWNTs were grown on a quartz substrate following a procedure adapted

from Kocabas et al. (Small, 1, 1110 (2005)). A tuneable wavelength laser beam is

linearly polarized and introduced into an index-matched sample geometry at an angle

using a darkfield condenser (DCW), and is reflected completely by total internal

reflection at the uppermost glass/air interface and subsequently directed back down

through the condenser. This geometry prevents the excitation laser from entering the

detection optics. The light scattered from SWNTs is collected by an objective lens (OL)

(UIS2 PlanSAPO 40x, N.A. 0.95) and focused onto a CCD camera by the tube lens (TL).

For further details, see reference 27 of the main text.

Figure S1. On-chip Rayleigh imaging setup. See text above for abbreviations.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 10: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

4 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

4

Measuring Spatial Distribution of Scattered Light (Fraunhofer Imaging)

Fig. S2 describes the experimental geometry used to measure the spatial

distribution of the scattered light from individual SWNTs. The scattered light was

collected by a high numerical aperture (N.A. = 0.95) objective lens (OL). An iris was

placed at the image forming plane of the tube lens (TL) to observe only one SWNT at a

time and then an additional lens was used to collect the spatial distribution of scattered

light from the same SWNT, which was imaged using a CCD camera.

Figure S2: Schematic of SWNT Fraunhofer imaging. See text above for

abbreviations.

Polarization Dependence

The Rayleigh scattering amplitude for SWNTs is strongly polarization dependent,

reaching a maximum when the electric field is polarized parallel to the nanotube axis.

This effect was demonstrated in previous Rayleigh scattering experiments (Science, 306,

1540 (2004)), and we show polarization-dependent behavior in Fig. S3 that is consistent

with these previous measurements. In Fig. S3 we show optical images of a SWNT in

which the polarization of the excitation laser (Pex) and that of an analyzer placed after the

objective lens (Pdet) are parallel (||) and/or perpendicular () to the nanotube axis. The

data show that the nanotube is only visible when both Pex and Pdet are parallel to the

SWNT axis.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 11: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 5

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

5

Figure S3: Polarization dependence for SWNT Rayleigh scattering. See text

above for abbreviations. Scale bar 4, m.

Spectral response for bent SWNTs

In this section, we show how the spectral response for SWNTs is mainly

controlled by the intrinsic properties of the material, while the spatial propagation of the

scattered field is independently determined by their shape. Fig. S4a shows an AFM

image of an initially straight SWNT that eventually curves along its length. By

recording Rayleigh spectra in curved and straight regions (sample spectra shown in Fig.

S4d), we observe similar wavelength dependence along the length of the nanotube.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 12: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

6 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

6

Figure S4: Optical scattering from curved vs. straight regions of an isolated

SWNT. a, AFM image of the SWNT (nanotube highlighted in white for clarity).

Scale bar, 5 m. b, c, Corresponding Rayleigh images taken at 1.88 eV (b) and

2.15 eV (c). d, Rayleigh spectra recorded at the locations marked by Greek

symbols in (a).

Although generally we observe uniform wavelength-dependence for this nanotube in both

straight and curved regions, a closer look at spatial profiles (Figs. S4b and S4c) of the

measured Rayleigh scattering intensity (at resonance) reveals distinct orientation-

dependence (i.e. ex dependence); for instance, the measured Rayleigh scattering intensity

for curved region is lower than that of straight region for both resonances at 1.88 eV

and 2.15 eV. We attribute this decrease in intensity to two major factors: 1) polarization-

dependent scattering and 2) radiation directivity.

To further illustrate this point, we first show the AFM image of a different

nanotube (dAFM = 1.46 nm) that bends much more sharply along its length (Fig. S5a) than

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 13: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 7

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

7

the SWNT shown in Fig S4. Fig. S5b spatially maps the measured Rayleigh scattering

intensity. The optical signal of the SWNT is brightest in regions where ex = 90o and

disappears completely in the middle region where ex ≠ 90o. Since the polarization-

dependent scattering amplitude of SWNTs is proportional to cos2(90o - ex), this predicts

that the SWNT’s optical intensity in the outlined region should be approximately one-half

that of the bright regions, which is not observed. The further reduction of the optical signal

in the outlined region can be attributed to the directed radiation pattern of SWNTs: when ex

is much steeper than 90o, the resulting radiation cone may lie outside of the detection angle of

our optics (Fig. S5e).

Figure S5: Radiation directivity dependence of the observed Rayleigh scattering

intensity. a, b, Composite AFM image (a) and optical image (b) of the SWNT,

with the bent region outlined in red. Scale bar, 2 m. c, Laser excitation

geometry of SWNTs d, e, Direction-dependent scattering of SWNTs. When ex

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 14: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

8 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

8

becomes much different than 90o, the radiation cone may lie outside of our

detection optics (e).

Patterning SWNTs with a Laser

The Rayleigh images of the sample containing aligned SWNTs were first

obtained to locate the nanotubes before any cutting was performed. Next, high-intensity

supercontuum light (NKT Photonics, Koheras SuperK Compact) was focused onto the

sample plane (glycerol-quartz interface) after passing through a wideband hot mirror (W)

(Thorlabs FM01) such that the peak laser pulse power after hot mirror is ~420 W. The

stage was then raster-scanned in the x-y plane to pattern the tubes into periodic segments

as shown in Fig. 2.

Figure S6: Schematic for laser-induced cutting of SWNTs. See text above for

abbreviations.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 15: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 9

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

9

Derivation of Equation (1)

For an infinite straight wire, the total radiated power per unit length (Prad) averaged over

time is given by 218 | ( ) |radP Zk I , where I() is the surface current (Abbott, T. A. &

Griffiths, D. J. Am. J. Phys. 53, 1203-1211 (1985), ref [29] ). Furthermore, if the incident

beam has an electric field ( ) i toE E e , the total scattered light power is

2 1102 | |scatt scattP E Z . In this regime, the two expressions for power are equivalent. Equating

the two expressions for power yields 3 2 2 2 112 | ( ) |scatt NTZ d . This expression is

consistent with Bachilo, S. et al. Science 298, 2361-2366 (2002), ref [12], except for its

dependence on nanotube diameter: this is due to our treatment of nanotubes as long hollow

cylinders (or wires), instead of long solid cylinders.

Chiral Index Assignments for SWNTs in Figure 3c and Figure 4

For SWNTs shown in Figs. 3c & 4, we provide a table of possible (n, m) assignments

based on Rayleigh scattering and AFM measurements. We assigned unique (n, m) values for

about half of the SWNTs studied, while for the other half they were narrowed down to two or

more possible indices due to the high density of optical resonances for larger diameter nanotubes.

The labels for data points in Fig. S7, which is identical to Fig. 3c except for the labels, correspond

to those in Table S1. The possible structural (n, m) index assignments for the SWNTs in Fig. 4

are indicated by L and R at the end of the table.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 16: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

10 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

10

Figure S7: Data points from Fig. 3c with labels corresponding to those in Table S1.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 17: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 11

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

11

Table S1: SWNT diameters (dexp), experimentally determined resonant energies, electronic

transition, and possible (n, m) values corresponding to the data points shown in Fig. S7. The last

two rows show the same information for the SWNTs in Fig. 4 of the main text.

Measurement of scattering cross section width on SWNT resonance

The scattering cross section width of an individual SWNT was measured by first

identifying its resonance wavelength from its Rayleigh spectrum. In our setup, the light

scattered by a SWNT is detected with a detection angle (det) of 80.6 out of 360 degrees.

Thus, the total scattering detection efficiency in our setup is given by (80.6/360)

×TOE()×QECCD(), where TOE is the combined transmission efficiency of all optical

elements and QECCD the quantum efficiency of the CCD, both measured at the resonance

wavelength, . This enabled us to calculate the resonance scattering intensity (Iscatt,

[number of photons/um×sec]) per unit length of the SWNT. We then determined the

power density per unit area of the excitation laser at the SWNT position (Ilaser, number of

photons/um2×sec]) by fitting the illumination profile to a 2D Gaussian. Dividing Iscatt by

Ilaser yields the scattering cross section width, scatt.

Models for Dynamic Conductivity and Radiative Coupling of SWNTs

Here we will briefly describe the method used to calculate the conductivity data in

Fig. 3. Our starting point is the Kubo formula for the conductivity:

2

2 2

| | | | 2( )( )2 ( )i RL E E i

J

.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 18: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

12 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

12

| indicates the ground state, | is an excited state, E is the energy of the excitation,

is the width of the excitation, R is the nanotube radius, L is the length of the tube, and

is the frequency of the laser.

The low energy properties of carbon nanotubes can be calculated from an

effective Dirac Hamiltonian (Phys. Rev. Lett., 78, 1932 (1997)). Imposing periodic

boundary conditions on the linear k p approximation to the tight-binding model near the

K and K' points of graphene generates a series of hyperbolic bands with the dispersion

relation

2 2( ) ( ) ( )F nE k v R kR

where 300Fv c is the Fermi velocity. The band gaps are 2 ( )n Fv R with on n .

n is an integer, and o 0 for metallic nanotubes (giving rise to a linear band for 0n )

and o 1/ 3 in semiconducting nanotubes. Each band is 4-fold degenerate: a factor of

2 for spin and an additional factor of 2 for the degenerate K and K' points.

o FE v R defines a natural energy scale for the nanotube. oE also defines an effective

mass: * 2oFm v E , so that *

Fm v R .

With these definitions, the Kubo formula can be expressed in terms of a

dimensionless oscillator strength:

2* | | | | ,zvf m

E

The expression for the conductivity becomes

2

* 2 2

2( )( ) .2 ( )

e fi RL m E i

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 19: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 13

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

13

The optical properties of carbon nanotubes are strongly influenced by electron

interactions. The Coulomb interaction leads to a large self energy that increases the

transport gap. It also leads to the formation of strongly bound electron-hole pairs –

excitons – below the band edge. Most of the spectral weight of an absorption or

scattering peak is localized in a single exciton transition.

We calculated the exciton wave functions and energies using a model first

developed by Ando (J. Phys. Soc. Jpn. 66, 1066 (1997)). In this model, the free electron

and hole energies and wave functions are calculated using the effective Dirac

Hamiltonian. We modified the Dirac Hamiltonian to include trigonal warping effects

near the band edge. These energies and wave functions are used to calculate the

polarization and static dielectric function within the random phase approximation, which

leads to a screened Coulomb interaction ( )V k . The electron and hole self energies,

( )e k and ( )h k , are calculated within the screened Hartree Fock approximation. The

exciton wave functions and energies are obtained as solutions of a Bethe Salpeter

equation:

( ) [ ( ) ( ) ( ) ( )] ( ) ( ) ( )p p h hq

E k k k k k k V q k q

We solved this equation in a two-band model. This is a good approximation for

excitons in the lowest band of a semiconductor 11( )SE , but excitons in higher bands can

dissociate into unbound electron-hole pairs in lower bands. This gives the exciton a finite

lifetime, which we calculate using the golden rule following Kane and Mele (Phys. Rev.

Lett. 90, 207401 (2003)). Since the experiments were performed in glycerol (n1.46), we

used a static dielectric constant of = n2 = 2.1 to account for static screening of the

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 20: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

14 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

14

Coulomb interaction. RPA screening roughly doubles this value and leads to an exciton

self energy of o0.05E for all excitons. This is several orders of magnitude larger

than the line width due to radiative recombination, so we ignore the latter.

To obtain the data in Fig. 3, we used the calculated exciton wave functions,

energies, and lifetimes in the Kubo formula for the conductivity.

A. Peak Conductivity

The exciton model predicts a peak conductivity that is independent of the

nanotube diameter. This can be inferred from the Kubo formula when a large fraction of

the oscillator strength is concentrated in a single transition, as is the case for excitons in

carbon nanotubes. The oscillator strength in the Dirac model of noninteracting electrons

satisfies a sum rule:

2Lf

R

Using this sum rule, the conductivity may be expressed as

2o

2 * 2 2o

2( )1( ) ,(2 ) ( )

Eei R m E E i

where a

f f is the fraction of the total oscillator strength in the transition to | .

The sum is a dimensionless number and the prefactor sets the scale of the AC

conductivity. Using the definitions of Eo and *m given above, the conductivity is

2o

2 2

2( )1( )2 ( )

Eeh i E i

.

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 21: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 15

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

15

We can place an upper bound on the conductivity by assuming all of the oscillator

strength is concentrated in a single transition: X . This is a good approximation to

the exciton model, where 0.6 0.8x , and the lowest exciton peak is well separated in

energy from the rest of the spectrum. In this case,

2o

2 2

2( )1( ) ,2 ( )X X

Eeh i E i

and the conductivity on resonance is

2 2o

2 2F X

PX

E ve eh h R

If the exciton lifetime X is proportional to the nanotube radius, then the peak

conductivity will be independent of the radius.

Using the exciton linewidth due to dissociation into free electrons and holes

o( 0.05 )X E gives an upper limit on the peak conductivity of 23.2P e h per

exciton.

The model also predicts that the peak conductivities in some metallic nanotubes

should be roughly twice as large as those in semiconducting nanotubes. The n = ±1 bands

in a metallic armchair nanotube are degenerate. In chiral nanotubes, trigonal warping lifts

the degeneracy so that individual peaks can be resolved. The data show little difference in

the peak conductivity between semiconducting and metallic nanotubes, which suggests

that all five of the metallic nanotubes in the sample were chiral where the two slightly

shifted resonances were unresolved, and thus form a single peak with its height similar to

the uniform value 8 e2/h in the range of wavelengths probed in the experiment. We expect

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 22: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

16 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.248

16

a wider range of peak conductivities in samples containing metallic nanotubes of all

chiralities.

While other optical experiments have demonstrated the necessity of an exciton

picture of the excitations, a peak conductivity that is independent of the nanotube radius

is not specific to the exciton model here. We have calculated the peak conductivity in a

free particle model with a constant scattering rate and a free particle model with acoustic

phonon scattering, models in which the oscillator strength is not concentrated in a single

transition. The peak conductivity is independent of the radius as long as the lifetime is

inversely proportional to the nanotube radius. (This will be described elsewhere.)

B. Radiative Coupling

Fig. 4 shows a strong radiative coupling between two parallel nanotubes with

similar Rayleigh spectra. As the laser frequency shifts from one resonance to the other,

there is a large transfer of spectral weight between the nanotubes. A classical model of

two parallel conducting wires can reproduce the qualitative features of this crossover.

However, the experimental values of the conductivity are too small to induce a detectable

effect. This suggests that a theory of correlated local fluctuations is necessary to fully

understand the mechanism responsible for the abrupt transition observed in the

experiment. Here we describe the radiative coupling between two parallel wires with

overlapping resonances.

The vector potential r due to an arbitrary current density J(x) is

| |1( ) ( )| |

ikedc

r x

A r x J xr x

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 23: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

nature nanotechnology | www.nature.com/naturenanotechnology 17

SUPPLEMENTARY INFORMATIONdoi: 10.1038/nnano.2010.248

17

and the radiated field at r is ( ) (1/ ) ( ) kE r A r . We consider an infinite wire with

conductivity driven by an oscillating field 0() applied along its axis. The field

induces a uniform surface current 0( ) 2 ( ) ( ) I R E . The scattered field at a point

distance D away from the wire is parallel to the applied field:

02

1( ) 2 ( ) ( ) ( )

SiE R kD E

D

where ( ) kD is a dimensionless geometric factor obtained by integrating along the

length of the wire:

22 2

12 5/ 2 2 2 2 3/ 2

2 2( ) ( )(1 ) (1 ) (1 )

ikD xikD k DkD dx ex x x

The scattered field is proportional to the applied field: 0( ) ( ) ( ) SE E .

If two parallel wires A and B are driven by the same applied field, then the

electric field at one wire is a superposition of the applied field and the scattered field of

the other wire. This leads to a set of coupled equations:

0( ) ( )[ ( ) ( )] A A BE E E

0( ) ( )[ ( ) ( )] iB B AE E e E ,

where accounts for the phase difference in the applied field between the two wires. For

instance, if the light is incident on the plane of the two wires at angle , then sin kd .

The scattered fields are

01 ( )( ) ( ) ( )

1 ( ) ( )

iB

A AA B

eE E

01 ( )( ) ( ) ( )

1 ( ) ( )

iA

B BA B

eE E

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 24: Single-walled carbon nanotubes as excitonic optical wiressyju/PDF/13.pdf · 2016. 2. 16. · Single-walled carbon nanotubes as excitonic optical wires Daniel Y. Joh1,2†,JesseKinder2†,LihongH.Herman1†,Sang-YongJu2,

18 nature nanotechnology | www.nature.com/naturenanotechnology

SUPPLEMENTARY INFORMATION doi: 10.1038/nnano.2010.24818

The scattered fields are affected by many different phases: the geometric phase

from ( ) kD , the phase due to the angle of incidence of the driving field, and the phase of

the conductivity as the frequency of the applied field passes through a resonance.

When ,| ( ) | 1 A B , the scattered fields are those of independent wires. However,

when ,| ( ) |~ 1 A B , the scattered field of a wire is strongly affected by that of the other

wire. For the wires in Fig. 4, ≈ 650 nm, D ≈ 300 nm, and kD ≈ 3. The geometric factor

( ) kD is on the order of 1, so the coupling strength is determined by the conductivity.

For a peak conductivity of 10 e2/h,

25

2( ) 10 10

e R

D

This is much too small to modify the scattered fields when two wires run parallel to one

another, and shows that a mean field description of the scattered fields cannot explain the

observed coupling. The simple mean field description of light scattering used to calculate

the conductivity, which correctly describes the optical scattering properties of single

nanotubes, ignores effects such as spatial and temporal fluctuations in the surface

currents and correlations between the two nanotubes. The size of the surface currents

required to reproduce the qualitative features of Fig. 4 suggests that large fluctuations on

short length and time scales may be responsible for the strong radiative coupling. Since

there is currently no theoretical model for these effects, we treated the effective

conductivity as a free parameter in analyzing the data. If the surface currents are

enhanced by a factor of 1000, it is possible to reproduce the qualitative features of the

sharp transition in Fig. 4. The parameters used to generate the plot were D = 300 nm, n =

1.47, EA = 1.85 eV, EB = 1.98 eV, A = 0.12 eV, B = 0.15 eV, and sin = 1.33/1.47.

© 2010 Macmillan Publishers Limited. All rights reserved.