<<

Model for quantitative tip-enhanced spectroscopy and the extraction of nanoscale-resolved optical constants

Alexander S. McLeod,1, ∗ P. Kelly,1 M. D. Goldflam,1 Z. Gainsforth,2 A. J. Westphal,2 Gerardo Dominguez,3, 1 Mark H. Thiemens,1 Michael M. Fogler,1 and D. N. Basov1 1University of California, San Diego 2Space Laboratory, University of California at Berkeley 3California State University San Marcos (Dated: October 29, 2018) Near-field infrared spectroscopy by elastic scattering of light from a probe tip resolves optical contrasts in materials at dramatically sub-wavelength scales across a broad energy range, with the demonstrated capacity for chemical identification at the nanoscale. However, current models of probe-sample near-field interactions still cannot provide a sufficiently quantitatively interpreta- tion of measured near-field contrasts, especially in the case of materials supporting strong surface phonons. We present a model of near-field spectroscopy derived from basic principles and verified by finite-element simulations, demonstrating superb predictive agreement both with tunable quantum cascade laser near-field spectroscopy of SiO2 thin films and with newly presented nanoscale Fourier transform infrared (nanoFTIR) spectroscopy of crystalline SiC. We discuss the role of probe ge- ometry, field retardation, and surface mode dispersion in shaping the measured near-field response. This treatment enables a route to quantitatively determine nano-resolved optical constants, as we demonstrate by inverting newly presented nanoFTIR spectra of an SiO2 thin film into the frequency dependent dielectric function of its mid-infrared optical phonon. Our formalism further enables tip- enhanced spectroscopy as a potent diagnostic tool for quantitative nano-scale spectroscopy.

I. INTRODUCTION ple point dipole model25,26, in which radiation scattered from a small polarizable sphere of radius a illuminated by Since Synge’s 1928 letter to Einstein proposing a bold an incident field Einc is modulated through electrostatic method for optical imaging beyond the diffraction limit1, interaction with a material surface located a distance d sub-wavelength optical characterization techniques have away in the z-direction: remained subjects of intensive interest and fierce debate owing to their transformative potential. Among such α αeff Pz /Einc = (1) techniques, apertureless near-field scanning optical mi- ≡ 1 αβ/16π(a + d)3 2,3 − croscopy (ANSOM) has shattered the diffraction limit,  1 achieving optical resolutions better than λ/1000 at in- with α 4πa3 and β − . ≡ ≡  + 1 frared and THz frequencies4–6. Recent coupling of ANSOM to a broadband coherent Here α denotes the “bare” polarizability of the sphere infrared light source and asymmetric Michelson inter- producing a vertical dipole moment Pz, and β denotes ferometer has enabled Fourier transform infrared spec- the quasi-static limit of the Fresnel coefficient rp(q, ω). troscopy at the nanometer length scales (nanoFTIR)7–9, A function of both frequency ω and in-plane momentum in switchable combination with single-frequency imaging q, the Fresnel coefficient describes the relative magnitude by the pseudo-heterodyne (PSHet) detection scheme10,11. and phase of p-polarized light reflected from the surface These novel interferometric techniques detect both am- of material with frequency-dependent dielectric function plitude and phase12–14 of the probe-scattered “near-field (ω). signal”, which encodes nano-scale near-field optical con- While important theoretical advances have trasts from the sample and transmits them to the far- brought near-field spectroscopy beyond qualitative field. While applications to nanoscale chemical sensing descriptions27–32, available models describing the probe- at vibrational “fingerprint” energies are obvious9,15,16, sample near-field interaction remain beset by critical the utility of this instrument for fundamental nano- limitations: arXiv:1308.1784v3 [cond-mat.mtrl-sci] 30 Jun 2014 scale studies of correlated electron systems are equally 17–24 1. Many general formulations, although formally ex- compelling . act, prove cumbersome to implement for practical ANSOM employs a conductive or dielectric AFM calculation beyond reduction to the simple point probe as both an intense near-field source and scatterer dipole model33,34. Field retardation and antenna of light into the far-field. The mechanism of optical con- effects of the probe are explored formally, but not trast has long been understood intuitively via the sim- quantitatively. 2. Although the near-field interaction may be de- scribed as an exact scattering problem, many so- ∗ [email protected] lution methods rely on perturbation expansions 2

in powers of the sample response factor β or sponse to an incident illumination field, together with the 9,32,34,35 rp. One can show that such series are di- near-field of the proximate sample, we begin by forming vergent beyond modest values of rp (the “strong an expression for this instantaneous charge distribution. coupling” regime), leaving this method inapplica- Constraining our attention to nearly axisymmetric ble for the analysis of crystalline solids and strongly probe geometries, the charge distribution is succinctly 31 resonant plasmonic systems. expressed through a linear λQ(z) dQ/dz(z), Q denoting charge and z the probe’s axial≡ 3. A number of tunable geometric parameters with ad coordinate. In the quasi-static approximation, the field hoc or empirical values are introduced to quantita- Erad re-radiated or back-scattered from the probe is pro- tively fit measured data. These include fractional portional to its induced dipole moment weights of relevant probe surface charge,30,31 effec- tive probe size and geometry,28–30 the bare tip po- Z larizability, etc.20 The multitude of ad hoc tunable Pz = dz z λQ(z). (2) parameters provides an unreliable recipe for pre- dictive modeling or quantitative interpretation of Appendix E presents how the radiated field is obtained data. from λQ(z) when electrodynamic phenomena are of fun- damental importance, i.e. when the size of the scatterer To address these extant shortcomings, the aim of the is comparable to the light wavelength. This relationship present work is threefold. We first present a new model highlights the central role of the induced charge distri- of probe-sample near-field interaction, the rod bution in determining the measured observables of near- model, whose generality allows exploring the influence of field spectroscopy. both probe geometry and electrodynamic effects, while λQ(z) can be written as the sum of charges induced remaining formally exact in both theory and implemen- by the incident field and those differential contributions tation. Mathematical reduction of this formalism back to dλQ nf (z) induced by reflection of probe-generated near- the point dipole model will make clear that unnecessary fields off the proximate sample: ad hoc assumptions underpin prevailing models30,31, and that geometric and electrodynamic considerations must Z ultimately play a role in their vindication. λQ(z) = Einc Λ0(z) + dλQ nf (z). (3) Second, we demonstrate this model’s capability to pre- dict spectroscopic near-field contrast in the case of lay- Here Λ0(z) denotes the induced charge per unit field re- ered structures, which exhibit a strongly momentum- sulting from incident illumination. Its functional form dependent optical response, as well as strongly reso- depends on the nature of the incident field and the probe nant systems, through comparison with near-field spectra geometry, but its contribution to λQ(z) scales with the measured on thin films of silicon dioxide (SiO2) and bulk magnitude of the incident field Einc. The induced charge silicon carbide (SiC). Our measurement apparatus is a elements dλQ nf (z) take the form: novel infrared near-field equipped for both ZZ ∞ PSHet imaging and broadband nanoFTIR spectroscopy, 0 dλQ nf (z) = dQ dq dq (q)Γt→s(q) described in Appendix A. 0 G Finally, we present a method to invert the lightning rod 0 0 0 R(q, q )Γs→t(q ) Λ(q , z). (4) model to extract a material’s complex dielectric function × with nanoscale resolution, which we demonstrate explic- Here q and q0 denote in-sample-plane momenta for itly for an SiO2 thin film sample. This procedure, com- Fourier components of the near-field reflected by the bined with the unifying formalism of the lightning rod sample in response to the polarized probe. Provided a model, provides a powerful diagnostic tool for quantita- planar sample geometry, this parameterization offers a tively studying the nano-resolved optical properties of sparse basis in which to efficiently solve the problem, in molecular systems, phase-separated materials, and con- contrast with real-space treatments (e.g. the finite ele- fined nanostructures using ANSOM.17 ment method). Eq. (4) can be understood through the physical mechanism shown schematically in Fig. 1b and described as follows in terms of field emission from the II. THE LIGHTNING ROD MODEL probe and sample reflection. 0 0 Charge elements dQ = dz λQ(z ) on the probe form Our model describes the near-field interaction between rings with radii z0 along its surface. Considered in the an ANSOM probe and a sample surface through a general angular spectrumR decomposition (Appendix B), each ring formalism that is in principle exact, without appealing to emits a distribution of axisymmetric p-polarized evanes- empirical or ad hoc parameters. The chief observable of cent fields whose Fourier components are weighted by ANSOM is the radiation field of a polarized probe placed (q) = q J0(q z0 ). Ji(...) is the Bessel function of the in proximity to a sample (experimental details described firstG kind at orderR i, with cylindrical coordinates ρ and z. in Appendix A). Since the field originates from reorgani- These emitted fields (so-called evanescent Bessel beams) zation of charges developing on the probe surface in re- reach the sample surface a distance d below the tip apex 3

Λ(q, z)" a) b) c) [ norm.]

+½" 0" 0"

-½"

-1" Eq! 2" 5 µm 250 nm 1" 0" -1" -2"

FIG. 1. a) Scanning electron micrograph of a typical commercial near-field probe exhibiting a conical geometric profile and characteristic length scales (probe length and tip radius) separated by nearly three orders of magnitude. The surface profile (blue) is considered in Sec. VI. b) Schematic description of the probe-sample near-field interaction, involving emission of cylindrical evanescent fields from charge elements in the probe and their reflection by the sample. c) Probe response function

Λ(q, z) (defined in the main text) computed by the boundary element method (Appendix C) for evanescent fields E~q of increasing momenta q. Dashed curves indicate the geometric profile of the probe, and surface charge distribution profiles are normalized by their minimum values for viewing purposes.

−q(z0+d) (z = 0) via propagator Γt→s(q) = e and in the distributing a distance δz = 1/q along the probe surface. empty tip-sample gap d < z < 0 take the divergence- The momentum-dependent lightning rod effect is criti- free form (per unit charge):− cally absent from models of the probe-sample near-field interaction lacking a faithful geometric description. q(z−z0) E~q(~r) = (q)(J0(qρ)z ˆ + J1(qρ)ρ ˆ) e . (5) Confining our attention to planar sample geometries, G Eqs. (3), (4), and (6) together describe a self-consistent In general, the sample may subsequently scatter quasi-one-dimensional scattering process: evanescent fields with momentum q into distinct Fourier 0 Z ∞ components q as described through a differential sample −2qd 0 λQ(z) = Einc Λ0(z) dq λeQ(q) q e rp(q) Λ(q, z) response function R(q, q ). For samples with continu- − 0 · ous in-plane translational symmetry (e.g. flat surfaces, (7) layered structures) this response function reduces to the Z L 0 0 −qz0 Fresnel reflection coefficient for p-polarized light, with λeQ(q) dz λQ(z ) e J0(q z0 ). (8) ≡ 0 · R 0 0 R(q, q ) = rp(q) δ(q q ), (6) − The integral transform in Eq. (8) denotes summation of written here as a function of the in-plane momentum q of near-fields emitted from charges along the entire length of incident light, with δ(...) denoting the Dirac delta dis- the probe, 0 < z < L. A similar integral transformation tribution. These scattered fields extend from the sam- z s applied to λ(z) and Λ(q, z) in Eq. (7) yields an 0 −q0(z+d) → ple surface via propagator Γs→t(q ) = e and integral equation in λe(s): re-polarize the probe, inducing a linear charge density Z ∞ (per unit field) described by a probe response function λ (s) = E Λ (s) dq λ (q) q e−2qd r (q) Λ(q, s). 0 eQ inc e0 eQ p e Λ(q , z). This formalism accommodates the non-trivial − 0 · influence of realistic tip geometries on the functional form (9) of illumination- and sample-induced charge distributions Eq. (9) resembles the Lippman-Schwinger equation of 0 36 Λ0(z) and Λ(q , z). scattering theory . Here Λe0(s) and Λ(e q, s) play the role Induced charge densities can be pre-computed for an of in- and out-going “scattering states.” Tractability in axisymmetric probe of arbitrary geometry using a simple this scattering formalism is afforded by the axisymmet- boundary element method (Appendices C and D). Fig. ric approximation, which preserves most fundamental as- 1c displays Λ(q, z) for several values of q, computed for pects of the system geometry. the case of a hyperboloidal (conical with rounded tip) Provided knowledge of the probe response functions 0 probe geometry with tip curvature radius a. As shown Λ0(z) and Λ(q , z), Eq. (9) is soluble by traditional in Fig. 1c, the density of charge dramatically accumulat- methods37 after discretizing q to a set of Gauss-Legendre ing at the probe apex - the celebrated lightning rod effect nodes qi . We found evaluation at Nq 200 such nodes - increases roughly as 1/q. This results from the req- sufficient{ } for numerical accuracy to within≈ 1%. Only a uisite screening of evanescent fields by induced charges finite range of momenta 0 q qmax need be considered ≤ ≤ 4

−1 in practice, since Λe0(s) and Λ(e q, s) drop precipitously in decay in magnitude near s a . In this case, Eq. (9) ∼ magnitude above a cutoff momentum s 1/a, with a yields λeQ(s) in closed form owing to the separability of the smallest length scale relevant to the probe∼ geometry, Λ(e q, s): in this case the radius of curvature at the probe apex, a 30 nm for many commercial probe tips. This reflects a3 s e−sa ≈ −1 λ (s) = . (17) the inability of strongly confined fields (e.g. q nm ) eQ 3 R ∞− 2 −2q(d+a) 1 a dq q e rp(q) to efficiently polarize the probe. ∼ − 0 The solution to Eq. (9) is then The sphere’s polarization is obtained through Eqs. (7) ~ and (2) as ~ Λ0 λQ = Einc, (10) I ΛG a3 − α P /E = . (18) where the denominator is taken in the sense of matrix in- eff z inc 3 R ∞ 2 −2q(d+a) ≡ 1 a 0 dq q e rp(q) version, vectors imply functional evaluation at momenta − qi , I is the identity operator, and other matrices are If the sample material is weakly dispersive for q ω/c, { }  defined as rp(q) β and Eq. (1) is recovered. Such≈ simplifications are instructive, but this work −2qid Λij Λ(e qj, si) and Gij qi e rp(qi) δqi δij. ≡ ≡ − makes full implementation of Eq. (13) without recourse (11) to approximation, thus revealing aspects of the probe- Here δqi is the measure of qi and δij denotes the Kro- sample near-field interaction unresolved by the point necker delta function. Defining similarly a vector of dipole model. While perturbative expansions and the ~  charge distribution functions Λ(z) i Λ(qi, z) together point-dipole model may be attractive for their relative with their associated contributions to≡ the radiated field   simplicity, they are certainly not expected to be accu- ~erad Erad Λ~ (see Appendix E), the total induced ≡ rate. In particular, for large β, nothing short of the full charge and consequent back-scattered field are provided numerical solution to Eq. (11) is acceptable for predict- through Eq. (7) as: ing experimental observables with quantitative reliabil- Λ~ ity. Our procedure for doing so is detailed in the following λ (z) /E = Λ (z) + Λ(~ z) G 0 (12) sections. Q inc 0 I ΛG · The near-field experiments presented in this work uti- −~ Λ0 lize lock-in detection of the probe’s back-scattered field at Erad /Einc = erad, 0 + ~erad G . (13) · I ΛG harmonics n of the probe tapping frequency Ω to suppress − Note that dependence on both the tip-sample distance d noise and unwanted background. Simulating this tech- and the local optical properties of the sample enter these nique, the probe’s back-scattered field Erad (Eq. (13)) is expressions through G, whereas geometric properties of connected to experimentally observed amplitude Sn and the probe enter separately through Λ. When applied to phase φn signals through a sine transform under sinu- a realistic probe geometry, these expressions constitute soidally varying tip-sample distance d: the lightning rod model of probe-sample interaction, so Z π/Ω named for its quantitative description of the strong elec-  sn(ω) = (ω) dt sin (nΩt) Erad d, rp(q, ω) (19) tric fields localized by an elongated geometry to a pointed I −π/Ω apex. The product of Λ and G signifies that strong near- with d = A (1 + sin (Ωt)) , fields from the probe multiplicatively enhance optical in- teractions with the sample surface. Expanding the Eq. Sn(ω) sn(ω) , and φn arg sn(ω) . (20) (13) inverse matrix as a geometric series reveals an infi- ≡ | | ≡ { } nite sequence of probe polarization and sample reflection Here Ω and A are the tapping frequency and amplitude of events, equivalent to the perturbation expansions pre- the near-field probe, respectively, and (ω) denotes the sented elsewhere9,32,34,35. frequency-dependent instrumental responseI of the collec- Eq. (13) can also recover the point dipole model (Eq. tion optics, interferometer and detector used for the mea- (1)). After simplifying the probe geometry to a metallic surement. This factor can be removed by normalizing ex- sphere of radius a and assuming that all center-evaluated perimental sn(ω) to “reference” near-field signal values, (z = a) fields polarize like the homogeneous incident field as collected from a uniformly reflective sample material Einc, we have: such as gold or undoped silicon. This normalization pro- cess is further discussed in Sec. IV. Λ(q, z) = 3/2 (z a) e−qa, (14) − A prediction of near-field contrast using the lightning Λ(e q, s) = a3 s e−(s+q)a, (15) rod model therefore requires calculation of Eq. (13) at − Z 2a several values of d; in practice we find 20 such values suf- 3 −qa and pz(q) = dz z Λ(q, z) = a e . (16) ficient, with evaluation of Eq. (13) for each requiring sev- 0 eral milliseconds on a single 2.7 GHz processor. Cumu- Λ(e q, s) is obtained from semicircular z in Eq. (8), and latively, the calculation remains both realistic and fast, exhibits a characteristic maximum followedR by a sharp more so than previously reported semi-analytic solutions 5 for realistic probe geometries.38,39 For example, typical calculations of a demodulated and normalized near-field spectrum across 100 distinct frequencies require less than 10 seconds of computation time. We conclude this formal introduction with a concep- L tual clarification. Although the geometry and material composition of the near-field probe implicitly determine its response function Λ(q, z), the formalism embodied by Eq. (13) is general and outwardly irrespective of specific a properties of the probe. Also, while plasmonic enhance- SiO2 phonon! ment may be encompassed in Λ(q, z), it is not a prereq- uisite for effective near-field enhancement at the probe apex. Near-field enhancement is attainable through a combination of three distinct mechanisms:40 1. the lightning rod effect proper, due to accumulation of charge at geometric singularities, an essentially electrostatic effect, FIG. 2. Spectral near-field contrast between the 1130 cm−1 2. plasmonic enhancement, due to the correlated mo- surface phonon polariton of SiO2 and silicon (providing nor- tion of surface charges near the plasma frequency malization) as predicted by the lightning rod model for an of , ellipsoidal probe in the quasi-static approximation. Contrast increases monotonically beyond experimentally observed lev- 3. and antenna resonances, in which the size of an els as the probe length is increased. optical antenna correlates with the incident wave- length in a resonant fashion, a purely electrody- namic effect. previous work38, demonstrating the capacity of near-field The quasi-static boundary element utilized in this work spectroscopy to resolve the ω 1130 cm−1 surface opti- ≈ (Appendix C) reproduces the first of these mechanisms cal phonon of thermal oxide films as thin as 2 nm. We by way of Λ(q, z), whereas its electrodynamic counterpart extend the theoretical study presented therein to inves- (Appendix D) reproduces all three. Although plasmonic tigate the influence of the probe length L on the ampli- enhancements are scarcely attainable in metallic probes tude of experimentally measurable back-scattered near- at infrared frequencies, Secs. III and IV establish the field signal S3(ω) (normalized to silicon) predicted by the important influence of both the lightning rod effect and lightning rod model. The outcome is presented in Fig. 2. antenna resonances in near-field spectroscopy. The probe tapping amplitude is 80 nm in these simu- lated experiments, and the radius of curvature a at the probe apex (equal to the inverse surface concavity) is held III. THE QUASI-STATIC CASE constant at 30 nm, typical of experiments with commer- cially available near-field probes. The minimum probe- We now apply Eq. (13) to realistic probe geome- sample distance is taken as d = 0 nm throughout (viz. tries in the quasi-static limit to investigate whether the physical contact, consistent with the established descrip- quasi-static approximation is appropriate for quantita- tion of tapping mode AFM). We describe the thin-film tive prediction of near-field contrasts. By reducing the optical response with a momentum-dependent Fresnel physical system to , this approximation is coefficient41 (further discussed in Sec. V) using optical strictly justified only in treating light-matter interactions constants of thermal oxide taken from literature42. at length scales much smaller than the wavelength of The probe response function Λ(q, z) is computed in the light, whereas a typical near-field probe consists of an quasi-static approximation once for each probe geometry AFM tip tens of microns in height (Fig. 1a), compara- according to a simple boundary element method. Math- ble to typical wavelengths encountered in infrared near- ematical details are provided in Appendix C. Whereas field spectroscopy. Consequently, for the assumptions of in this work we present calculations only for ideally con- a quasi-static probe-sample interaction to remain valid, ducting metallic probes, Appendix C presents also the the emergent behavior of a realistic near-field probe must general formulation suitable for application to dielec- be shown nearly equivalent to those of a deeply sub- tric probes. Consequently, the case of a dielectric probe wavelength one. presents no formal difficulty for the model presented here. To test this assumption, we consider the near-field in- However, previously reported models present a suitably teraction between a metallic ellipsoidal probe oriented simpler description of the “weak coupling regime” in vertically over a planar sample of 300 nm of thermal sil- which externally excited near-fields may be mapped non- 13,43,44 icon dioxide (SiO2) on silicon substrate. This sample perturbatively . We identify this as the regime in system and model probe geometry were considered in a which a perturbation expansion of Eq. (13) is found to 6

a) b)

λ = 10 µm |Esca / Einc| 1" Esca Einc

Einc θ L

0"

FIG. 3. a) Scattered field of a realistic near-field probe geometry under plane wave illumination (incident along the viewing angle) as computed by the finite-element method. Oscillatory fields near the tip apex are associated with standing wave-like surface charge densities resulting from field retardation. b) Lower panel: Field enhancement at the tip apex computed quasi-statically (QS) and electrodynamically (ED) for two probe geometries of varying size illuminated perpendicular to their principle axes. Upper panel: Near-field S3 contrast between SiO2 at the surface optical phonon resonance (ω SO) and silicon simulated by the fully retarded lightning rod model. The vertical dashed line indicates the length of a typical near-field probe, L ≈ 19 µm. converge, and several terms therein might be summed for IV. THE ELECTRODYNAMIC CASE: a sufficient estimate of near-field scattering. NEAR-FIELD PROBE AS ANTENNA As shown in Fig. 2, the most dramatic feature of our quasi-static calculations is the strong variation in normal- Near-field microscopy is occasionally described as an ized scattering amplitude with increasing probe length at antenna-based technique, in which the antenna-like near- the probe-sample polariton resonance. The implication is field probe efficiently converts incident light into strongly worrisome: there is no clear rational choice for “effective confined fields at the probe-sample feed-gap17,45–47. The probe length” when computing the strength of probe- antenna’s scattering cross section is consequently modu- sample near-field interaction in the quasi-static approxi- lated through strong interactions with the sample sur- mation. With a free-space wavelength of light λ 10 µm, face to provide the nano-resolved optical contrasts of although the largest credibly quasi-static probe∼ length ANSOM4,48. At a formal level, these considerations leave L λ/10 (or L 20a) provides reasonable qualitative the mathematical form of the lightning rod model unal- agreement∼ with data∼ acquired by nanoFTIR under identi- tered; nevertheless, the probe response function Λ(q, z) cal experimental conditions (Fig. 7), quantitative agree- must encapsulate the probe’s role as an antenna, particu- ment is clearly only attainable a posteriori, for example larly in the probe’s response Λ0(z) to illumination fields. by fitting to agreeable values of L. Furthermore, the As for any antenna, due to retardation and radiative extreme dependence on probe geometry exhibited here effects, the field scattered by a near-field probe is mani- discredits the utility of quantitative “fits” to experimen- festly dependent on both its size relative to the free space tal data. The ill-posed description of near-field coupling light wavelength as well as its geometric profile relative afforded by this quasi-static treatment lacks clear predic- to the incident light polarization. Such electrodynamic tive power. We are therefore compelled towards a con- effects have been demonstrated experimentally47,49. To sistent electrodynamic treatment, which as we will show characterize how these length scales influence the observ- provides an unambiguous description of near-field inter- ables of near-field spectroscopy, the full electrodynamic actions with wavelength-scale probes – a case applicable responses of two probe geometries were computed numer- to the vast majority of near-field experiments at infrared ically as a function of their overall length L relative to and THz frequencies. the free-space wavelength of incident light. 7

A fully retarded boundary element method taking ac- geometries47. Due to the axially polarized incident field, count of field retardation and radiative forcing (math- Resonant enhancement modes of the hyperboloidal probe ematical details provided in Appendix D) was used to are less pronounced and more complicated in character; calculate charge distributions Λ0(z) induced on metallic we discuss them here in no further detail. Finally, it is ellipsoidal and hyperboloidal probe geometries by inci- clear that quasi-static predictions depart from their elec- dent 10 µm wavelength light (ω = 1000 cm−1). We con- trodynamic counterparts near a probe length L λ/10, sider here the hyperboloidal geometry to faithfully reflect precisely where quasi-static approximations might∼ be ex- the cone-like structure of conventional near-field probes pected to falter lacking the antenna enhancement mecha- which exhibit a taper angle θ 20◦ relative to their nism. Field retardation halts subsequent increases in en- axis in our experiments. A similar≈ hyperboloid probe ge- hancement from the quasi-static lightning rod effect, con- ometry was applied previously by Behr and Raschke to ferring a practical limit to realistically attainable near- explore plasmonic field enhancements39. However, their field enhancements outside the plasmonic regime. fully analytic treatment necessitates a semi-infinite probe Similarly, the onset of antenna modes is expected geometry treated in the quasi-static approximation, re- to modulate the intensity of frequency-dependent back- quiring an unconventional field normalization method to scattered radiation from wavelength-scale near-field obtain finite values for the probe response. Their for- probes, opening the possibility to optimally enhance ab- malism also left back-scattering from the ANSOM probe solute near-field signals through application-driven de- unexplored. For our examination, we explore the explicit sign of novel probe geometries. However, the need for a electrodynamics of probes with lengths between L = 60 broadband and normalizable probe response is equally nm (rendering a sphere in the ellipsoidal case) and 30 µm, crucial for spectroscopy applications51. A typical in- with the apex curvature radius held constant at a = 30 frared near-field spectroscopy experiment involves nor- nm. malizing acquired signals to a reference material that The axisymmetry favored by the lightning rod model exhibits a nominally flat optical response (e.g. gold was maintained throughout these calculations by ap- or undoped silicon) in order to remove the influence proximating plane wave illumination by an inwards- of instrumental sensitivities9,20,38, including the probe’s propagating cylindrical field bearing a local phase veloc- frequency-dependent antenna response. Normalizability ity angled towards the tip apex at 60◦ from the probe of this response is typically assumed, but we predict here axis (see Appendix D). Validity of this axisymmetric ap- for the first time the breakdown of this assumption in the proximation was confirmed through comparison of resul- vicinity of strong antenna resonances. tant surface charge density profiles with those predicted Fig. 3b (upper panel) displays the result of fully elec- by full finite-element simulations (Comsol Multiphysics), trodynamic lightning rod model predictions for the near- ◦ consisting of a realistic metallic probe geometry (θ = 20 field signal S3 SO obtained at the peak probe-sample reso- and L = 19 µm) including AFM cantilever, subject to nance frequency (ω 1130 cm−1) induced by the strong ≈ plane-wave illumination. Differences in charge density SiO2 surface optical phonon, normalized to the signal were found to be negligible within microns of the tip from silicon. Whereas an increase in absolute back- apex, suggesting the robustness of key near-field param- scattered signal is expected near the onset of a (radiative) eters to fine details of the extended probe geometry. Fig. antenna mode, this evidently accompanies a remarkable 3a displays finite-element predictions for the magnitude decrease in relative material contrast. The effect re- of the probe’s scattered field E~sca illustrating the charac- sults from strong cross-talk between the implicit probe teristically standing wave-like pattern of charge density response coincident with that of a resonant sample. along the probe’s conical surface, a consequence of field The explanation becomes clear when considering that retardation. an antenna’s resonance can be strongly detuned by its The resultant field enhancement at the probe apex dielectric environment.52,53 The point dipole model (Eq. in the absence of a sample calculated by our fully re- (1)) admits interpretation as a dipole interacting with its tarded boundary element method is shown in Fig. 3b mirror image projected from the sample surface. Extend- (lower panel) in comparison with the quasi-static case, ing this interpretation to an antenna, the electrodynamic demonstrating several key phenomena: First, quasi- system consists then not of a single antenna, but of a cou- static probe geometries exhibit field enhancements that pled antenna-mirror pair, and it is well established that increase monotonically with the geometric “sharpness” coupling an antenna with an exact mirror copy induces L/a due to the electrostatic lightning rod effect, origi- a resonance red-shift54,55. Whereas the mirror coupling nating the divergent quasi-static near-field contrast dis- scales with the inverse dimer gap size in the case of physi- played in Fig. 2. Second, at lengths L = m λ/2 cal antenna pairs, this coupling scales with rp in our case, for odd integers m 1, the electrodynamic ellipsoidal and could be appropriately considered a case of dielectric probe exhibits resonant≥ enhancement, whereas minima loading.44,52 are observed for m even. These features signify antenna Consequently, an SiO2 film is expected to detune an- modes with antisymmetric and symmetric surface charge tenna resonances more strongly at ωSO than a Si sub- densities50, respectively, such as those experimentally strate, rendering their respective probe back-scattering characterized among similarly elongated near-field probe signals potentially incomparable even when collected at 8

a) 22 nm! b) SiO2! Esca Einc

2 nm! SiO2 Silicon Si ! SiO2! t

min.! S3 max.!

FIG. 4. a) Near-field response of SiO2 thin films etched to varying thicknesses on a silicon substrate measured by tunable QCL spectroscopy and normalized to silicon38 (see text). The faint curves are provided as guides to the eye. Inset: Sample −1 near-field signal S3 at ω = 1130 cm overlaid on simultaneously acquired AFM topography. b) Near-field S3 spectra predicted by the lightning rod model using optical constants from literature42. Data points from the 300 nm film are superimposed for point of comparison. Our model captures the key features of the data; we infer that discrepancies with ultra-thin film data result from substantial variations in optical properties. the same frequency, since there is no clear way to normal- in this work are electrodynamic and calculated in this ize out the effect. Stated another way, interaction with fashion unless otherwise indicated. a resonant sample can not only enhance the strength Nevertheless, this examination tells a cautionary tale of a probe’s antenna mode, it can modify the antenna concerning the use of strongly resonant probes47 for behavior outright, driving the probe towards a regime quantitative near-field spectroscopy, wherein convolution of destructive radiative interference. Normalized S3 sig- of the probe’s antenna response may not be easily re- nals calculated for the electrodynamic ellipsoid (Fig. 3a, moved. However, the resonant enhancement of back- solid blue line) therefore resemble a quotient of two res- scattered fields by L λ/2 probes can provide a for- onance functions, oscillatory but shifted versus the light tunate trade-off, with∼ encouraging applications to reso- frequency relative to one another. For this extreme nantly enhanced THz near-field imaging experiments. case, we might conclude that fluctuations observed in the frequency-dependent near-field signal radiated from the probe could associate more with variable dielectric load- V. MOMENTUM-DEPENDENT ing of the antenna response than with genuine near-field LIGHT-MATTER COUPLING contrast. Antenna detuning is considerably moderated in the To test the lightning rod model description of systems case of the hyperboloidal probe, whose normalized near- exhibiting explicit momentum-dependent light-matter field response at frequencies λ < L exhibits much weaker coupling, we consider a thin film of phonon-resonant SiO2 dependence on the probe length (or, complementarily, on on silicon substrate. The film thickness t introduces a probing frequency). The normalization procedure there- characteristic length scale to the sample geometry, asso- fore appears sufficient for systematic removal of the probe ciated with a characteristic crossover momentum q t−1. sensitivity at the 20% level in the absence of strong an- Incident evanescent fields exceeding this momentum∼ are tenna resonances. Furthermore, given the clear asymp- reflected much as though bulk SiO2 were present, whereas totic character of near-field contrast for the broadband lower momentum fields can penetrate the film and reflect hyperboloidal probe, it would appear acceptable to quan- from the substrate38. With the lightning rod model we titatively model normalized near-field signals from such consider this momentum dependence exactly and directly a probe geometry using electrodynamic charge distri- compare its predictions to near-field spectroscopy mea- butions Λ(q, z) computed only for a single characteris- surements performed using the experimental apparatus tic frequency. In the case of weak antenna resonances, described in Appendix A. this renders implementation of the fully retarded light- Mid-infrared near-field images of SiO2 thin films of ning rod model no more complex than the quasi-static varying thickness were acquired with a tunable QCL at a version. Therefore, all following calculations presented probe tapping amplitude of 50 nm, taking signal from the 9

a) 1300 3 as good as earlier quasi-static predictions with an ad- t = 100 nm hoc probe geometry38. In contrast to the prediction of !LO" a blue-shifting phonon resonance with decreasing film thickness (originating entirely in the Fresnel formula 1200

] Eq. 21), experimental data indicate a slight red-shift 1 | −

r among ultra-thin films. This discrepancy should not be

!SO" p

| counted against our model: although identical optical cm [ 1100 constants were employed for predictions at all film thick- ω ! -1" TO" nesses, a growing body of experimental and ab initio q = t evidence suggests legitimate phonon confinement effects can modify the intrinsic optical properties of nanostruc- 1000 tured samples56. Clear discernment of these effects by .25 b) 1.0 ≤ near-field spectroscopy opens the possibility for quanti- q-vector tatively evaluating the optical properties of nanostruc-

) tures that exhibit and utilize bona fide three-dimensional q 57 ( 0.5 confinement . E δ A clear physical description of the depth sensitivity exhibited in Fig. 4 proves just as valuable as quanti- 0.0 2 1 0 1 tative agreement. The onset of a dramatic decrease in 10− 10− 10 10 q [a 1] near-field signal at the phonon resonance near t a can − be understood on the basis of the momentum decompo-∼ sition of electric fields emitted by the near-field probe. FIG. 5. a) Example of strong surface optical phonon dis- A straightforward analysis building on Eq. (12) reveals persion for a 100 nm thick SiO2 film on silicon computed the following decomposition for probe-generated electric by the Fresnel reflection coefficient rp(q, ω) (Eq. (21)). b) fields by their momenta in the plane of the sample (the The momentum-dependent distribution of electric fields at basis given by Eq. (5)): the sample surface (dashed line) calculated by the lightning rod model at the tip-sample phonon resonance for a conical tip in full contact. " # Λ~ δE(q ) /E = Γ ΛG 0 δq (22) i inc t→s I ΛG underlying silicon substrate for normalization (Fig. 4a). − i 38 h i These data were first presented in an earlier work . Con- −qid with Γt→s e δij, trolled film thicknesses were produced through selective ij ≡ etching (NT-MDT Co.) and confirmed by AFM height measurements acquired simultaneously with the collec- tion of near-field images. Spectroscopy was obtained from area-averaged near-field contrast levels. where d is the tip-sample distance and δE(q)/δq is un- Momentum-dependent Fresnel reflection coefficients derstood in the sense of a distribution function. 41 were used to describe these systems and provided to Fig. 5b displays δE(q) calculated on resonance with the lightning rod model in order to predict spectroscopic the SiO phonon in comparison with the example dis- near-field contrast: 2 persion of a 100 nm SiO2 film on silicon shown in Fig. 2ikz,1 t 5a. The surface optical phonon is evident at ω , ρ1 + ρ2 e SO rp(q, ω) = 2ik t (21) characteristically centered in the Restrahlen band be- 1 + ρ1ρ2 e z,1 tween the transverse optical (ωTO) and longitudinal op- i kz,i−1 i−1 kz,i with ρi − , tical (ωLO) phonon frequencies. Given that our SiO2 ≡ i kz,i−1 + i−1 kz,i forms an amorphous layer, indication of these phonon fre- p 2 2 and kz,i i (ω/c) q . quencies is approximate. Nanoscale thickness introduces ≡ − considerable momentum dependence in the regime rele- Here numeric subscripts 0, 1, 2 correspond with air, vant to probe-sample near-field interactions (q a−1), ∼ SiO2, and silicon, respectively, i denotes the complex effecting a strong phonon response only for momenta frequency-dependent dielectric function of the relevant q > t−1 as mentioned earlier. The spectroscopic char- material (ellipsometric optical constants for thermal ox- acter of the probe-sample near-field response can there- ide taken from literature42), and t denotes the oxide film fore be inferred from the momentum-space integral of thickness. δE(q) rp(q, ω). Note however the explicit rp- and d- Lightning rod model predictions are presented in Fig. dependence× of δE(q) by way of G in Eq. (22) amounts 4b for comparison with measured data. Agreement is su- to a near-field response strongly super-linear in the sam- perior to that of the simple dipole model and at least ple’s intrinsic surface response. 10 a) b)

Esca Einc

SiC Au Au SiC

890"cm(1"

FIG. 6. a) Amplitude S3 and phase φ3 of the back-scattered near-field signal from a 6H SiC crystal, measured in the vicinity of the surface optical phonon and referenced to a surface-deposited gold film, as obtained in a single acquisition by nanoFTIR. Left inset: Visible light image (width 60 µm) above the near-field probe at the SiC/gold interface. Right inset: PSHet near-field S3 image (width 1 µm) of the interface with sample and reference nanoFTIR locations indicated. b) Lightning rod model predictions for near-field signal from SiC using optical constants measured by in-house ellipsometry. While insensitive to details of probe geometry (see text), fully retarded (Ret.) calculations provide superior agreement to the experimental spectra than the quasi-static (QS) approximation.

VI. THE STRONGLY RESONANT LIMIT: sured across the interface by pseudoheterodyne (PSHet) 11 −1 SILICON CARBIDE imaging with a CO2 laser tuned to 890 cm , with nanoFTIR acquisition positions indicated. As confirmed We can critically evaluate the generality of the light- by nanoFTIR, near-field resonance with the SiC sur- face optical phonon produces a stronger signal than gold ning rod model formalism through comparison with mea- −1 surements of crystalline SiC, a strongly resonant mate- across a considerable energy range, 800-940 cm . Such rial in the mid-infrared owing to an exceedingly strong strong near-field resonances enable potential technolog- surface optical phonon at ω 950 cm−1. Here we find ical applications for guiding and switching of confined ≈ infrared light within nanostructured polar crystals, as the limit at which contingent assumptions for alternative 60 near-field models9,26,32,34,35 are expected to break down, suggested in related work . since resonant materials can interact non-perturbatively Predicted spectra presented in Fig. 6b reveal that ex- along the entire length of the near-field probe. This plicit consideration of field retardation effects according breakdown signals the importance of both probe geom- to the findings of Sec. IV (spectra labeled Ret.) sig- etry and field retardation effects. Lacking these con- nificantly improves quantitative agreement with exper- siderations, previous models have dramatically overesti- imental spectra in contrast to the quasi-static predic- 17,58 mated the near-field contrast generated by SiC , leav- tion (labeled QS), which drastically overestimates the ing the estimation of optical properties through quanti- near-field contrast of SiC up to a factor of 20 over gold. tative analysis of near-field observables quite ambiguous. The QS curve additionally reflects an excessive red-shift Fig. 6 displays quantitative agreement between newly of the probe-sample resonance peak on account of the presented nanoFTIR spectroscopy of a 6H SiC crystal overly predominant low-momentum phonon excitations and lightning rod model predictions. Asymmetry in the permitted in the quasi-static approximation; these re- observed phonon-induced probe-sample resonance spec- side at lower energy due to the typical positive group trum mimics that of the underlying surface response velocity of surface phonon polaritons. We furthermore function β(ω). To ensure unambiguous comparison be- explored the influence of particular probe geometries on tween experiment and theory, uniaxial optical constants the predicted near-field spectrum by employing charge of our crystal were directly determined by in-house in- distributions Λ(q, z) calculated for both the ideal hyper- frared ellipsometry and were found consistent with lit- boloidal probe geometry as well as for the actual profile erature data for similar crystals59. A 100 nm gold film of an used probe tip, obtained from an SEM micrograph was deposited onto the crystal surface to provide a nor- (displayed as the blue curve in Fig. 1a). The Fig. 6b malization material for nanoFTIR measurements, which comparison of SiC spectra predicted with these geome- were conducted at 60 nm tapping amplitude across the tries reveals that only essential features of the probe ge- SiC-gold interface. The right inset of Fig. 6a displays ometry, such as the overall conical shape and taper angle strong interfacial contrast in near-field amplitude mea- (θ 20◦) shared by both, are relevant for predicting ≈ 11 near-field contrasts at the 10% level of accuracy. Fur- ther quantitative refinements to near-field spectroscopy Esca Einc will therefore benefit from the standardization of repro- ducible probe geometries61. SiO2 300nm Si

VII. NANO-RESOLVED EXTRACTION OF OPTICAL CONSTANTS

Systematic improvements in the light sources and detection methods available for near-field spectroscopy now enable sufficiently high signal-to-noise levels and fast acquisition times for routine, reproducible measurements7,8. Fig. 7 displays newly presented nanoFTIR measurements acquired on a 300 nm SiO2 film with silicon used for normalization, displaying both the amplitude S and phase φ of the probe’s back-scattered FIG. 7. Amplitude S and phase φ of the back-scattered near- radiation demodulated at the 2nd and 3rd harmonics of field signal from a 300 nm SiO2 film, measured in the vicinity the probe frequency, collected at 60 nm tapping ampli- of the surface optical phonon and referenced to the silicon substrate, as obtained in a single acquisition by nanoFTIR. tude. Such broadband data are ideally eligible for the quantitative extraction of SiO2 optical constants in the vicinity of the transverse optical phonon (ωTO 1075 harmonic oscillator equilibrating to s = NF(): cm−1). ≈ Using the lightning rod model, a method requiring min- d2 d δs + 2ζ Ω δs + Ω2 δs = 0 (24) imal computational effort was developed to solve the in- dω2 dω verse problem of near-field spectroscopy, proceeding as with δs(ω) s(ω) NF(ω). follows: The connection between optical properties of a ≡ − sample material (e.g. the complex dielectric function, Here ζ denotes a damping constant tuned to induce crit-  = 1 +i2) and near-field observables (e.g. S and φ, or ical damping (ζ=1), and Ω is a force constant ensuring equivalently the real and imaginary parts of the complex decay to equilibrium over an interval δω = 2π/Ω compa- back-scattered signal s = s1 + is2 at a given harmonic rable to the frequency resolution of measurement. This n 2) is described by a smooth map NF : C C, ≥ → equation of motion enables adiabatic tracking of exper- with C the set of complex numbers. A “trajectory” s(ω) imentally observed signal values while both penalizing through the space of observable near-field signals there- deviations δs and dissipating their energy. Eq. (24) fore corresponds to a trajectory (ω) through the space may alternatively be parametrized by an auxiliary in- of possible optical constants. The uniqueness of this cor- dependent variable x for which ω(x) increments only respondence was confirmed for bulk and layered sample when δs(ω) < δsthresh, a threshold value ensuring sys- geometries by computing s = NF() across the parameter tem equilibration| | arbitrarily close to the measured signal range of interest for real materials (2 > 0) and ensuring value at each ω. This also ensures solutions to Eq. (24) local invertibility of the map, conditional on the deter- are relatively insensitive to the “guessed” initial condi- minant of the Jacobian matrix of NF: tion (ω0), amounting to a robust relaxation method. Our inversion of measured s(ω) consists of numerically ∂(s1, s2) 62 J(1, 2) > 0 with J(1, 2) = . (23) solving Eq. (24) for  by finite difference techniques . | | ∂(1, 2) This requires at least five evaluations of NF per ω- or x-step in order to estimate local first and second deriva- Because parameters internal to the operation of NF tives of NF with respect to real and imaginary parts of are frequently variable (e.g. sample thickness, tip radius . Although consequently the procedure is more com- of the probe, tapping amplitude), instead of establish- putationally costly than forward evaluation by the light- ing the inverse map NF−1 as a “look-up table” by brute ning rod model, it is at least as efficient in principle as computation, we instead introduce a method for nucle- global nonlinear least-squares methods (e.g. Levenberg- ated growth of the trajectory (ω) which optimizes con- Marquardt63) and often considerably faster, furthermore sistency with the forward mapping s = NF() beginning requiring no a priori knowledge for the form of the fit- at some initial frequency ω0. We re-imagine the problem ting function. This is considerably advantageous in cases as a particle navigating -space under the influence of ex- where spectra are not available in a sufficiently wide fre- ternal forces penalizing displacements δs from measured quency range to permit well-determined fitting to (ω) signal values s(ω). The trajectory (ω) for such a particle by Kramers-Kronig-consistent oscillators64. solves, for example, the equation of motion for a damped We applied our inversion technique to the spectro- 12

a) b) a) b)

FIG. 8. a) Typical variation in optical constants among thermal oxide thin films taken from literature ellipsometry. Pairs of red and blue curves with identical line style are associated with distinct thin film samples.42 b) Optical constants of a 300 nm SiO2 film extracted from near-field spectra S2(ω) and S3(ω) following the method of Eq. (24). scopic data displayed in Fig. 7 by parametrizing NF the existing technique relies on a polynomial expansion with the reflection coefficient of an “unknown” 300 nm in β strictly limited to weakly resonant samples, vis. the layer (film thickness determined by AFM) on silicon perturbative limit of Eq. (13), and employs a model substrate. For mapping the film’s optical constants with tunable ad hoc parameters30,32. Our procedure re- film(ω) = 1(ω) + i 2(ω) to a measurable, normalized moves both shortcomings. These advantages make Eq. near-field spectrum sn(ω), the form for NF used here is (24) a suitable technique for the unconditional on-line that given by the lightning rod model, namely: analysis of near-field spectroscopy data in a diagnos- tic setting. Combining for the first time the powerful  film Si NF 1(ω), 2(ω) = sn (ω)/sn (25) nanoFTIR instrumentation with a quantitative inver- sion methodology unlimited by sample characteristics, with this procedure makes possible potent new applications π/Ω of nano-spectroscopy to the quantitative optical study Z 17,18 s film(ω) = dt sin (nΩt) E film(d, ω), (26) of phase-separated materials and nano-engineered n rad 21,22 −π/Ω devices , as well for the nano-resolved chemical iden- ~ tification of structures in biological or surface film Λ0 7,9,16,32 E rad (d, ω) = ~erad G film(d, ω) , (27) applications . · I ΛG film(d, ω) − and d = A (1 + sin (Ωt)). Describing the near-field re- VIII. CONCLUSIONS AND OUTLOOK sponse of the film, G film(d, ω) is given by Eq. (11) in film terms of the film reflection coefficient rp (q, ω), which is in turn a function of film(ω) via Eq. (21). The sil- The lightning rod model provides a general quantitative Si icon normalization signal sn is computed analogously, formalism for predicting and interpreting the experimen- but using the reflection coefficient for a bulk surface with tal observables of near-field spectroscopy. Simplified de- frequency-independent dielectric constant Si 11.7. All scriptions of the probe-sample near-field interaction such other parameters are defined as detailed in Sec.≈ II. as the point dipole model can be obtained as special cases In Fig. 8 we present the favorable comparison of resulting from convenient though unnecessary physical our extracted film(ω) with typical literature optical con- assumptions. In particular, the choice of effective probe stants for three thermal oxide films measured by conven- length L30,32 was shown to be ad hoc in the quasi-static tional infrared ellipsometry42. Fig. 8a makes clear the approximation, and consequently susceptible to dubious typical variation in optical constants expected among ox- a posteriori fitting to experimental data. ide films grown even under nominally fixed conditions. We find a fully electrodynamic treatment renders the Furthermore, our extraction technique produced virtu- effective length construct unnecessary, since field retarda- ally identical output when conducted on both 2nd and tion effects modify the distribution of probe charge inter- 3rd harmonic near-field spectra (s2(ω) and s3(ω)), at- acting with the sample. While this provides a resolution testing to the internal consistency of the lightning rod to problems of convergence inherent to the quasi-static model. treatment, sample-induced dielectric loading of strong Although near-field inversion has been very recently antenna resonances (e.g. for the long ellipsoidal probe) demonstrated on measurements of prepared polymers, was found to deceivingly modulate relative material con- 13 trasts predicted in the vicinity of sample resonances, ACKNOWLEDGMENTS such as the surface optical phonon of SiO2, an impor- tant caveat and consideration for the rational design of 47 This work was supported through NASA grants optimized spectroscopic probes . Nevertheless, fine de- NNX08AI15G and NNX11AF24G. Alexander S. McLeod tails of the probe geometry for realistic conical probe acknowledges support from a U. S. Department of Energy geometries are predicted to impact observable near-field Office of Science Graduate Fellowship. material contrasts at or below a 10% level of variation. Using the fully retarded lightning rod model with a real- istic probe geometry, we obtain quantitatively predictive Appendix A: Experimental methods agreement compared both with tunable QCL near-field spectroscopy of SiO2 films with varying thickness and with newly presented nanoFTIR spectroscopy measure- In the following sections, we apply the lightning rod ments of the strongly resonant polar material SiC. This model in comparison with near-field spectra measured exhibits our model’s proper momentum-space description for SiO2 thin films and SiC, acquired with the follow- of the probe-sample optical interaction, as well as its suit- ing experimental apparatus. Infrared nano-imaging and ability for the truly quantitative description of strongly nano-spectroscopy measurements were performed with a resonant near-field interactions, in contrast with the ca- NeaSNOM scanning near-field optical microscope (Nea- pabilities or implementations of the alternative models spec GmBH ) by scanning a platinum silicide AFM probe heretofore demonstrated.34,35,58 (PtSi-NCH, NanoAndMore USA; cantilever resonance frequency 300 kHz, nominal radius of curvature 20-30 Finally, we present a deterministic method to invert nanometers) in tapping mode over the sample while il- the lightning rod model without recourse to ad hoc pa- luminating with a focused infrared laser beam, result- rameters or over-simplifications. This rather general ing in back-scattered radiation modulated at the probe technique flexibly solves the inverse problem of near- tapping frequency Ω and harmonics thereof. In our field spectroscopy at a computational cost significantly pseudo-heterodyne detection setup, this back-scattered lower than exhaustive lookup-tables or oscillator fit- radiation interferes at a mercury-cadmium-telluride de- ting methods, offering exciting opportunities for the on- tector (Kolmar Inc.) with a reference beam line interpretation of nano-resolved near-field spectra ac- whose phase is continuously modulated by reflection from quired in a diagnostic setting. We envision the in- a mirror piezoelectrically oscillated at a frequency δΩ verse lightning rod model employed quantitatively for ( 300 Hz). Demodulation of the detector signal at deeply sub-wavelength optical studies of naturally or frequency≈ side-bands nΩ m δΩ for integral m supplies artificially heterogeneous and phase-separated materi- ± the background-free amplitude Sn and phase φn of the als, promising further novel applications to systems like infrared signal at harmonics n of the probe’s tapping 65 energy storage nanostructures, transition oxide frequency.4,10,11 heterostructures,66 and single- or multi-layered graphene 20,21 The super-linear dependence of near-field interactions plasmonic devices . versus the tip-sample separation distance implies that, in There remain outstanding challenges for the present the case of harmonic tapping motion, signal harmonics at model, including its extension to cases where deviations n 2 are directly attributable to near-field polarization ≥ 48 from axisymmetry are crucial, as for s-polarization of of the tip . Contrasts in near-field signal intensity and incident light, or for probe geometries with strong ro- phase at these harmonics therefore correspond to vari- tational asymmetry. We envision an expansion of our ations in local optical properties of the sample26. Tun- boundary element methods and of the lightning rod able fixed-frequency CW quantum cascade lasers (QCLs, model into basis components with differing rotational Daylight Solutions Inc.) and a tunable CO2 laser (Ac- “quantum numbers”67 to capture the features of irrota- cess Laser Co.) were used for imaging and spectroscopy tional geometries in a computationally inexpensive fash- of SiO2 films and SiC, respectively. ion. Furthermore, the explicit application of our model NanoFTIR spectroscopy78 was enabled by illumination to dielectric probes, particularly in the plasmonic regime, from a broadband mid-infrared laser producing tunable is an undertaking of great potential interest for which the radiance across the frequency range 700-2400 cm−1. This extension of our electrodynamic boundary element (Ap- coherent mid-infrared illumination is generated through pendix D) to materials of non-negligible skin depth might the nonlinear difference-frequency combination of beams play a crucial role. However, even at its present stage from two near-infrared erbium-doped fiber lasers – one the quantitative scattering formalism presented here also at 5400 cm−1 and the other a tunable supercontinuum lays a solid foundation for the rational analysis and opti- near-infrared laser (TOPTICA Photonics Inc.) – result- mization of tip-enhanced optical phenomena in an ever- ing in 100 fs pulses at a repetition rate of 40 MHz. An growing number of exciting experimental applications, asymmetric Michelson interferometer with 1.5 millimeter including single-molecule Raman spectroscopy15,75 and travel range translating mirror enables collection of de- other novel partnerings of optics with scanning tunnel- modulated near-field amplitude Sn(ω) and phase φn(ω) ing microscopy.76,77 spectra with 3 cm−1 resolution. 14

Appendix B: Resolution of the field from a charged The scalar potential V (~r) finds contributions from both ring into evanescent waves incident (external) fields E~inc, originating as from dis- tant free charges, as well as from bound charges at the The xy-plane Fourier decomposition of the Coulomb dielectric boundary. Taking the bound charge ρb as the field of a point charge Q located at the origin is well product of a surface density σQ (distinguished from elec- known68: trical conductivity σ) with the surface Dirac delta func- ZZ ∞   tion, the latter contribution comprises a surface integral ~ Q kxxˆ + kyyˆ E(~r) = dkx dky i +z ˆ on the boundary S: −2π −∞ q i(kxx+ky y)+qz Z e 0 σQ(~r ) Z ∞ × Vb(~r) = dS 0 (C3) qz S ~r ~r = Q dq q (J0(qρ)ˆz + J1(qρ)ˆρ) e (B1) | − 0| − 0 Evaluating the discontinuous surface normal electric field q 2 2 for z < 0 and with q k + k . This decomposition nˆ12 Vb from this contribution involves: ≡ x y − · ∇ can be applied similarly to a ring of charge with radius lim nˆ12 (1/ ~r + t nˆ12 ~r ) = 2πδ(~r ~r ) F (~r, ~r ) , centered in a plane through the origin with z-axis t→0+ − · ∇ | − 0| − 0 − 0 R normal: nˆ (~r ~r ) with F (~r, ~r ) 12 . (C4) Q Z 2π Z ∞ Z 2π · −3 0 E~ (~r) = dφ0 dq dφ ~ (q, ~r, φ0) 0 ≡ − ~r ~r R 2 R | − 0| 4π 0 0 0 E   Gauss’s law (Eq. C1) therefore yields an integral equa- ~ kxxˆ + kyyˆ iq(ρ cos φ−R cos (φ−φ0))+qz R zˆ + i e . tion in the surface bound charge density σ (~r): E ≡ − q Q 0  Here φ is an angular integration variable about the cir- 12 4π σQ(~r) = 2 nˆ12 E~inc(~r) + 2πσQ(~r) cumference of the ring. We obtain 1 2 · Z ∞ − Z  qz 0 E~R(~r) = Q dq q J0(q )(J0(qρ)ˆz + J1(qρ)ˆρ) e . dS F (~r, ~r ) σQ(~r ) , (C5) − 0 R − S 0 0 (B2) The total field is thus a sum of axisymmetric p- which upon consolidation yields: polarized evanescent waves weighted by the geometry- Z 1 + 2 0 induced prefactor q J0(q ). Eq. (B2) constitutes the 2π σQ(~r) =n ˆ12 E~inc dS F (~r, ~r ) σQ(~r ). central result of this section.R 1 2 · − S 0 0 − (C6) Without loss of generality, this equation can be utilized Appendix C: Quasi-static boundary element method to pre-compute the quasi-electrostatic response of an ax- for an axisymmetric dielectric and conductor isymmetric body of dielectric constant 2 to incident fields, taking 1 = 1 as air, parametrizing the integral ker- nel F by axial and surface radial coordinates z and z, A tractable electrostatic boundary element method ap- R plicable to systems of homogeneous dielectrics can be de- respectively, and expressingn ˆ12 through axial derivatives veloped as follows67. Gauss’s law constrains the density of the latter. However, to unambiguously present our method of so- of bound charge ρb at the boundaries between dielectric media as: lution to equations like Eq. C6 and to promote its ap- plication for the description of metallic near-field probes, D~ we confine our attention specifically to the ideally con- E~ = = 4πρb ∇ · ∇ ·  ducting limit, wherein 2 is divergent. For Eq. C1 to   1 1 ~ hold with finite normal displacement in Eq. C2 there- ∴ 4πρb = δS nˆ12 D (C1) 2 − 1 · fore requires a vanishing normal gradient of the total po- tential V just inside the probe surface. Lacking free or Eq. C1 follows in the case that free charge is absent at bound charges within its volume, the probe interior and dielectric boundaries such that D~ = 0, and δS is a sur- ∇· surface therefore reside at constant total potential, sig- face Dirac delta function associated with the boundary nifying zero internal field and perfect screening by the between media of dielectric constant 1 and 2, withn ˆ12 surface: the unit vector perpendicular to this boundary oriented from medium 1 to medium 2. Continuity of the surface Vinc(~r) + Vb(~r) = V0 on S. (C7) normal displacement field across the dielectric interface permits its evaluation at positions ~r along the boundary This criterion follows equivalently from Eq. (C6) in the as a limit taken infinitesimally inside medium 2: limit 2 1 through reverse application of Eq. (C4). The incident potential of an axisymmetric evanescent nˆ12 D~ (~r) = 2 lim nˆ12 V (~r + t nˆ12) . (C2) · t→0+ − · ∇ field is given in cylindrical coordinates ρ, φ, and z by 15

−qz Vinc(~r) = J0(qρ)/q e . The potential Vb is gener- Appendix D: Electrodynamic boundary element ated by the surface charge density σQ(~r), which may method for an axisymmetric conductor be divided into a continuum of rings, each with charge dQ = λQ(z) dz and radius z: As in the quasi-static case, the charge distribution in- R Z duced on a nearly perfectly conducting object by an in- σQ(~r ) V (~r) = dS0 0 cident electrodynamic field oscillating at frequency ω re- b ~r ~r S sides exclusively at the object’s surface. To compute this Z ∞ | 0 − | 0 0 0 distribution, we begin with detailed force balance at the = dz Φ(~r, z ) λQ(z ) (C8) 0 boundary S along directions tangential to the surface. Z 2π dφ0 1 Assuming axisymmetry, we need only consider without Φ ˆ p 0 2 2 2 0 loss of generality the surface tangential directions ξ or- ≡ 0 2π (z z) + ρ + 0 2ρ z0 cos φ − Rz − R thogonal to φˆ that possess positivez ˆ-component: 2K( 4 ρ Rz0 ) (ρ−R 0 )2+(z−z0)2 = − z . (C9) ˆ  ~ ~  ~ p 2 0 2 ξ Einc + Eobj = 0 on S. (D1) π (ρ z0 ) + (z z ) · − R − Here Φ constitutes the Coulomb kernel for a ring of Since E~ = V + iωA~ for scalar and vector potentials charge, and K(...) denotes the elliptic integral of the V and A~, we−∇ have first kind. Evaluating V at the boundary of the object obj Z   (ρ = z) and discretizing z in Eqs. (C7) and (C8) as by E (~r) = ∂ dV (~r) iω ξˆ dA~ (~r) on S R inc ξ ξ obj obj Gauss-Legendre quadrature, we obtain the linear system S − · Z L ~ ~ 0h 0 0 M λQ = V0 Vinc (C10) = dz ∂ξΦ(z, z ) λQ(z ) − 0 with Mij Φ(zi, zj) δzj. ≡ 0 0 i iω ξ(z, z ) I(z ) , (D2) Vectors denote evaluation at positions zi, (zi) . The − A { R } condition of overall charge neutrality fixes the value of where we have parametrized points on S by the object’s V : 0 axial coordinate 0 < z < L; meanwhile λQ(z) dQ/dz ≡ X ~ X −1   denotes the linear charge density and I(z) denotes the λQ δzi = 0 = δzi M V0 V~inc − total current passing along the object surface through a i i h i zˆ-normal plane at z. Φ and ξ denote integration kernels P −1 ~ A i δzi M Vinc for the scalar and vector potentials, respectively. i ∴ V0 = h i . (C11) The continuity equation for charge implies ∂zI(z) = P −1 ~ i δzi M I iω λ (z), and since current is forbidden to flow from i Q the hypothetically isolated object, integration by parts Here I~ denotes a vector with all entries unity. While Eq. yields: (C10) would appear to be directly solvable, such Fred- Z L holm integral equations of the first kind are notoriously 0 0 h 0 Einc ξ(z) = dz λQ(z ) ∂ξΦ(z, z ) ill-conditioned. Consequently, we adopt regularization 0 6970 methods to invert the integral operator (matrix) M, Z z0 2 0 0 i yielding smooth functions λ(z) in accord with standard ω dz ξ(z, z ) (D3) quasi-static solutions for well-studied geometries like the − 0 A conducting sphere and ellipsoid. (It is worth noting that, In terms of the azimuthal angle φ and surface radial co- since Eq. (C6) presents a well-conditioned Fredholm in- ordinate at z denoted z, the scalar potential kernel is tegral equation of the second kind, no such regularization given by R of the solution is required in the case of a dielectric solid.) 0 −1 Z 2π 0 iω/c ∆(z,z ,φ) Once the inverse operator M has been computed for a 0 dφ e Φ(z, z ) = 0 (D4) given geometry, calculation of λ(z) for arbitrary Vinc(~r) 0 2π ∆(z, z , φ) is fast and trivial. q 0 2 2 2 with ∆ (z z ) + + 0 2 z z0 cos φ, For an axisymmetric system, Eqs. (C10) and (C11) ≡ − Rz Rz − R R together with this solution method are sufficient to cal- which may be computed straightforwardly for a given ob- culate the linear charge density induced on a conducting ject geometry by adaptive quadrature. The exponential body due to an incident quasi-static field, and constitute phase ensures the integrand is evaluated at retarded time. the central result of this section. In practice, the con- The vector potential kernel may be established from verged calculation of M−1 for a particular axisymmetric geometry takes no longer than a few tens of seconds on a 1 Z K~ (~r ) A~(~r) = dS0 0 eiω/c |~r−~r0| (D5) single 2.7 GHz processor. Calculation of λ(z) for a range c2 ~r ~r of q values sufficient for converged lightning rod model | − 0| ~ calculations requires only several seconds using the same 0 ˆ dA and ξ(z, z ) ξz (z), (D6) processor. A ≡ · dz0 16

ˆ with K~ (z) I(z)/2π z ξ denoting the local surface cur- As in Appendix C, discretizing z results in a linear ≡ R0 0 p 0 0 rent. Noting that dS = 2π z 1 + ∂z0 z dz and that system ˆ R R the direction of ξ is manifestly z- and φ-dependent (ex-  2  ω T ¯ ~ ~ pressed here as ξˆ ), we obtain Φ W W WλQ = V0 Vinc, (D10) zφ − c2 A − 0 2π 0 Z ˆ ˆ 0 0 dAξ p 0 I(z ) dφ ξzφ ξz φ i ... where Φ Φ(z , z ), W diag δz , ¯ ¯ (z , z ), (z) = 1 + ∂z0 · e , ij i j i ij ξ i j dz0 z c2 2π ~0 ≡ ≡ { } A ≡ A R 0 ~r r Wij δzi θ(j i), and θ(...) denotes the Heaviside unit | − | step function.≡ − The superscript T denotes matrix trans- with the exponential factor unchanged. The φ- pose. Vectors again denote evaluation at axial and radial dependence of Aξ is rendered moot on account of ax- coordinates zi, (zi) along the object surface. Self- isymmetry, and so is suppressed. { R } consistency requires a value of V0 ensuring charge neu- Finally, since the surface tangential unit vector at trality. Taking M to be the full integral operator preced- height z and azimuthal angle φ is expressed in terms of ing ~λ in the linear system above, V is again given by the radial coordinate and the radial unit vectorρ ˆ as Q 0 z φ ~ R Eq. (C11), and λQ is obtained via inversion of M. Note that the particular selections of lower integration bounds ˆ ∂z z ρˆφ +z ˆ 0 ξzφ = R , withρ ˆφ ρˆφ0 = cos(φ φ ), p 2 on Vinc in Eq. (D8) and ¯ in Eq. (D9) are naturally 1 + ∂z · − Rz rendered arbitrary when thisA condition is satisfied. As we obtain the vector potential kernel as: in the quasi-static case, once M−1 has been computed for a given geometry (less than one minute of computa- " Z 2π 0 iω/c ∆(z,z0,φ0) 0 1 dφ e tion on a 2.7 GHz processor), the calculation of λQ(z) ξ(z, z ) = 2 0 0 + ~ A c 0 2π ∆(z, z , φ ) for arbitrary Einc(~r) is both fast and trivial (several mil- liseconds). To emulate plane wave illumination from an Z 2π 0 iω/c ∆(z,z0,φ0) # dφ e 0 inclination angle θ with respect to the z-axis, in this work 0 0 ∂z z ∂z z 0 0 cos φ R R 0 2π ∆(z, z , φ ) we substitute the axisymmetric analog 1 p . (D7)  k2 q2  p 2 E~inc(~r) = J0(qρ)z ˆ + i − J1(qρ)ρ ˆ × 1 + ∂z Rz q Here ∆ is defined as in Eq. (D4), and note that the first e−i√k2−q2z (D11) × term within brackets in fact equates with the scalar po- with q k sin θ and k ω/c. tential kernel. Only the second term must be computed ≡ ≡ anew, and similarly by adaptive quadrature. This field profile equates with a rotational sum of θ- We now define a convenient quasi-potential function directed plane waves inbound from all azimuthal angles for the incident field: φ. Z For an axisymmetric system, Eqs. (D4), (D7), and p 2 ˆ ~ Vinc(z) dz 1 + ∂z z ξz Einc(z) (D10) are sufficient to calculate the linear charge density ≡ − R · induced on a conducting body due to an incident elec- Z trodynamic field, and constitute the central result of this = dz (∂z z Einc ρ + Einc z) . (D8) − R section. In practice, the converged electrodynamic cal- culation of M−1 for a particular axisymmetric geometry Proceeding with Eq. (D3), we relabel z z before ap- → takes only twice as long as for the quasi-static case. R ξz R z q 2 plying the operation dξ = dz 1 + ∂z to both 0 0 Rz sides, resulting in: Appendix E: Radiation from an axisymmetric Z L " 2 Z z Z z0 # conductor 0 0 0 ω 0 0 dz λQ(z ) Φ(z, z ) dz dz ¯ξ(z, z ) − c2 A 0 0 0 The far-field radiation profile from an arbitrary cur- = V0 Vinc(z). (D9) rent distribution can be obtained by integrating the far- − field contribution ←→G FF to the Green’s dyadic function Here we have applied Eq. (D8) and taken V0 as a con- ←→G 71 from infinitesimal current elements at positions ~r , stant of integration. Furthermore we have defined a new 0 ¯ 0 p 2 0 here for demonstration considered oriented along thez ˆ- vector potential kernel ξ(z, z ) 1 + ∂z z ξ(z, z ), which is now symmetricA in its≡ two arguments.R A Note direction, as that the first term in brackets in Eq. (D9) accounts ~ iω ~ 0 for the retarded Coulomb force among surface charges, dErad(~r) = GFF,z(~r, ~r ) jz(~r ) dV (E1) 4π 0 0 whereas the second term describes radiative forces with iω/c |~r−~r0| 2 1 e ˆ strength of order (L/λ) produced by conduction cur- with G~ FF,z(~r, ~r ) sin θ θ, rents, where λ theO free-space wavelength of light. 0 ≡ −c2 ~r ~r | − 0| 17

of the ρ-polarized contribution similarly results in: ω eiω/c δr(z) dE~ = cos θ I (z) dξ rad,ρ 4πc2 δr(z) ρ ˆ J1(ω/c z sin θ) θ. (E3) × R Current I(z) flows on the surface of an axisymmetric conductor along a surface tangent vector

∂z z ρˆ +z ˆ ξˆ = R . p 2 1 + ∂z Rz Consequently, the total radiation from the axisymmetric body of length L is given by a commensurate sum ofz ˆ- andρ ˆ-polarized contributions: FIG. 9. (Left) The field radiated from an axisymmetric body iω eiω/c ∆r Z L to an observation point at inclination angle θ is constructed ~ ˆ Erad(θ) = 2 dz (z, θ) I(z) θ (E4) from the contributions of currents (shown in red) through in- −4πc ∆r 0 E finitesimal surface annuli. (Right) The two independent po- −iω/c z cos θh with = e sin θ J0(ω/c z sin θ) larizations composing any annular axisymmetric current dis- E R tribution, with associated angular radiation profiles (Eq. E1) i + i ∂z z cos θ J1(ω/c z sin θ) . (E5) shown schematically in blue. R R p 2 Note the factor 1/ 1 + ∂z z has been absorbed by the integration measure dz. HereR ∆r δr(z) + z cos θ is exhibiting the familiar field profile of a radiating dipole. the distance from one apex of the≡ object (at z = 0) Here θ denotes the inclination angle of the observation to the observation point, and we have applied the ap- point ~r from the z-axis in a spherical coordinate system. proximation δr(z)−1 ∆r−1 appropriate for distances ∆r L. After applying≈ the continuity equation for The dimension of a nearly perfect conductor is by def-  inition much greater than the magnetic skin depth of the charge ∂zI(z) = iω λQ(z) (with λQ the linear charge constituent material. Consequently, volume integration density) together with the fact that the current van- reduces to an integral over surface current contributions ishes at the extrema of a hypothetically isolated body dS K~ (~r). In an axisymmetric object, these contributions (at z = 0,L), integration by parts yields: are associated with surface annuli located at axial coor- 2 iω/c ∆r Z L Z z ~ ω e 0 0 ˆ dinates z and radii z, for which dS = 2π z dξ with Erad(θ) = 2 dz λQ(z) dz (z , θ) θ. p 2 R R −4πc ∆r 0 0 E dξ 1 + ∂z dz. We first evaluate the radiated field ≡ Rz (E6) from such an annulus, considering contributions from the Provided an electrodynamically consistent charge dis- two independently allowed polarizations of axisymmetric tribution λQ(z) calculated at frequency ω, Eq. (E6) can current separately, as depicted in Fig. 9. be evaluated straightforwardly for a given object geom- We define the total current as Iα(z) = 2π z Kα(z) for etry by quadrature. In the notation of Appendix D, the R polarizations α = z, ρ. For the z-polarized contribution, complex amplitude of the θˆ-polarized radiation field be- integrating Eq. E1 through an annulus about azimuthal comes: angle φ obtains 2 iω/c ∆r ω e ~ T Erad(θ) = λ W W ~θ. (E7) −4πc2 ∆r Q E iω eiω/c δr(z) dE~rad,z = sin θ Kz(z) dξ ~ −4πc2 δr(z) Here the superscript T denotes vector transpose, and θ indicates evaluation of (z, θ) at the chosen observationE Z 2π ˆ angle θ. Together withE Eq. (E5), this expression is suf- dφ z exp (iω/c z cos φ sin θ) θ × 0 R R ficient to compute the electric field radiated from a con- iω eiω/c δr(z) ducting axisymmetric system, and constitutes the central = sin θ Iz(z) dξ result of this section. −4πc2 δr(z) Projected onto a detector sensitive to θˆ-polarized light, ˆ J0(ω/c z sin θ) θ. (E2) the radiation contributions [~e ] utilized in Sec. II are × R rad i computed by expressing each single-momentum probe re- Here δr(z) denotes the distance from the center of the sponse function (linear charge density) Λ(qi, z) in dis- ~  z-located annulus to the observation point, and we have cretized real space representation λQ Λ(qi, zj) and j ≡ applied the approximation ~r ~r −1 δr(z)−1 valid for applying Eq. (E7), taking θ 60◦ relative to the z-axis | − 0| ≈ ≈ δr(z) z. An elementary analysis accounting for ro- of the near-field probe as the typical collection angle of  R tation of the radiant polarization vector in the integrand experimental detection optics. 18

[1] L. Novotny, Progress in Optics 50, 137 (2007). tro Neto, C. N. Lau, F. Keilmann, and D. N. Basov, [2] U. C. Fischer and D. W. Pohl, Phys. Rev. Lett. 62, 458 Nature 487, 82 (2012). (1989). [22] J. Chen, M. Badioli, P. Alonso-Gonzalez, S. Thongrat- [3] Y. Inouye and S. Kawata, Opt. Lett. 19, 159 (1994). tanasiri, F. Huth, J. Osmond, M. Spasenovic, A. Cen- [4] B. Knoll and F. Keilmann, Optics 182, teno, A. Pesquera, P. Godignon, A. Zurutuza Elorza, 321 (2000). N. Camara, F. J. Garcia de Abajo, R. Hillenbrand, and [5] F. Keilmann and R. Hillenbrand, Philosophical Transac- F. H. L. Koppens, Nature 487, 77 (2012). tions of the Royal Society of London. Series A: Math- [23] A. Charnukha, A. Cvitkovic, T. Prokscha, D. Pr¨opper, ematical, Physical and Engineering Sciences 362, 787 N. Ocelic, A. Suter, Z. Salman, E. Morenzoni, J. Deisen- (2004). hofer, V. Tsurkan, A. Loidl, B. Keimer, and A. V. Boris, [6] A. J. Huber, F. Keilmann, J. Wittborn, Phys. Rev. Lett. 109, 017003 (2012). J. Aizpurua, and R. Hillenbrand, Nano [24] M. Wagner, Z. Fei, A. S. McLeod, A. S. Rodin, Letters 8, 3766 (2008), pMID: 18837565, W. Bao, E. G. Iwinski, Z. Zhao, M. Goldflam, M. Liu, http://pubs.acs.org/doi/pdf/10.1021/nl802086x. G. Dominguez, M. Thiemens, M. M. Fogler, A. H. [7] S. Amarie and F. Keilmann, Phys. Rev. B 83, 045404 Castro-Neto, C. N. Lau, S. Amarie, F. Keilmann, and (2011). D. N. Basov, In submission (2013). [8] M. Brehm, A. Schliesser, and F. Keilmann, Opt. Express [25] G. W. Ford and W. H. Weber, Physics Reports 113, 195 14, 11222 (2006). (1984). [9] F. Huth, A. Govyadinov, S. Amarie, W. Nuansing, [26] R. Hillenbrand, B. Knoll, and F. Keilmann, Journal of F. Keilmann, and R. Hillenbrand, Nano Letters 12, 3973 Microscopy 202, 77 (2001). (2012), http://pubs.acs.org/doi/pdf/10.1021/nl301159v. [27] J. Aizpurua, T. Taubner, F. J. G. de Abajo, M. Brehm, [10] L. Gomez, R. Bachelot, A. Bouhelier, G. P. Wiederrecht, and R. Hillenbrand, Opt. Express 16, 1529 (2008). S. hui Chang, S. K. Gray, F. Hua, S. Jeon, J. A. Rogers, [28] J. A. Porto, P. Johansson, S. P. Apell, and T. L´opez- M. E. Castro, S. Blaize, I. Stefanon, G. Lerondel, and R´ıos,Phys. Rev. B 67, 085409 (2003). P. Royer, J. Opt. Soc. Am. B 23, 823 (2006). [29] K. Moon, E. Jung, M. Lim, Y. Do, and H. Han, Opt. [11] N. Ocelic, A. Huber, and R. Hillenbrand, Applied Express 19, 11539 (2011). Physics Letters 89, 101124 (2006). [30] A. Cvitkovic, N. Ocelic, and R. Hillenbrand, Opt. Ex- [12] R. Hillenbrand and F. Keilmann, Phys. Rev. Lett. 85, press 15, 8550 (2007). 3029 (2000). [31] B. Hauer, A. P. Engelhardt, and T. Taubner, Opt. Ex- [13] M. Schnell, A. Garcia-Etxarri, A. J. Huber, K. B. press 20, 13173 (2012). Crozier, A. Borisov, J. Aizpurua, and R. Hillenbrand, [32] A. A. Govyadinov, I. Amenabar, F. Huth, P. S. Carney, The Journal of Physical Chemistry C 114, 7341 (2010), and R. Hillenbrand, The Journal of Physical Chemistry http://pubs.acs.org/doi/pdf/10.1021/jp909252z. Letters 4, 1526 (2013). [14] P. S. Carney, B. Deutsch, A. A. Govyadinov, [33] R. Carminati and J. J. S´aenz,Phys. Rev. Lett. 84, 5156 and R. Hillenbrand, ACS Nano 6, 8 (2012), (2000). http://pubs.acs.org/doi/pdf/10.1021/nn205008y. [34] M. Esslinger and R. Vogelge- [15] X. G. Xu, M. Rang, I. M. Craig, and M. B. Raschke, The sang, ACS Nano 6, 8173 (2012), Journal of Physical Chemistry Letters 3, 1836 (2012), http://pubs.acs.org/doi/pdf/10.1021/nn302864d. http://pubs.acs.org/doi/pdf/10.1021/jz300463d. [35] P. J. Valle, R. Carminati, and J.-J. Greffet, Ultrami- [16] S. Amarie, P. Zaslansky, Y. Kajihara, E. Griesshaber, croscopy 71, 39 (1998). W. W. Schmahl, and F. Keilmann, Beilstein Journal of [36] R. Pike and P. Sabatier, eds., Scattering and Inverse Nanotechnology 3, 312 (2012). Scattering in Pure and Applied Science, 1st ed., Vol. 1 [17] J. M. Atkin, S. Berweger, A. C. Jones, and M. B. (Academic Press, 525 B Street, Suite 1900, San Diego, Raschke, Advances in Physics 61, 745 (2012). California 92101-4495, 2002). [18] M. M. Qazilbash, M. Brehm, B.-G. Chae, P.- [37] E. Nystr¨om, Acta Mathematica 54, 185 (1930), C. Ho, G. O. Andreev, B.-J. Kim, S. J. Yun, 10.1007/BF02547521. A. V. Balatsky, M. B. Maple, F. Keilmann, H.- [38] L. M. Zhang, G. O. Andreev, Z. Fei, A. S. McLeod, T. Kim, and D. N. Basov, Science 318, 1750 (2007), G. Dominguez, M. Thiemens, A. H. Castro-Neto, D. N. http://www.sciencemag.org/content/318/5857/1750.full.pdf. Basov, and M. M. Fogler, Phys. Rev. B 85, 075419 [19] A. C. Jones, S. Berweger, J. Wei, D. Cobden, (2012). and M. B. Raschke, Nano Letters 10, 1574 (2010), [39] N. Behr and M. B. Raschke, J. Phys. Chem. C 112, 3766 http://pubs.acs.org/doi/pdf/10.1021/nl903765h. (2008). [20] Z. Fei, G. O. Andreev, W. Bao, L. M. Zhang, [40] A. Hartschuh, Angewandte Chemie International Edition A. S. McLeod, C. Wang, M. K. Stewart, Z. Zhao, 47, 8178 (2008). G. Dominguez, M. Thiemens, M. M. Fogler, M. J. [41] M. Kildemo, Appl. Opt. 37, 113 (1998). Tauber, A. H. Castro-Neto, C. N. Lau, F. Keilmann, [42] A. Kuˇc´ırkov´aand K. Navr´atil,Appl. Spectrosc. 48, 113 and D. N. Basov, Nano Letters 11, 4701 (2011), (1994). http://pubs.acs.org/doi/pdf/10.1021/nl202362d. [43] A. Garc´ıa-Etxarri, I. Romero, F. J. Garc´ıa de Abajo, [21] Z. Fei, A. S. Rodin, G. O. Andreev, W. Bao, R. Hillenbrand, and J. Aizpurua, Phys. Rev. B 79 A. S. McLeod, M. Wagner, L. M. Zhang, Z. Zhao, (2009). M. Thiemens, G. Dominguez, M. M. Fogler, A. H. Cas- [44] M. Schell, A. Garc´ıa-Etxarri, A. J. Huber, K. Crozier, 19

J. Aizpurua, and R. Hillenbrand, Nat. Photonics 3, 287 J. Alexander, J. Randall, and J. Lyding, Nature Com- (2009). munications 3 (2012), 10.1038/ncomms1907. [45] L. Novotny and S. J. Stranick, Annual Review of Physical [62] J. Kiusalaas, Numerical Methods in Engineering with Chemistry 57, 303 (2006), pMID: 16599813. Python, 2nd ed. (Cambridge University Press, New York, [46] M. Brehm, A. Schliesser, F. Cajko, I. Tsukerman, and NY, USA, 2010) pp. 246–262 & 216–229. F. Keilmann, Opt. Express 16, 11203 (2008). [63] D. W. Marquardt, Journal of the Society for Industrial [47] F. Huth, A. Chuvilin, M. Schnell, I. Amenabar, and Applied Mathematics 11, pp. 431 (1963). R. Krutokhvostov, S. Lopatin, and R. Hil- [64] A. B. Kuzmenko, Review of Scientific Instruments 76, lenbrand, Nano Letters 13, 1065 (2013), 083108 (2005). http://pubs.acs.org/doi/pdf/10.1021/nl304289g. [65] A. S. Arico, P. Bruce, B. Scrosati, J.-M. Tarascon, and [48] M. B. Raschke and C. Lienau, Applied Physics Letters W. van Schalkwijk, Nat. Mater. , 366 (2005). 83, 5089 (2003). [66] H. Y. Hwang, Y. Iwasa, M. Kawasaki, B. Keimer, N. Na- [49] L. Aigouy, A. Lahrech, S. G. Asillon, H. Cory, A. C. gaosa, and Y. Tokura, Nat. Mater. , 103 (2012). Boccara, and J. C. Rivoal, Opt. Lett. 24, 187 (1999). [67] F. J. Garc´ıade Abajo and J. Aizpurua, Phys. Rev. B 56, [50] P. Biagioni, J.-S. Huang, and B. Hecht, Reports on 15873 (1997). Progress in Physics 75, 024402 (2012). [68] J. L. Agudin and A. M. Platzeck, J. Opt. Soc. Am. 70, [51] W. Bao, M. Staffaroni, J. Bokor, M. B. Salmeron, 1329 (1980). E. Yablonovitch, S. Cabrini, A. Weber-Bargioni, and [69] S. Twomey, J. ACM 10, 97 (1963). P. J. Schuck, Opt. Express 21, 8166 (2013). [70] D. L. Phillips, J. ACM 9, 84 (1962). [52] Y. Alaverdyan, N. Vamivakas, J. Barnes, C. Lebouteiller, [71] L. Novotny and B. Hecht, Principles of Nano-Optics J. Hare, and M. Atat¨ure,Opt. Express 19, 18175 (2011). (Cambridge University Press, 2006). [53] N. Large, M. Abb, J. Aizpurua, and O. L. [72] L. Novotny, E. J. Sanchez, and X. S. Xie, Ultrami- Muskens, Nano Letters 10, 1741 (2010), croscopy 71, 21 (1998). http://pubs.acs.org/doi/pdf/10.1021/nl1001636. [73] T. Taubner, F. Keilmann, and R. Hillenbrand, Opt. Ex- [54] D. P. Fromm, A. Sundaramurthy, P. J. Schuck, G. Kino, press 13, 8893 (2005). and W. E. Moerner, Nano Letters 4, 957 (2004), [74] L. M. Zhang, B. Y. Jiang, and M. M. Fogler, in prepa- http://pubs.acs.org/doi/pdf/10.1021/nl049951r. ration (2013). [55] P. Alonso-Gonz´alez,P. Albella, F. Golmar, L. Arzubiaga, [75] R. Zhang, Y. Zhang, Z. C. Dong, S. Jian, C. Zhang, L. G. F. Casanova, L. E. Hueso, J. Aizpurua, and R. Hillen- Chen, L. Zhang, Y. Liao, J. Aizpurua, Y. Luo, J. L. Yang, brand, Opt. Express 21, 1270 (2013). and J. G. Hou, Nature , 82 (2013). [56] F. Giustino and A. Pasquarello, Phys. Rev. Lett. 95, [76] X. H. Qiu, G. V. Nazin, and 187402 (2005). W. Ho, Science 299, 542 (2003), [57] A. A. Balandin, E. P. Pokatilov, and D. Nika, Journal http://www.sciencemag.org/content/299/5606/542.full.pdf. of Nanoelectronics and Optoelectronics 2, 140 (2007). [77] I. V. Pechenezhskiy, X. Hong, G. D. Nguyen, J. E. P. [58] A. Huber, N. Ocelic, T. Taubner, and R. Hillen- Dahl, R. M. K. Carlson, F. Wang, and M. F. Crommie, brand, Nano Letters 6, 774 (2006), pMID: 16608282, Phys. Rev. Lett. 111, 126101 (2013). http://pubs.acs.org/doi/pdf/10.1021/nl060092b. [78] J. D. Jackson, Classical Electrodynamics Third Edition, [59] H. Mutschke, A. Andersen, D. Clement, T. Henning, and 3rd ed. (Wiley, 1998). G. Peiter, arXiv:astro-ph/9903031 (1999). [79] H. W. Wyld, Mathematical Methods for Physics (Perseus [60] R. Hillenbrand, Ultramicroscopy 100, 421 (2004). Books Publishing, Reading, Massachussetts, 1999). [61] S. Schmucker, N. Kumar, J. Abelson, S. Daly, G. Giro- lami, M. Bischof, D. Jaeger, R. Reidy, B. Gorman, J. B.