On-chip phonon-enhanced ir near-field detection of molecular vibrations

On-chip phonon-enhanced ir near-field detection of molecular vibrations

Play all audios:

Loading...

ABSTRACT Phonon polaritons – quasiparticles formed by strong coupling of infrared (IR) light with lattice vibrations in polar materials – can be utilized for surface-enhanced infrared


absorption (SEIRA) spectroscopy and even for vibrational strong coupling with nanoscale amounts of molecules. Here, we introduce and demonstrate a compact on-chip phononic SEIRA spectroscopy


platform, which is based on an h-BN/graphene/h-BN heterostructure on top of a metal split-gate creating a p-n junction in graphene. The metal split-gate concentrates the incident light and


launches hyperbolic phonon polaritons (HPhPs) in the heterostructure, which serves simultaneously as SEIRA substrate and room-temperature infrared detector. When thin organic layers are


deposited directly on top of the heterostructure, we observe a photocurrent encoding the layer’s molecular vibrational fingerprint, which is strongly enhanced compared to that observed in


standard far-field absorption spectroscopy. A detailed theoretical analysis supports our results, further predicting an additional sensitivity enhancement as the molecular layers approach


deep subwavelength scales. Future on-chip integration of infrared light sources such as quantum cascade lasers or even electrical generation of the HPhPs could lead to fully on-chip phononic


SEIRA sensors for molecular and gas sensing. SIMILAR CONTENT BEING VIEWED BY OTHERS REAL-SPACE OBSERVATION OF VIBRATIONAL STRONG COUPLING BETWEEN PROPAGATING PHONON POLARITONS AND ORGANIC


MOLECULES Article 23 November 2020 TERAHERTZ PHONON ENGINEERING WITH VAN DER WAALS HETEROSTRUCTURES Article 26 June 2024 HIGHLY-EFFICIENT ELECTRICALLY-DRIVEN LOCALIZED SURFACE PLASMON SOURCE


ENABLED BY RESONANT INELASTIC ELECTRON TUNNELING Article Open access 25 May 2021 INTRODUCTION Mid-infrared (mid-IR) spectroscopy is a versatile analytical tool for label-free and


non-destructive identification of materials via their infrared-vibrational fingerprint1. However, the small infrared extinction cross-sections challenge the detection of minute amounts or


concentrations of molecules. To overcome this limitation, surface-enhanced infrared absorption (SEIRA) spectroscopy has been developed2,3,4. In SEIRA spectroscopy, the sensitivity is


improved by placing molecules on substrates designed to enhance the incident infrared field, which is often based on rough metal surfaces5,6,7 or metallic nanostructures8,9,10,11,12


exhibiting surface plasmon polariton resonances. Further improvement can be achieved by exploiting the extraordinary infrared field confinement of plasmon polaritons in graphene13,14,15,16.


However, plasmonic resonances typically exhibit low-quality factors, which limits field enhancements and makes it challenging to achieve vibrational strong coupling for further boosting


sensitivity4. Higher quality factors can be achieved with dielectric resonators17,18,19, however, the rather long infrared wavelengths in dielectric materials prevent the development of deep


nanoscale resonator and waveguide structures for the utmost concentration of infrared fields. Phonon polaritons in polar materials, particularly in two-dimensional (2D)


materials20,21,22,23,24, offer promising opportunities for SEIRA spectroscopy, owing to their long lifetimes paired with ultra-small wavelengths (i.e., confinement). Specifically, hyperbolic


phonon polaritons (HPhPs) in hexagonal boron nitride (h-BN) have already proven their utility for SEIRA spectroscopy of nanometer-thin molecular layers25 and phonon-enhanced mid-IR gas


sensing26. The high field confinement and lifetimes of HPhPs allow for even achieving vibrational strong coupling of organic molecules with localized27 and propagating modes28, as


demonstrated with h-BN nanoresonators and unstructured h-BN flakes, respectively. Beyond h-BN, other polar materials such as MoO322 and V2O523 exhibit long-lived HPhPs, promising to expand


the material basis and spectral ranges of phononic SEIRA spectroscopy in the future. Phononic SEIRA spectroscopy is based on far-field extinction measurements (Fig. 1a). It thus requires the


conversion of HPhPs into photons, which is an inefficient process due to the short HPhP wavelengths compared to that of photons. Here, we circumvent this problem by introducing on-chip


detection of phononic SEIRA (Fig. 1b). Previously, on-chip detection of plasmonic SEIRA has been demonstrated exploiting plasmon polaritons in a hybrid graphene-metamaterial detector array


covering a large area of 600 × 600 μm2 29. In contrast, our phononic SEIRA implementation not only utilizes HPhPs instead of plasmon polaritons but also employs a single detector as small as


7 × 4 μm2, establishing an efficient and compact SEIRA platform. In our specific demonstration, we deposited organic molecules on a h-BN/graphene/h-BN heterostructure on top of a metal


split-gate creating a p–n junction (split-gate detector)30,31,32. The metal split gate concentrates the incident light and launches HPhPs in the heterostructure. These effects together


provide strong mid-IR field enhancement at the position of the molecules and the graphene p–n junction (Fig. 1c), enhancing both the molecular vibrational absorption and the photocurrent


created at the p–n junction31,32,33. We observe dips in PC spectra at the molecular vibrational frequencies, which, furthermore, are much stronger than the ones observed in a standard


far-field transmission experiment, in good agreement with numerical simulations. The sensitivity enhancement increases for thinner molecular layers, owing to the strong confinement of the


polariton-enhanced near fields inside and at the surface of the h-BN layers. Our simulations further predict that deep subwavelength-scale areas of thin molecular layers can be detected with


enhanced sensitivity as compared to diffraction-limited far-field mid-IR spectroscopy, thus underpinning the potential of our on-chip phononic SEIRA concept to become a compact chemical


sensor. RESULTS CONCEPT AND THEORETICAL ANALYSIS Figure 2a illustrates the graphene split-gate detector and its application for on-chip phononic SEIRA detection of molecular vibrations. The


detector itself consists of a heterostructure comprising a monolayer of exfoliated graphene that is encapsulated between two h-BN layers and placed on top of two metal gates separated by a


narrow gap of about 160 nm (see fabrication details in the “Methods” section and Supplementary Note 1). By applying voltages of opposite signs to the gates (0.3 V to the left gate and −0.28 


V to the right gate), a p–n junction is created in graphene in a small region above the gap. Fig. 2b illustrates the p–n junction by showing a simulated carrier concentration profile across


the gap (Supplementary Note 5.1). When the detector is illuminated by mid-IR light, the graphene p–n junction is heated up, and via the photo-thermoelectric effect, a photovoltage _V_PC is


generated between the source and drain contacts31,32,33 (see the electrical and optical detector characterization in Supplementary Notes 2, 4). Importantly, the metal gates concentrate the


incident light into the gap area through the lightning-rod effect. The enhanced fields at the gate edges launch HPhPs (manifested by bright rays in the simulated field distribution shown in


Fig. 2e) in the h-BN/graphene/h-BN heterostructure, which exist between the transverse and longitudinal optical phonon frequencies of h-BN, _ω_TO = 1363 cm−1 and _ω_LO = 1617 cm−1, where the


in-plane permittivity of h-BN is negative20,21. Both of these effects strongly increase the mid-IR field enhancement in the graphene p–n junction and, thus, the absorption in graphene in a


small region above the gap area, subsequently enhancing the photovoltage _V_PC between the source and drain contacts31,32,33. In Fig. 2c, we illustrate the spectral response of the detector


with a simulated spectrum of the absorption in graphene, \({\alpha }_{0}\), (black curve), which in good approximation describes the photovoltage31,32 (see the simulation details in the


“Methods” section and Supplementary Note 5). We clearly observe an absorption maximum slightly above the _ω_TO (indicated by the vertical dashed red line), which can be attributed to


localized resonant HPhP mode in the gap between the gates (further discussion see below and in Supplementary Note 6). To demonstrate that the graphene split-gate detector can be used for


on-chip detection of molecular vibrations, we simulated the absorption spectrum of graphene, \({\alpha }_{{{{\rm{CBP}}}}}\left(\omega \right)\), when a 10-nm-thick layer of


4,4′-bis(_N_-carbazolyl)-1,1′-biphenyl (CBP) molecules are placed on top of the detector (as illustrated in Fig. 2a). Fig. 2c and d show \({\alpha }_{{{{\rm{CBP}}}}}\) and the normalized


spectrum, \({\alpha }_{{{{\rm{CBP}}}}}/{\alpha }_{0}\), (red curves), respectively, where \({\alpha }_{0}\) is the reference (i.e., background) spectrum of the bare detector. This


normalization procedure is analog to the one used in standard spectroscopy experiments to eliminate the spectral characteristic of the light source, spectrometer and detector. We note that


the small dips in \({\alpha }_{{{{\rm{CBP}}}}}/{\alpha }_{0}\) around 1400 cm−1 in the spectra when HPhPs are present in the h-BN are attributed to the higher-order slab modes in the


h-BN27,34. These modes exhibit distinct redshifts due to the presence of the thin molecular layer on top of the detector thus leading to small variations in the normalized spectra.


Importantly, in both spectra, we can clearly observe dips at the molecular vibrational frequencies (marked by three vertical dashed blue lines). To elucidate the impact of HPhPs on the


sensitivity of our on-chip molecular-vibrational spectroscopy concept, we calculated the absorption spectra of graphene in the absence of HPhPs (gray and light red curves in Fig. 2c, d; see


the “Methods” section). In this case, the bright rays in the field distribution vanish (Fig. 2e). In the non-normalized spectra (represented by the gray and light red curves in Fig. 2c), we


observe a relatively flat spectral profile and a generally reduced photocurrent. Particularly, we see that the maximum slightly above _ω_TO (indicated by the vertical dashed red line) is not


present. Further, the depth of the molecular-vibrational resonance dips, \(\delta\), is reduced (light red curve in Fig. 2c). We conclude that HPhPs do not only enhance the absolute


photocurrent but also the photocurrent signatures associated with the molecular-vibrational resonances of the molecular layer on top of the detector (see the additional simulation in Suppl.


Note 6). Interestingly, we also observe that the depth of the molecular-vibrational resonance dip in the normalized spectrum (i.e., the molecular-vibrational contrast), \(\varDelta\), is


enhanced by the HPhPs (Fig. 2d). Note that, generally, the enhancement of a molecular-vibrational contrast cannot be explained by merely an increased detector sensitivity, but by an enhanced


light-molecule interaction mediated by a strong optical field enhancement at surfaces or nanostructures in proximity to the molecules. We thus explain the enhancement of \(\varDelta\) in


our specific on-chip spectroscopy concept by the interaction of the molecules on top of the detector with the strong near fields of the h-BN HPhPs. EXPERIMENTAL DEMONSTRATION For an


experimental demonstration of on-chip phononic SEIRA detection of molecular vibrations, we implemented the graphene split-gate detector (bare and covered with CBP molecules) into a


custom-made Fourier transform (FT) spectrometer, where it replaced the standard IR detector (see the “Methods” section and Supplementary Note 3). As a light source, we used a mid-IR


continuum laser covering the spectral range from 1200 to 1800 cm−1 with an average power of about 0.6 mW. We applied a voltage of _V_L = 0.15 V to the left gate and a voltage of _V_R = −0.4 


V to the right gate of the graphene split-gate detector, where the maximum spectrally integrated photocurrent was obtained (Supplementary Note 4). The frequency-resolved


photocurrent\(,\,{I}_{{{{\rm{PC}}}}}(\omega )\), was obtained by standard FT spectrometer operation and interferogram processing (see the “Methods” section). We first measured the PC


spectrum of the bare detector, \({I}_{{{{\rm{PC}}}}}^{0}\), which serves as a reference (i.e. background) spectrum to obtain the normalized PC spectra of the molecule-covered detector,


\({I}_{{{{\rm{PC}}}}}^{{{{\rm{CBP}}}}}\)/\({I}_{{{{\rm{PC}}}}}^{0}\), where the spectral response characteristics of the laser, spectrometer and bare detector are eliminated. After the


reference measurement, we evaporated (see the “Methods” section) a 10-nm-thick layer of CBP molecules directly on top of the detector (as shown in the sketch in Fig. 3a) and measured the PC


spectrum, \({I}_{{{{\rm{PC}}}}}^{{{{\rm{CBP}}}}}\). Next, we repeatedly evaporated CBP molecules on top of the detector and measured the PC spectrum, until a CBP layer thickness of 100 nm


was reached. The red, blue, and green curves in Fig. 3b show the normalized PC spectra, \({I}_{{{{\rm{PC}}}}}^{{{{\rm{CBP}}}}}/{I}_{{{{\rm{PC}}}}}^{0}\), for the 10, 40, and 100 nm-thick CBP


layers, respectively (the complete set of spectra is shown in Supplementary Fig. 13). In all spectra, we observe a pronounced peak around 1370 cm−1 (indicated by the gray circle), followed


by a photocurrent that tends to increase with increasing frequency. Further, we observe three dips in the PC spectra (marked by vertical dashed blue lines), whose depths increase with the


thickness of the CBP layer. Most importantly, the spectral dip positions correspond to the frequencies of molecular vibrations of CBP, which can be recognized as peaks in the imaginary part


of the dielectric function of CBP25 (Fig. 3d), demonstrating that molecular vibrational fingerprints of nanometer-thin molecular layers of lateral sizes around _λ_IR can be detected. The


experimental spectra are qualitatively well reproduced by calculated normalized absorption spectra, \({\alpha }_{{{{\rm{CBP}}}}}/{\alpha }_{0}\) (Fig. 3c), apart from a slightly rising


background observed only experimentally. This discrepancy may stem from assuming a 2D geometry for the detector in the simulations, that is, the detector is considered infinite in the


_y_-direction. In contrast to resonant SEIRA spectra, where typically Fano-type line shapes are observed3, the spectral dips occur at the molecular vibrational resonances (marked by vertical


dashed lines) and in the simulations exhibit nearly symmetric line shapes. We attribute this observation to the simultaneous coupling of the molecular vibrations to a variety of resonant,


non-resonant modes and dark polariton modes (the latter not observed in far-field SEIRA spectroscopy), yielding in average dip-like line shapes. Further systematic studies are certainly


needed for in-depth clarification of this interesting phenomenon, which, however, goes beyond the scope of this work. We also note that in both experiment and calculation we observe a peak


around 1370 cm−1 and a baseline, which are not related to the molecular vibrational resonances. They are caused by the slight frequency shift of the absorption spectrum of the


molecule-covered detector compared to that of the bare detector, which can be seen by comparing the red and black curves in Fig. 2c. This shift can be explained by the background


permittivity of the CBP molecules, \({\varepsilon }_{\infty }\) = 2.8 (see Supplementary Note 7 for additional simulations), acting as a dielectric load on the detector. The detector, thus,


can also be applied for pure refractive-index sensing. COMPARISON WITH FAR-FIELD INFRARED TRANSMISSION SPECTROSCOPY For evaluating the spectroscopy benefit of placing the molecular layer


directly on top of the graphene split-gate detector, we compare the PC spectra of Fig. 3b with standard far-field FTIR transmission spectra of equally thick CBP layers (Fig. 3f). To that


end, we deposited CBP molecules onto a CaF2 substrate and placed them in a distance of more than 10 cm in front of the bare graphene split-gate detector (illustrated in Fig. 3e). Note that


for this experiment we used the same graphene split-gate detector used for recording the spectra of Fig. 3b but before any molecule layer was deposited onto the detector itself (i.e. before


recording the spectra of Fig. 3b). The normalized transmission spectra \({T}_{{{{\rm{CBP}}}}}/{T}_{0}\) (where \({T}_{0}\) is the reference spectrum obtained from transmission through the


bare CaF2 substrate) show the typical absorption dips at the molecular vibrational frequencies (marked by vertical dashed lines), in excellent agreement with calculated normalized


transmission spectra (Fig. 3g) (see the “Methods” section). However, the dip depths in the transmission spectra in Fig. 3f, g are much less pronounced compared to Fig. 3b, c, particularly


for thin layers. Remarkably, for the 10 nm thin CBP layer, we observe the molecular vibrational fingerprint only when the layer is deposited directly onto the detector, clearly demonstrating


the superior sensitivity due to the polariton-enhanced near fields in the vicinity of the split gate detector. It is important to note that the same detector was used for both experiments,


ensuring that the key detector characteristics, such as sensitivity, detectivity, and signal-to-noise ratio, remained constant. As a result, it allows for a reliable comparison and


verification of the superior performance of the on-chip phononic SEIRA detection of molecular vibrations compared to a standard far-field approach employing the same detector. To quantify


the enhancement of the molecular-vibrational contrast when the molecules are placed directly on top of the detector, we measured the depth of the molecular-vibrational contrast at 1508 cm−1


in the on-chip phononic SEIRA and far-field FTIR transmission spectra (denoted \(\varDelta\) and \({\varDelta }_{{{{\rm{t}}}}}\), respectively, and illustrated in Fig. 3b, f), and plotted


them in Fig. 4a as a function of the layer thickness _d_CBP (the full data set of spectra are shown in Supplementary Note 9 in Supplementary Fig. 13). In Fig. 4b we show the ratio


\(\varDelta /{\varDelta }_{{{{\rm{t}}}}}\) (i.e. the enhancement of molecular-vibrational contrast). We find that \(\varDelta\) (red symbols in Fig. 4a, on-chip phononic SEIRA experiment) is


enhanced by a factor of two compared to \(\varDelta\)t (blue symbols in Fig. 4a, far-field experiment) at _d_CBP = 100 nm, increasing to a factor of about 5 at _d_CBP = 40 nm (Fig. 4b).


Most important, for the 10 and 20 nm-thick molecular layers we can obtain measurable values only for \(\varDelta\) but not for \(\varDelta\)t. The enhancement of \(\varDelta\) relative to


\(\varDelta\)t is qualitatively confirmed by numerical simulations (solid lines in Fig. 4a and b; for details, see the “Methods” section). For quantitative agreement, we have to divide the


simulated \(\varDelta\)-values by a factor of 1.5, which we attribute to the approximations made in the simulations, such as assuming a plane wave illumination and a 2D detector geometry


(i.e., the detector is infinite in the _y_-direction, see Supplementary Note 5). For molecular layers with _d_CBP < 40 nm, the simulations predict large enhancements of


molecular-vibrational contrast, \(\varDelta\)/\(\varDelta\)t, reaching well above one order of magnitude for _d_CBP < 10 nm (Fig. 4b), which we explain by the strong field concentration


in the vicinity of the detector surface (illustrated in the inset of Fig. 4b). SENSITIVITY TO MOLECULAR LAYERS WITH NANOSCALE LATERAL SIZE Finally, we predict on-chip phononic SEIRA


detection of deep subwavelength-scale wide molecular layers (i.e., molecular stripes). To that end, we calculated on-chip phononic SEIRA and far-field transmission spectra, \({\alpha


}_{{{{\rm{CBP}}}}}/{\alpha }_{0}\) (Fig. 5a) and \({T}_{{{{\rm{CBP}}}}}/{T}_{0}\) (Fig. 5b), respectively, for 300 nm-wide (about twice the gap width) molecular stripes of different


thicknesses placed either directly above the detector´s p–n junction (illustrated by the inset of Fig. 5a) or on a CaF2 substrate (illustrated by the inset of Fig. 5b). For a conservative


estimation of the far-field transmission spectra, we assumed a plane wave illumination on a width of 3 μm (corresponding to a diffraction-limited focus diameter). For all thicknesses, we


observe a larger enhancement of the molecular-vibrational contrast, \(\varDelta /{\varDelta }_{{{{\rm{t}}}}}\), compared to Fig. 3. For example, we find \(\varDelta /{\varDelta


}_{{{{\rm{t}}}}}\) = 80 for the 10 nm-thick stripe (compare red spectra in Fig. 5a, b), which is 5 times larger than for the 10 nm-thick layer (compare red spectra in Fig. 3c, g). We


attribute this additional contrast enhancement to the strong localization of the polaritonic field enhancement near the p–n junction. As a consequence, when the width of the molecular layer


decreases, the molecular vibrational absorption contrast, \(\varDelta\), is less reduced compared to that observed in far-field transmission spectroscopy, \({\varDelta }_{{{{\rm{t}}}}}\),


where the IR field distribution is considered to be homogeneous. These simulations demonstrate the extraordinary sensitivity of on-chip phononic SEIRA detection to molecular stripe,


surpassing the limitations of conventional diffraction-limited far-field transmission spectroscopy. DISCUSSION We note that the sensitivity of on-chip SEIRA detection has not yet reached the


state-of-the-art far-field resonant SEIRA experiments utilizing nitrogen-cooled mercury cadmium telluride (MCT) detector, where a sensitivity of about 500 molecules has been demonstrated35.


This can be attributed to the use of a non-resonant split-gate design, the mismatch between the HPhP resonance localized in the gap of split-gate and the molecular vibrational resonances,


and the lower sensitivity of the room temperature graphene split-gate detector compared to cooled MCT detectors. Future optimization of the device design for resonant detection could


significantly boost sensitivity. This optimization includes adjusting the h-BN layer thicknesses, modifying bottom gate lengths, and fine-tuning the split-gate gap size. Importantly, the


room-temperature operation and the complementary metal-oxide-semiconductor (CMOS) compatibility of graphene detectors36 represent progress toward compact on-chip sensors functioning under


ambient conditions. Our work demonstrates that the near fields above the p–n junction of a graphene split-gate detector—which is enhanced by h-BN HPhPs—can be used for the mid-IR on-chip


phononic SEIRA detection of wavelength-size nanometer-thin layers of organic molecules deposited directly on top of the detector. We observe dips in the PC spectra at the molecular


vibrational frequencies, which are much stronger than those measured in standard far-field FTIR transmission experiments, confirming the enhanced sensitivity of on-chip phononic SEIRA


detection to nanometer-thin molecular layers. Numerical simulations predicted a significant further sensitivity enhancement compared to standard far-field transmission spectroscopy when the


lateral size of the molecular layer is reduced to a deep subwavelength scale. In the future, the applicability of on-chip phononic SEIRA detection can be expanded to a wider frequency range


and different analytes by employing several approaches. One approach could be the combination of different phononic van der Waals (vdW) materials in heterostructures for multispectral


applications37, each material supporting phonon polaritons in its own Reststrahlen bands (RBs). Corresponding materials include MoO3 (three RBs between 600 and 1000 cm−1)22,38, V2O5 (three


RBs between 500 and 1000 cm−1)23, and LiV2O5 (three RBs between 950 to 1025 cm−1)39. Further, thin layers of conventional 3D materials such as SiC40 could be incorporated. We also envision


exploiting plasmon polaritons in the detector´s active graphene layer or by incorporating additional graphene layers41. The molecule-covered detector may also be used for the PC near-field


spectroscopy42. Furthermore, the detector can be potentially combined with a transparent microfluidic system for biosensing in aqueous solutions43. Moreover, nano- and Å-scale channels


fabricated with vdW materials44,45 could also be integrated above the graphene p–n junction to enable aqueous molecular on-chip sensing. Finally, we envision either integrating tunable


quantum cascade lasers on the same chip or enabling the electrical generation of polaritons46,47 in the various layers comprising the detector, which could lead to the development of


compact, fully on-chip phononic SEIRA spectroscopy devices. METHODS DETECTOR FABRICATION First, the metallic split-gate on a Si/SiO2 substrate was fabricated via electron beam lithography


(EBL). To that end, 300 nm of PMMA resist was spin-coated onto the substrate. After EBL patterning and subsequent development of the resist in a solution of methyl isobutyl ketone


(MIBK)/isopropanol (IPA), a 2 nm-thick Ti layer and an 8 nm-thick Au layer were thermally evaporated onto the sample, followed by a lift-off procedure in acetone48. The dimensions of the


fabricated split-gate were determined from the scanning electron microscopy (SEM) image shown in Suppl. Fig. 1b and summarized in Supplementary Fig. 1a. Second, graphene was encapsulated


between two thin layers of h-BN. To that end, two flakes of h-BN and single-layer graphene were mechanically cleaved and exfoliated onto freshly cleaned Si/SiO2 substrates. Then, a


heterostructure stack (h-BN/graphene/h-BN) was fabricated using the vdW assembly technique48,49,50 and placed onto the metal split-gate. The thicknesses of the top and bottom h-BN layers


were 3 and 4.5 nm, respectively (measured by atomic force microscope (AFM)). Third, the graphene was structured into a rectangular shape to avoid sample inhomogeneity and reduce the


potential leakage between the source–drain electrodes with the local gates. To that end, a 300 nm-thick PMMA layer was spin-coated on top of the h-BN/graphene/h-BN heterostructure on top of


the split gate. Then, EBL was used to pattern an etching mask with a subsequent development step of the PMMA in the MIBK/IPA solution. The top h-BN was selectively etched using SF6 gas for 1


 min at a pressure of ≈40 cm3 STP min−1 32. The graphene was etched using O2 gas at a pressure of ≈30 cm3 STP min−1 for ≈1 min. Finally, metal contacts (for measuring the photocurrent in the


graphene layer) were added. To that end, 300 nm of PMMA resist was spin-coated onto the etched sample. Source and drain contacts were patterned using EBL, followed by resist development and


etching of h-BN and graphene (same as described above). Then, a 5 nm-thick layer of Cr and a 60 nm-thick layer of Au were evaporated. The process was completed with a lift-off in acetone.


Before using the detector for photocurrent measurements, we performed a mechanical cleaning technique known as “brooming”51, typically used to remove lithographic residues, such as PMMA,


from the surface. Brooming is done by scanning an AFM tip in contact mode across the sample area to be cleaned, here the p–n junction of the detector. Optical and AFM images of the broomed


detector are shown in Supplementary Fig. 1c, d, respectively. The electrical characterization of the fabricated detector is described in Supplementary Note 2. MOLECULE DEPOSITION


4,4′-bis(_N_-carbazolyl)-1,1′-biphenyl (CBP) of sublimed quality (99.9%) (Sigma-Aldrich, Saint Louis, MO, USA) was thermally evaporated in a high-vacuum evaporator chamber (base pressure 


< 10−9 mbar), at a rate of 0.1 nm s−1 using a Knudsen cell. To prevent the agglomeration of molecules into islands during the initial growth of molecular layers within the first 10 nm,


the Si/SiO2 substrate with the graphene split-gate detector and bare CaF2 substrate, respectively, were placed onto the cold finger, which was cooled down to a temperature of ~77–80 K by


liquid nitrogen. FREQUENCY-RESOLVED PHOTOCURRENT MEASUREMENT USING THE FOURIER TRANSFORM SPECTROSCOPY PRINCIPLE A schematic of the spectroscopy setup described in the following is shown in


Supplementary Fig. 3b. The graphene split-gate detector was illuminated by a broadband mid-IR laser (supercontinuum laser, Femtofiber pro-IR from Toptica, Gräfelfing, Germany) via a


parabolic mirror with a numerical aperture of about 0.3. The laser was tuned to emit in the frequency range 1200–1700 cm−1 with an average power of about 0.6 mW. The PC signal from the


graphene split-gate detector was measured with a low-noise current amplifier with a gain factor of 106 (DLPCA-200 from Femto, Germany). The gate voltages were applied using a DAQ card


(USB–6218 from National Instruments, USA). For frequency-resolved PC measurements, we employed the FT spectroscopy principle: the illumination of the detector was done via a Michelson


interferometer comprising a ZnSe beamsplitter and two planar mirrors. The position of one mirror (scanned mirror) was controlled by a piezo-electric scanner. The PC signal was recorded as a


function of the mirror position, yielding interferograms. Each interferogram had a maximum length of 0.8 mm. For the apodization of the interferograms, a Hann window function was applied.


After zero-filling (4× padding), the interferograms were Fourier transformed to obtain the PC spectra with a spectral resolution of about 6.5 cm−1. We first used the setup described above to


measure the IR transmission spectra of thin CBP molecule layers on a double-sided polished CaF2 substrate (Fig. 3f and Supplementary Fig. 13b), which can be considered as conventional FTIR


transmission spectra. To that end, a 10 nm-thick layer of CBP was deposited onto one-half of the CaF2 substrate (see the “Methods” section) by using a shadow mask to protect the other half


from molecular deposition. After depositing the molecules, the CaF2 substrate was placed into the beam path of our spectroscopy setup, positioning it in front of the parabolic mirror (i.e.,


outside the interferometer). First, we measure the reference transmission spectrum by aligning the substrate such that all laser radiation passes through the bare (uncoated) half of the


substrate. Then, we moved the substrate such that the laser radiation passed through the molecule-covered half of the CaF2 substrate, enabling us to measure the transmission spectrum of a


thin layer of CBP molecules. The deposition/spectroscopy cycle was repeated 5 times, ensuring that one half of the substrate remained bare (uncoated), until a cumulative CBP layer thickness


of 100 nm was reached. The thickness of the deposited layer in the first cycle was 10 and 20 nm in subsequent cycles. Each spectrum is shown in Fig. 3f and Supplementary Fig. 13b was


obtained by FT of an average of 20 interferograms recorded with 1024 pixels (positions of the scanned mirror) and an integration time of 25 ms per pixel. For the on-chip SEIRA experiments


(Fig. 3b and Supplementary Fig. 13a), we first measured the PC spectrum of the bare detector, which served as the reference spectrum for all subsequent measurements. Then, the detector was


removed from the setup and placed into the evaporation chamber for the deposition of a 10 nm-thick layer of CBP molecules directly onto the detector surface (see above). After deposition,


the detector was placed back into the spectroscopy setup to measure the PC spectrum of the molecule-covered detector. The deposition/spectroscopy cycle was repeated five times until a CBP


layer thickness of 100 nm was reached. The thickness of the deposited layer in the first cycle was 10 and 20 nm in subsequent cycles. Each spectrum shown in Fig. 3b and Supplementary Fig. 


12a was obtained by FT of an average of 20 interferograms recorded with 1024 pixels (positions of the scanned mirror) and an integration time of 20 ms per pixel. For the measurements shown


in Supplementary Note 4 (Supplementary Fig. 4a, b), the position of the scanning mirror of the interferometer was fixed, yielding a spectrally integrated photocurrent. The broadband laser


was chopped, and the output of the current amplifier was connected to a lock-in amplifier (7280 Perkin Elmer, USA). The photocurrent _I_PC was derived from the output signal of the lock-in


amplifier, _V_LIA, using the equation \({I}_{{{{\rm{PC}}}}}\,=\frac{2{{{\rm{\pi }}}}\sqrt{2}}{4\xi }\,{V}_{{{{\rm{LIA}}}}}\) (see refs. 32,52), where \(\xi\) = 106 is the gain factor of the


current amplifier in AV-1. We note that the setup was implemented with optical, electronic and hardware components of a commercial scattering-type scanning near-field optical microscope


(s-SNOM) comprising a nano-FTIR module (Neaspec/Attocube, Germany). The commercial nano-FTIR module consists of a ZnSe beamsplitter and one planar mirror (scanned mirror), whose position is


controlled by a piezo-electric scanner. In the standard s-SNOM setup, as shown in Supplementary Fig. 3a, the light from the laser is split into two beams (reference and illumination beams)


by a ZnSe beamsplitter. The reference beam is reflected in a scanned mirror. The illumination beam is focused on an AFM tip by a parabolic mirror, where the illumination beam is then


scattered. At an HgCdTe (MCT) detector, the light backscattered from the AFM tip interferes with the back-reflected reference beam. The MCT detector signal is recorded as a function of the


position of the scanned mirror that is linearly translated. To perform our frequency-resolved photocurrent measurement, we modified the setup by removing the MCT detector and the AFM tip,


rotating the nano-FTIR module, and adding a second planar mirror to build the Michelson interferometer, as shown in Supplementary Fig. 3b. NUMERICAL SIMULATIONS Full-wave numerical


simulations using the finite-element method (COMSOL) were performed to simulate the electrical field enhancement in the gap region of the detector and to calculate the absorption in graphene


(see Supplementary Note 5). The dielectric permittivities of h-BN, Au, CBP, SiO2, and Si are provided in Supplementary Note 5.2. The simulations shown in Fig. 2, which were performed in the


absence of HPhPs, employed a frequency-independent permittivity tensor for h-BN, where \({\varepsilon }_{{{{\rm{h}}}}-{{{\rm{BN}}}}}^{\perp }\left(\omega \right)={\varepsilon }_{\infty


}^{\perp }=\)4.98 and \({\varepsilon }_{{{{\rm{h}}}}-{{{\rm{BN}}}}}^{\parallel }\left(\omega \right)={\varepsilon }_{\infty }^{\parallel }=\)2.95. DATA AVAILABILITY The data that support the


findings of this study are available from the corresponding author upon request. The raw interferograms generated in this study are available from Zenodo53. Source data are provided with


this paper. REFERENCES * Griffiths, P. R. & Haseth, J. A. _Fourier Transform Infrared Spectrometry_ (John Wiley & Sons, Inc., Hoboken, NJ, 2007). * Masatoshi, O. Surface-enhanced


infrared absorption. In _Near-Field Optics and Surface Plasmon Polaritons_ (ed. Kawata, S.) 163–187 (Springer, 2001). * Neubrech, F., Huck, C., Weber, K., Pucci, A. & Giessen, H.


Surface-enhanced infrared spectroscopy using resonant nanoantennas. _Chem. Rev._ 117, 5110–5145 (2017). Article  CAS  PubMed  Google Scholar  * John-Herpin, A. et al. Metasurface-enhanced


infrared spectroscopy: an abundance of materials and functionalities. _Adv. Mater._ 35, 2110163 (2023). Article  CAS  Google Scholar  * Kundu, J., Le, F., Nordlander, P. & Halas, N. J.


Surface enhanced infrared absorption (SEIRA) spectroscopy on nanoshell aggregate substrates. _Chem. Phys. Lett._ 452, 115–119 (2008). Article  ADS  CAS  Google Scholar  * Ataka, K. &


Heberle, J. Biochemical applications of surface-enhanced infrared absorption spectroscopy. _Anal. Bioanal. Chem._ 388, 47–54 (2007). Article  CAS  PubMed  PubMed Central  Google Scholar  *


Ataka, K., Stripp, S. T. & Heberle, J. Surface-enhanced infrared absorption spectroscopy (SEIRAS) to probe monolayers of membrane proteins. _Biochim. Biophys. Acta-Biomembr._ 1828,


2283–2293 (2013). Article  CAS  Google Scholar  * Neubrech, F. et al. Resonant plasmonic and vibrational coupling in a tailored nanoantenna for infrared detection. _Phys. Rev. Lett._ 101,


157403 (2008). Article  ADS  PubMed  Google Scholar  * Semenyshyn, R. et al. Pushing down the limit: in vitro detection of a polypeptide monolayer on a single infrared resonant nanoantenna.


_ACS Photonics_ 6, 2636–2642 (2019). Article  CAS  Google Scholar  * Brown, L. V. et al. Surface-enhanced infrared absorption using individual cross antennas tailored to chemical moieties.


_J. Am. Chem. Soc._ 135, 3688–3695 (2013). Article  CAS  PubMed  Google Scholar  * Aouani, H. et al. Ultrasensitive broadband probing of molecular vibrational modes with multifrequency


optical antennas. _ACS Nano_ 7, 669–675 (2013). Article  CAS  PubMed  Google Scholar  * Pfitzner, E., Seki, H., Schlesinger, R., Ataka, K. & Heberle, J. Disc antenna enhanced infrared


spectroscopy: from self-assembled monolayers to membrane proteins. _ACS Sens._ 3, 984–991 (2018). Article  CAS  PubMed  Google Scholar  * Rodrigo, D. et al. Mid-infrared plasmonic biosensing


with graphene. _Science_ 349, 165–168 (2015). Article  ADS  CAS  PubMed  Google Scholar  * Hu, H. et al. Gas identification with graphene plasmons. _Nat. Commun._ 10, 1131 (2019). Article 


ADS  PubMed  PubMed Central  Google Scholar  * Bareza, N. J., Gopalan, K. K., Alani, R., Paulillo, B. & Pruneri, V. Mid-infrared gas sensing using graphene plasmons tuned by reversible


chemical doping. _ACS Photonics_ 7, 879–884 (2020). Article  CAS  Google Scholar  * Lee, I. H., Yoo, D., Avouris, P., Low, T. & Oh, S. H. Graphene acoustic plasmon resonator for


ultrasensitive infrared spectroscopy. _Nat. Nanotechnol._ 14, 313–319 (2019). Article  ADS  CAS  PubMed  Google Scholar  * Tittl, A. et al. Imaging-based molecular barcoding with pixelated


dielectric metasurfaces. _Science_ 360, 1105–1109 (2018). Article  ADS  MathSciNet  CAS  PubMed  Google Scholar  * Leitis, A. et al. Angle-multiplexed all-dielectric metasurfaces for


broadband molecular fingerprint retrieval. _Sci. Adv._ 5, eaaw2871 (2019). Article  ADS  PubMed  PubMed Central  Google Scholar  * Tseng, M. L., Jahani, Y., Leitis, A. & Altug, H.


Dielectric metasurfaces enabling advanced optical biosensors. _ACS Photonics_ 8, 47–60 (2021). Article  CAS  Google Scholar  * Dai, S. et al. Tunable phonon polaritons in atomically thin van


der Waals crystals of boron nitride. _Science_ 343, 1125–1129 (2014). Article  ADS  CAS  PubMed  Google Scholar  * Caldwell, J. D. et al. Sub-diffractional volume-confined polaritons in the


natural hyperbolic material hexagonal boron nitride. _Nat. Commun._ 5, 5221 (2014). Article  ADS  CAS  PubMed  Google Scholar  * Ma, W. et al. In-plane anisotropic and ultra-low-loss


polaritons in a natural van der Waals crystal. _Nature_ 562, 557–562 (2018). Article  ADS  CAS  PubMed  Google Scholar  * Taboada-Gutiérrez, J. et al. Broad spectral tuning of ultra-low-loss


polaritons in a van der Waals crystal by intercalation. _Nat. Mater._ 19, 964–968 (2020). Article  ADS  PubMed  Google Scholar  * Hu, G., Shen, J., Qiu, C. W., Alù, A. & Dai, S. Phonon


polaritons and hyperbolic response in van der Waals materials. _Adv. Opt. Mater._ 8, 1901393 (2020). Article  CAS  Google Scholar  * Autore, M. et al. Boron nitride nanoresonators for


phonon-enhanced molecular vibrational spectroscopy at the strong coupling limit. _Light Sci. Appl._ 7, 17172–17178 (2018). Article  CAS  PubMed  PubMed Central  Google Scholar  * Bareza, N.


J. et al. Phonon-enhanced mid-infrared CO2 gas sensing using boron nitride nanoresonators. _ACS Photonics_ 9, 34–42 (2022). Article  CAS  Google Scholar  * Autore, M. et al. Enhanced


light–matter interaction in 10B monoisotopic boron nitride infrared nanoresonators. _Adv. Opt. Mater._ 9, 2001958 (2021). Article  CAS  Google Scholar  * Bylinkin, A. et al. Real-space


observation of vibrational strong coupling between propagating phonon polaritons and organic molecules. _Nat. Photonics_ 15, 197–202 (2021). Article  ADS  CAS  Google Scholar  * Nash, G. R.,


Luxmoore, I. J., Liu, P. Q., Li, P. & Faist, J. Graphene—metamaterial photodetectors for integrated infrared sensing. _ACS Photonics_ 3, 936–941 (2016). Article  Google Scholar  *


Lundeberg, M. B. et al. Thermoelectric detection and imaging of propagating graphene plasmons. _Nat. Mater._ 16, 204–207 (2017). Article  ADS  CAS  PubMed  Google Scholar  * Woessner, A. et


al. Electrical detection of hyperbolic phonon-polaritons in heterostructures of graphene and boron nitride. _npj 2D Mater. Appl._ 1, 25 (2017). Article  Google Scholar  * Castilla, S. et al.


Plasmonic antenna coupling to hyperbolic phonon-polaritons for sensitive and fast mid-infrared photodetection with graphene. _Nat. Commun._ 11, 4872 (2020). Article  ADS  CAS  PubMed 


PubMed Central  Google Scholar  * Badioli, M. et al. Phonon-mediated mid-infrared photoresponse of graphene. _Nano Lett._ 14, 6374–6381 (2014). Article  ADS  CAS  PubMed  Google Scholar  *


Dolado, I. et al. Remote near-field spectroscopy of vibrational strong coupling between organic molecules and phononic nanoresonators. _Nat. Commun._ 13, 6850 (2022). Article  ADS  CAS 


PubMed  PubMed Central  Google Scholar  * Dong, L. et al. Nanogapped Au antennas for ultrasensitive surface-enhanced infrared absorption spectroscopy. _Nano Lett._ 17, 5768–5774 (2017).


Article  ADS  CAS  PubMed  Google Scholar  * Goossens, S. et al. Broadband image sensor array based on graphene–CMOS integration. _Nat. Photonics_ 11, 366–371 (2017). Article  ADS  CAS 


Google Scholar  * Qin, T., Ma, W., Wang, T. & Li, P. Phonon polaritons in van der Waals polar heterostructures for broadband strong light-matter interactions. _Nanoscale_ 15, 12000–12007


(2023). Article  CAS  PubMed  Google Scholar  * Álvarez-Pérez, G. et al. Infrared permittivity of the biaxial van der Waals semiconductor α-MoO3 from near- and far-field correlative


studies. _Adv. Mater._ 32, 1908176 (2020). Article  Google Scholar  * Tresguerres-Mata, A. I. et al. Observation of naturally canalized phonon polaritons in LiV2O5 thin layers. _Nat.


Commun._ 15, 2696 (2024). Article  ADS  Google Scholar  * Mancini, A. et al. Near-field retrieval of the surface phonon polariton dispersion in free-standing silicon carbide thin films. _ACS


Photonics_ 9, 3696–3704 (2022). Article  CAS  Google Scholar  * Doukas, S., Sharma, P., Goykhman, I. & Lidorikis, E. Electrical detection of graphene plasmons for mid-infrared


photodetection and chemical sensing: a computational study. _Appl. Phys. Lett._ 121, 051103 (2022). Article  ADS  CAS  Google Scholar  * Ulrich, G. et al. Thermoelectric nanospectroscopy for


the imaging of molecular fingerprints. _Nanophotonics_ 9, 4347–4354 (2020). Article  CAS  Google Scholar  * Wu, C. et al. Ultrasensitive mid-infrared biosensing in aqueous solutions with


graphene plasmons. _Adv. Mater._ 34, 2110525 (2022). Article  CAS  Google Scholar  * Keerthi, A. et al. Water friction in nanofluidic channels made from two-dimensional crystals. _Nat.


Commun._ 12, 3092 (2021). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Bhardwaj, A. et al. Fabrication of angstrom-scale two-dimensional channels for mass transport. _Nat.


Commun._ 12, 240–280 (2024). Google Scholar  * De Vega, S. & García De Abajo, F. J. Plasmon generation through electron tunneling in graphene. _ACS Photonics_ 4, 2367–2375 (2017).


Article  Google Scholar  * Enaldiev, V., Bylinkin, A. & Svintsov, D. Plasmon-assisted resonant tunneling in graphene-based heterostructures. _Phys. Rev. B_ 96, 125437 (2017). * Wang, L.


et al. One-dimensional electrical contact to a two-dimensional material. _Science_ 342, 614–617 (2013). Article  ADS  CAS  PubMed  Google Scholar  * Pizzocchero, F. et al. The hot pick-up


technique for batch assembly of van der Waals heterostructures. _Nat. Commun._ 7, 11894 (2016). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Purdie, D. G. et al. Cleaning


interfaces in layered materials heterostructures. _Nat. Commun._ 9, 5387 (2018). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Lindvall, N., Kalabukhov, A. & Yurgens, A.


Cleaning graphene using atomic force microscope. _J. Appl. Phys._ 111, 064904 (2012). Article  ADS  Google Scholar  * Vicarelli, L. et al. Graphene field-effect transistors as


room-temperature terahertz detectors. _Nat. Mater._ 11, 865–871 (2012). Article  ADS  CAS  PubMed  Google Scholar  * Andrei, B. et al. On-chip phonon-enhanced IR near-field detection of


molecular vibrations. _Zenodo_. https://doi.org/10.5281/zenodo.13757138 (2024). Article  Google Scholar  Download references ACKNOWLEDGEMENTS The work was financially supported by the


European Union’s Horizon 2020 research and innovation program under grant agreement Nos. 785219 and 881603 (GrapheneCore2 and GrapheneCore3 of the Graphene Flagship). R.H., F. Casanova, and


L.E.H. acknowledge funding by the Spanish MICIU/AEI/10.13039/501100011033 (grant CEX2020-001038-M). R.H. acknowledges funding by the Spanish MICIU/AEI/10.13039/501100011033 and ERDF/EU


(grants RTI2018-094830-B-I00 and PID2021-123949OB-I00). F. Casanova and L.E.H. acknowledge funding by the Spanish MICIU/AEI/10.13039/501100011033 and ERDF/EU (grant PID2021-122511OB-I00).


F.H.L.K. acknowledges funding by the Spanish MICIU/AEI/10.13039/501100011033 (grants FIS2016-81044, PID2019-106875GB-100, CEX2019-000910-S, PCI2021-122020-2A and PDC2022-133844-100). A.N.


acknowledges funding by the Spanish MICIU/AEI/10.13039/501100011033 and ERDF/EU (grants PID2020-115221GB-C42 and PID2023-147676NB-I00); by the Department of Education of the Basque


Government (grant PIBA-2023-1-0007). L.M.-M and T.S. acknowledge funding by the Spanish MICIU/AEI/10.13039/501100011033 (grants PID2020-115221GB-C41 and CEX2023-001286-S); by the Government


of Aragon through Project Q-MAD. F.H.L.K, S.C., and V.P. acknowledge funding by the European Union (ERC, POLARSENSE, 101123421). Views and opinions expressed are, however, those of the


author(s) only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held


responsible for them. AUTHOR INFORMATION AUTHORS AND AFFILIATIONS * CIC nanoGUNE BRTA, 20018, Donostia-San Sebastián, Spain Andrei Bylinkin, Francesco Calavalle, Marta Autore, Fèlix


Casanova & Luis E. Hueso * Donostia International Physics Center (DIPC), 20018, Donostia-San Sebastián, Spain Andrei Bylinkin & Alexey Y. Nikitin * ICFO-Institut de Ciències


Fotòniques, The Barcelona Institute of Science and Technology, Av. Carl Friedrich Gauss 3, 08860, Castelldefels (Barcelona), Spain Sebastián Castilla, Varun-Varma Pusapati & Frank H. L.


Koppens * Instituto de Nanociencia y Materiales de Aragon (INMA), CSIC-Universidad de Zaragoza, 50009, Zaragoza, Spain Tetiana M. Slipchenko & Luis Martín-Moreno * Departamento de Fisica


de la Materia Condensada, Universidad de Zaragoza, Zaragoza, 50009, Spain Tetiana M. Slipchenko & Luis Martín-Moreno * Donostia International Physics Center (DIPC) and EHU/UPV, 20018,


Donostia-San Sebastián, Spain Kateryna Domina * IKERBASQUE, Basque Foundation for Science, 48009, Bilbao, Spain Fèlix Casanova, Luis E. Hueso, Alexey Y. Nikitin & Rainer Hillenbrand *


ICREA—Institució Catalana de Recerca i Estudis Avançats, Barcelona, 08010, Spain Frank H. L. Koppens * CIC nanoGUNE BRTA and EHU/UPV, 20018, Donostia-San Sebastián, Spain Rainer Hillenbrand


Authors * Andrei Bylinkin View author publications You can also search for this author inPubMed Google Scholar * Sebastián Castilla View author publications You can also search for this


author inPubMed Google Scholar * Tetiana M. Slipchenko View author publications You can also search for this author inPubMed Google Scholar * Kateryna Domina View author publications You can


also search for this author inPubMed Google Scholar * Francesco Calavalle View author publications You can also search for this author inPubMed Google Scholar * Varun-Varma Pusapati View


author publications You can also search for this author inPubMed Google Scholar * Marta Autore View author publications You can also search for this author inPubMed Google Scholar * Fèlix


Casanova View author publications You can also search for this author inPubMed Google Scholar * Luis E. Hueso View author publications You can also search for this author inPubMed Google


Scholar * Luis Martín-Moreno View author publications You can also search for this author inPubMed Google Scholar * Alexey Y. Nikitin View author publications You can also search for this


author inPubMed Google Scholar * Frank H. L. Koppens View author publications You can also search for this author inPubMed Google Scholar * Rainer Hillenbrand View author publications You


can also search for this author inPubMed Google Scholar CONTRIBUTIONS R.H. conceived the initial study and supervised the work. S.C. fabricated the detector with the help of V.-V.P.


supervised by F.H.L.K. A.B. performed the experiments and data analysis. S.C. participated in the experiments and data analysis. M.A. contributed to the initial experiments. F. Calavalle


performed molecular deposition supervised by F. Casanova and L.E.H. K.D., T.M.S., and A.B. performed numerical simulations supervised by L.M.-M. and A.Y.N. A.B. and R.H. wrote the manuscript


with input from S.C., T.M.S., L.M.-M., A.Y.N., and F.K. All authors contributed to the scientific discussion and paper revisions. CORRESPONDING AUTHOR Correspondence to Rainer Hillenbrand.


ETHICS DECLARATIONS COMPETING INTERESTS R.H. is a co-founder of Neaspec GmbH, now part of Attocube AG, a company producing scattering-type scanning near-field optical microscope systems,


such as the one described in this work. Apart from this, the authors declare no competing interests. PEER REVIEW PEER REVIEW INFORMATION _Nature Communications_ thanks Patrick Bouchon,


Sang-Hyun Oh, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available. ADDITIONAL INFORMATION PUBLISHER’S NOTE Springer


Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION SOURCE DATA TRANSPARENT PEER


REVIEW FILE SOURCE DATA RIGHTS AND PERMISSIONS OPEN ACCESS This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation,


distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and


indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to


the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will


need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE


CITE THIS ARTICLE Bylinkin, A., Castilla, S., Slipchenko, T.M. _et al._ On-chip phonon-enhanced IR near-field detection of molecular vibrations. _Nat Commun_ 15, 8907 (2024).


https://doi.org/10.1038/s41467-024-53182-9 Download citation * Received: 09 July 2024 * Accepted: 04 October 2024 * Published: 16 October 2024 * DOI:


https://doi.org/10.1038/s41467-024-53182-9 SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not


currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative