Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Stress-Dependent Optical Extinction in LPCVD Silicon Nitride Measured by Nanomechanical Photothermal Sensing

Kostas Kanellopulos    Robert G. West    Stefan Emminger    Paolo Martini [    Markus Sauer    Annette Foelske [    Silvan Schmid [ silvan.schmid@tuwien.ac.at
Abstract

Understanding optical absorption in silicon nitride is crucial for cutting-edge technologies like photonic integrated circuits, nanomechanical photothermal infrared sensing and spectroscopy, and cavity optomechanics. Yet, the origin of its strong dependence on film deposition and fabrication process is not fully understood. This Letter leverages nanomechanical photothermal sensing to investigate optical extinction κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT at 632.8 nm wavelength in LPCVD SiN strings across a wide range of deposition-related tensile stresses (200850200850200-850200 - 850 MPa). Measurements reveal a reduction in κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT from 103 to 101 ppm with increasing stress, correlated to variations in Si/N content ratio. Within the band-fluctuations framework, this trend indicates an increase of the energy bandgap with the stress, ultimately reducing absorption. Overall, this study showcases the power and simplicity of nanomechanical photothermal sensing for low absorption measurements, offering a sensitive, scattering-free platform for material analysis in nanophotonics and nanomechanics.

keywords:
American Chemical Society,

TU Wien] Institute of Sensor and Actuator Systems, TU Wien, 1040 Vienna, Austria. TU Wien] Analytical Instrumentation Center, TU Wien, 1060 Vienna, Austria TU Wien] Institute of Sensor and Actuator Systems, TU Wien, 1040 Vienna, Austria. \abbreviationsVIS,IR,NPAS

Investigating the optical properties of solid-state materials is essential for both fundamental and applied science. Optical absorption, in particular, is critical in various fields, including photonic integrated circuits (PIC) for quantum information,1 and the design of nanomechanical resonant sensors for infrared (IR) light detection, 2 photothermal spectromicroscopy, 3, 4, 5 and cavity optomechanics 6, 7, 8, 9, 10.

In IR sensing, high optical absorption is desired to enhance the sensor’s specific detectivity. 11, 2 Conversely, applications like PICs, cavity optomechanics, and nanomechanical photothermal spectromicroscopy require minimal absorption to realize high-confinement waveguides,12 to prevent mechanical instability13, 10 and cavity bistability,14 and to mitigate photothermal back-action frequency noise introduced in the resonator 15, respectively.

Silicon nitride (SiN) holds a prominent position in these fields for its excellent mechanical, thermal, and optical properties.16, 17 Its extensive use in photonics stems from its broad transparency window (0.48μm0.48𝜇m0.4-8\ \mathrm{\mu m}0.4 - 8 italic_μ roman_m), which, however, strongly depends on the film deposition and fabrication process, for which the underlying mechanisms are not fully understood. This has been observed by means of various characterization techniques, such as ellipsometry 18, 19, 20, 21, 22, 23, direct single-pass absorption spectroscopy 24, FTIR interferometry 25, 26, cutback 27, outscattered light method 28, 29, 30, 31, 32, prism coupling 33, photoluminescence 34, photothermal common-path interferometry 35, and cavity-enhanced absorption spectroscopy 36, 16, 12. However, these approaches often suffer from scattering losses and slow measurement time, which obscure the true absorption of SiN and make analyses prone to parasitic heating of the surroundings.

Within this context, nanomechanical photothermal spectroscopy offers a robust solution to these challenges.37, 38, 39, 40, 4, 41, 42 Here, a nanomechanical resonator detects directly absorption via resonance frequency shifts due to photothermal heating, insensitive to scattering. Upon illumination, the resultant temperature rise makes the initial tensile stress relax, leading to frequency detuning. Due to its high power sensitivity, fast thermal response, and versatility in sensor design, this technique has significantly advanced the characterization of low-loss materials 5.

In this Letter, nanomechanical photothermal sensing is employed to elucidate the relationship between absorption and residual tensile stress in low-pressure chemical vapour deposition (LPCVD) deposited SiN thin films. The extinction coefficient at 632.8 nm wavelength is measured from low stress (200absent200\approx 200≈ 200 MPa), relevant to photothermal-based applications,5 to high stress (>800absent800>800> 800 MPa), relevant to cavity optomechanics 43, 44 and PIC design.12 The thin films are patterned in a string geometry, which ensures high photothermal responsivity and fast response, as previously demonstrated.15 The experimental results reveal a reduction in extinction (from 103 to 101 ppm) with increasingly higher tensile stress, consistent with previously observed trends.23 The measurements are analyzed within the framework of the band-fluctuations model,45 attributing the observed reduction to a blue-shift in the energy bandgap caused by a decrease in silicon-to-nitrogen (Si/N) content ratio. Overall, this study underscores the power and simplicity of nanomechanical photothermal sensing for the characterization of low-loss materials in nanophotonics and nanomechanics.

Refer to caption
Figure 1: (a) Sketch of the experimental set-up (LDV, Polytech GmbH MSA-500). A laser of wavelength λ𝜆\lambdaitalic_λ and input power P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT impinges on the resonator, of absorption coefficient αabs(λ)subscript𝛼abs𝜆\alpha_{\mathrm{abs}}(\lambda)italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( italic_λ ). This causes a frequency detuning of the nanomechanical resonator. BS: beam-splitter. BC: Bragg cell. PD: photodetector. (b) Mechanical frequency detuning measured by monitoring the shift of the thermomechanical noise peak of the string’s fundamental mode as a function of P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (c) Optical micrograph of the SiN strings used in the present study. Orange/light blue regions are made of SiN; the grey regions are the Si substrate. (d) Photo of the cm-scale copper thermal equilibrium chamber used for the characterization of the linear coefficient of thermal expansion. A thermoelectric module is glued beneath to heat up the whole oven (thick red electrical connections) to guarantee a uniform temperature rise of the chips. The temperature is monitored and kept constant with a PID controller.

In the present setup, the nanomechanical resonator is operated in a custom-made vacuum chamber at high-vacuum conditions (p<105mbar𝑝superscript105mbarp<10^{-5}\ \mathrm{mbar}italic_p < 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT roman_mbar) to reduce gas damping and thermal convection losses.4 The mechanical displacement is transduced optically via laser-Doppler vibrometry (LDV, Polytec GmbH MSA-500) equipped with a HeNe laser at 632.8 nm wavelength, with a beam waist of 1.5μmabsent1.5𝜇m\approx 1.5\ \mathrm{\mu m}≈ 1.5 italic_μ roman_m (Fig. 1). The same laser also probes SiN absorption, which simplifies the measurement procedure. For each structure, the frequency shift of the thermomechanical noise peak for the fundamental resonance mode is recorded at various optical powers (6120μW6120𝜇W6-120\ \mathrm{\mu W}6 - 120 italic_μ roman_W)46, 15, as schematically shown in Fig. 1b. A minimum of five resonators is evaluated for each stress and length. Fig. 1c shows a representative image of the characterized strings.

The absorption coefficient αabssubscript𝛼abs\alpha_{\mathrm{abs}}italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT is determined via comparison between the theoretical PsubscriptP\mathcal{R}_{\mathrm{P}}caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT and the experimental P0subscriptsubscriptP0\mathcal{R}_{\mathrm{P_{0}}}caligraphic_R start_POSTSUBSCRIPT roman_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT relative power responsivity for the string resonators.47, 5, 15 On the one hand, PsubscriptP\mathcal{R}_{\mathrm{P}}caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT denotes the relative frequency change per absorbed power P(λ)=αabs(λ)P0𝑃𝜆subscript𝛼abs𝜆subscript𝑃0P(\lambda)=\alpha_{\mathrm{abs}}(\lambda)P_{0}italic_P ( italic_λ ) = italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( italic_λ ) italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and is expressed as 47

P(ω)=1f0f0P=TG|Hth(ω)|,subscriptP𝜔1subscript𝑓0subscript𝑓0𝑃subscriptT𝐺subscript𝐻th𝜔\mathcal{R}_{\mathrm{P}}(\omega)=\frac{1}{f_{0}}\frac{\partial f_{0}}{\partial P% }=\frac{\mathcal{R}_{\mathrm{T}}}{G}\left|H_{\mathrm{th}}(\omega)\right|,caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_P end_ARG = divide start_ARG caligraphic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT end_ARG start_ARG italic_G end_ARG | italic_H start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT ( italic_ω ) | , (1)

where TsubscriptT\mathcal{R}_{\mathrm{T}}caligraphic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT, G𝐺Gitalic_G, and Hth(ω)=(1+iωτth)1subscript𝐻th𝜔superscript1i𝜔subscript𝜏th1H_{\mathrm{th}}(\omega)=(1+\mathrm{i}\omega\tau_{\mathrm{th}})^{-1}italic_H start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT ( italic_ω ) = ( 1 + roman_i italic_ω italic_τ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT denote the temperature responsivity in units [1/K], the thermal conductance in units [W/K], and the thermal response of the resonator, with τthsubscript𝜏th\tau_{\mathrm{th}}italic_τ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT its thermal time constant, respectively.47, 15 All the measurements have been performed far from any thermal transient (ωτth1much-less-than𝜔superscriptsubscript𝜏th1\omega\ll\tau_{\mathrm{th}}^{-1}italic_ω ≪ italic_τ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT). Hence, the ω𝜔\omegaitalic_ω-dependence is dropped in the following.

For a string resonator, Eq. (1) is given by 47

P=αthE2σ0[8hwLκ+8LwϵradσSBT03]1,subscriptPsubscript𝛼th𝐸2subscript𝜎0superscriptdelimited-[]8𝑤𝐿𝜅8𝐿𝑤subscriptitalic-ϵradsubscript𝜎SBsuperscriptsubscript𝑇031\mathcal{R}_{\mathrm{P}}=-\frac{\alpha_{\mathrm{th}}E}{2\sigma_{0}}\left[8% \frac{hw}{L}\kappa+8Lw\epsilon_{\mathrm{rad}}\sigma_{\mathrm{SB}}T_{0}^{3}% \right]^{-1},caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT = - divide start_ARG italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT italic_E end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG [ 8 divide start_ARG italic_h italic_w end_ARG start_ARG italic_L end_ARG italic_κ + 8 italic_L italic_w italic_ϵ start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT roman_SB end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (2)

with αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT, E𝐸Eitalic_E, σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, hhitalic_h, w𝑤witalic_w, L𝐿Litalic_L, κ𝜅\kappaitalic_κ, ϵradsubscriptitalic-ϵrad\epsilon_{\mathrm{rad}}italic_ϵ start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT, σSBsubscript𝜎SB\sigma_{\mathrm{SB}}italic_σ start_POSTSUBSCRIPT roman_SB end_POSTSUBSCRIPT and T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denoting the resonator’s linear coefficient of thermal expansion, Young’s modulus, tensile stress, thickness, width, length, thermal conductivity, emissivity, Stefan-Boltzmann constant, and bath temperature, respectively. The thermal conductance G𝐺Gitalic_G includes the thermal dissipation through the surrounding frame Gcondsubscript𝐺condG_{\mathrm{cond}}italic_G start_POSTSUBSCRIPT roman_cond end_POSTSUBSCRIPT (first addend in brackets), and thermal radiation to the environment Gradsubscript𝐺radG_{\mathrm{rad}}italic_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT (second addend in brackets).47, 15

P0subscriptsubscriptP0\mathcal{R}_{\mathrm{P_{0}}}caligraphic_R start_POSTSUBSCRIPT roman_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is obtained by directly measuring the relative frequency shift per impinging power P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (as schematically shown in Fig. 1b). It relates to PsubscriptP\mathcal{R}_{\mathrm{P}}caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT through the absorption coefficient αabs(λ)subscript𝛼abs𝜆\alpha_{\mathrm{abs}}(\lambda)italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( italic_λ ) as follows

P0(λ)=1f0f0P0=1f0f0PPP0=αabs(λ)P,subscriptsubscriptP0𝜆1subscript𝑓0subscript𝑓0subscript𝑃01subscript𝑓0subscript𝑓0𝑃𝑃subscript𝑃0subscript𝛼abs𝜆subscriptP\mathcal{R}_{\mathrm{P_{0}}}(\lambda)=\frac{1}{f_{0}}\frac{\partial f_{0}}{% \partial P_{0}}=\frac{1}{f_{0}}\frac{\partial f_{0}}{\partial P}\frac{\partial P% }{\partial P_{0}}=\alpha_{\mathrm{abs}}(\lambda)\ \mathcal{R}_{\mathrm{P}},caligraphic_R start_POSTSUBSCRIPT roman_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_λ ) = divide start_ARG 1 end_ARG start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_P end_ARG divide start_ARG ∂ italic_P end_ARG start_ARG ∂ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( italic_λ ) caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT , (3)

with λ𝜆\lambdaitalic_λ denoting the probe optical wavelength. From Eq. (3), it is possible to directly evaluate the optical absorption coefficient as αabs=P0/Psubscript𝛼abssubscriptsubscriptP0subscriptP\alpha_{\mathrm{abs}}=\mathcal{R}_{\mathrm{P_{0}}}/\mathcal{R}_{\mathrm{P}}italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT = caligraphic_R start_POSTSUBSCRIPT roman_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT / caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT.

The experimental power responsivity (3) across the different stresses is displayed in Fig. 2a as a function of the resonators’ length L𝐿Litalic_L (circles), together with the theoretical calculations (2) (solid curves). The scale is given in terms of absorbed power P𝑃Pitalic_P. It is worth noting that PsubscriptP\mathcal{R}_{\mathrm{P}}caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT grows for longer strings in the conduction-limited regime (L<1𝐿1L<1italic_L < 1 mm), as GGcondsimilar-to-or-equals𝐺subscript𝐺condG\simeq G_{\mathrm{cond}}italic_G ≃ italic_G start_POSTSUBSCRIPT roman_cond end_POSTSUBSCRIPT is inversely proportional to the length L𝐿Litalic_L (see Eq. (2)), leading to better thermal insulation from the environment.15 Conversely, increasingly longer resonators (L>1𝐿1L>1italic_L > 1 mm) enter the radiation-limited regime (GGradsimilar-to-or-equals𝐺subscript𝐺radG\simeq G_{\mathrm{rad}}italic_G ≃ italic_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT), leading to a drop of PsubscriptP\mathcal{R}_{\mathrm{P}}caligraphic_R start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT as the radiating surface increases.

Refer to caption
Figure 2: (a) Psubscript𝑃\mathcal{R}_{P}caligraphic_R start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT for different SiN string structures. Circles: experimental responsivity (3), divided by the corresponding mean absorption coefficient αabssubscript𝛼abs\alpha_{\mathrm{abs}}italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT. Solid curve: theoretical model (2). Material parameter assumed: ρ=3000kg/m3𝜌3000kgsuperscriptm3\rho=3000\ \mathrm{kg/m^{3}}italic_ρ = 3000 roman_kg / roman_m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, κ=3𝜅3\kappa=3italic_κ = 3 W/(m K). Emissivity values are calculated from data reported in Ref. 25: 0.05 (h=5656h=56italic_h = 56 nm), 0.13 (h=157157h=157italic_h = 157 nm), 0.133 (h=177177h=177italic_h = 177 nm), 0.171 (h=312312h=312italic_h = 312 nm), 0.176 (h=340340h=340italic_h = 340 nm). (b) Example of Young’s modulus estimation, following the procedure of Ref. 48. (c) Experimental Young’s modulus E as a function of the prestress σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (d) Experimental linear coefficient of thermal expansion αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT as a function of the prestress σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.
Table 1: String resonators’ geometrical (hhitalic_h), mechanical (σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, E𝐸Eitalic_E, αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT), compositional (Si/N), and optical (η𝜂\etaitalic_η, κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT, Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) properties.
σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (MPaMPa\mathrm{MPa}roman_MPa) h (nmnm\mathrm{nm}roman_nm) E (GPaGPa\mathrm{GPa}roman_GPa) αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT (ppm/KppmK\mathrm{ppm/K}roman_ppm / roman_K) η𝜂\etaitalic_η κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT (ppm) Si/N Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT (eVeV\mathrm{eV}roman_eV) β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (meVmeV\mathrm{meV}roman_meV)
174 177 200 1.45 1.215 606 0.96 3.23 201
275 340 243 1.28 1.105 588 0.98 3.09 183
370 56 214 1.06 1.022 176 0.89 3.62 212
775 56 214 1.51 1.022 38 0.86 3.90 208
815 157 173 1.29 1.273 20 0.83 4.21 227
834 312 227 1.55 1.268 21 0.84 4.10 217

The analyzed resonators have thicknesses of h=56340nm56340nmh=56-340\ \mathrm{nm}italic_h = 56 - 340 roman_nm and widths of w=550μm𝑤550𝜇mw=5-50\ \mathrm{\mu m}italic_w = 5 - 50 italic_μ roman_m, ensuring minimal thermal dissipation. As indicated in Eq. (2), Gcondhwproportional-tosubscript𝐺cond𝑤G_{\mathrm{cond}}\propto hwitalic_G start_POSTSUBSCRIPT roman_cond end_POSTSUBSCRIPT ∝ italic_h italic_w and Gradwproportional-tosubscript𝐺rad𝑤G_{\mathrm{rad}}\propto witalic_G start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT ∝ italic_w, making these strings highly responsive to photothermal heating. Furthermore, the length L𝐿Litalic_L varies in the range 0.120.120.1-20.1 - 2 mm, making the resonator’s power response mainly thermal conduction limited.15 The current experimental approach is, therefore, less influenced by the SiN emissivity, which, according to Kirchhoff’s law, equals the optical absorption49—the parameter under scrutiny in this study. Moreover, the temperature responsivity TsubscriptT\mathcal{R}_{\mathrm{T}}caligraphic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT appearing in Eq. (2) is independent of the Poisson’s ratio ν𝜈\nuitalic_ν, opposite to what occurs in, e.g., membrane resonators,47, 15 which further reduces the uncertainty in the absorption measurement stemming from its dependence on other material parameters. Hence, the resonators employed here make this approach highly robust for solid-state material absorption characterization.

In this regard, the Young’s modulus E𝐸Eitalic_E and the linear coefficient of thermal expansion αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT have been measured to reduce the uncertainty on the estimation of the absorption coefficient αabssubscript𝛼abs\alpha_{\mathrm{abs}}italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT. E𝐸Eitalic_E has been estimated following the procedure of Ref. 48, upon recording of the strings’ eigenmode spectrum (an example is displayed in Fig. 2b, where Δ=|mn|Δ𝑚𝑛\Delta=|m-n|roman_Δ = | italic_m - italic_n |, with nm𝑛𝑚n\neq mitalic_n ≠ italic_m being modal numbers). The experimental results are displayed in Fig. 2c and Table 1, with values in the range 170250170250170-250170 - 250 GPa, consistent with previously reported data.48, 17

αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT has been measured by recording the frequency shift of the thermomechanical noise peak as a function of controlled temperature rises (ΔT=010Δ𝑇010\Delta T=0-10roman_Δ italic_T = 0 - 10 K),50 through the relation αth,SiN=αth,Si2Tσ0/Esubscript𝛼thSiNsubscript𝛼thSi2subscriptTsubscript𝜎0𝐸\alpha_{\mathrm{th,SiN}}=\alpha_{\mathrm{th,Si}}-2\mathcal{R}_{\mathrm{T}}% \sigma_{0}/Eitalic_α start_POSTSUBSCRIPT roman_th , roman_SiN end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT roman_th , roman_Si end_POSTSUBSCRIPT - 2 caligraphic_R start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E. For that, a thermoelectric module (GM200-127-10-15, Adaptive Power Management) has been used to heat up the resonators, while monitoring and keeping the temperature at the desired value via a PID controller (TEC-1092, Meerstetter Engineering). The chips have been enclosed inside a cm-scale copper thermal bath to guarantee thermal equilibrium through radiative heat transfer of the string with the environment (see Fig. 1d). Fig. 2d and Table 1 show the corresponding results, with αthsubscript𝛼th\alpha_{\mathrm{th}}italic_α start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT lying in the range 11.6ppm/K11.6ppmK1-1.6\ \mathrm{ppm/K}1 - 1.6 roman_ppm / roman_K (assuming αth,Si=2.6subscript𝛼thSi2.6\alpha_{\mathrm{th,Si}}=2.6italic_α start_POSTSUBSCRIPT roman_th , roman_Si end_POSTSUBSCRIPT = 2.6 ppm/K), which are consistent with previously reported values.50, 17 No stress dependence has been observed for the two material parameters.

From the absorption measurements, the extinction coefficient for these thin films can be evaluated as 51

κext(λ)=λαabs(λ)4πhη.subscript𝜅ext𝜆𝜆subscript𝛼abs𝜆4𝜋𝜂\kappa_{\mathrm{ext}}(\lambda)=\ \frac{\lambda\ \alpha_{\mathrm{abs}}(\lambda)% }{4\pi\ h\ \eta}.italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ( italic_λ ) = divide start_ARG italic_λ italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( italic_λ ) end_ARG start_ARG 4 italic_π italic_h italic_η end_ARG . (4)

with η𝜂\etaitalic_η denoting a dimensionless factor that accounts for possible interference inside the thin SiN slab.42 For the film thicknesses analyzed here, η11.27𝜂11.27\eta\approx 1-1.27italic_η ≈ 1 - 1.27 at 632.8 nm wavelength (see Table 1 and Supporting Information). Fig. 3a shows the nanomechanical photothermal results of κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT as a function of the resonators’ tensile stress σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (black circles). κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT decreases from 103absentsuperscript103\approx 10^{3}≈ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ppm for the lowest stress, to 101absentsuperscript101\approx 10^{1}≈ 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ppm for the highest. These findings are compared with previously reported values of optical extinction for LPCVD (colored circles), as well as PECVD (colored diamonds), and ECR-CVD (colored squares) deposited SiN films (see Supporting Information for details on their deposition dependencies). The variance in magnitude among the compiled data for σ0850subscript𝜎0850\sigma_{0}\geq 850italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≥ 850 MPa can be partially attributed also to the inability of some of the considered techniques to differentiate between true absorption and scattering losses (in particular cutback and outscattered light 12). Overall, a general trend emerges in Fig. 3a, with κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT decreasing for increasingly higher SiN deposition-related tensile stress.

Refer to caption
Figure 3: (a) κextsubscript𝜅𝑒𝑥𝑡\kappa_{ext}italic_κ start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT for different SiN string’s tensile stresses at an excitation wavelength of λ=(632.8±30)𝜆plus-or-minus632.830\lambda=(632.8\pm 30)italic_λ = ( 632.8 ± 30 ) nm. Characterization techniques included in the figure are: nanomechanical photothermal absorption spectroscopy (NPAS) 42, direct absorption spectroscopy (DAS) in waveguides 52, 53, 54, 55, cavity absorption spectroscopy in microrings resonators (CAS-µring) 12, 56, cutback 27, ellipsometry 19, and prism coupling 33. Markers refer to LPCVD (circles), plasma-enhanced CVD (PECVD, diamonds), and electron-cyclotron resonance CVD (ECR-CVD, squares) deposited SiN films. For the reported values, the vertical lines indicate a relationship with stress σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (intersection with the bottom x-axis) or Si/N (intersection with the top x-axis), explicitly given in (solid lines) or derived from (dashed lines) the original article. When none of these values could be extracted, a stress error bar has been used (σ0=8651365subscript𝜎08651365\sigma_{0}=865-1365italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 865 - 1365 MPa). (b) Absorption coefficient in the band-fluctuations model. The dashed blue and red curves represent the absorption due to electronic transition between extended states (Tauc regime), and absorption due to disorder-induced localized to extended state transitions (Urbach regime), respectively. (c) Energy bandgap Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT as a function of the Si/N ratio. The solid curve is a fitting function of the displayed reported values of the form f(x)=aebx+c𝑓𝑥𝑎superscript𝑒𝑏𝑥𝑐f(x)=ae^{bx}+citalic_f ( italic_x ) = italic_a italic_e start_POSTSUPERSCRIPT italic_b italic_x end_POSTSUPERSCRIPT + italic_c, with a=95.94𝑎95.94a=95.94italic_a = 95.94 eV, b=4.356𝑏4.356b=4.356italic_b = 4.356, and c=1.633𝑐1.633c=1.633italic_c = 1.633 eV. Only LPCDV SiN films have been considered. Compilation: darkcyan, Ref. 57; blue, Ref. 12; purple, Ref. 23; orange, Ref. 21. Dashed vertical lines indicate the Si/N ratios measured in this study with XPS. Intersections with the fitting curve are given in Table 1. (d) Corresponding Urbach energy β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT of the thin films analyzed in this study (black circles). For comparison, data from Ref. 57 (darkcyan) and Ref. 12 (blues) are displayed.

The measurements are analyzed within the framework of the band-fluctuations model,45 which describes the absorption coefficient in units of [dB/m] as a function of the excitation energy ωPlanck-constant-over-2-pi𝜔\hbar\omegaroman_ℏ italic_ω for amorphous materials as

αabs(ω)=α0ω1β2𝒥cv(β(ωEg)),subscript𝛼absPlanck-constant-over-2-pi𝜔subscript𝛼0Planck-constant-over-2-pi𝜔1superscript𝛽2subscript𝒥cv𝛽Planck-constant-over-2-pi𝜔subscript𝐸g\alpha_{\mathrm{abs}}(\hbar\omega)=\ \frac{\alpha_{0}}{\hbar\omega}\frac{1}{% \beta^{2}}\mathcal{J}_{\mathrm{cv}}(\beta(\hbar\omega-E_{\mathrm{g}})),italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ( roman_ℏ italic_ω ) = divide start_ARG italic_α start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_ℏ italic_ω end_ARG divide start_ARG 1 end_ARG start_ARG italic_β start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG caligraphic_J start_POSTSUBSCRIPT roman_cv end_POSTSUBSCRIPT ( italic_β ( roman_ℏ italic_ω - italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT ) ) , (5)

where α0subscript𝛼0\alpha_{0}italic_α start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, β𝛽\betaitalic_β, Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, and 𝒥cvsubscript𝒥cv\mathcal{J}_{\mathrm{cv}}caligraphic_J start_POSTSUBSCRIPT roman_cv end_POSTSUBSCRIPT denote a coefficient collecting physical constants in unit [(meV)1superscriptmeV1\mathrm{(m\ eV)^{-1}}( roman_m roman_eV ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT], the Urbach slope in units [eV1superscripteV1\mathrm{eV}^{-1}roman_eV start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT], the energy bandgap in units [eV], and a dimensionless joint electronic density of states (DOS), respectively. Fig. 3b displays its functional form. For excitation energies ω>EgPlanck-constant-over-2-pi𝜔subscript𝐸g\hbar\omega>E_{\mathrm{g}}roman_ℏ italic_ω > italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, Eq. (5) converges to the Tauc regime 23, where only fundamental electronic transitions between extended states are considered (dashed blue curve); for ω<EgPlanck-constant-over-2-pi𝜔subscript𝐸g\hbar\omega<E_{\mathrm{g}}roman_ℏ italic_ω < italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, the model converges to the empirical Urbach tail, where electronic defect-induced absorption follows αabseβωproportional-tosubscript𝛼abssuperscript𝑒𝛽Planck-constant-over-2-pi𝜔\alpha_{\mathrm{abs}}\propto e^{\beta\hbar\omega}italic_α start_POSTSUBSCRIPT roman_abs end_POSTSUBSCRIPT ∝ italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT (dashed red curve) 45.

The model input parameters Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT and the Urbach energy β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT depend on the film deposition process through the residual tensile stress present in the films. In turn, this dependence is underpinned by the underlying correlation between the stress and the corresponding Si/N ratio, with the former increasing as the latter is reduced (see Table 1), as observed in LPCVD, as well as PECVD and ECR-CVD deposited SiN films 58, 21, 23. Hence, the optical extinction reduction observed in Fig. 3a for increasing tensile stress has to be related to the difference in the chemical composition of the thin films.

In this regard, the Si/N ratio of each chip has been experimentally characterized by X-ray photoelectron spectroscopy (XPS, PHI Versa Probe III-spectrometer) equipped with a monochromatic Al-KαsubscriptK𝛼\mathrm{K_{\alpha}}roman_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT X-ray source and a hemispherical analyser. Data analysis was performed using CASA XPS and Multipak software packages. The results are displayed in Table 1, and are consistent with those reported in previous works for similar tensile stress range.23, 58, 26 These values are also shown in the top x-axis of Fig. 3a, to highlight how SiN extinction increases with Si/N.

Finally, the energy bandgap Egsubscript𝐸gE_{\mathrm{g}}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT of each thin film has been extracted by means of the fitting curve constructed from the compilation of previous works on LPCVD SiN only,57, 12, 21, 23 which are shown in Fig. 3c. The XPS data (dashed vertical lines) are shown for clarity, and fall in the region of strongest dependence on Si/N. The corresponding energy bandgap (Table 1) has been found to increase from 3absent3\approx 3≈ 3 eV, for the highest relative Si concentration, to 4.2absent4.2\approx 4.2≈ 4.2 eV, for the lowest. All these values exceed the probing energy used in this study (ω=1.96eVPlanck-constant-over-2-pi𝜔1.96eV\hbar\omega=1.96\ \mathrm{eV}roman_ℏ italic_ω = 1.96 roman_eV), indicating that the absorption results from localized-to-extended electronic transitions of disorder-induced tail states, as it occurs typically in amorphous semiconductors.45

With the energy bandgap defined for each thin film, the corresponding Urbach energy β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT has been determined by matching the experimental absorption to the band-fluctuations model. The results are shown in Fig. 3d and Table 1, and are consistent with previously reported studies of LPCVD SiN (β1200superscript𝛽1200\beta^{-1}\approx 200italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≈ 200 meV).57, 12 β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT slightly decreases with increasing Si/N ratios, as it has been observed also for PECVD deposited SiN, but at lower values (see the Supporting Information for a comparison) 59, 60. Hence, lowering the Si/N ratio has the main effect of shifting the bandgap Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT to higher energies, broadening the SiN transparency window. Conversely, the Urbach energy β1superscript𝛽1\beta^{-1}italic_β start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT does not vary significantly among these thin films, indicating that the reduction in extinction coefficient κextsubscript𝜅ext\kappa_{\mathrm{ext}}italic_κ start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT is driven by an exponential decrease in the disorder-induced electronic tail DOS at the probing energy of 1.96 eV.

In conclusion, it has been shown that nanomechanical photothermal spectroscopy represents a highly sensitive, simple, and scattering-free platform for the optical characterization of low-loss materials. In this study, its capabilities have been explored using nanostring resonators made of LPCVD deposited SiN. Upon meticulous characterization of their mechanical and thermomechanical properties, it has been shown that SiN intrinsic extinction coefficient decreases with increasingly higher thin film tensile stress. This trend is attributed to a blue-shift in energy bandgap as a function of material composition. Therefore, varying the Si/N ratio provides a degree of freedom to tune the optical properties of SiN, advancing the understanding of this ubiquitous material.

{suppinfo}

The Supporting information is available free of charge.

  • Calculation of the emissivity and factor η𝜂\etaitalic_η for different thicknesses; compilation of data (table); derivation of Si/N ratios for the compiled data; comparison of the measured Urbach energies with reported data for LPCVD and PECVD SiN films.

{acknowledgement}

The authors thank Johannes Hiesberger for the help with the experimental set-up, and Hajrudin Besic, Nicola Cavalleri and Niklas Luhmann for useful discussions. This project received funding from the Novo Nordisk Foundation under project MASMONADE with project number NNF22OC0077964. Moreover, the Austrian Research Promotion Agency (FFG) is gratefully acknowledged for funding of the used XPS infrastructure (FFG project number: 884672).

References

  • Moody et al. 2022 Moody, G. et al. 2022 Roadmap on integrated quantum photonics. J. Phys. Photonics 2022, 4, 012501.
  • Piller et al. 2023 Piller, M.; Hiesberger, J.; Wistrela, E.; Martini, P.; Luhmann, N.; Schmid, S. Thermal IR Detection With Nanoelectromechanical Silicon Nitride Trampoline Resonators. IEEE Sens. J. 2023, 23, 1066–1071.
  • Kirchhof et al. 2023 Kirchhof, J. N.; Yu, Y.; Yagodkin, D.; Stetzuhn, N.; de Araújo, D. B.; Kanellopulos, K.; Manas-Valero, S.; Coronado, E.; van der Zant, H.; Reich, S.; Schmid, S.; Bolotin, K. I. Nanomechanical absorption spectroscopy of 2D materials with femtowatt sensitivity. 2D Mater. 2023, 10, 035012.
  • Kanellopulos et al. 2023 Kanellopulos, K.; West, R. G.; Schmid, S. Nanomechanical Photothermal Near Infrared Spectromicroscopy of Individual Nanorods. ACS Photonics 2023, 10, 3730–3739.
  • West et al. 2023 West, R. G.; Kanellopulos, K.; Schmid, S. Photothermal Microscopy and Spectroscopy with Nanomechanical Resonators. J. Phys. Chem. C 2023, 127, 21915–21929.
  • Wilson 2012 Wilson, D. J. Cavity Optomechanics with High-Stress Silicon Nitride Films; California Institute of Technology, 2012; Ph.D. Dissertation.
  • Reinhardt et al. 2016 Reinhardt, C.; Müller, T.; Bourassa, A.; Sankey, J. C. Ultralow-Noise SiN Trampoline Resonators for Sensing and Optomechanics. Phys. Rev. X 2016, 6, 021001.
  • Gärtner et al. 2018 Gärtner, C.; Moura, J. P.; Haaxman, W.; Norte, R. A.; Gröblacher, S. Integrated Optomechanical Arrays of Two High Reflectivity SiN Membranes. Nano Lett. 2018, 18, 7171–7175.
  • Mason et al. 2019 Mason, D.; Chen, J.; Rossi, M.; Tsaturyan, Y.; Schliesser, A. Continuous force and displacement measurement below the standard quantum limit. Nat. Phys. 2019, 15, 745–749.
  • Huang et al. 2024 Huang, G.; Beccari, A.; Engelsen, N. J.; Kippenberg, T. J. Room-temperature quantum optomechanics using an ultralow noise cavity. Nature 2024, 626, 512–516.
  • Snell et al. 2022 Snell, N.; Zhang, C.; Mu, G.; Bouchard, A.; St-Gelais, R. Heat Transport in Silicon Nitride Drum Resonators and its Influence on Thermal Fluctuation-Induced Frequency Noise. Phys. Rev. Appl. 2022, 17, 044019.
  • Corato-Zanarella et al. 2024 Corato-Zanarella, M.; Ji, X.; Mohanty, A.; Lipson, M. Absorption and scattering limits of silicon nitride integrated photonics in the visible spectrum. Opt. Express 2024, 32, 5718.
  • Metzger et al. 2008 Metzger, C.; Ludwig, M.; Neuenhahn, C.; Ortlieb, A.; Favero, I.; Karrai, K.; Marquardt, F. Self-Induced Oscillations in an Optomechanical System Driven by Bolometric Backaction. Phys. Rev. Lett. 2008, 101, 133903.
  • An et al. 1997 An, K.; Sones, B. A.; Fang-Yen, C.; Dasari, R. R.; Feld, M. S. Optical bistability induced by mirror absorption: measurement of absorption coefficients at the sub-ppm level. Opt. Lett. 1997, 22, 1433.
  • Kanellopulos et al. 2024 Kanellopulos, K.; Ladinig, F.; Emminger, S.; Martini, P.; West, R. G.; Schmid, S. Comparative Analysis of Nanomechanical Resonators: Sensitivity, Response Time, and Practical Considerations in Photothermal Sensing. 2024.
  • Ji et al. 2023 Ji, X.; Okawachi, Y.; Gil-Molina, A.; Corato-Zanarella, M.; Roberts, S.; Gaeta, A. L.; Lipson, M. Ultra-Low-Loss Silicon Nitride Photonics Based on Deposited Films Compatible with Foundries. Laser Photonics Rev. 2023, 17, 2200544.
  • Kaloyeros et al. 2020 Kaloyeros, A. E.; Pan, Y.; Goff, J.; Arkles, B. Review—Silicon Nitride and Silicon Nitride-Rich Thin Film Technologies: State-of-the-Art Processing Technologies, Properties, and Applications. ECS J. Solid State Sci. Technol. 2020, 9, 063006.
  • Makino 1983 Makino, T. Composition and Structure Control by Source Gas Ratio in LPCVD SiNx. J. Electrochem. Soc. 1983, 130, 450.
  • Poenar and Wolffenbuttel 1997 Poenar, D. P.; Wolffenbuttel, R. F. Optical properties of thin-film silicon-compatible materials. Appl. Opt. 1997, 36, 5122.
  • Lin et al. 2013 Lin, P. T.; Singh, V.; Lin, H. Y. G.; Tiwald, T.; Kimerling, L. C.; Agarwal, A. M. Low-stress silicon nitride platform for mid-infrared broadband and monolithically integrated microphotonics. Adv. Opt. Mater. 2013, 1, 732–739.
  • Krückel et al. 2017 Krückel, C. J.; Fülöp, A.; Ye, Z.; Andrekson, P. A.; Torres-Company, V. Optical bandgap engineering in nonlinear silicon nitride waveguides. Opt. Express 2017, 25, 15370.
  • Resende et al. 2021 Resende, J.; Fuard, D.; Le Cunff, D.; Tortai, J.-H.; Pelissier, B. Hybridization of ellipsometry and energy loss spectra from XPS for bandgap and optical constants determination in SiON thin films. Mater. Chem. Phys. 2021, 259, 124000.
  • Beliaev et al. 2022 Beliaev, L. Y.; Shkondin, E.; Lavrinenko, A. V.; Takayama, O. Optical, structural and composition properties of silicon nitride films deposited by reactive radio-frequency sputtering, low pressure and plasma-enhanced chemical vapor deposition. Thin Solid Films 2022, 763, 139568.
  • Philipp 1973 Philipp, H. R. Optical Properties of Silicon Nitride. J. Electrochem. Soc. 1973, 120, 295.
  • Cataldo et al. 2012 Cataldo, G.; Beall, J. A.; Cho, H.-M.; McAndrew, B.; Niemack, M. D.; Wollack, E. J. Infrared dielectric properties of low-stress silicon nitride. Opt. Lett. 2012, 37, 4200–4202.
  • Yang and Pham 2018 Yang, C.; Pham, J. Characteristic Study of Silicon Nitride Films Deposited by LPCVD and PECVD. Silicon 2018, 10, 2561–2567.
  • Sorace-Agaskar et al. 2019 Sorace-Agaskar, C.; Kharas, D.; Yegnanarayanan, S.; Maxson, R. T.; West, G. N.; Loh, W.; Bramhavar, S.; Ram, R. J.; Chiaverini, J.; Sage, J.; Juodawlkis, P. Versatile Silicon Nitride and Alumina Integrated Photonic Platforms for the Ultraviolet to Short-Wave Infrared. IEEE J. Sel. Top. Quantum Electron. 2019, 25, 1–15.
  • Smith et al. 2023 Smith, J. A.; Francis, H.; Navickaite, G.; Strain, M. J. SiN foundry platform for high performance visible light integrated photonics. Opt. Mater. Express 2023, 13, 458.
  • Blasco-Solvas et al. 2024 Blasco-Solvas, M.; Fernández-Vior, B.; Sabek, J.; Fernández-Gavela, A.; Domínguez-Bucio, T.; Gardes, F. Y.; Domínguez-Horna, C.; Faneca, J. Silicon Nitride Building Blocks in the Visible Range of the Spectrum. J. Light. Technol. 2024, 1–10.
  • Lelit et al. 2022 Lelit, M.; Słowikowski, M.; Filipiak, M.; Juchniewicz, M.; Stonio, B.; Michalak, B.; Pavłov, K.; Myśliwiec, M.; Wiśniewski, P.; Kaźmierczak, A.; Anders, K.; Stopiński, S.; Beck, R. B.; Piramidowicz, R. Passive Photonic Integrated Circuits Elements Fabricated on a Silicon Nitride Platform. Materials 2022, 15.
  • Mashayekh et al. 2021 Mashayekh, A. T.; Klos, T.; Geuzebroek, D.; Klein, E.; Veenstra, T.; Büscher, M.; Merget, F.; Leisching, P.; Witzens, J. Silicon nitride PIC-based multi-color laser engines for life science applications. Opt. Express 2021, 29, 8635.
  • Gorin et al. 2008 Gorin, A.; Jaouad, A.; Grondin, E.; Aimez, V.; Charette, P. Fabrication of silicon nitride waveguides for visible-light using PECVD: a study of the effect of plasma frequency on optical properties. Opt. Express 2008, 16, 13509–13516.
  • Bonneville et al. 2021 Bonneville, D. B.; Miller, J. W.; Smyth, C.; Mascher, P.; Bradley, J. D. Low-temperature and low-pressure silicon nitride deposition by ecr-pecvd for optical waveguides. Appl. Sci. (Switzerland) 2021, 11, 1–11.
  • Museur et al. 2016 Museur, L.; Zerr, A.; Kanaev, A. Photoluminescence and electronic transitions in cubic silicon nitride. Sci. Rep. 2016, 6, 18523.
  • Steinlechner et al. 2017 Steinlechner, J.; Krüger, C.; Martin, I. W.; Bell, A.; Hough, J.; Kaufer, H.; Rowan, S.; Schnabel, R.; Steinlechner, S. Optical absorption of silicon nitride membranes at 1064 nm and at 1550 nm. Phys. Rev. D 2017, 96, 022007.
  • Stambaugh et al. 2014 Stambaugh, C.; Durand, M.; Kemiktarak, U.; Lawall, J. Cavity-enhanced measurements for determining dielectric-membrane thickness and complex index of refraction. Appl. Opt. 2014, 53, 4930.
  • Yamada et al. 2013 Yamada, S.; Schmid, S.; Larsen, T.; Hansen, O.; Boisen, A. Photothermal infrared spectroscopy of airborne samples with mechanical string resonators. Anal. Chem. 2013, 85, 10531–10535.
  • Bose et al. 2014 Bose, S.; Schmid, S.; Larsen, T.; Keller, S. S.; Sommer-Larsen, P.; Boisen, A.; Almdal, K. Micromechanical string resonators: Analytical tool for thermal characterization of polymers. ACS Macro Lett. 2014, 3, 55–58.
  • Andersen et al. 2016 Andersen, A. J.; Yamada, S.; Pramodkumar, E. K.; Andresen, T. L.; Boisen, A.; Schmid, S. Nanomechanical IR spectroscopy for fast analysis of liquid-dispersed engineered nanomaterials. Sens. Actuators B Chem. 2016, 233, 667–673.
  • Kurek et al. 2017 Kurek, M.; Carnoy, M.; Larsen, P. E.; Nielsen, L. H.; Hansen, O.; Rades, T.; Schmid, S.; Boisen, A. Nanomechanical Infrared Spectroscopy with Vibrating Filters for Pharmaceutical Analysis. Angew. Chem., Int. Ed. 2017, 56, 3901–3905.
  • Luhmann et al. 2023 Luhmann, N.; West, R. G.; Lafleur, J. P.; Schmid, S. Nanoelectromechanical Infrared Spectroscopy with In Situ Separation by Thermal Desorption: NEMS-IR-TD. ACS Sens. 2023, 8, 1462–1470.
  • Land et al. 0 Land, A. T.; Dey Chowdhury, M.; Agrawal, A. R.; Wilson, D. J. Sub-ppm Nanomechanical Absorption Spectroscopy of Silicon Nitride. Nano Lett. 0, 0, null, PMID: 38742810.
  • Sementilli et al. 2022 Sementilli, L.; Romero, E.; Bowen, W. P. Nanomechanical Dissipation and Strain Engineering. Adv. Funct. Mater. 2022, 32, 2105247.
  • Engelsen et al. 2024 Engelsen, N. J.; Beccari, A.; Kippenberg, T. J. Ultrahigh-quality-factor micro- and nanomechanical resonators using dissipation dilution. Nat. Nanotechnol. 2024,
  • Guerra et al. 2019 Guerra, J. A.; Tejada, A.; Töfflinger, J. A.; Grieseler, R.; Korte, L. Band-fluctuations model for the fundamental absorption of crystalline and amorphous semiconductors: A dimensionless joint density of states analysis. J. Phys. D: Appl. Phys. 2019, 52, 105303.
  • Sadeghi et al. 2020 Sadeghi, P.; Tanzer, M.; Luhmann, N.; Piller, M.; Chien, M. H.; Schmid, S. Thermal Transport and Frequency Response of Localized Modes on Low-Stress Nanomechanical Silicon Nitride Drums Featuring a Phononic-Band-Gap Structure. Phys. Rev. Appl. 2020, 14, 024068.
  • Schmid et al. 2023 Schmid, S.; Villanueva, L. G.; Roukes, M. L. Fundamentals of Nanomechanical Resonators; Springer International Publishing, 2023.
  • Klaß et al. 2022 Klaß, Y. S.; Doster, J.; Bückle, M.; Braive, R.; Weig, E. M. Determining Young’s modulus via the eigenmode spectrum of a nanomechanical string resonator. Appl. Phys. Lett. 2022, 121, 083501.
  • Bergman et al. 2017 Bergman, T. L.; Lavine, A.; Incropera, F. P. Fundamentals of heat and mass transfer.; John Wiley, 2017; p 966.
  • Larsen et al. 2011 Larsen, T.; Schmid, S.; Grönberg, L.; Niskanen, A. O.; Hassel, J.; Dohn, S.; Boisen, A. Ultrasensitive string-based temperature sensors. Appl. Phys. Lett. 2011, 98, 121901.
  • Macleod 2010 Macleod, H. A. Thin-Film Optical Filters; CRC Press, 2010.
  • Inukai and Ono 1994 Inukai, T. I. T.; Ono, K. O. K. Optical Characteristics of Amorphous Silicon Nitride Thin Films Prepared by Electron Cyclotron Resonance Plasma Chemical Vapor Deposition. Jpn. J. Appl. Phys. 1994, 33, 2593.
  • Bulla et al. 1999 Bulla, D.; Borges, B.; Romero, M.; Morimoto, N.; Neto, L.; Cortes, A. Design and fabrication of SiO/sub 2//Si/sub 3/N/sub 4/ CVD optical waveguides. 1999 SBMO/IEEE MTT-S International Microwave and Optoelectronics Conference. 1999; pp 454–457 vol. 2.
  • Daldosso et al. 2004 Daldosso, N.; Melchiorri, M.; Riboli, F.; Sbrana, F.; Pavesi, L.; Pucker, G.; Kompocholis, C.; Crivellari, M.; Bellutti, P.; Lui, A. Fabrication and optical characterization of thin two-dimensional Si 3N4 waveguides. Mater. Sci. Semicond. Process. 2004, 7, 453–458.
  • Sacher et al. 2019 Sacher, W. D.; Luo, X.; Yang, Y.; Chen, F.-D.; Lordello, T.; Mak, J. C. C.; Liu, X.; Hu, T.; Xue, T.; Lo, P. G.-Q.; Roukes, M. L.; Poon, J. K. S. Visible-light silicon nitride waveguide devices and implantable neurophotonic probes on thinned 200 mm silicon wafers. Opt. Express 2019, 27, 37400.
  • Worhoff et al. 2008 Worhoff, K.; Klein, E.; Hussein, G.; Driessen, A. Silicon oxynitride based photonics. 2008 10th Anniversary International Conference on Transparent Optical Networks. 2008; pp 266–269.
  • Bauer 1977 Bauer, J. Optical properties, band gap, and surface roughness of Si3N4. Phys. Status Solidi A 1977, 39, 411–418.
  • Temple-Boyer et al. 1998 Temple-Boyer, P.; Rossi, C.; Saint-Etienne, E.; Scheid, E. Residual stress in low pressure chemical vapor deposition SiNx films deposited from silane and ammonia. J. Vac. Sci. Technol. A 1998, 16, 2003–2007.
  • Garcia et al. 1995 Garcia, S.; Bravo, D.; Fernandez, M.; Martil, I.; López, F. J. Role of oxygen on the dangling bond configuration of low oxygen content SiNx:H films deposited at room temperature. Appl. Phys. Lett. 1995, 67, 3263.
  • Kato et al. 2003 Kato, H.; Kashio, N.; Ohki, Y.; Seol, K. S.; Noma, T. Band-tail photoluminescence in hydrogenated amorphous silicon oxynitride and silicon nitride films. J. Appl. Phys. 2003, 93, 239–244.