Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Metasurface-assisted amorphous germanium-tin waveguide bolometer for mid-infrared photodetection

Open Access Open Access

Abstract

An amorphous germanium-tin (a-Ge0.83Sn0.17) waveguide bolometer featuring a one-dimension (1D) metasurface absorber is proposed for mid-infrared photodetection at room-temperature. The device is based on the germanium-on-silicon (GOS) photonic platform. The impacts of the 1D metasurface on the performances of the waveguide bolometer are investigated. The responsivity of the a-Ge0.83Sn0.17 waveguide bolometer could be significantly enhanced by the metasurface. A responsivity of around −3.17%/µW within the 4.1 ∼ 4.3 µm wavelength range is achieved. In addition, a 3-dB roll-off frequency higher than 10 kHz is obtained.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Mid-infrared (MIR) integrated optical sensors have received widespread attentions because many gas molecules exhibit characteristic rotational or vibrational transition bands in the MIR spectrum (3 ∼ 14 µm) [1]. Waveguide photodetector is a key component for an integrated MIR gas sensor. The MIR photodetector can be classified into two categories: photon detectors and thermal detectors based on the working principle [2]. Compared with photon detectors, thermal detectors can operate at room temperature, thus they do not require the expensive and bulky cryogenic equipment. One of the most widely used thermal detectors is bolometer, which generally comprises an absorber and a thermistor. The light sensing capability of the bolometer is achieved by detecting the resistance variation due to opto-thermal conversion [35]. Remarkable progress has been made in the development of waveguide bolometers on silicon-on-insulator (SOI) over the last decade [613]. Nonetheless, the transparent window of SOI platform limits the detection of carbon dioxide (CO2) and other mid-infrared gases at the wavelengths beyond 4 µm. The germanium-on-silicon (GOS) platform has a transparent window from 3 to 14 µm [1417]. Therefore, bolometers based on GOS waveguide platform is a potential solution of MIR integrated gas sensors.

One of the figures of merit for a bolometer is responsivity, which defined as the change in the resistance value per unit incident light power. There are two feasible methods to improve a bolometer’s responsivity. The first approach is to utilize thermal sensitive materials with high temperature coefficient of resistance (TCR). Amorphous silicon (a-Si) and vanadium oxide (VOx) are commonly used as the thermal sensitive materials of thermistors. Though the intrinsic a-Si has a high TCR of −10%/K, the high electrical resistivity (∼109 Ω·cm) has set a barrier to measure the resistance changes by read-out integrated circuits. The resistivity of a-Si can be suppressed by doping. However, its TCR is reduced accordingly to a much lower value of −2.8%/K [18]. The other material VOx has relatively low compatibility with the complementary metal-oxide-semiconductor (CMOS) process. Amorphous germanium tin (a-Ge1-ySny) is a promising candidate material for bolometers, as it has a high TCR, low resistivity and is CMOS-compatible. When the Sn concentration is at 17%, the a-Ge0.83Sn0.17 has a high TCR of −3.96%/K and relatively low resistivity of 164 Ω·cm [19,20]. The second approach is to implement the resonance structures on the waveguide to enhance light absorption efficiency. When light frequency is matched with the resonance frequency of the metasurface, the local surface plasmon resonance (LSPR) effect greatly enhances the electric field. This increases light absorption which finally enhances the photo-thermal conversion efficiency.

In this work, a room-temperature a-Ge0.83Sn0.17 waveguide bolometer on the GOS platform with high photodetection responsivity working at mid-infrared wavelength is proposed. This device features with a 1D tungsten (W) metasurface on top of the a-Ge0.83Sn0.17 waveguide. The metasurface induced LSPR effect results in strong light absorption, based on which the responsivity of the bolometer has been significantly increased. An absorption efficiency of nearly 93% at λ = 4.2 µm is achieved, and the responsivity is improved to −3.17%/µW. In addition, a 3-dB roll-off frequency of beyond 10 kHz is obtained.

2. Device design

Figure 1(a) is a three-dimensional (3D) schematic diagram of an a-Ge0.83Sn0.17 waveguide bolometer on the GOS platform. The bolometer consists of a 1D metasurface light absorber, a suspended a-Ge0.83Sn0.17 waveguide and an a-Ge0.83Sn0.17 thermistor. Within the 1D metasurface, the meta-atoms have a radius r = 180 nm, thickness t = 20 nm and pitch p = 400 nm, as shown in the inset of Fig. 1(a). The meta-atom length (lm) varies from 300 to 400 nm to investigate the lm impact on the bolometer’s performance. Tungsten is adopted as the metasurface material due to its corrosion resistance, high melting point and CMOS compatibility. All the meta-atoms are arranged uniformly on the suspended a-Ge0.83Sn0.17 waveguide with width wGeSn = 2.5 µm, thickness tGeSn = 0.5 µm and length lw = 15 µm. At the end of the waveguide, there is a thermistor with length lth = 5 µm. There is a gap lg = 1 µm between the metasurface and the thermistor. The bolometer is built on the GOS waveguide, which enables the photodetection beyond λ = 4 µm as compared to those on the SOI platform. The Ge waveguide has 2 µm width (wGe) and 1.35 µm thickness (tGe). A bilevel taper coupler with a tip width wtip = 800 nm and coupling length lc = 40 µm is utilized between the Ge and the a-Ge0.83Sn0.17 waveguides. The tapered structure is designed on both a-Ge0.93Sn0.17 and Ge waveguides. Similar structures have been reported in Ref. [21]. Such bilevel taper coupler can be used to improve the light coupling efficiency between the Ge and a-Ge0.93Sn0.17 waveguides [22]. In addition, it can reduce the multimode effects to an almost negligible level [23]. We have designed two types of waveguide bolometers: Device Ai (the subscript i represents the number of meta-atoms) and Device Bi. The a-Ge0.83Sn0.17 thermistor of Device Ai is suspended, whereas the one of Device Bi is on top of SiO2, as shown in Fig. 1(b). The responsivity as well as response speed of the two types of bolometers are investigated by numerical simulation.

 figure: Fig. 1.

Fig. 1. (a) 3D schematic of the 1D metasurface assisted a-Ge0.83Sn0.17 waveguide bolometer (Device Ai). (b) The top view and the right-side view for Device Ai and Device Bi.

Download Full Size | PDF

The electromagnetic (EM) field distribution of the GOS waveguide at λ = 4.2 µm is obtained by numerical simulation using the commercial software Ansys Lumerical. The mode profile of the fundamental transverse electric mode (TE0) of the GOS waveguide is depicted in Fig. 2(a), which demonstrates a decent light confinement within the waveguide. To couple the light from the Ge waveguide into the a-Ge0.83Sn0.17 waveguide, a bilayer taper coupler is designed by finite-difference time domain (FDTD) method. The refractive index n of a-Ge0.83Sn0.17 is obtained by a linear interpolation method based on the refractive indices of c-Ge1-ySny [24]. In Figs. 2(b) and 2(c), the n value of a-Ge0.83Sn0.17 is 4.28 and k value is assumed to be 0. The cross-sectional electric field (E) profile of the coupler at λ = 4.2 µm is shown in Fig. 2(b), and a power coupling efficiency of 80% is achieved. With a proper design of the W meta-atoms, the LSPR could be excited by the EM field traveling in the a-Ge0.83Sn0.17 waveguide, as depicted in Fig. 2(c). Owing to the ohmic losses in the metal metasurface, the incident EM energy is effectively converted into heat [25]. After that, the heat is conducted to the a-Ge0.83Sn0.17 waveguide, which results in a temperature change of the a-Ge0.83Sn0.17 thermistor. This leads to the change of the electrical resistance of the thermistor, which is connected in a bridge configuration between two contact pads. Thus, the measurement of resistance change could be conducted by applying voltage to the pads and measuring the current change.

 figure: Fig. 2.

Fig. 2. (a) TE0 mode distribution of the Ge waveguide at 4.2 µm wavelength. (b) X-Z cross-sectional view of the E-field at the bilevel coupler region. (c) Near-field E-field profile of 10 meta-atoms with r = 180 nm, t = 20 nm, p = 400 nm at λ = 4.2 µm.

Download Full Size | PDF

3. Results and discussion

3.1 Design of the meta-atoms

To analyze light absorption (A) of the bolometer, 3D simulation is carried out by FDTD method. The n of W at mid-infrared wavelengths are obtained from Ref. [26]. It should be noted that the n values of a-Ge1-ySny at mid-infrared are seldomly reported. According to experimental data from Ref. [27], the n values of a-Ge1-ySny are close to those of crystalline c-Ge1-ySny. In this work, therefore, the refractive indices of a-Ge0.83Sn0.17 is approximated using a linear interpolation method based on the refractive indices of c-Ge1-ySny at mid-infrared [24]. The absorption coefficient (α) of a-Ge0.83Sn0.17 near λ = 4.2 µm is another important parameter in device simulation. Although it has been shown experimentally that the absorption cut-off of c-Ge0.83Sn0.17 is about 4.13 µm [28], unfortunately, the experimental α of a-Ge0.83Sn0.17 near 4.2 µm wavelength has not been reported yet. One should be noted that the absorption cutoff of a-Ge might be shorter [29] or longer [30] than that of c-Ge. Analogous to a-Ge0.83Sn0.17, it might be ambiguous if the α of a-Ge0.83Sn0.17 near λ = 4.2 µm is negligible or not. In this work, therefore, we have simulated and compared the waveguide bolometer performances with various α of a-Ge0.83Sn0.17, i.e. 0, 10, 100, and 1000 cm−1.

To achieve high absorption, the dipole oscillation frequency in the metasurface should be designed to match the input light frequency ωl. The dipole oscillation frequency is determined by the geometry of the metasurface. Here, the α of a-Ge0.83Sn0.17 is set as 0 cm−1. The lw and lg are fixed at 15 and 1 µm, respectively. Figure 3(a) shows the energy fraction results at λ = 4.2 µm as the number of meta-atoms increases from 1 to 20. The absorption A increases as more meta-atoms are added, exceeding 90% when the number of meta-atoms is larger than 10. Meanwhile, the scattering (S), reflection (R) and transmission (T) are decreased. The enhancement of A origins from the excited collective oscillation mode of the W metasurface [31], which results in a strong confinement of light in meta-atoms. Due to the high absorption of metal W, the confinement of light leads to the strong absorption. The absorption rates A as a function of lm at 4.2 µm wavelength are extracted and presented in Fig. 3(b). The A increases at first, continuously comes to a “peak value” and then decreases with the increase of lm. In general, the strength of the E-field in metasurface affects the absorption rate. Thus, the variation trend of the field enhancement factor |E/E0| (E0 is the source E-field amplitude in numerical simulations [32,33]) with meta-atom length is consistent with the absorption, as shown in the inset of Fig. 3(b). The field enhancement factor reaches 7.5 at lm = 350 nm, while an absorption rate of 93% is achieved.

 figure: Fig. 3.

Fig. 3. (a) Energy fraction of A, T, R, S as a function of the number of meta-atoms at 4.2 µm wavelength. The lw is 15 µm and the lg is 1 µm. (b) Absorption of meta-atoms with different lengths lm at 4.2 µm wavelength. The inset shows the field enhancement factor with different lengths lm. (c) The absorption spectra by 3D FDTD simulation (solid lines) and by TCMT curve fitting (dashed lines) in frequency domains. A decent matching between the TCMT results and the FDTD curves can be observed. (d) Extracted values of η and −γ/(γ2 + β2) for different devices.

Download Full Size | PDF

Figure 3(c) shows the absorption spectra for lengths lm ranging from 300 nm to 400 nm, while keeping the r, t and p fixed at 180, 20 and 400 nm, respectively. Since the dipole moments in the two ends of the meta-atoms are in opposite direction, increasing the lm weakens the restoring force between them, resulting in a red-shift of the resonance peak [34]. It is noteworthy that the resonant peak of lm = 360 nm is at the wavelength around 4.2 ∼ 4.3 µm, which is corresponded to the absorption peak of CO2.

To gain the underlying mechanism of the absorption, a fast and less compute-intensive temporal coupled-mode theory (TCMT) model is adopted, the absorption can be obtained by [32,33]:

$$A = 1 - T - R - S \cong 1 - \frac{{{{|{{s_{2 - }}} |}^2}}}{{{{|{{s_{1 + }}} |}^2}}} = 1 - [{|{{t_0}} |^2} + \frac{{ - 2\beta \tau _{te}^{ - 1}({\omega _l} - {\omega _0}) + \tau _{te}^{ - 2} + 2\gamma {\tau ^{ - 1}}\tau _{te}^{ - 1}}}{{{{({\omega _l} - {\omega _0})}^2} + {\tau ^{ - 2}}}}],$$
where ωl is the frequency of input light, ω0 is the resonance frequency of the resonator, 1/τ is the total damping rate of resonance cavity, 1/τte presents the total external damping rate, γ is the broadband damping and β is the phase shift. The resonance perturbation coefficient t0 = γ + βj are caused by the non-resonant interaction of the metasurface. All of those parameter values can be obtained by fitting the absorption curves in Fig. 3(c). The model in this work is formed by a single resonant cavity with three light ports, including two output waveguide (s1-, s2-) and one input waveguide (s1+). The absorption A is obtained by A = 1 − TRS. The R and S can be ignored according to Fig. 3(a). The TCMT results agree well with the FDTD simulations, as shown in Fig. 3(c). When lm = 350 nm, the difference between η (defined as η = τ/τte) and −γ/(γ2 + β2) is minimal as shown in Fig. 3(d). This indicates that the maximum absorption of the 1D metasurface occurs when lm = 350 nm [32], which is consistent with the results in Fig. 3(c). Noting that the non-resonant metallic absorption (1 − |t0|2) is typically 30% [32,33], it becomes evident that nearly all energy from the a-Ge0.83Sn0.17 waveguide mode is coupled into the metasurface. This strong coupling efficiency in conjunction with the high field intensity within the metasurface provides a substantial signal amplification for sensing applications.

Next, the effects of various a-Ge0.83Sn0.17 absorption coefficients αGeSn on the bolometer absorption are studied, as shown in Table 1. The entire FDTD absorption simulation region includes the a-Ge0.17Sn0.83 waveguide and the metasurface. The wavelength is set as 4.2 µm. The 1D metasurface consists of 10 meta-atoms. The lm of meta-atoms is 360 nm. The AGeSn is the absorption rate of a-Ge0.83Sn0.17. The total absorption Atotal represents the sum of the absorption of the waveguide and the metasurface. The parameter Am, which equals to (Atotal -AGeSn)/Atotal, is the ratio of the absorption rate of the metasurface to the total absorption. In Table 1, the AGeSn reaches 86.2% when αGeSn is 1000 cm−1, which corresponds to an Am of 3.1%. It should be noted that a high αGeSn may lead to large optical loss within the bilayer taper as well as the a-Ge0.17Sn0.83 waveguide. This inevitably degrades the opto-thermal energy conversion efficiency. Therefore, the bolometer designed in this work is more suitable at the wavelengths where the a-Ge1−ySny has low absorption coefficient.

Tables Icon

Table 1. Absorption of the bolometers in the FDTD simulation

3.2 Thermal simulation of bolometers

Simulations based on the Ansys Lumerical Heat Transport (HEAT) solver are carried out to investigate the thermal and electrical properties of the bolometer. In the thermal simulation, the setting of the heat source is based on the FDTD absorption simulation results. The lm of meta-atoms is 360 nm. The input light power is set as 5 µW. The boundary temperature is fixed at room temperature of 293.15 K. A constant convection boundary condition is added between the surface of the bolometer and air. The convective heat flux coefficient of air is 5 W/(m2·K). In addition, the voltage boundary conditions are applied to the contact pads. The thermal properties of different materials are summarized in Table 2. In this table, the specific heat and density of a-Ge0.83Sn0.17 is obtained by linearly interpolation method based on the value of Ge [35] and Sn [36]. Besides, the thermal conductivity of a-Ge0.83Sn0.17 is obtained through an empirical formula in Ref. [37].

Tables Icon

Table 2. Thermal properties of the materials in device simulation

The temperature change in the bolometer can be described as [10]:

$$\Delta T = \frac{{A{\varphi _0}}}{{{{(G_{th}^2 + {f^2}C_{th}^2)}^{1/2}}}},$$
where ΔT is the average temperature change of the thermistor, φ0 and f are the optical power amplitude and the modulating frequency of incident radiation, respectively. In addition, the Cth is the thermal capacitance and Gth is the thermal conductance. The formulation presents that if the other parameters are fixed, the lower the values of Gth and Cth, the larger the temperature change. Moreover, the thermal resistance Rth (Rth = 1/Gth) is determined by Gth. A low Rth together with a low heat capacity (Cth) will lead to a small thermal time constant (τ = Cth/Gth) and thus a fast device response speed.

Temperature distributions of the bolometers with different number of meta-atoms (Device A5, A10, A15), and device structures (Device A10, B10) are compared in Fig. 4. The cross-sectional temperature distributions of the thermistor along the orange dashed lines in Figs. 4(a)-(d) are described in Figs. 4(e)-(h), respectively. From Fig. 4(c), the heat is mainly concentrated in the a-Ge0.83Sn0.17 waveguide due to its low Gth. This leads to a maximum temperature Tmax above 305.75 K within this region. The heat is then conducted to both sides along the a-Ge0.83Sn0.17 waveguide. The a-Ge0.83Sn0.17 thermistor experiences a maximum temperature of 301.97 K, as shown in Fig. 4(g). However, the Tmax in the metasurface region of Device A10 is reduced to 304.27 K due to the relatively low absorption rate A of the W metasurface. Correspondingly, the T'max of its thermistor is reduced to 301.14 K, as shown in Fig. 4(f). Similarly, the Tmax and the T'max in Device A5 are 298.71 and 297.25 K, respectively. Furthermore, the high thermal conductivity of the SiO2 beneath the a-Ge0.83Sn0.17 thermistor results in a smaller temperature change. As can be seen from Fig. 4(d) that the Tmax of Device B10 is only 300.29 K, which is lower than that of Device A10.

 figure: Fig. 4.

Fig. 4. Top view (a), (b), (c), (d) and cross-sectional view (e), (f), (g), (h) of the temperature distribution for Device A5/A10/A15/B10. The Pin is 5 µW and the λ is 4.2 µm.

Download Full Size | PDF

The current-voltage (I - V) characteristics of Device A10 are simulated using finite element method with various input power Pin ranging from 0 to 10 µW, as shown in Fig. 5(a). It can be seen that a higher input light power leads to a larger slope of the I - V curve. Moreover, the resistance response (ΔR/R0, where R0 is the resistance without light illumination) for the four types of bolometers are shown in Fig. 5(b). It should be noted that the nonlinear change process of Gth and Cth with temperature can be approximated by a linear model, when the temperature range is small [10]. Based on this concept, Device A10 and Device A15 shows a similar responsivity of around −3.17%/µW, which is higher than those of the other bolometers. The metasurfaces of Device A10 and A15 can absorb almost all of the incident light power, resulting in the generation of more heat, which leads to a significant resistance change of the thermistor. Due to the presence of SiO2 under the thermistor which increases the thermal conductance, the responsivity of Device B10 is degraded to −0.645%/µW. In addition, the responsivity of Device A10 is analyzed at the wavelength range from 4.1 to 4.3 µm with Pin = 5 µW, as plotted in Fig. 5(c). The nearly flat responsivity spectrum indicates a relatively broadband property of the metasurface-enhanced bolometer, which is beneficial for the detection of CO2 gas molecules with the fingerprint absorption peak at around λ = 4.2 µm.

 figure: Fig. 5.

Fig. 5. (a) I - V curves of the Device A10 with various Pin from 0 to 10 µW. The ΔT represents the average temperature change in the thermistor. (b) The resistance response as a function of Pin for each bolometer. (c) Wavelength-dependence of the responsivity for waveguide bolometer A10.

Download Full Size | PDF

The impacts of αGeSn on the bolometer responsivity are investigated, as summarized in Table 3. The responsivity of the bolometer with metasurface (Device A10) decreases as the αGeSn increases, due to the fact that a-Ge1-ySny waveguide with high αGeSn absorbs more EM energy, resulting in a lower opto-thermal conversion efficiency. In addition, the responsivity of the bolometer without metasurface is −0.042%/µW when the αGeSn = 1000 cm−1. It can be seen from Table 3 that a maximum responsivity of −3.13%/µW can be achieved when αGeSn = 0 cm−1. Thus, reducing the αGeSn is one of the feasible ways to achieve a high responsivity waveguide bolometer.

Tables Icon

Table 3. Responsivities of the Device A10 with different αGeSn

Furthermore, the Gth of each device is calculated according to Eq. (2). The absorption rates A are 36.7%, 71.6%, 82.0%, and 71.6% for Devices A5, A10, A15, B10, respectively. The temperature responsivities (ΔT/ΔPin) are 1.186, 1.974, 2.249 and 0.455 K/µW for the above-mentioned devices. In a steady state, the frequency-dependent term (${f^2}C_{th}^2$) can be ignored. Therefore, Gth is calculated to be 3.094 × 10−7, 3.627 × 10−7, 3.646 × 10−7 and 1.560 × 10−6 W/K for Device A5, A10, A15, B10, respectively.

Another critical metric of a bolometer is the light response speed. The temporal response of waveguide bolometers at λ = 4.2 µm, Pin = 5 µW and modulation frequency f = 1 kHz is studied, as shown in Fig. 6(a). The rise response time trise is defined as the time needed to increase the temperature from 10% to 90% of its maximum, and vice versa for the fall response time tfall. It is observed that trise (tfall) becomes larger when the number of meta-atoms i is increased from 5 to 15. This could be explained by the increasing of Cth resulting from the larger number of meta-atoms. The normalized temperature change as a function of f is shown in Fig. 6(b). The 3 dB roll-off frequency (f3 dB) of Device A5, A10, A15 and B10 are 16, 14, 12 and 25 kHz, respectively. It is noted that f3 dB of Device B10 is higher than those of Device Ai in this work. The fast response speed of Device B10 could be attributed to its higher thermal conductance as the thermal conductivity of SiO2 is much larger than that of air. With further optimizing the device design, e.g., by decreasing the thermal capacity, the f3 dB of the bolometer could be significantly improved, which is the subject of our future works.

 figure: Fig. 6.

Fig. 6. (a) Temporal responses of ΔT as a function of time t. The optical power and bias voltage are 5 µW and 1 V, respectively. (b) Normalized temperature change with respect to the modulating frequency f of the heat power in the simulation.

Download Full Size | PDF

In addition, a possible fabrication process for the proposed waveguide bolometer is discussed. The starting substrate for the a-Ge0.93Sn0.17 waveguide bolometer could be a commercially-available GOS wafer. The Ge waveguide could be formed by electron beam lithography (EBL) and inductively coupled plasma (ICP) etching. After that, the SiO2 film could be deposited by plasma enhanced chemical vapor deposition (PECVD). To expose the surface of Ge waveguide, the chemical-mechanical polishing (CMP) could be used to polish the SiO2 film. Then, the a-Ge0.93Sn0.17 waveguide could be formed by magnetron sputtering, EBL and ICP etching. After that, the metal contact pads and 1D metasurface could be formed by lift-off process. Finally, the suspended a-Ge0.93Sn0.17 structure could be formed by selective removal of SiO2, possibly using the diluted hydrofluoric (DHF) chemical or vapor hydrofluoric (VHF) etching process.

There are a few literatures regarding the design of Si-based waveguide bolometer in infrared region. The comparison between our device and the reported waveguide bolometers are summarized in Table 4. As a proof of concept, our designed waveguide bolometer exhibits an improved responsivity in comparison to the previously reported devices. This could be attributed to the large TCR and low thermal conductance of the suspended a-Ge0.83Sn0.17 waveguide, as well as the enhanced absorption from the 1D metasurface. In addition, a decent 3 dB bandwidth of beyond 10 kHz can be achieved, which should be enough for a variety of sensor applications.

Tables Icon

Table 4. Benchmark between the reported Si-based waveguide bolometers

4. Conclusion

In summary, we have proposed a mid-infrared metasurface assisted a-Ge0.83Sn0.17 waveguide bolometer with high responsivity. The designed bolometer takes advantage of the LSPR effect from the 1D metasurface, exhibiting a high responsivity of around −3.17%/µW within the 4.1 ∼ 4.3 µm wavelength range. In addition, a f3 dB beyond 10 kHz is realized. The proposed bolometer in this work paves the way for high responsivity and high-speed MIR integrated photonic sensors on Si substrates.

Funding

Shanghai Collaborative Innovation Center of Intelligent Sensing Chip Technology; Shanghai Key Laboratory of Chips and Systems for Intelligent Connected Vehicle; National Natural Science Foundation of China (62204149, 62205193, U23A20356); National Key Laboratory of Infrared Detection Technologies.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. X. Tan, H. Zhang, J. Li, et al., “Non-dispersive infrared multi-gas sensing via nanoantenna integrated narrowband detectors,” Nat. Commun. 11(1), 5245 (2020). [CrossRef]  

2. C. Chen, C. Li, S. Min, et al., “Ultrafast Silicon Nanomembrane Microbolometer for Long-Wavelength Infrared Light Detection,” Nano Lett. 21(19), 8385–8392 (2021). [CrossRef]  

3. A. Rogalski, P. Martyniuk, and M. Kopytko, “Challenges of small-pixel infrared detectors: a review,” Rep. Prog. Phys. 79(4), 046501 (2016). [CrossRef]  

4. M. Kimata, “Uncooled infrared focal plane arrays,” IEEJ Trans. Electr. Electron. Eng. 13(1), 4–12 (2018). [CrossRef]  

5. L. Yu, Y. Guo, H. Zhu, et al., “Low-Cost Microbolometer Type Infrared Detectors,” Micromachines 11(9), 800 (2020). [CrossRef]  

6. A. Kumar, J. Gosciniak, T. B. Andersen, et al., “Power monitoring in dielectric-loaded surface plasmon-polariton waveguides,” Opt. Express 19(4), 2972–2978 (2011). [CrossRef]  

7. J. Gosciniak, M. G. Nielsen, L. Markey, et al., “Power monitoring in dielectric-loaded plasmonic waveguides with internal Wheatstone bridges,” Opt. Express 21(5), 5300–5308 (2013). [CrossRef]  

8. H. Wu, K. Ma, Y. Shi, et al., “Ultracompact on-chip photothermal power monitor based on silicon hybrid plasmonic waveguides,” Nanophotonics 6(5), 1121–1131 (2017). [CrossRef]  

9. Y. Li, A. Chen, T. Dai, et al., “Size effects in a hybrid plasmonic waveguide based power monitor with a bolometric method,” J. Phys. D: Appl. Phys. 54(35), 355107 (2021). [CrossRef]  

10. Y. Wu, Z. Qu, A. Osman, et al., “Mid-Infrared Nanometallic Antenna Assisted Silicon Waveguide Based Bolometers,” ACS Photonics 6(12), 3253–3260 (2019). [CrossRef]  

11. Y. Wu, Z. Qu, A. Osman, et al., “Nanometallic antenna-assisted amorphous silicon waveguide integrated bolometer for mid-infrared,” Opt. Lett. 46(3), 677–680 (2021). [CrossRef]  

12. J. Shim, J. Lim, D.-M. Geum, et al., “TiOx/Ti/TiOx Tri-Layer Film-Based Waveguide Bolometric Detector for On-Chip Si Photonic Sensor,” IEEE Trans. Electron Devices 69(4), 2151–2158 (2022). [CrossRef]  

13. J. Shim, J. Lim, D.-M. Geum, et al., “High-responsivity waveguide-integrated bolometer based on free-carrier absorption for Si photonic sensors,” Opt. Express 30(23), 42663–42677 (2022). [CrossRef]  

14. M. Nedeljkovic, J. S. Penades, C. J. Mitchell, et al., “Surface-Grating-Coupled Low-Loss Ge-on-Si Rib Waveguides and Multimode Interferometers,” IEEE Photonics Technol. Lett. 27(10), 1040–1043 (2015). [CrossRef]  

15. A. Malik, M. Muneeb, S. Pathak, et al., “Germanium-on-Silicon Mid-Infrared Arrayed Waveguide Grating Multiplexers,” IEEE Photonics Technol. Lett. 25(18), 1805–1808 (2013). [CrossRef]  

16. Y. C. Chang, V. Paeder, L. Hvozdara, et al., “Low-loss germanium strip waveguides on silicon for the mid-infrared,” Opt. Lett. 37(14), 2883–2885 (2012). [CrossRef]  

17. M. Nedeljkovic, J. S. Penades, V. Mittal, et al., “Germanium-on-silicon waveguides operating at mid-infrared wavelengths up to 8.5 µm,” Opt. Express 25(22), 27431–27441 (2017). [CrossRef]  

18. C. Calleja, A. Torres, P. Rosales-Quintero, et al., “Amorphous Silicon-Germanium Films with Embedded Nanocrystals for Thermal Detectors with Very High Sensitivity,” J. Nanotechnol. 2016, 1–9 (2016). [CrossRef]  

19. M. Abdel-Rahman, M. Alduraibi, M. Hezam, et al., “Sputter deposited GeSn alloy: A candidate material for temperature sensing layers in uncooled microbolometers,” Infrared Phys. Technol. 97, 376–380 (2019). [CrossRef]  

20. E. Bahaidra, N. Al-Khalli, M. Hezam, et al., “Electro-Optical Characterization of an Amorphous Germanium-Tin (Ge1-xSnx) Microbolometer,” J. Infrared, Millimeter, Terahertz Waves 44(3-4), 233–244 (2023). [CrossRef]  

21. X. She, D. Wang, Y. Zhao, et al., “Broadband and CMOS compatible polarization splitter–rotator based on a bi-level taper,” Appl. Opt. 60(31), 9619–9623 (2021). [CrossRef]  

22. D. Dai, S. He, and H.-K. Tsang, “Bilevel mode converter between a silicon nanowire waveguide and a larger waveguide,” J. Lightwave Technol. 24(6), 2428–2433 (2006). [CrossRef]  

23. D. Dai, J.-J. He, and S. He, “Elimination of multimode effects in a silicon-on-insulator etched diffraction grating demultiplexer with bi-level taper structure,” IEEE J. Sel. Top. Quantum Electron. 11(2), 439–443 (2005). [CrossRef]  

24. D. Imbrenda, R. Hickey, R. A. Carrasco, et al., “Infrared dielectric response, index of refraction, and absorption of germanium-tin alloys with tin contents up to 27% deposited by molecular beam epitaxy,” Appl. Phys. Lett. 113(12), 122104 (2018). [CrossRef]  

25. J. Ge, Y. Zhang, H. Li, et al., “Ultra-Broadband, Tunable, and Transparent Microwave Meta-Absorber Using ITO and Water Substrate,” Adv. Opt. Mater. 11(10), 2202873 (2023). [CrossRef]  

26. A. D. Rakić, A. B. Djurisic, J. M. Elazar, et al., “Optical properties of metallic films for vertical-cavity optoelectronic devices,” Appl. Opt. 37(22), 5271–5283 (1998). [CrossRef]  

27. R. R. Lieten, C. Fleischmann, S. Peters, et al., “Structural and Optical Properties of Amorphous and Crystalline GeSn Layers on Si,” ECS J. Solid State Sci. Technol. 3(12), P403–P408 (2014). [CrossRef]  

28. C. Xu, P. M. Wallace, D. A. Ringwala, et al., “Mid-infrared (3-8 µm) Ge1-ySny alloys (0.15 < y < 0.30): Synthesis, structural, and optical properties,” Appl. Phys. Lett. 114(21), 2104 (2019). [CrossRef]  

29. A. Terrasi, S. Cosentino, I. Crupi, et al., “Ge nanostructures for harvesting and detection of light,” in Nanotechnology and Photovoltaic Devices: Light Energy Harvesting with Group IV Nanostructures, J. ValentaS. Mirabella, eds. (Jenny Stanford Publishing, 2015), pp. 317–353.

30. D. Beaglehole and M. Zavetova, “The fundamental absorption of amorphous Ge, Si and GeSi alloys,” J. Non-Cryst. Solids 4, 272–278 (1970). [CrossRef]  

31. M. Février, P. Gogol, A. Aassime, et al., “Giant coupling effect between metal nanoparticle chain and optical waveguide,” Nano Lett. 12(2), 1032–1037 (2012). [CrossRef]  

32. C. Chen, D. A. Mohr, H.-K. Choi, et al., “Waveguide-Integrated Compact Plasmonic Resonators for On-Chip Mid-Infrared Laser Spectroscopy,” Nano Lett. 18(12), 7601–7608 (2018). [CrossRef]  

33. C. Chen, S.-H. Oh, and M. Li, “Coupled-mode theory for plasmonic resonators integrated with silicon waveguides towards mid-infrared spectroscopic sensing,” Opt. Express 28(2), 2020–2036 (2020). [CrossRef]  

34. M. P. Hokmabadi, J.-H. Kim, E. Rivera, et al., “Impact of Substrate and Bright Resonances on Group Velocity in Metamaterial without Dark Resonator,” Sci. Rep. 5(1), 14373 (2015). [CrossRef]  

35. W. K. Rhim and T. Ishikawa, “Thermophysical properties of molten germanium measured by a high-temperature electrostatic levitator,” Int. J. Thermophys 21(2), 429–443 (2000). [CrossRef]  

36. B. Peng, H. Zhang, H. Shao, et al., “Low lattice thermal conductivity of stanene,” Sci. Rep. 6(1), 20225 (2016). [CrossRef]  

37. D. Spirito, N. V. D. Driesch, C. L. Manganelli, et al., “Thermoelectric Efficiency of Epitaxial GeSn Alloys for Integrated Si-Based Applications: Assessing the Lattice Thermal Conductivity by Raman Thermometry,” ACS Appl. Energy Mater. 4(7), 7385–7392 (2021). [CrossRef]  

38. C. J. S. Glassbrenner and A. Glen, “Thermal Conductivity of Silicon and Germanium from 3 K to the Melting Point,” Phys. Rev. 2016(4A), 1–9 (2016). [CrossRef]  

39. Y. Z. Deng, S.-F. Tang, H. Y. Zeng, et al., “Experiments on Temperature Changes of Microbolometer under Blackbody Radiation and Predictions Using Thermal Modeling by COMSOL Multiphysics Simulator,” Sensors 18(8), 2593 (2018). [CrossRef]  

40. J. G. Hust and A. B. Lankford, Thermal Conductivity of Aluminum, Copper, Iron, and Tungsten for Temperatures from 1 K to the Melting Point, NBSIR 84-3007 (National Bureau of Standards, U.S. Department of Commerce, Boulder, CO, 1884).

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) 3D schematic of the 1D metasurface assisted a-Ge0.83Sn0.17 waveguide bolometer (Device Ai). (b) The top view and the right-side view for Device Ai and Device Bi.
Fig. 2.
Fig. 2. (a) TE0 mode distribution of the Ge waveguide at 4.2 µm wavelength. (b) X-Z cross-sectional view of the E-field at the bilevel coupler region. (c) Near-field E-field profile of 10 meta-atoms with r = 180 nm, t = 20 nm, p = 400 nm at λ = 4.2 µm.
Fig. 3.
Fig. 3. (a) Energy fraction of A, T, R, S as a function of the number of meta-atoms at 4.2 µm wavelength. The lw is 15 µm and the lg is 1 µm. (b) Absorption of meta-atoms with different lengths lm at 4.2 µm wavelength. The inset shows the field enhancement factor with different lengths lm. (c) The absorption spectra by 3D FDTD simulation (solid lines) and by TCMT curve fitting (dashed lines) in frequency domains. A decent matching between the TCMT results and the FDTD curves can be observed. (d) Extracted values of η and −γ/(γ2 + β2) for different devices.
Fig. 4.
Fig. 4. Top view (a), (b), (c), (d) and cross-sectional view (e), (f), (g), (h) of the temperature distribution for Device A5/A10/A15/B10. The Pin is 5 µW and the λ is 4.2 µm.
Fig. 5.
Fig. 5. (a) I - V curves of the Device A10 with various Pin from 0 to 10 µW. The ΔT represents the average temperature change in the thermistor. (b) The resistance response as a function of Pin for each bolometer. (c) Wavelength-dependence of the responsivity for waveguide bolometer A10.
Fig. 6.
Fig. 6. (a) Temporal responses of ΔT as a function of time t. The optical power and bias voltage are 5 µW and 1 V, respectively. (b) Normalized temperature change with respect to the modulating frequency f of the heat power in the simulation.

Tables (4)

Tables Icon

Table 1. Absorption of the bolometers in the FDTD simulation

Tables Icon

Table 2. Thermal properties of the materials in device simulation

Tables Icon

Table 3. Responsivities of the Device A10 with different αGeSn

Tables Icon

Table 4. Benchmark between the reported Si-based waveguide bolometers

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

A = 1 T R S 1 | s 2 | 2 | s 1 + | 2 = 1 [ | t 0 | 2 + 2 β τ t e 1 ( ω l ω 0 ) + τ t e 2 + 2 γ τ 1 τ t e 1 ( ω l ω 0 ) 2 + τ 2 ] ,
Δ T = A φ 0 ( G t h 2 + f 2 C t h 2 ) 1 / 2 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.