• Photonics Research
  • Vol. 13, Issue 6, 1497 (2025)
Naiquan Yan1,2,†, Feng Shi3,†, Xiaomeng Xue1,†, Kenan Zhang2..., Cheng Huo1,2 and Menglu Chen1,2,4,5,*|Show fewer author(s)
Author Affiliations
  • 1School of Optics and Photonics, Beijing Institute of Technology, Beijing 100081, China
  • 2Zhejiang Key Laboratory of 3D Micro/Nano Fabrication and Characterization, Westlake Institute for Optoelectronics, Hangzhou 311421, China
  • 3Laboratory of Science and Technology on Integrated Logistics Support, Changsha 410073, China
  • 4State Key Laboratory of Materials for Integrated Circuits, Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Sciences, Shanghai 200050, China
  • 5National Key Laboratory on Near-Surface Detection, Beijing 100012, China
  • show less
    DOI: 10.1364/PRJ.546383 Cite this Article Set citation alerts
    Naiquan Yan, Feng Shi, Xiaomeng Xue, Kenan Zhang, Cheng Huo, Menglu Chen, "Cavity-enhanced infrared quantum dot homojunction arrays," Photonics Res. 13, 1497 (2025) Copy Citation Text show less

    Abstract

    Infrared spectroscopy has wide applications in the medical field, industry, agriculture, and other areas. Although the traditional infrared spectrometers are well developed, they face the challenge of miniaturization and cost reduction. Advances in nanomaterials and nanotechnology offer new methods for miniaturizing spectrometers. However, most research on nanomaterial-based spectrometers is limited to the visible wavelength or near infrared region. Here, we propose an infrared spectrometer based on diffraction gratings and colloidal quantum dot (CQD) homojunction photodetector arrays. Coupled with a Fabry-Perot cavity, the CQD photodetector covers the 1.4–2.5 μm spectral range, with specific detectivity 4.64×1011 Jones at 2.5 μm at room temperature. The assembled spectrometer has 256 channels, with total area 2.8mm×40mm. By optimizing the response matrix from machine learning algorithms, the CQD spectrometer shows high-resolution spectral reconstruction with a resolution of approximately 7 nm covering the short-wave infrared.

    1. INTRODUCTION

    Optical spectrometry can analyze substance characteristics at different wavelengths, which is widely applied in food safety, chemical material analysis, disease detection, environmental monitoring, and other fields [15]. Infrared spectroscopy especially provides crucial data [6,7] distinguishing materials [810], due to the fingerprint features of substances in the infrared spectrum. By far, traditional infrared spectrometers are well developed. For example, Fourier transform infrared spectroscopy provides rapid acquisition of high-resolution spectral data, widely applied in chemical analysis and materials science [11,12]. Although the set-up significantly enhances convenience and efficiency in laboratory work, it is bulky and costly. There is a growing demand for miniaturized, low-cost, and high-performance infrared spectrometers [1317]. Many attempts have been made by researchers to miniaturize spectrometers. In 2001, Kong et al. developed a dual-silicon planar grating chip spectrometer consisting of two wafers bonded vertically [18]. The top silicon wafer is scribed with a slit to act as a planar transmission grating, and a line array detector is integrated on the bottom silicon wafer to miniaturize the spectrometer. In 2012, a miniature spectrometer based on linear variable optical filters was proposed by Emadi et al. [19]. This design scheme can circumvent the optical range limitation, thus making it possible to achieve a more compact spectrometer.

    The design of discrete dispersion elements, mechanical movable parts, and photodetectors directly affects the miniaturization. The development of nanotechnologies [2024] and computer technologies [2528] provides a new method [13,15]. Nanomaterials exhibit many novel optical and electrical properties due to the quantum confinement. Therefore, they show attractive prospects in electrical and optical devices and have attracted wide attention for spectral filters and the photodetectors in spectrometers. For instance, Bao et al. utilized 195 different colloidal quantum dot (CQD) filters, pioneering CQD integrated spectrometers [29]. Utilizing adjustable absorption characteristics of CQD, they have developed fine-transmission filters to replace interferometric optical devices. These filters are integrated with charge coupled device (CCD) cameras, covering the 390–690 nm visible spectrum. Later, Zhu et al. integrated an array of 361 CQD-embedded film filters with a CCD camera, achieving a spectral resolution of 1.6  nm over a broad range from 250 to 1000 nm [30]. In addition, research on CQD-based detector arrays is rapidly advancing. In 2018, Chu et al. explored the application of two nanocrystalline materials, PbS and HgTe, in the short-wave infrared (SWIR) region and successfully integrated HgTe nanocrystal films into multi-pixel devices, demonstrating their potential for high-performance imaging systems [31]. In 2024, Ahn et al. proposed a novel hybrid photodetector by combining graphene with PbS quantum dots [32]. This device achieves spectral analysis through an alternating stacked structure, where photosensitive layers at different positions exhibit distinct spectral responses, thereby broadening the functional scope of the detector.

    These achievements present the potential of CQD in high-performance micro spectrometer applications. Recently, the advancement of computer technology has promoted computational spectrometry [2528]. The reconstruction spectrometer systems typically consist of a set of detectors with unique spectral response characteristics. Yang et al. have proven that utilizing machine learning algorithms for spectral reconstruction significantly reduces the size of spectrometers [33] based on single nanowires.

    However, since the nanomaterials are only used as filters, most nanomaterial-based spectrometers are limited in the visible wavelength band due to the detectable wavelength of silicon photodetectors. The configuration of infrared photodetectors in linear or planar arrays for spectral analysis heavily relies on narrow band semiconductor materials. Unfortunately, the epitaxial growth and the silicon incompatibility do come with increased costs from the manufacturing complexity and the specialized equipment. Recent research shows great progress on CQD infrared photodetection, covering near infrared [34], short-wave infrared [35], mid-wave infrared [3638], long-wave infrared [39,40], and up to very long-wave infrared [40]. The solution processability also promotes the development of infrared spectrometers with CQD photodetector arrays.

    2. RESULTS AND DISCUSSION

    In this work, we developed the CQD homojunction photodetector arrays with resonant cavities, with a detectivity of 4.64×1011 Jones at 2.5 μm at room temperature. Compared to a heterojunction CQD detector, homojunction demonstrates higher quantum efficiency and responsivity due to the well-designed band alignment [38]. What is more, resonant cavity structures have been used to further improve the optical collection efficiency of the CQD film [41]. We then built an infrared spectrometer based on diffraction gratings, CQD photodetector arrays, and computing methods. Diffraction gratings can offer better resolution in diffraction spectrometers because of their superior dispersion capability [42]. The spectrometer was equipped with a total of 256 detector channels. As shown in Fig. 5 (Appendix B), the structure of each pixel and the device fabrication process are identical, and the CQD and SU-8 solutions are spin-coated. An array of contact electrodes is integrated onto the wafer and the entire wafer is then cut into individual array modules. This approach ensures the uniformity of the CQD film and resonant cavity layers and improves the uniformity of the optical response over 256 pixels. The area of a single pixel is 100  μm×400  μm. The spectrometer displays a detection spectrum spanning 1.4–2.5 μm, and a spectral resolution capability reaching 7 nm confirms the spectrometer’s capabilities.

    Figure 1 shows the characterization of a single-pixel photodetector. The device structure design is shown in Fig. 1(a), combining a homojunction photodiode and a Fabry-Perot cavity. For a PIN photodiode, N-type and P-type CQD films are used as the electron and hole transport layers, respectively. The homojunction design not only avoids carrier lost at the interface but also benefits a larger built-in electric field. The Fabry-Perot cavity mainly consists of a gold contact electrode (lower gold layer), an optical spacer layer (SU-8 photoresist), and a high-reflectivity upper gold back-reflector. The bottom lighting is used to maximize the light collection efficiency. As the SEM image shows in Fig. 1(a), an indium tin oxide (ITO) layer about 50 nm thick is used as the bottom contact on the Al2O3 substrate. Stacked in sequence to form the PIN homojunction are layers of N-type CQD with 50 nm, intrinsic with 300 nm, and P-type with 50 nm [43]. Then, the resonant cavity structure was prepared on top of the CQD film, which includes a 10 nm thick contact gold electrode, 1.3 μm SU-8 photoresist, and 40 nm upper gold as the reflective layer. Detailed information on the device fabrication can be found in Appendix A.

    Characterization of one resonant-cavity-enhanced CQD homojunction photodetector. (a) Schematic structure of the detector is depicted on the left. The upper-right inset displays a cross-sectional scanning electron microscopy (SEM) image of the device, and the bottom-right inset shows a magnified image of the CQD layer. Scale bar: 500 nm. (b) Electric field distribution of the detector. (c) Simulated absorption of CQD layer with different spacer thicknesses. (d) Simulated absorption of CQD layer with different gold contact thicknesses. (e) Simulated absorption of CQD layers with different upper gold thicknesses. (f) Simulated absorption of CQD layer with different CQD thicknesses.

    Figure 1.Characterization of one resonant-cavity-enhanced CQD homojunction photodetector. (a) Schematic structure of the detector is depicted on the left. The upper-right inset displays a cross-sectional scanning electron microscopy (SEM) image of the device, and the bottom-right inset shows a magnified image of the CQD layer. Scale bar: 500 nm. (b) Electric field distribution of the detector. (c) Simulated absorption of CQD layer with different spacer thicknesses. (d) Simulated absorption of CQD layer with different gold contact thicknesses. (e) Simulated absorption of CQD layers with different upper gold thicknesses. (f) Simulated absorption of CQD layer with different CQD thicknesses.

    COMSOL simulation is used to determine the parameters of the device, mainly the thickness of the SU-8 photoresist, the lower gold layer, the upper gold, and the CQD layer. The left and right sides were set as periodic boundary conditions. For the CQD layer, we set the real part of the refractive index at 2.3 and the value of the imaginary part changed with the wavelength, determined by extracting the distribution of extinction coefficients from the absorption. The refractive index of the air was 1+0i. The thickness of CQD was pre-set to 400 nm, determined by the typical thickness used. With the SU-8 photoresist thickness of 1.3 μm, the electric field distribution of the device under resonant conditions is shown in Fig. 1(b). The incident light was confined within the Fabry-Perot cavity, forming standing waves, which caused a strong spatial overlap between the CQD absorber and the optical field [44,45].

    The resonant cavity structure is spectrally selective. To achieve the optimal device performance, we explored the variation in optical spacer layers and CQD thicknesses. As shown in Fig. 1(c), by tuning the optical spacer layers thicknesses from 1 to 2.1 μm, the peak absorption would be tuned from 5800 to 6100  cm1. The peak absorption was increased from 45% (no cavity) to 85% (1.3 μm cavity thickness). In the simulation, sharp absorption peaks appeared at 3800  cm1, 5000  cm1, and 6600  cm1, which may be from the absorption of the two layers of gold in the device. Figures 1(d) and 1(e) show the effect of different thicknesses of the lower and upper gold layers on the simulated absorption of the device when the SU-8 photoresist is 1.3 μm and the CQD layer is 400 nm. From the results, when the thickness of the lower gold layer was varied, there was a notable change in the sharp absorption at 5000  cm1, with a small blue shift. When the lower gold layer was fixed at 10 nm, changing the thickness of the upper gold layer, there was a significant absorption at 5000  cm1 and the intensity of absorption was maximum at 40 nm of the upper gold layer, while the absorption at other positions was almost unchanged. Figure 1(f) shows the simulation varying the CQD thickness, which determined the cavity length and the resonance peak absorption from 4800 to 6500  cm1. As expected [46,47], reducing the CQD thickness caused a blue shift of the resonance peak. The 400 nm CQD film with the resonant cavity exhibits a more significant absorption enhancement effect. Based on the simulation, we have prepared resonant-cavity-enhanced CQD homojunction arrays, with the length of 4 cm and the width of 2.8 mm containing 256 pixels.

    Figure 2 shows the photodetection property on the device. The 600°C blackbody is used as the light source with an incident power P of 1.24 μW. Figure 2(a) shows the IV curves of the detector with and without a resonant cavity. At zero bias, the photocurrent (Iph) of the device with the resonant cavity is two-fold that without the cavity, which are 1.44 μA and 0.68 μA, respectively. Responsivity R=IphP is 1.16 A/W and 0.54 A/W with and without the resonant cavities, respectively. The two-fold difference is consistent with the simulation results where the resonant cavity increases the total photon absorption of the QD film. The noise of the photodetector with a cavity is increased by a factor of 1.6 compared to the photodetector without a cavity as shown in Fig. 2(b); the measured noise of PIN homojunction detectors with and without a resonant cavity is 0.05  pA  Hz1/2 and 0.03  pA  Hz1/2, respectively. The 50 Hz industrial frequency interference noise may come from the test instrument. In our case, the noise spectrum is measured in the dark without an external bias, indicating the main noise is from Johnson noise and generated complex noise. The Johnson noise is related to the device dark resistance, expressed as IJohnson=4KBTB/R, where KB is the Boltzmann constant, R the resistance, B the bandwidth, and T the temperature. As Fig. 2(a) shows, the slope of the dark current, which is the dark resistance, increases with the FP cavity. The increasement of dark resistance is possibly from the change of the internal electric field with the cavity, since the metal-insulator-metal layer basically forms a capacitor. Besides, the process of the SU-8 layer and additional Au layer may also slightly affect the QD layers’ property. Although the noise increases with the cavity, the specific detectivity (D*) is calculated as D*=RAIn, where In is the current noise, and A is the area of the device.

    Resonant cavity devices increase the transmission efficiency and gain of light, enhancing their detectivities, which are 4.64×1011 Jones and 3.6×1011 Jones for devices with and without a resonant cavity, respectively. The external quantum efficiency of the photodetector is defined as EQE=1.24Rλ(μm). The device can achieve an EQE of more than 57%, which is comparable to epitaxially grown HgCdTe detectors at the same wavelength. The response speed of the single-pixel detector is within 10 μs as shown in Fig. 6 (Appendix B). The spectrometer could work as high as 50 kHz with the current design.

    We have prepared detectors with different CQD thicknesses to further verify the simulation results, as shown in Fig. 7 (Appendix B). The photocurrents of the homojunction resonance-enhanced CQD detectors with 300 nm, 500 nm, and 600 nm thicknesses were all increased compared to the detectors without the resonant cavity, while the 400 nm CQD thickness device exhibits the most superior performance. We also verified the IV curves of other channels among the 256 pixels, and both the photocurrent and rectification characteristics showed high consistency, which is attributed to our fabrication method, as shown in Fig. 8 (Appendix B). Figure 2(c) shows the spectral response of detectors with and without the resonant cavity at the cut-off of 2.5 μm. The peak response of the device enhanced by the resonant cavity is nearly twice that without the cavity, which also matches our simulation. The spectral response shows two smooth absorptions. To test the model reliability, we also show that the CQD devices’ simulation results are consistent with the actual measurement response spectra at the cut-off of 2.1 μm in Fig. 9 (Appendix B).

    Resonance-enhanced CQD homojunction device characterization. (a) I–V curves comparing detectors with and without resonant-cavity-enhanced microstructures. (b) Noise spectra comparing detectors with and without resonant-cavity-enhanced microstructures. (c) Response spectra comparing detectors with and without resonant-cavity-enhanced microstructures.

    Figure 2.Resonance-enhanced CQD homojunction device characterization. (a) IV curves comparing detectors with and without resonant-cavity-enhanced microstructures. (b) Noise spectra comparing detectors with and without resonant-cavity-enhanced microstructures. (c) Response spectra comparing detectors with and without resonant-cavity-enhanced microstructures.

    Figure 3 shows a CQD homojunction photovoltaic array. A grating is used as a compact dispersive element, with the microscope image shown in Fig. 10 (Appendix B). The spectral reconstruction process includes learning, processing, and reconstruction stages [48,49]. The basic steps of spectrum reconstruction are as shown in Fig. 3(a), where the right figure shows the circuit diagram of the spectrometer operation. This work uses a 16-bit, 256-channel ADC with parallel sampling and current inputs, which allows for the simultaneous measurement of up to 256 channels of current signals. The physical picture is shown in Fig. 11 (Appendix B).

    HgTe CQD spectrometer. (a) The detector line array and data processing section, and the schematic diagram of the line array post-processing circuit is shown on the right. (b1) Voltage spectra calculated following Eq. (1). (b2) Directly measured photovoltage spectra. (c) Response spectra of the detector at different positions. (d) Spectrogram after spectral differentiation from (c).

    Figure 3.HgTe CQD spectrometer. (a) The detector line array and data processing section, and the schematic diagram of the line array post-processing circuit is shown on the right. (b1) Voltage spectra calculated following Eq. (1). (b2) Directly measured photovoltage spectra. (c) Response spectra of the detector at different positions. (d) Spectrogram after spectral differentiation from (c).

    We outline the spectrum reconstruction process using mathematical formulations, beginning with computational methods for spectral reconstruction [13,33,50]. In spectrometer operation, r(λ) is the normalized spectral response of our homemade photodetector, which is consistent for each pixel. f(xn,λ) denotes the spectral distribution of the certain light shined on the nth photodetector. G(xn) represents photovoltage of the nth photodetector under certain illumination. The function f can be described as a linear combination of basic functions. Since the shape of the Gaussian function is similar to the shape of the contours of the source spectral lines, we choose the Gaussian function u(λ) for fitting. To reduce the effect of noise, we introduce Tikhonov regularization along a generalized cross-validation approach to determine the appropriate parameters. Refer to the support information diagram for a detailed algorithm flowchart in Fig. 12 (Appendix B):G(xn)=λminλmaxr(λ)f(xn,λ)dλ,n=1,2,,256,setting the wavelength range for spectrometer operation: G=RF.Here, G represents the measured photovoltage, R denotes the mapping matrix, and F denotes the initial spectrum.

    In the reconstruction process of the spectrometer described in this work, the initial step involves a learning process. The spectral distribution f of a particular light beam incident on the nth photodetector is a function of the position x and the wavelength λ. During this phase, the learning data consist of photovoltages (G) and the spectral distribution f recorded at position (x) under light excitation. The photovoltage G(x) is the integral of the product of the incident spectral distribution and the responsivity over the entire wavelength range λminλmaxG(xn)=λminλmaxr(λ)f(xn,λ)dλ, where xn=x1, x2,..., x256. Figure 3(c) shows schematically the incident spectral distribution at different positions. At different positions x, optical voltages can be measured for each channel of the spectrometer, where each xn corresponds to a specific voltage value, as depicted in Fig. 3(b2). In this learning process, we derive a total of n equations from x1 to xn, which can be expressed as matrix equations: (G(x1),G(x2),,G(xn))=(r(λ1),r(λ2),,r(λc))(f(x1,λ1)f(x2,λ1)f(xn,λ1)f(x1,λ2)f(x2,λ2)f(xn,λ2)f(x1,λc)f(x2,λc)f(xn,λc)).

    To achieve greater accuracy in r(λ), we employ an iterative algorithm [51]. Since the photovoltage G(x) is the integral of the product of the incident spectral distribution and the responsivity over the wavelength range, we display its integral result, as shown in Fig. 3(b1). Subsequently, we substitute the obtained integral result back into Eq. (1) and solve iteratively until r(λ) stabilizes. For two adjacent spectral curves, the main difference is the band edge of the spectral response, lying mainly in the specific spectral distribution of the neighboring photodetectors. The main bodies between the two adjacent spectral curves are approximately the same, so differentiation is naturally considered to visualize more intuitively the difference between neighboring f(xn,λ). Here, we perform a spectral difference discretization of its two neighboring f(xn,λ)r(λ), as shown in Eq. (4). Since both sides of Eq. (4) have the same r(λ), the spectral distribution difference map can be obtained as shown in Fig. 3(d): h(xn,λ)r(λ)=f(xn+1,λ)r(λ)f(xn,λ)r(λ).

    By solving Eq. (2), the mapping matrix r(λ) can be obtained. Here, the response vector terminates at λc, determined by the detector. Noise present during the operation of the spectrometer may affect the reconstruction results. To solve matrix Eq. (3) and mitigate the impact of noise and errors on the reconstruction results, a reliable algorithm is necessary. Previous studies have demonstrated various reconstruction algorithms such as simulated annealing, nonnegative least squares, sparse optimization, and machine learning [5256]. In this work, nonnegative constraint Tikhonov regularization and machine learning algorithms are employed to address these issues. In this learning process, 256 different positions are utilized to generate spectral responsivity vectors. Subsequently, the spectral mapping matrix R is reconstructed from these 256 response vectors. Figure 3(c) illustrates that the spectrometer’s cut-off wavelength in this study spans approximately 1.4–2.5 μm.

    Next, we used a machine learning approach. Our goal was to read the light information under different positions by spectral measurements. A training database was first created that contained datasets of photovoltaic and spectral eigenvalues from 256 detectors. The spectral data were measured using a commercial Fourier transform spectrometer (FTIR) (as shown in Figs. 13 and 14, Appendix B) [54,5759]. The known incident data are used to train a multilayer artificial neural network (ANN), consisting of input, hidden, and output layers. We used a one-dimensional convolutional neural network followed by a fully connected network. Spectra are fed to the input layer of the network, and the hidden layer converts the input data into a form that favors the prediction of the output. Besides, in the output layer, each bit sequence is attributed to a neuron. Therefore, the input-output mapping for predicting spectral eigenvalues can be built directly by providing enough training data to the ANN. Additionally, the machine learning algorithm incorporates a feedback adjustment module. The mapping matrix r(λ) obtained from the training dataset can be adjusted in time to further optimize the reconstructed spectra.

    To achieve higher reconstruction accuracy, we conducted a deeper investigation into our reconstruction methods. The training dataset directly affects the reconstruction accuracy. By making subtle adjustments to the positions of the detector array, we can alter the light signals received by each channel of the detector array, thereby changing the information captured in each channel. So, we adjust the position of the detector arrays to obtain an additional dataset. The calculated spectra could be expressed as f(λ)=1Ni=1Nfi(λ). N is the number of the detector array movement. Then the error between the calculated spectra and measured spectra could be converged with N. In this work, we measured the responsivity of the photodetectors at new positions [denoted as f(xn,λ) and the photovoltage g(xn) at these positions]. More details can be found in Fig. 15 (Appendix B).

    We present the spectral reconstruction process for broadband spectra, as depicted in Fig. 4. Initially, the spectral responsivity matrix r(λ) is obtained during the learning phase, which is put into our processing chip. Then, we measured the photovoltage as a function of the position x. In the absence of samples, the line array optical voltages were measured at the x and x positions (respectively denoted as reference@x, reference@x). Then, with the sample positioned in the path of the optical beam, we again measured the photovoltage at positions x and x (denoted as sample@x, sample@x). These photovoltage values are functions of x, as illustrated in Fig. 4(a). Here, photovoltage g(xn)=λminλmaxf(xn,λ)·r(λ)dλ, where r(λ) denotes the mapping matrix to be determined. Schematic spectral response curves for 256 different positions xn are shown in Fig. 4(b). Reconstruction of the unknown spectra is achieved by fitting the photocurrent data to the relationship with position xn akin to the initial learning phase.

    Reconstructed spectrum. (a) Plot of photovoltage as a function of x at different positions before and after placing the sample. (b) Schematic spectral response curves for 256 different positions xn. (c) Comparison of spectra obtained with a commercial Fourier transform spectrometer and reconstructed spectra.

    Figure 4.Reconstructed spectrum. (a) Plot of photovoltage as a function of x at different positions before and after placing the sample. (b) Schematic spectral response curves for 256 different positions xn. (c) Comparison of spectra obtained with a commercial Fourier transform spectrometer and reconstructed spectra.

    As shown in Fig. 4(c), we compared the differences between spectra obtained from the commercial Fourier transform spectrometer and our reconstructed spectra. The dashed line corresponds to the measurement results from the commercial spectrometer, while the solid line depicts the results from our spectrometer. It is evident that our reconstruction captures all major spectral features. In addition, by expanding the channel of the detector, it is theoretically possible to obtain higher reconstruction accuracy.

    3. CONCLUSION

    In this work, we have developed a miniaturized reconstruction spectrometer. A grating is used to spatially separate the optical information. The reconstruction of the spectrum is achieved through a CQD detector line array. The performance of the device is further improved by the resonant-cavity-enhanced homojunction structure. In addition, solution processing of our detector arrays significantly reduces the complexity of epitaxial growth processing and significantly reduces device preparation costs. In the spectral reconstruction process, we use a machine learning algorithm to train the corresponding matrices and reduce the effect of noise through a regularization law. The spectrometer covers the spectral range from 1.4 to 2.5 μm and achieves high-resolution spectral reconstruction with a resolution of about 7 nm. The resolutions are comparable to those of the reported high-performance NIR miniature spectrometers (Table 1). This research proposed a new path for the development of miniaturized spectroscopic systems, which represents significant advances in spectral analysis and in practical applications.

    Comparison of NIR Spectrometers between Reported Works and This Work

    SpectrometerSensorSpectral Range (nm)Spectral Resolution (nm)Refs.
    AvaSpec-NIR (Avantes)InGaAs linear array200–11004.4–85[60]
    FieldSpec4 (ASD)Photodiode array350–250010[60]
    Luminar 5030 (Brimrose)InGaAs600–11002–10[60]
    SCiO (Consumer Physics)Photodiode array740–107028[61]
    NIRscan (Texas Instruments)Single InGaAs900–170010[61]
    MicroNIR Pro ES 1700 (VIAVI)InGaAs diode array908–167612.5[61]
    NeoSpectra (Si-Ware Systems)Single InGaAs1350–250016[61]
    Vis-NIR spectrometerCMOS300–170010[62]
    Single-photon spectrometerNbTiN nanowire1200–17005[26]
    Fourier transform microspectrometerInGaAS1100–17007.5[63]
    Photonic crystal film spectrometerCMOS450–6505[64]
    QD-SWIR spectrometerHgTe CQDs array1400–25007This work

    MATERIALS AND METHODS

    Synthesis of Colloidal Nanomaterials

    The synthesis of HgTe CQDs nanocrystals is similar to those reported in Ref. [38]. Place 0.15 mmol of mercury chloride in a 20 mL glass vial and dissolve it in 4 mL of oleylamine. Heat the mixture to 100°C in a glove box while stirring for 60 min. Subsequently, rapidly inject diluted bis(trimethylsilyl)telluride (TMSTe) into the HgCl2 solution, which will promptly turn from transparent to black. Finally, quench the reaction by adding 4 mL of tetrachloroethylene.

    Prior to drop-casting, the CQD required cleaning. The CQD solution was placed in a centrifuge tube, and 30 mL of isopropanol (IPA) was added for washing. The mixture was then centrifuged at 5000 r/min for 5 min. The supernatant was discarded, and the quantum dots were dried with nitrogen before redissolving them in 5 mL n-hexane.

    Mixed-phase ligand exchange is used to modulate the properties and doping of CQD. 5 mL of HgTe CQD was dissolved in hexane and then mixed with 200 μL of β-mercaptoethanol (β-me) dissolved in N,N-dimethylformamide (DMF). The process was accelerated by adding 15 mg of didodecyldimethylammonium bromide (DDAB). Typically, P-type HgTe CQDs were prepared by adding 5 mg of (NH4)2S to the CQD/DMF solution. For N-type HgTe CQD, 20 mg of HgCl2 salt was added. Meanwhile, intrinsic HgTe CQDs were obtained by simply adding 10 mg of HgCl2.

    Device Fabrication

    A cavity-enhanced infrared detector was fabricated on a sapphire substrate. A 50 nm thick layer of ITO was deposited on the substrate via magnetron sputtering, followed by a 15 min annealing process at 300°C. Prior to the dispensing of CQD, the surface was treated with 3-mercaptopropyltrimethoxysilane (MPTS) for 30 s and subsequently rinsed with isopropyl alcohol (IPA); then the CQD solution was spin-coated onto the substrate, with each layer of CQD undergoing solid-phase ligand exchange in an ethanedithiol (EDT)/hydrochloride acid/IPA (1:1:50 by volume) solution for 10 s. Afterward, the layers were rinsed with IPA and dried with nitrogen. The N-type CQD layer of 100 nm, intrinsic CQD layer of 200 nm, and P-type CQD layer of 100 nm were sequentially prepared through continuous deposition of CQD. The 10 nm gold contact electrode was deposited on top of the CQD film via a coating machine. An optical spacer layer (SU-8) was then spin-coated over the gold contact electrode with 1.3 μm thickness. Finally, the gold upper reflective layer of 40 nm was deposited over the optical spacer layer.

    Photodetector Characterization

    The effective sensing area of the detector is 0.1  mm×0.4  mm. Single-pixel detector IV curves were recorded using a Keithley 2602B source meter with a 600°C blackbody as the light source. Noise spectra were captured by a Stanford Research Systems SR770 noise analyzer. The CQD detector line array uses the AKDC256A chip for testing, which can simultaneously measure 256 channels of nA-level input current signals. Each channel of the AKDC256A includes a low-noise, low-power capacitive transimpedance amplifier (CTIA) that captures the input charge and converts it into a voltage signal. Each channel also includes a 16-bit ADC to convert the integrator output to a digital signal. Then, the 16-bit conversion results of each channel are encoded and sent off-chip through the output interface to be captured by the computer.

    Training Database

    The training database for machine learning was obtained using FTIR (FOLI20, Ying Sa Optical Instruments) measurements. The photovoltage data were amplified by a transimpedance amplifier and then amplified with a voltage amplifier.

    SUPPLEMENTARY INFORMATION

    Figures 515 supplement the main text. Figure 5 shows the preparation process of the wire array device. Figures 68 show the performance characterization of the devices. Figure 9 shows the supplementary information of the simulation. Figure 10 shows the microscopic image of the grating. Figure 11 shows the physical image of the spectroscopic system. Figure 12 shows the flowchart of the algorithm. Figures 13 and 14 show the data training library. Figure 15 shows more details after moving the line array.

    Picture of the detector array. (a) 4-inch wafer without drop-casting HgTe CQD. (b) 4-inch wafer after drop-casting HgTe CQD. (c) Line array detector with 256 pixels.

    Figure 5.Picture of the detector array. (a) 4-inch wafer without drop-casting HgTe CQD. (b) 4-inch wafer after drop-casting HgTe CQD. (c) Line array detector with 256 pixels.

    Response speed of single pixel device.

    Figure 6.Response speed of single pixel device.

    The I–V curves and CQD layer thickness tests. (a) I–V curves and 340 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (b) I–V curves and 470 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (c) I–V curves and 600 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (d) I–V curves and 400 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity.

    Figure 7.The IV curves and CQD layer thickness tests. (a) IV curves and 340 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (b) IV curves and 470 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (c) IV curves and 600 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity. (d) IV curves and 400 nm QD layer thickness tests for a resonant cavity-enhanced detector and a microstructured detector without a resonant cavity.

    The I–V curves of other channels among the 256 pixels.

    Figure 8.The IV curves of other channels among the 256 pixels.

    Simulation of quantum dots with different response wavelengths for devices of the same structure. (a) Simulated absorption of HgTe layers with different spacer thicknesses. (b) Simulated absorption of HgTe layers with different quantum dot thicknesses. (c) Spectral response of devices with and without resonant cavity enhancement.

    Figure 9.Simulation of quantum dots with different response wavelengths for devices of the same structure. (a) Simulated absorption of HgTe layers with different spacer thicknesses. (b) Simulated absorption of HgTe layers with different quantum dot thicknesses. (c) Spectral response of devices with and without resonant cavity enhancement.

    Pictures of grating microscope observations. (a) 1 mm/100, (b) 1 mm/300, (c) 1 mm/600.

    Figure 10.Pictures of grating microscope observations. (a) 1 mm/100, (b) 1 mm/300, (c) 1 mm/600.

    Physical picture of the detector array and the grating elements.

    Figure 11.Physical picture of the detector array and the grating elements.

    Algorithm flow chart.

    Figure 12.Algorithm flow chart.

    Training datasets comprising photovoltage and spectral eigenvalue. (a) Spectral response of a detector with resonant cavity at different xn, (b) spectral integration, (c) measured voltage. (d) Transmittance of filter 1. (e) Spectral response of a detector with resonant cavity enhancement under the condition of filter 1 at different positions, (f) measured voltage. (g) Transmittance of filter 2. (h) Spectral response with resonant cavity enhanced detector under the condition of filter 2 at different positions, (i) measured voltage.

    Figure 13.Training datasets comprising photovoltage and spectral eigenvalue. (a) Spectral response of a detector with resonant cavity at different xn, (b) spectral integration, (c) measured voltage. (d) Transmittance of filter 1. (e) Spectral response of a detector with resonant cavity enhancement under the condition of filter 1 at different positions, (f) measured voltage. (g) Transmittance of filter 2. (h) Spectral response with resonant cavity enhanced detector under the condition of filter 2 at different positions, (i) measured voltage.

    Neural network self-learning filter transmittance library with (a) transmittance peaks of about 2.3 μm, (b) transmittance peaks of about 1.7 μm, and (c) transmittance peaks of about 1.9 μm.

    Figure 14.Neural network self-learning filter transmittance library with (a) transmittance peaks of about 2.3 μm, (b) transmittance peaks of about 1.7 μm, and (c) transmittance peaks of about 1.9 μm.

    Schematic of the detector line array movement and the data processing section. (a) Schematic of the detector line array movement. (b) Spectral response curve when the line array detector is located at xn′. (c) Integral plot of spectral curves. (d) Spectrogram of panel (b) after performing spectral differencing.

    Figure 15.Schematic of the detector line array movement and the data processing section. (a) Schematic of the detector line array movement. (b) Spectral response curve when the line array detector is located at xn. (c) Integral plot of spectral curves. (d) Spectrogram of panel (b) after performing spectral differencing.

    References

    [1] C. P. Bacon, Y. Mattley, R. DeFrece. Miniature spectroscopic instrumentation: applications to biology and chemistry. Rev Sci. Instrum., 75, 1-16(2004).

    [2] G. Elmasry, M. Kamruzzaman, D. W. Sun. Principles and applications of hyperspectral imaging in quality evaluation of agro-food products: a review. Crit. Rev. Food Sci. Nutr., 52, 999-1023(2012).

    [3] J. X. Cheng, X. S. Xie. Vibrational spectroscopic imaging of living systems: an emerging platform for biology and medicine. Science, 350, aaa8870(2015).

    [4] L. Cabernard, L. Roscher, C. Lorenz. Comparison of Raman and Fourier transform infrared spectroscopy for the quantification of microplastics in the aquatic environment. Environ. Sci. Technol., 52, 13279-13288(2018).

    [5] P. Edwards, C. Zhang, B. Zhang. Smartphone based optical spectrometer for diffusive reflectance spectroscopic measurement of hemoglobin. Sci. Rep., 7, 12224(2017).

    [6] D. Romanini, M. Chenevier, S. Kassi. Optical–feedback cavity–enhanced absorption: a compact spectrometer for real–time measurement of atmospheric methane. Appl. Phys., 83, 659-667(2006).

    [7] A. J. S. McGonigle, T. C. Wilkes, T. D. Pering. Smartphone spectrometers. Sensors, 18, 223(2018).

    [8] M. Manley. Near-infrared spectroscopy and hyperspectral imaging: non-destructive analysis of biological materials. Chem. Soc. Rev., 43, 8200-8214(2014).

    [9] S. Kang, Z. Qian, V. Rajaram. Ultra-narrowband metamaterial absorbers for high spectral resolution infrared spectroscopy. Adv. Opt. Mater., 7, 1801236(2019).

    [10] F. Hase, M. Frey, T. Blumenstock. Application of portable FTIR spectrometers for detecting greenhouse gas emissions of the major city Berlin. Atmos. Meas. Tech., 8, 3059-3068(2015).

    [11] M. J. Baker, J. Trevisan, P. Bassan. Using Fourier transform IR spectroscopy to analyze biological materials. Nat. Protoc., 9, 1771-1791(2014).

    [12] I. Amenabar, S. Poly, M. Goikoetxea. Hyperspectral infrared nanoimaging of organic samples based on Fourier transform infrared nanospectroscopy. Nat. Commun., 8, 14402(2017).

    [13] Z. Yang, T. Albrow-Owen, W. Cai. Miniaturization of optical spectrometers. Science, 371, eabe0722(2021).

    [14] B. Redding, S. F. Liew, R. Sarma. Compact spectrometer based on a disordered photonic chip. Nat. Photonics, 7, 746-751(2013).

    [15] H. Meng, Y. Gao, X. Wang. Quantum dot-enabled infrared hyperspectral imaging with single-pixel detection. Light Sci. Appl., 13, 121(2024).

    [16] G. Cai, Y. Li, Y. Zhang. Compact angle-resolved metasurface spectrometer. Nat. Mater., 23, 71-78(2024).

    [17] H. Li, L. Bian, K. Gu. A near-infrared miniature quantum dot spectrometer. Adv. Opt. Mater., 9, 2100376(2021).

    [18] S. H. Kong, D. D. L. Wijngaards, R. F. Wolffenbuttel. Infrared micro-spectrometer based on a diffraction grating. Sens. Actuators A, 92, 88-95(2001).

    [19] E. Arvin, H. Wu, G. de Graaf. Design and implementation of a sub-nm resolution microspectrometer based on a linear-variable optical filter. Opt. Express, 20, 489-507(2012).

    [20] X. Xue, H. Lv, Y. Qiu. Thermally stable high carrier mobility nanocomposite infrared photodetector. APL Photonics, 9, 046101(2024).

    [21] Y. Qiu, Y. Yu, S. Wang. Advances in quantum dot direct photolithographic patterning. ACS Mater. Lett., 6, 3176-3189(2024).

    [22] Y. Qiu, X. Zhou, X. Tang. Micro spectrometers based on materials nanoarchitectonics. Mater. (Basel), 16, 2253(2023).

    [23] N. Yan, Y. Qiu, X. He. Plasmonic enhanced nanocrystal infrared photodetectors. Materials, 16, 3216(2023).

    [24] Y. Qiu, N. Yan, H. Yao. Plasmon-enhanced HgTe colloidal quantum dot infrared photodetectors. Infrared Phys. Technol., 135, 104980(2023).

    [25] J. Wang, B. Pan, Z. Wang. Single-pixel p-graded-n junction spectrometers. Nat. Commun., 15, 1773(2024).

    [26] Y. Xiao, S. Wei, J. Xu. Superconducting single-photon spectrometer with 3D-printed photonic-crystal filters. ACS Photonics, 9, 3450-3456(2022).

    [27] Y. Ye, J. Zhang, Y. Xu. Miniature spectrometer based on gold nanorod-polyvinylpyrrolidone film. ACS Photonics, 11, 512-519(2024).

    [28] E. Huang, Q. Ma, Z. Liu. Etalon array reconstructive spectrometry. Sci. Rep., 7, 40693(2017).

    [29] J. Bao, M. G. Bawendi. A colloidal quantum dot spectrometer. Nature, 523, 67-70(2015).

    [30] X. Zhu, L. Bian, H. Fu. Broadband perovskite quantum dot spectrometer beyond human visual resolution. Light Sci. Appl., 9, 73(2020).

    [31] A. Chu, B. Martinez, S. Ferré. HgTe nanocrystals for SWIR detection and their integration up to the focal plane array. ACS Appl. Mater. Interfaces, 11, 33116-33123(2019).

    [32] S. Ahn, J. Y. Shang, S. K. Patel. Intercalated graphene and colloidal quantum dots for multispectral photodetection. Funct. Mater., 34, 2409523(2024).

    [33] Z. Yang, T. Albrow-Owen, H. Cui. Single nanowire spectrometers. Science, 365, 1017-1020(2019).

    [34] J. Liu, P. Liu, T. Shi. Flexible and broadband colloidal quantum dots photodiode array for pixel-level X-ray to near-infrared image fusion. Nat. Commun., 14, 5352(2023).

    [35] R. Alchaar, A. Khalili, N. Ledos. Focal plane array based on HgTe nanocrystals with photovoltaic operation in the short-wave infrared. Appl. Phys. Lett., 123, 111103(2023).

    [36] S. Keuleyan, E. Lhuillier, V. Brajuskovic. Mid-infrared HgTe colloidal quantum dot photodetectors. Nat. Photonics, 5, 489-493(2011).

    [37] X. Lan, M. Chen, M. H. Hudson. Quantum dot solids showing state-resolved band-like transport. Nat. Mater., 19, 323-329(2020).

    [38] X. Xue, M. Chen, Y. Luo. High-operating-temperature mid-infrared photodetectors via quantum dot gradient homojunction. Light Sci. Appl., 12, 2(2023).

    [39] H. Zhang, J. C. Peterson, P. Guyot-Sionnest. Intraband transition of HgTe nanocrystals for long-wave infrared detection at 12  μm. ACS Nano, 17, 7530-7538(2023).

    [40] X. Xue, Q. Hao, M. Chen. Very long wave infrared quantum dot photodetector up to 18  μm. Light Sci. Appl., 13, 89(2024).

    [41] Y. Luo, S. Zhang, X. Tang. Resonant cavity-enhanced colloidal quantum-dot dual-band infrared photodetectors. J. Mater. Chem., 10, 8218-8225(2022).

    [42] M. Faraji-Dana, E. Arbabi, A. Arbabi. Compact folded metasurface spectrometer. Nat. Commun., 9, 4196(2018).

    [43] M. Chen, X. Xue, T. Qin. Universal homojunction design for colloidal quantum dot infrared photodetectors. Adv. Mater. Technol., 8, 2300315(2023).

    [44] Y.-C. Tai, S. An, P.-R. Huang. Transfer-printing-enabled GeSn flexible resonant-cavity-enhanced photodetectors with strain-amplified mid-infrared optical responses. Nanoscale, 15, 7745-7754(2023).

    [45] S. Ghosh, K.-C. Lin, C.-H. Tsai. Resonant-cavity-enhanced responsivity in germanium-on-insulator photodetectors. Opt. Express, 28, 23739-23747(2020).

    [46] O. Ouellette, N. Hossain, B. R. Sutherland. Optical resonance engineering for infrared colloidal quantum dot photovoltaics. ACS Energy Lett., 1, 852-857(2016).

    [47] F. L. Pedrotti, L. M. Pedrotti, L. S. Pedrotti. Introduction to Optics(2017).

    [48] S. Yuan, D. Naveh, K. Watanabe. A wavelength-scale black phosphorus spectrometer. Nat. Photonics, 15, 601-607(2021).

    [49] C. Yao, K. Xu, W. Zhang. Integrated reconstructive spectrometer with programmable photonic circuits. Nat. Commun., 14, 6376(2023).

    [50] J. Meng, J. J. Cadusch, K. B. Crozier. Detector-only spectrometer based on structurally colored silicon nanowires and a reconstruction algorithm. Nano Lett., 20, 320-328(2020).

    [51] L. Huang, R. Luo, X. Liu. Spectral imaging with deep learning. Light Sci. Appl., 11, 61(2022).

    [52] Z. Wang, Z. Yu. Spectral analysis based on compressive sensing in nanophotonic structures. Opt. Express, 22, 25608-25614(2014).

    [53] X.-L. Wang, Y. Chen, Y. Chu. Spectrum reconstruction with filter-free photodetectors based on graded-band-gap perovskite quantum dot heterojunctions. ACS Appl. Mater. Interfaces, 14, 14455-14465(2022).

    [54] J. Wen, L. Hao, C. Gao. Deep learning-based miniaturized all-dielectric ultracompact film spectrometer. ACS Photonics, 10, 225-233(2022).

    [55] C. Yao, M. Chen, T. Yan. Broadband picometer-scale resolution on-chip spectrometer with reconfigurable photonics. Light Sci. Appl., 12, 156(2023).

    [56] C. C. Chang, H. Y. Lin. Spectrum reconstruction for on-chip spectrum sensor array using a novel blind nonuniformity correction method. IEEE Sens. J., 12, 2586-2592(2012).

    [57] Y. LeCun, Y. Bengio, G. Hinton. Deep learning. Nature, 521, 436-444(2015).

    [58] M. Yao, Z. Xiong, L. Wang. Spectral-depth imaging with deep learning based reconstruction. Opt. Express, 27, 38312-38325(2019).

    [59] L. Kades, J. M. Pawlowski, A. Rothkopf. Spectral reconstruction with deep neural networks. Phys. Rev., 102, 096001(2020).

    [60] C. A. T. Dos Santos, M. Lopo, R. N. Páscoa. A review on the applications of portable near-infrared spectrometers in the agro-food industry. Appl. Spectrosc., 67, 1215-1233(2013).

    [61] K. B. Beć, J. Grabska, H. W. Siesler. Handheld near-infrared spectrometers: where are we heading?. NIR News, 31, 28-35(2020).

    [62] G. Y. Belay, W. Hoving, A. Putdervan. Miniaturized broadband spectrometer based on a three-segment diffraction grating for spectral tissue sensing. Opt. Lasers Eng., 134, 106157(2020).

    [63] Y. M. Eltagoury, Y. M. Sabry, D. A. Khalil. All-silicon double-cavity Fourier-transform infrared spectrometer on-chip. Adv. Mater. Technol., 4, 1900441(2019).

    [64] T. Yang, M. AhmedAizaz, G. Zhang. Miniaturized spectrometers based on graded photonic crystal films. Opt. Express, 32, 25830-25838(2024).