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

Design of a bandgap-engineered barrier-blocking HOT HgCdTe long-wavelength infrared avalanche photodiode

Open Access Open Access

Abstract

The performance of high-operating-temperature (HOT) longwavelength infrared (LWIR) HgCdTe avalanche photodiodes (APDs) is significantly limited by the increasing dark current related to temperature. In this paper, a novel barrier-blocking LWIR pBp-APD structure is proposed and studied, and the results show that the dark current of pBp-APD is significantly restricted compared with conventional APD without sacrificing the gain at high temperature. Furthermore, the reduction of avalanche dark current is found to be the key points of the significant suppression of dark current. The physical essence of this reduction is revealed to be the depletion of carriers in the absorption region, and the feasibility of the improved structure is further confirmed by the analysis of its energy band and electric field distribution. In addition, the reduction of gain-normalized dark current (GNDC) does not need to sacrifice the gain. The proposed LWIR pBp-APD paves the way for development of high operation temperature infrared APDs.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

APD can realize high-speed, weak signal and even single photon detection due to its signal amplification effect, and has been widely used for decades in optical fiber communication, 3D lidar, astronomical observation and atmospheric detection [14]. Due to the huge difference of the ionization coefficient between electron and hole, HgCdTe (MCT) has been found to be the most powerful solution of low excess noise infrared APDs [511]. However, the application of HgCdTe APDs are significantly limited by its high dark current [1214]. Recent studies on HgCdTe APDs have shown that band-to-band tunneling (BBT) and trap-assisted tunneling (TAT) may be the main factors of dark current at low temperatures because of the narrow band gap and material quality [1517]. Typically, the dark current characteristics of HgCdTe APDs are much more complex in HOT regime. With the increase of operating temperature, the bandgap of HgCdTe will have a nonnegligible broadening, and some dark current components closely related to temperature also have a sharp increase. Therefore, the analysis of the dark current mechanisms of HgCdTe APDs at high temperature and reducing its dark current will be an important way of the developing of HOT HgCdTe APDs.

A potential choice for HOT HgCdTe detection is an unipolar barrier-blocking structure, which was introduced into HgCdTe in 2011 [1823]. The application of barrier-blocking devices on HgCdTe has been widely studied, and its inhibition effect on dark current has shown a broad prospect in the development of HOT HgCdTe detection [2431]. By designing the energy band structure of the barrier layer, the supplement of carriers in the absorption region could be blocked and suppress the Auger current, which is the main dark current component of HgCdTe photodetector at high temperature [3235]. However, pBp HgCdTe devices generally operate at the bias voltage below - 1 V, which is far lower than the high bias required for APD avalanche multiplication.

In this paper, a novel APD structure of pBp-APD based on barrier-blocking band engineering is proposed. By coupling barrier layer structure in APD, the carrier concentration in the absorption region can be effectively reduced. This makes all the temperature related dark current mechanisms to be less effective, allowing for a substantial increase of the operating temperature. By analyzing its physics characteristics such as the band structure and dark current, we proved that the introduction of the barrier layer does not affect the avalanche gain characteristics of the device, while the avalanche dark current component could be effectively suppressed.

2. Model and principle

In this paper, a novel pBp-APD Hg1-xCdxTe structure is designed and simulated by commercial software Sentaurus-TCAD, and x refers to the Cd fraction shown in Table 1. The drift-diffusion method is used in our calculations, thus the classical Poisson and continuity equations are self-coupled. The total dark current mechanisms consist of Shockley-Read-Hall (SRH), Auger, radiative, band-to-band tunneling, trap-assisted tunneling and impact ionization [36,37]:

$${R_{SRH}} = \frac{{pn - n_i^2}}{{{\tau _p}\left[ {n + {n_i}exp \left( {\frac{{{E_t} - {E_i}}}{{kT}}} \right)} \right] + {\tau _n}\left[ {p + {n_i}exp\left( {\frac{{{E_i} - {E_t}}}{{kT}}} \right)} \right]}}$$
$${R_{Auger}} = ({{B_1}n + {B_2}p} )({np - n_i^2} )$$
$${R_{Rad}} = 0.58 \times {10^{ - 12}}\sqrt \varepsilon {\left( {\frac{{{m_0}}}{{m_h^{\ast } + m_e^{\ast }}}} \right)^{ - 1.5}}\left( {1 + \frac{1}{{m_h^{\ast }}} + \frac{1}{{m_e^{\ast }}}} \right){\left( {\frac{{300}}{T}} \right)^{1.5}}({E_g^2 + 3kT{E_g} + 3.75k_B^2{T^2}} )$$
$${G_{BBT}} = \frac{{{q^2}\sqrt {2m_e^{\ast }} {E^2}({{V_{bias}} - {V_d}} )}}{{4{\pi ^3}{\hbar ^2}\sqrt {{E_g}} }}exp\left( { - \frac{{\pi \sqrt {m_e^{\ast }/2} E_g^{3/2}}}{{2qE\hbar }}} \right)$$
$${R_{TAT}} = \frac{{pn - n_i^2}}{{\frac{{{\tau _p}}}{{1 + {\Gamma _p}}}\left[ {n + {n_i}exp \left( {\frac{{{E_t} - {E_i}}}{{kT}}} \right)} \right] + \frac{{{\tau _n}}}{{1 + {\Gamma _n}}}\left[ {p + {n_i}exp\left( {\frac{{{E_i} - {E_t}}}{{kT}}} \right)} \right]}}$$
where E is the electric field,${\; }{\tau _{n,p}}$is the doping-temperature-field dependent lifetime, B1,2 is the temperature-dependent Auger coefficient, Eg is the bandgap, and Et and Nt are the trap level and density respectively. The extended formulas for the coefficients ${\Gamma _p}$ and ${\Gamma _n}$ could be found in Ref. [16]. The impact ionization model is considered by the Okuto–Crowel model, which has been demonstrated to be in agreement with the experimental results in HgCdTe APDs, given by [16]:
$${G_{avalanche}} = {\alpha _n}n{\nu _n} + {\alpha _p}p{\nu _p}$$
$${\alpha _{n,p}} = {a_{n,p}}{E^c}exp({ - b/E} )$$
where the temperature-dependent coefficient ${a_{n,p}}$ and b are referred to the slope of the gain curve and the scaling of the gain curve as a function of junction width at the low bias, respectively. The fitting coefficient c equal to 0.6 is reported to give good results [38].

Tables Icon

Table 1. Key parameters of pBp-APD and conventional mesa-APD used in the simulations.

The barrier-blocking pBp structure is coupled into APD in the improved pBp-APD, and they share the same absorption region, as shown in Fig. 1(a), and Fig. 1(b) shows the structure of conventional mesa APD for comparison.

 figure: Fig. 1.

Fig. 1. Schematic of the simulated (a) conventional APD and (b) pBp-APD, and the simulated energy bands of (c) conventional APD and (d) pBp-APD at -7V, where V denotes the reveres bias. The insert of (c) shows the comparison between the experimental results and the simulation results of the model of conventional APDs demonstrating the validity of the numerical simulations.

Download Full Size | PDF

Commonly, the working bias of nBn or pBp is not higher than - 1 V. Higher reverse bias will significantly reduce the conduction barrier and destroy the electron blocking effect. However, a much higher bias is required in APD to obtain a high enough built-in electric field in the multiplication layer for a high avalanche gain. The band diagrams of conventional APD and pBp-APD structure at -7 V are shown in Figs. 1(c) and 1(d). It could be seen that the conduction barrier of the barrier layer of pBp-APD is about 0.13 eV at - 7 V reverse bias, which can ensure that the structure can maintain excellent electronic blocking effect and suppress dark current even at high avalanche bias. This prevents the replenishment of electrons into the absorption region, and the electrons in the absorption region are extracted to the multiplication layer and avalanched. Also, the holes in the absorption region flow to the electrode layer under bias, and can hardly be recovered from the heavily n-doped contact region. Thus, the carrier concentration in the absorption layer could be reduced and reduce the dark current. What is shown in the insert of Fig. 1(c) is the simulation results of the model of conventional APDs built compared with our experimental results, and it shows that the simulation model fits well with our experiment. The detailed structure parameters used in the simulations are shown in Table 1.

Research shows that TAT and Shockley-Read-Hall (SRH) currents dominate the dark current of APDs at the low reverse bias at 77K, while BBT current dominates at the higher bias [39]. However, the temperature-related generation-recombination (G-R) and avalanche current should not be neglected as the operating temperature increases. The temperature-dependent dark current characteristics of the conventional APD are shown in Fig. 2(a). The dark current of APD increases monotonously with the increase of temperature at the lower bias voltage, but the trend is chaotic under the higher reverse bias. It could be easily understood because the bandgap of HgCdTe will widen with the increase of temperature, which will lead to a significant reduction of the tunneling current. At higher reverse bias, the dark temperature mechanism of APD will be the competition between the tunneling current which decreases with the increase of temperature and other dark current compositions which commonly increase at a higher temperature. However, at the operating temperature above 180K, the dark current increases steadily with the increase of temperature, and the tunneling current starts to lose its dominance. This is due to that high temperature leads to higher carrier concentration in the absorption region, and significantly increases the avalanche dark current.

 figure: Fig. 2.

Fig. 2. Dark currents of (a) conventional HgCdTe APD and (b) HgCdTe pBp-APD under different temperatures in the whole bias range. The dark currents at -0.5 V and-7 V are listed separately in (c) and (d), and the insert shows the comparison of conventional HgCdTe APD and pBp-APD at 260 K.

Download Full Size | PDF

However, the avalanche current is significantly suppressed, and tunneling current absolutely dominates the total dark current under high bias. At lower reverse bias, the generation-recombination current increases significantly with the increase of temperature and gradually dominate the dark current, and the dark current increases significantly with the increase of operating temperature, as shown in Fig. 2(c). However, at higher bias, the tunneling current always dominates dark current because of the suppression of generation-recombination and avalanche dark current. This is due to the electron depletion in the absorption layer, which will be shown in Fig. 4. The band gap broadening caused by increasing temperature will lead to the decrease of tunneling, so the dark current of the device exhibits a negative temperature gradient at high bias voltage.

To further understand the dark current reducing characteristics of the advanced pBp-APD structure, the vertical distribution of the tunneling generation and impact ionization rate of conventional HgCdTe APDs and pBp-APD at 260K are shown in Fig. 3. It could be seen that both the avalanches and tunneling of conventional APD and pBp-APD occur in the junction region of the multiplication layer. However, the impact ionization of pBp-APD has a significant reduction compared with the conventional APD. This further proves that the principle of pBp-APD reducing dark current is to reduce the avalanche current.

 figure: Fig. 3.

Fig. 3. Vertical distributions of tunneling generation and impact ionization of (a) conventional APD and (b) pBp-APD at -7 V.

Download Full Size | PDF

Previous research shows that one of the factors determining the avalanche current in APD is the number of carriers injected into the multiplication layer [40]. Therefore, an important method to suppress the avalanche dark-current of MWIR/LWIR HgCdTe APD devices is to reduce the electron density in the absorption region. Figure 4 shows the vertical distribution of electron and hole density of conventional APD and pBp-APD at - 7V. It could be obviously seen that the electrons in the absorption region of the pBp-APD structure are significantly depleted to even far lower than the intrinsic carrier concentration, thus reducing the avalanche current at high temperatures. The vertical distribution of electric field and potential are also shown in Fig. 4. It could be seen that there exists a very high electric field in the multiplication layer, which is necessary for APD devices to multiply electrons in the multiplication layer. Moreover, the potential difference is mostly distributed in the multiplication layer, and only a very small potential difference is applied to the barrier layer. Overall, it could be proved that the addition of the barrier layer can significantly reduce the dark current by depleting the electrons in the absorption region, and the high bias required by APD does not affect the barrier layer conduction barrier.

 figure: Fig. 4.

Fig. 4. Simulated vertical distributions of carrier densities, electric field, and electric potential for (a) conventional APD and (b) pBp-APD at -7 V.

Download Full Size | PDF

Finally, the gain M and GNDC (defined as ${I_{dark}}/M$) characteristics of the two structures are both calculated in Fig. 5. The existence of valence band-offset in the barrier layer will affect the flow of photogenerated holes, therefore reduce the gain of pBp-APD. However, the results at -7 V show that the dark current could be reduced without sacrificing the gain. This is because that the valence band-offset could be flattened at higher bias so as not to block the photogenerated holes, and the electric field concentration in the multiplication layer as shown in Fig. 3 could provide enough avalanche gain.

 figure: Fig. 5.

Fig. 5. (a) Gain and (b) GNDC characteristics of conventional APD and pBp-APD.

Download Full Size | PDF

3. Conclusion

In this paper, a novel pBp-APD structure is designed and numerically analyzed in detail. The results show that the dark current of conventional APD at high operating temperature could be significantly reduced by coupling barrier blocking layer. The analysis of the energy band and electric field of pBp-APD show that the conduction barrier of pBp is not destroyed by the high bias voltage, and the electric field in the multiplication layer is maintained to be large enough to produce multiplication. By discussing the mechanism of dark current of pBp-APD at different temperatures and bias voltages, and analyzing the temperature-dependent characteristics, it could be proved that the avalanche dark current is sufficiently suppressed at high temperature and reverse bias. In addition, the dark current reduction does not affect its gain performance. The results will provide a very potential solution for the development of MWIR or LWIR HOT APDs.

Funding

National Key Research and Development Program of China (2020YFB2009300, 2017YFA0205801); National Natural Science Foundation of China (61521005, 61725505, 61905266); Shanghai Sailing Program (19YF1454600); Key Research Project of Frontier Science of CAS (QYZDB-SSW-JSC031); Natural Science Foundation of Shanghai (18ZR1445800, 18ZR1445900, 19XD1404100).

Disclosures

The authors declare no conflicts of interest.

References

1. S. Velicu, R. Ashokan, and S. Sivananthan, “A model for dark current and multiplication in HgCdTe avalanche photodiodes,” J. Electron. Mater. 29(6), 823–827 (2000). [CrossRef]  

2. S. Cova, M. Ghioni, A. Lacaita, C. Samori, and F. Zappa, “Avalanche photodiodes and quenching circuits for single-photon detection,” Appl. Opt. 35(12), 1956 (1996). [CrossRef]  

3. B. F. Aull, A. H. Loomis, D. J. Young, A. Stern, B. J. Felton, P. J. Daniels, D. J. Landers, L. Retherford, D. D. Rathman, R. M. Heinrichs, R. M. Marino, D. G. Fouche, M. A. Albota, R. E. Hatch, G. S. Rowe, D. G. Kocher, J. G. Mooney, M. E. O’Brien, B. E. Player, B. C. Willard, Z. L. Liau, and J. J. Zayhowski, “Three-dimensional imaging with arrays of Geiger-mode avalanche photodiodes,” Proc. SPIE 5353, 105–116 (2004). [CrossRef]  

4. W. Hu, Q. Li, X. Chen, and W. Lu, “Recent progress on advanced infrared photodetectors,” Acta. Phys. Sin. 68(12), 7–41 (2019). [CrossRef]  

5. W. Lei, J. Antoszewski, and L. Faraone, “Progress, challenges, and opportunities for HgCdTe infrared materials and detectors,” Appl. Phys. Rev. 2(4), 041303 (2015). [CrossRef]  

6. V. S. Varavin, I. V. Sabinina, G. Y. Sidorov, D. V. Marin, V. G. Remesnik, A. V. Predein, S. A. Dvoretsky, V. V. Vasilyev, Y. G. Sidorov, M. V. Yakushev, and A. V. Latyshev, “Photodiodes based on p-on-n junctions formed in MBE-grown n-type MCT absorber layers for the spectral region 8 to 11 µm,” Infrared Phys. Technol. 105, 103182 (2020). [CrossRef]  

7. W. D. Lawson, S. Nielsen, E. H. Putley, and A. S. Young, “Preparation and properties of HgTe and mixed crystals of HgTe-CdTe,” J. Phys. Chem. Solids 9(3-4), 325–329 (1959). [CrossRef]  

8. A. Rogalski, “Recent progress in infrared detector technologies,” Infrared Phys. Technol. 54(3), 136–154 (2011). [CrossRef]  

9. A. Singh, V. Srivastav, and R. Pal, “HgCdTe avalanche photodiodes: A review,” Opt. Laser Technol. 43(7), 1358–1370 (2011). [CrossRef]  

10. J. D. Beck, C. Wan, M. A. Kinch, and J. E. Robinson, “MWIR HgCdTe avalanche photodiodes,” Proc. SPIE 4454, 188–197 (2001). [CrossRef]  

11. J. D. Beck, C. F. Wan, M. A. Kinch, J. E. Robinson, P. Mitra, R. Scritchfield, and J. C. Campbell, “The HgCdTe electron avalanche photodiode,” J. Electron. Mater. 35(6), 1166–1173 (2006). [CrossRef]  

12. G. Perrais, O. Gravrand, J. Baylet, G. Destefanis, and J. Rothman, “Gain and Dark Current Characteristics of Planar HgCdTe Avalanche Photo Diodes,” J. Electron. Mater. 36(8), 963–970 (2007). [CrossRef]  

13. J. Rothman, G. Perrais, G. Destefanis, J. Baylet, P. Castelein, and J. P. Chamonal, “High performance characteristics in pin MW HgCdTe e-APDs,” Proc. SPIE 6542, 654219 (2007). [CrossRef]  

14. J. D. Beck, R. Scritchfield, P. Mitra, W. W. Sullivan, A. D. Gleckler, R. Strittmatter, and R. J. Martin, “Linear mode photon counting with the noiseless gain HgCdTe e-avalanche photodiode,” Opt. Eng. 53(8), 081905 (2014). [CrossRef]  

15. S. Bailey, W. McKeag, J. Wang, M. Jack, and F. Amzajerdian, “Advances in HgCdTe APDs and LADAR receivers,” Proc. SPIE 7660, 76603I (2010). [CrossRef]  

16. W. Qiu, W. Hu, L. Chen, C. Lin, X. Cheng, X. Chen, and W. Lu, “Dark Current Transport and Avalanche Mechanism in HgCdTe Electron-Avalanche Photodiodes,” IEEE Trans. Electron Devices 62(6), 1926–1931 (2015). [CrossRef]  

17. X. Sun, J. B. Abshire, J. D. Beck, P. Mitra, K. Reiff, and G. Yang, “HgCdTe avalanche photodiode detectors for airborne and spaceborne lidar at infrared wavelengths,” Opt. Express 25(14), 16589 (2017). [CrossRef]  

18. P. Martyniuk and A. Rogalski, “MWIR barrier detectors versus HgCdTe photodiodes,” Infrared Phys. Technol. 70, 125–128 (2015). [CrossRef]  

19. A. M. Itsuno, J. D. Phillips, and S. Velicu, “Design of an Auger-Suppressed Unipolar HgCdTe NBνN Photodetector,” J. Electron. Mater. 41(10), 2886–2892 (2012). [CrossRef]  

20. S. Maimon and G. W. Wicks, “nBn detector, an infrared detector with reduced dark current and higher operating temperature,” Appl. Phys. Lett. 89(15), 151109 (2006). [CrossRef]  

21. W. C. Qiu, T. Jiang, and X. A. Cheng, “A bandgap-engineered HgCdTe PBπn long-wavelength infrared detector,” J. Appl. Phys. 118(12), 124504 (2015). [CrossRef]  

22. A. M. Itsuno, J. D. Phillips, and S. Velicu, “Design and Modeling of HgCdTe nBn Detectors,” J. Electron. Mater. 40(8), 1624–1629 (2011). [CrossRef]  

23. F. Uzgur and S. Kocaman, “Barrier engineering for HgCdTe unipolar detectors on alternative substrates,” Infrared Phys. Technol. 97, 123–128 (2019). [CrossRef]  

24. M. Kopytko, A. Keblowski, W. Gawron, P. Madejczyk, A. Kowalewski, and K. Joźwikowski, “High-operating temperature MWIR nBn HgCdTe detector grown by MOCVD,” Opto-Electron. Rev. 21(4), 402–405 (2013). [CrossRef]  

25. M. Kopytko, A. Kębłowski, W. Gawron, and W. Pusz, “LWIR HgCdTe barrier photodiode with Auger-suppression,” Semicond. Sci. Technol. 31(3), 035025 (2016). [CrossRef]  

26. M. Kopytko and K. Jóźwikowski, “Numerical Analysis of Current–Voltage Characteristics of LWIR nBn and p-on-n HgCdTe Photodetectors,” J. Electron. Mater. 42(11), 3211–3216 (2013). [CrossRef]  

27. N. D. Akhavan, G. A. Umana-Membreno, R. Gu, J. Antoszewski, and L. Faraone, “Delta Doping in HgCdTe-Based Unipolar Barrier Photodetectors,” IEEE T. Electron. Dev. 65(10), 4340–4345 (2018). [CrossRef]  

28. M. Kopytko, “Design and modelling of high-operating temperature MWIR HgCdTe nBn detector with n- and p-type barriers,” Infrared Phys. Technol. 64, 47–55 (2014). [CrossRef]  

29. P. Madejczyk, W. Gawron, P. Martyniuk, A. Keblowski, W. Pusz, J. Pawluczyk, M. Kopytko, J. Rutkowski, A. Rogalski, and J. Piotrowski, “Engineering steps for optimizing high temperature LWIR HgCdTe photodiodes,” Infrared Phys. Technol. 81, 276–281 (2017). [CrossRef]  

30. M. Kopytko, J. Wróbel, K. Jóźwikowski, A. Rogalski, J. Antoszewski, N. D. Akhavan, G. A. Umana-Membreno, L. Faraone, and C. R. Becker, “Engineering the Bandgap of Unipolar HgCdTe-Based nBn Infrared Photodetectors,” J. Electron. Mater. 44(1), 158–166 (2015). [CrossRef]  

31. M. Kopytko and K. Jozwikowski, “Generation-Recombination Effect in MWIR HgCdTe Barrier Detectors for High-Temperature Operation,” IEEE T. Electron. Dev. 62(7), 2278–2284 (2015). [CrossRef]  

32. M. Vallone, M. Goano, F. Bertazzi, G. Ghione, S. Hanna, D. Eich, A. Sieck, and H. Figgemeier, “Constraints and performance trade-offs in Auger-suppressed HgCdTe focal plane arrays,” Appl. Opt. 59(17), E1–E8 (2020). [CrossRef]  

33. A. Rogalski, M. Kopytko, and P. Martyniuk, “Performance prediction of p-i-n HgCdTe long-wavelength infrared HOT photodiodes,” Appl. Opt. 57(18), D11–D19 (2018). [CrossRef]  

34. M. Vallone, M. Goano, F. Bertazzi, G. Ghione, A. Palmieri, S. Hanna, D. Eich, and H. Figgemeier, “Reducing inter-pixel crosstalk in HgCdTe detectors,” Opt. Quantum Electron. 52(1), 25 (2020). [CrossRef]  

35. P. Wang, H. Xia, Q. Li, F. Wang, L. Zhang, T. Li, P. Martyniuk, A. Rogalski, and W. Hu, “Sensing Infrared Photons at Room Temperature: From Bulk Materials to Atomic Layers,” Small 15(46), 1904396 (2019). [CrossRef]  

36. W. Hu, Handbook of Optoelectronic Device Modelling and Simuation. CRC Press, 307–336 (2017).

37. R. Hall, “Recombination processes in semiconductors,” Proc. Inst. Electr. Eng., Part B 106(17S), 923–931 (1959). [CrossRef]  

38. J. Rothman, L. Mollard, S. Bosson, G. Vojetta, K. Foubert, S. Gatti, G. Bonnouvrier, F. Salveti, A. Kerlain, and O. Pacaud, “Short-Wave Infrared HgCdTe Avalanche Photodiodes,” J. Electron. Mater. 41(10), 2928–2936 (2012). [CrossRef]  

39. Q. Li, J. He, W. Hu, L. Chen, X. Chen, and W. Lu, “Influencing Sources for Dark Current Transport and Avalanche Mechanisms in Planar and Mesa HgCdTe p-i-n Electron-Avalanche Photodiodes,” IEEE T. Electron. Dev. 65(2), 572–576 (2018). [CrossRef]  

40. S. L. Miller, “Avalanche Breakdown in Gerimanium,” Phys. Rev. 99(4), 1234–1241 (1955). [CrossRef]  

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 (5)

Fig. 1.
Fig. 1. Schematic of the simulated (a) conventional APD and (b) pBp-APD, and the simulated energy bands of (c) conventional APD and (d) pBp-APD at -7V, where V denotes the reveres bias. The insert of (c) shows the comparison between the experimental results and the simulation results of the model of conventional APDs demonstrating the validity of the numerical simulations.
Fig. 2.
Fig. 2. Dark currents of (a) conventional HgCdTe APD and (b) HgCdTe pBp-APD under different temperatures in the whole bias range. The dark currents at -0.5 V and-7 V are listed separately in (c) and (d), and the insert shows the comparison of conventional HgCdTe APD and pBp-APD at 260 K.
Fig. 3.
Fig. 3. Vertical distributions of tunneling generation and impact ionization of (a) conventional APD and (b) pBp-APD at -7 V.
Fig. 4.
Fig. 4. Simulated vertical distributions of carrier densities, electric field, and electric potential for (a) conventional APD and (b) pBp-APD at -7 V.
Fig. 5.
Fig. 5. (a) Gain and (b) GNDC characteristics of conventional APD and pBp-APD.

Tables (1)

Tables Icon

Table 1. Key parameters of pBp-APD and conventional mesa-APD used in the simulations.

Equations (7)

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

R S R H = p n n i 2 τ p [ n + n i e x p ( E t E i k T ) ] + τ n [ p + n i e x p ( E i E t k T ) ]
R A u g e r = ( B 1 n + B 2 p ) ( n p n i 2 )
R R a d = 0.58 × 10 12 ε ( m 0 m h + m e ) 1.5 ( 1 + 1 m h + 1 m e ) ( 300 T ) 1.5 ( E g 2 + 3 k T E g + 3.75 k B 2 T 2 )
G B B T = q 2 2 m e E 2 ( V b i a s V d ) 4 π 3 2 E g e x p ( π m e / 2 E g 3 / 2 2 q E )
R T A T = p n n i 2 τ p 1 + Γ p [ n + n i e x p ( E t E i k T ) ] + τ n 1 + Γ n [ p + n i e x p ( E i E t k T ) ]
G a v a l a n c h e = α n n ν n + α p p ν p
α n , p = a n , p E c e x p ( b / E )
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.