Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Actively variable-spectrum optoelectronics with black phosphorus

Abstract

Room-temperature optoelectronic devices that operate at short-wavelength and mid-wavelength infrared ranges (one to eight micrometres) can be used for numerous applications1,2,3,4,5. To achieve the range of operating wavelengths needed for a given application, a combination of materials with different bandgaps (for example, superlattices or heterostructures)6,7 or variations in the composition of semiconductor alloys during growth8,9 are used. However, these materials are complex to fabricate, and the operating range is fixed after fabrication. Although wide-range, active and reversible tunability of the operating wavelengths in optoelectronic devices after fabrication is a highly desirable feature, no such platform has been yet developed. Here we demonstrate high-performance room-temperature infrared optoelectronics with actively variable spectra by presenting black phosphorus as an ideal candidate. Enabled by the highly strain-sensitive nature of its bandgap, which varies from 0.22 to 0.53 electronvolts, we show a continuous and reversible tuning of the operating wavelengths in light-emitting diodes and photodetectors composed of black phosphorus. Furthermore, we leverage this platform to demonstrate multiplexed nondispersive infrared gas sensing, whereby multiple gases (for example, carbon dioxide, methane and water vapour) are detected using a single light source. With its active spectral tunability while also retaining high performance, our work bridges a technological gap, presenting a potential way of meeting different requirements for emission and detection spectra in optoelectronic applications.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Strain-tunable bandgap in black phosphorus.
Fig. 2: Strain-induced modification of the bandgap in bP.
Fig. 3: Strain-tunable MWIR LEDs based on a bP–MoS2 heterostructure.
Fig. 4: NDIR gas sensing using strain-tunable LEDs.

Similar content being viewed by others

Data availability

All data generated or analysed during this study are included in this published article. Source data are provided with this paper.

References

  1. Kahn, J. M. & Barry, J. R. Wireless infrared communications. Proc. IEEE 85, 265–298 (1997).

    Article  Google Scholar 

  2. Vollmer, M. & Mollmann, K.-P. Infrared Thermal Imaging: Fundamentals, Research and Applications (2nd edn) (Wiley-VCH, Weinheim, 2018).

  3. Bagavathiappan, S., Lahiri, B. B., Saravanan, T., Philip, J. & Jayakumar, T. Infrared thermography for condition monitoring—a review. Infrared Phys. Technol. 60, 35–55 (2013).

    Article  ADS  Google Scholar 

  4. Baker, M. J. et al. Using Fourier transform IR spectroscopy to analyze biological materials. Nat. Protocols 9, 1771–1791 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Gibson, D. & Macgregor, C. A novel solid state non-dispersive infrared CO2 gas sensor compatible with wireless and portable deployment. Sensors (Basel) 13, 7079–7103 (2013).

    Article  ADS  CAS  Google Scholar 

  6. Haugan, H. J., Szmulowicz, F., Brown, G. J. & Mahalingam, K. Bandgap tuning of InAs/GaSb type-II superlattices for mid-infrared detection. J. Appl. Phys. 96, 2580–2585 (2004).

    Article  ADS  CAS  Google Scholar 

  7. Kang, J., Tongay, S., Zhou, J., Li, J. & Wu, J. Band offsets and heterostructures of two-dimensional semiconductors. Appl. Phys. Lett. 102, 012111 (2013).

    Article  ADS  CAS  Google Scholar 

  8. Wu, J. et al. Universal bandgap bowing in group-III nitride alloys. Solid State Commun. 127, 411–414 (2003).

    Article  ADS  CAS  Google Scholar 

  9. Ning, C.-Z., Dou, L. & Yang, P. Bandgap engineering in semiconductor alloy nanomaterials with widely tunable compositions. Nat. Rev. Mater. 2, 17070 (2017).

    Article  ADS  CAS  Google Scholar 

  10. Yang, Z. et al. Single-nanowire spectrometers. Science 365, 1017–1020 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  11. Kramer, I. J., Levina, L., Debnath, R., Zhitomirsky, D. & Sargent, E. H. Solar cells using quantum funnels. Nano Lett. 11, 3701–3706 (2011).

    Article  ADS  CAS  PubMed  Google Scholar 

  12. Whitney, W. S. et al. Field effect optoelectronic modulation of quantum-confined carriers in black phosphorus. Nano Lett. 17, 78–84 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Liu, Y. et al. Gate-tunable giant stark effect in few-layer black phosphorus. Nano Lett. 17, 1970–1977 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Yablonovitch, E. & Kane, E. O. Band structure engineering of semiconductor lasers for optical communications. J. Lightwave Technol. 6, 1292–1299 (1988).

    Article  ADS  CAS  Google Scholar 

  15. Thompson, S. E. et al. A 90-nm logic technology featuring strained-silicon. IEEE Trans. Electron Dev. 51, 1790–1797 (2004).

    Article  ADS  CAS  Google Scholar 

  16. Chen, Y. et al. Strain engineering and epitaxial stabilization of halide perovskites. Nature 577, 209 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Lee, C., Wei, X., Kysar, J. W. & Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 321, 385–388 (2008).

    Article  ADS  CAS  PubMed  Google Scholar 

  18. Bertolazzi, S., Brivio, J. & Kis, A. Stretching and breaking of ultrathin MoS2. ACS Nano 5, 9703–9709 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Rodin, A. S., Carvalho, A. & Neto, A. C. Strain-induced gap modification in black phosphorus. Phys. Rev. Lett. 112, 176801 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Quereda, J. et al. Strong modulation of optical properties in black phosphorus through strain-engineered rippling. Nano Lett. 16, 2931–2937 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Zhang, Z. et al. Strain-modulated bandgap and piezo-resistive effect in black phosphorus field-effect transistors. Nano Lett. 17, 6097–6103 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Zhang, G. et al. Infrared fingerprints of few-layer black phosphorus. Nat. Commun. 8, 14071 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Çakır, D., Sahin, H. & Peeters, F. M. Tuning of the electronic and optical properties of single-layer black phosphorus by strain. Phys. Rev. B Condens. Matter Mater. Phys. 90, 205421 (2014).

    Article  ADS  CAS  Google Scholar 

  24. Huang, S. et al. Strain-tunable van der Waals interactions in few-layer black phosphorus. Nat. Commun. 10, 2447 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  25. Ma, W. et al. Piezoelectricity in multilayer black phosphorus for piezotronics and nanogenerators. Adv. Mater. 32, 1905795 (2020).

    Article  CAS  Google Scholar 

  26. Sanchez-Perez, J. R. et al. Direct-bandgap light-emitting germanium in tensilely strained nanomembranes. Proc. Natl Acad. Sci. USA 108, 18893–18898 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Takei, K. et al. Quantum confinement effects in nanoscale-thickness InAs membranes. Nano Lett. 11, 5008–5012 (2011).

    Article  ADS  CAS  PubMed  Google Scholar 

  28. Ling, X., Wang, H., Huang, S., Xia, F. & Dresselhaus, M. S. The renaissance of black phosphorus. Proc. Natl Acad. Sci. USA 112, 4523–4530 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ge, S. et al. Dynamical evolution of anisotropic response in black phosphorus under ultrafast photoexcitation. Nano Lett. 15, 4650–4656 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Bhaskar, P., Achtstein, A. W., Vermeulen, M. J. W. & Siebbeles, L. D. A. Radiatively dominated charge carrier recombination in black phosphorus. J. Phys. Chem. C 120, 13836–13842 (2016).

    Article  CAS  Google Scholar 

  31. Chen, C. et al. Bright mid-infrared photoluminescence from thin-film black phosphorus. Nano Lett. 19, 1488–1493 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Du, Y. et al. Auxetic black phosphorus: a 2D material with negative Poisson’s ratio. Nano Lett. 16, 6701–6708 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  33. Wang, J. et al. Mid-infrared polarized emission from black phosphorus light-emitting diodes. Nano Lett. 20, 3651–3655 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  34. Zong, X. et al. Black phosphorus-based van der Waals heterostructures for mid-infrared light-emission applications. Light Sci. Appl. 9, 114 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  35. Chang, T.-Y. et al. Black phosphorus mid-infrared light-emitting diodes integrated with silicon photonic waveguides. Nano Lett. 20, 6824–6830 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Haug, A. Auger recombination in direct-gap semiconductors: band-structure effects. J. Phys. C 16, 4159 (1983).

    Article  ADS  CAS  Google Scholar 

  37. Kurtz, S. R., Biefeld, R. M. & Dawson, L. R. Modification of valence-band symmetry and Auger threshold energy in biaxially compressed InAs1–xSbx. Phys. Rev. B 51, 7310 (1995).

    Article  ADS  CAS  Google Scholar 

  38. Lee, D.-D. & Lee, D.-S. Environmental gas sensors. IEEE Sens. J. 1, 214–224 (2001).

    Article  ADS  CAS  Google Scholar 

  39. Dinh, T.-V., Choi, I.-Y., Son, Y.-S. & Kim, J.-C. A review on non-dispersive infrared gas sensors: improvement of sensor detection limit and interference correction. Sens. Actuators B 231, 529–538 (2016).

    Article  CAS  Google Scholar 

  40. Gomes, J., Rodrigues, J. J., Rabêlo, R. A., Kumar, N. & Kozlov, S. IoT-enabled gas sensors: technologies, applications, and opportunities. J. Sens. Actuator Netw. 8, 57 (2019).

    Article  Google Scholar 

  41. Deng, Y. et al. Black phosphorus–monolayer MoS2 van der Waals heterojunction p-n diode. ACS Nano 8, 8292–8299 (2014).

    Article  CAS  PubMed  Google Scholar 

  42. Youngblood, N., Chen, C., Koester, S. J. & Li, M. Waveguide-integrated black phosphorus photodetector with high responsivity and low dark current. Nat. Photonics 9, 247–252 (2015).

    Article  ADS  CAS  Google Scholar 

  43. Yuan, H. et al. Polarization-sensitive broadband photodetector using a black phosphorus vertical p–n junction. Nat. Nanotechnol. 10, 707–713 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Guo, Q. et al. Black phosphorus mid-infrared photodetectors with high gain. Nano Lett. 16, 4648–4655 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  45. Huang, M. et al. Broadband black-phosphorus photodetectors with high responsivity. Adv. Mater. 28, 3481–3485 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  46. Chen, X. et al. Widely tunable black phosphorus mid-infrared photodetector. Nat. Commun. 8, 1672 (2017).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  47. Bullock, J. et al. Polarization-resolved black phosphorus/molybdenum disulfide mid-wave infrared photodiodes with high detectivity at room temperature. Nat. Photonics 12, 601–607 (2018).

    Article  ADS  CAS  Google Scholar 

  48. Amani, M., Regan, E., Bullock, J., Ahn, G. H. & Javey, A. Mid-wave infrared photoconductors based on black phosphorus-arsenic alloys. ACS Nano 11, 11724–11731 (2017).

    Article  CAS  PubMed  Google Scholar 

  49. Wu, J., Mao, N., Xie, L., Xu, H. & Zhang, J. Identifying the crystalline orientation of black phosphorus using angle‐resolved polarized Raman spectroscopy. Angew. Chem. Int. Ed. 54, 2366–2369 (2015).

    Article  CAS  Google Scholar 

  50. Castellanos-Gomez, A. et al. Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping. 2D Mater. 1, 011002 (2014).

    Article  CAS  Google Scholar 

  51. Favron, A. et al. Photooxidation and quantum confinement effects in exfoliated black phosphorus. Nat. Mater. 14, 826–832 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Foo, E., Jaafar, M., Aziz, A. & Sim, L. C. Properties of spin coated epoxy/silica thin film composites: effect of nano- and micron-size fillers. Compos. Part A Appl. Sci. Manuf. 42, 1432–1437 (2011).

    Article  CAS  Google Scholar 

  53. Mei, H., Landis, C. M. & Huang, R. Concomitant wrinkling and buckle-delamination of elastic thin films on compliant substrates. Mech. Mater. 43, 627–642 (2011).

    Article  Google Scholar 

  54. Madelung, O. Semiconductors: Data Handbook (3rd edn) (Springer, Berlin, 2004).

  55. Kim, H. et al. Synthetic WSe2 monolayers with high photoluminescence quantum yield. Sci. Adv. 5, eaau4728 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  56. Gramling, H. M. et al. Spatially precise transfer of patterned monolayer WS2 and MoS2 with features larger than 104 μm2 directly from multilayer sources. ACS Appl. Electron. Mater. 1, 407–416 (2019).

    Article  CAS  Google Scholar 

  57. Nguyen, V. et al. Deterministic assembly of arrays of lithographically defined WS2 and MoS2 monolayer features directly from multilayer sources into van der Waals heterostructures. J. Micro Nano-Manuf. 7, 041006 (2019).

    Article  CAS  Google Scholar 

  58. Huang, Y. et al. Interaction of black phosphorus with oxygen and water. Chem. Mater. 28, 8330–8339 (2016).

    Article  CAS  Google Scholar 

  59. Qiao, J., Kong, X., Hu, Z. X., Yang, F. & Ji, W. High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus. Nat. Commun. 5, 4475 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  60. Desai, S. B. et al. Strain-induced indirect to direct bandgap transition in multilayer WSe2. Nano Lett. 14, 4592–4597 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  61. Matthewson, M. J., Kurkjian, C. R. & Gulati, S. T. Strength measurement of optical fibers by bending. J. Am. Ceram. Soc. 69, 815–821 (1986).

    Article  CAS  Google Scholar 

  62. Plechinger, G. et al. Control of biaxial strain in single-layer molybdenite using local thermal expansion of the substrate. 2D Mater. 2, 015006 (2015).

    Article  CAS  Google Scholar 

  63. Luo, W., Song, Q., Zhou, G., Tuschel, D. & Xia, G. Study of black phosphorus using angle-resolved polarized Raman spectroscopy with 442 nm excitation. Preprint at https://arxiv.org/abs/1610.03382 (2016).

  64. Zhang, Z. M., Lefever-Button, G. & Powell, F. R. Infrared refractive index and extinction coefficient of polyimide films. Int. J. Thermophys. 19, 905–916 (1998).

    Article  CAS  Google Scholar 

  65. Sherrott, M. C. et al. Anisotropic quantum well electro-optics in few-layer black phosphorus. Nano Lett. 19, 269–276 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  66. Beal, A. R. & Hughes, H. P. Kramers-Kronig analysis of the reflectivity spectra of 2H-MoS2, 2H-MoSe2 and 2H-MoTe2. J. Phys. Chem. 12, 881 (1979).

    CAS  Google Scholar 

  67. Hori, Y., Ando, Y., Miyamoto, Y. & Sugino, O. Effect of strain on band structure and electron transport in InAs. Solid-State Electron. 43, 1813–1816 (1999).

    Article  ADS  CAS  Google Scholar 

  68. Delimitis, A. et al. Strain distribution of thin InN epilayers grown on (0001) GaN templates by molecular beam epitaxy. Appl. Phys. Lett. 90, 061920 (2007).

    Article  ADS  CAS  Google Scholar 

  69. Orsal, G. et al. Bandgap energy bowing parameter of strained and relaxed InGaN layers. Opt. Mater. Express 4, 1030–1041 (2014).

    Article  ADS  CAS  Google Scholar 

  70. Varshni, Y. P. Band‐to‐band radiative recombination in groups IV, VI, and III‐V semiconductors (I). Phys. Status Solidi 19, 459 (1967).

    Article  CAS  Google Scholar 

  71. Vodopyanov, K. L., Graener, H., Phillips, C. C. & Tate, T. J. Picosecond carrier dynamics and studies of Auger recombination processes in indium arsenide at room temperature. Phys. Rev. B 46, 13194 (1992).

    Article  ADS  CAS  Google Scholar 

  72. Rogalski, A. & Jóźwikowski, K. The intrinsic carrier concentration in Pb1− xSnxTe, Pb1− xSnxSe, and PbS1− xSex. Phys. Status Solidi 111, 559 (1989).

    Article  ADS  CAS  Google Scholar 

  73. Klann, R., Höfer, T. & Buhleier, R. Fast recombination processes in lead chalcogenide semiconductors studied via transient optical nonlinearities. J. Appl. Phys. 77, 277 (1995).

    Article  ADS  CAS  Google Scholar 

  74. Marchetti, S., Martinelli, M. & Simili, R. The Auger recombination coefficient in InAs and GaSb derived from the infrared dynamical plasma reflectivity. J. Phys. Condens. Matter 14, 3653–3656 (2002).

    Article  ADS  CAS  Google Scholar 

Download references

Acknowledgements

We thank H. M. Fahad for help with setting up the gas-sensing instrument. Luminescence studies were funded by the US Department of Energy, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division under contract DE-AC02-05-CH11231 (EMAT program KC1201). Photodetector fabrication and testing were funded by the Defense Advanced Research Projects Agency under contract HR0011-16-1-0004. K.B.C. acknowledges funding from the Australian Research Council (grant numbers DP180104141 and FT140100577). H.K. acknowledges support from Samsung Scholarship.

Author information

Authors and Affiliations

Authors

Contributions

H.K. and A.J. conceived the idea for the project and designed the experiments. H.K., S.Z.U. and D.-H.L. performed optical measurements. H.K., M.Y. and T.K. fabricated devices. S.Z.U., N.S.A., S.B. and K.B.C. performed optical simulations. N.G. helped with gas-sensing experiments. Y.R. and C.P.G. helped with Raman measurements. H.K. wrote the manuscript. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Ali Javey.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Kah Wee Ang, Andres Castellanos-Gomez and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Strain applied in bP.

a, Photographic image of the two-point bending apparatus used here. An electrical linear actuator that can push/pull one point of the two-point bending apparatus applies a continuous and precise amount of uniaxial tensile strain to bP. b, Schematic of the two-point bending apparatus. Strain is calculated as ε = tsinθ/a, where ε is the amount of strain; t is the thickness of the substrate; a is the length of the substrate; and θ represents the angle of bending, which is equal to a/(2R) where R is the radius of the curvature60. Note that the circular arc approximation is not satisfied when θ is large at strains of 20% or more61. c, Raman spectra of the bP measured in Fig. 1 and Fig. 2. d, Schematic showing the atomic vibrations that correspond to Raman modes of \({{\rm{A}}}_{{\rm{g}}}^{1}\) (out-of-plane), \({{\rm{B}}}_{{\rm{2g}}}\) (in-plane; zigzag), and \({{\rm{A}}}_{{\rm{g}}}^{2}\) (in-plane; armchair). e, Optical image of the strained bP flake on the PETG substrate.

Extended Data Fig. 2 Detailed optical characterization of strained bP.

a, Photoluminescence (PL) peak wavelength as a function of transfer temperature. We characterized the photoluminescence peak wavelength as a function of biaxial compressive strain before and after the application of tensile strain (1.21%, zigzag). Each measurement was performed for five bP samples with thickness 20–22 nm. Samples were compressively strained by different amounts via different transfer temperatures (Ttr = 20 °C, 50 °C, 70 °C, 90 °C and 95 °C). The photoluminescence peak shift resulting from tensile strain increased with increasing biaxial compressive strain (as determined by transfer temperature). This is understood to be the result of the following. At high biaxial strain, the larger friction-induced resistance prevents the sliding of the 2D materials55. At lower transfer temperatures—that is, with reduced biaxial compressive strain—the bP is thus more likely to slip during bending of the substrate, such that the intended uniaxial tensile strain is not efficiently delivered to the bP. This could be because, at lower values of biaxial strain (that is, lower transfer temperatures), the friction-induced resistance that would prevent sliding of the bP is reduced. b, Laser power dependence of strain effect. We characterized the laser-induced heating effect on the bandgap shift by measuring photoluminescence peak wavelengths as a function of laser power. As the excitation spot size was similar or slightly smaller than the bP size, this helped to prevent the thermal expansion of the surrounding PETG by laser excitation (this thermal expansion could have elicited unexpected strain or slippage of bP from the PETG62). Regardless of the strain in bP, an excitation power higher than 1,500 W cm−2 always resulted in blueshift of the photoluminescence, attributed to thermal heating by the laser63. Although the photoluminescence peak position recovered after cooling to room temperature without excitation, when the laser intensity was even higher (higher than around 20 kW cm−2), there was visible damage to bP, which did not return to its original photoluminescence peak position. Therefore, we kept the laser incident power for photoluminescence measurements below 600 W cm−2, such that the photoluminescence peak of the exfoliated sample remained constant. This laser incident power is much less than that of the least powerful laser pump (roughly 20 kW cm−2), a value that is known to have a laser thermal effect in bP and MoS2 transferred on polyimide or PDMS62,63. c, Bandgap shift under different directions of strain with respect to the crystal orientation of bP. As the direction of tensile strain changed with respect to the crystal orientation of bP, there was no apparent difference in the strain-induced bandgap shift. This observation is consistent with previous results from a similar bending experiment performed on six-layer bP atop a polyethylene terephthalate (PET) substrate22. AC, armchair; ZZ, zigzag. d, Absorption at excitation wavelength for bP under zero strain, compressive strain and 1.21% of tensile strain. We found that strain had no notable effects on the absorption of light by bP at the photoluminescence excitation wavelength. Even though the bP bandgap was being modulated by strain, because our excitation wavelength was far from the absorption edge, the enhancement in photoluminescence cannot be attributed to increased absorption. e, Reversibility and repeatability of bandgap tuning in bP using compressive strain (0.6%) and tensile strain (1.2%). The photoluminescence peak from 20 nm bP shifts and recovers throughout ten cycles of bending and relaxation. At much higher strain, the PETG is subject to plastic deformation, exhibiting no return to its original state.

Extended Data Fig. 3 Detailed characterization of variable-spectrum bP LED.

a, Dependence on current density of the electroluminescence peak wavelength, showing the reliability of strain-tunable emission at different injection levels. To prevent degradation at high temperatures and to minimize the effect of localized hot spots on device performance, we used a polyimide film with high thermal conductivity, coupled with a Peltier module, to facilitate heat dissipation and to keep a constant temperature during operation. We also maintained a forward current density of less than 20% of the lowest injection level where thermal failure started to take place. When the current density was high, the devices failed sooner, and visible degradation was observed in the channel region. We therefore kept the current levels within the range shown (around 4 A cm−2 to 90 A cm−2) and the device showed stable operations over roughly 8 h (see Extended Data Fig. 5d, e). b, Distribution of angular intensity of the strain-tunable bP LED, calculated with finite-difference time domain (FDTD) simulations (FDTD Solutions, Lumerical). Precise computations await further study on changes in the refractive index of bP with strain (compressive and tensile), but here we simply calculated the angular distribution of the bP LED at two different peak wavelengths, using the published refractive indices of bP without strain. c, Left, schematic of the device architecture. Right, table showing published28,47,64,65,66 complex refractive indices of the polyimide substrate, bP and MoS2; these values were used for simulations. We found the angular distributions at two different wavelengths to be close enough that we could assume there was no discrepancy between the power collections at these two emission wavelengths using an objective lens with a fixed collection angle. d, IV curves of a strain-tunable bP–MoS2 LED measured at 0.20% of compressive strain and 1.06% of tensile strain.

Extended Data Fig. 4 Temperature-dependent performance of strain-tunable bP LED.

a, Electroluminescence spectra for the bP LED on a polyimide substrate, operating at a constant current density of 20 A cm−2 and at different temperatures, under compressive strain (0.2%) or tensile strain (1.0%). b, c, Peak wavelength (b) and peak intensity (c) of the electroluminescence from the bP LED under different strains and at different temperatures. Note that, to achieve heat dissipation and a uniform temperature control during device operation, dry N2 gas was consistently purged, and a mechanically flexible heat sink was installed, connected to the cold finger of the cryostat.

Extended Data Fig. 5 Tuning emission wavelengths to detect different gases.

a, Normalized electroluminescence spectra of bP LED with 0.2% compressive strain, 0.3% tensile strain and 1.0% tensile strain, for detecting CO2, CH4 and H2O, respectively. b, c, Sensor response from the device under 0.2% compressive strain in the presence of CH4 gas (b) and under 0.3% tensile strain in the presence of CO2 gas (c). Our approach showed minimal CH4 detection at a concentration of 2.5% under 0.2% compressive strain, and at 0.3% tensile strain, it could no longer detect CO2 gas. d, e, Stability of the gas-sensing setup for the bP LED measured under 0.2% compressive strain (d) and 0.3% tensile strain (e). Both measurements were performed at a current density of 20 A cm−2 with fmod = 1 kHz. Over 8 h of measurement, the device exhibited a maximum drift of 0.90% and 1.22% for 0.2% compressive strain and 0.3% tensile strain, respectively.

Extended Data Fig. 6 Strain-tunable photoconductors based on bP.

a, Schematic of a strain-tunable bP photoconductor. b, Schematic of the device architecture, showing the generation of a photocurrent at a bias voltage. c, Optical micrograph of the device. D, drain; S, source. d, Strain-dependent spectral photoresponse. A/W, amps/watts. e, Polarization-dependent responsivity at 4.0 μm and 2.0 μm for the device under 0.4% compressive strain and 1.0% tensile strain, respectively. f, IV curves for the strain-tunable bP photoconductor measured in the dark and under illumination by a 1,000 K black body. g, Rise and fall times (between photoresponses of 10% and 90%) under 0.4% compressive and 1.0% tensile strain, using a 1,650-nm laser at roughly 10 mW cm−2. All measurements were conducted at a bias voltage of 100 mV from a device comprising bP of thickness 22 nm.

Extended Data Fig. 7 Detailed characterization of a variable-spectrum bP photoconductor.

a, Spectral noise density under 0.4% compressive and 1.0% tensile strain. The dashed line indicates the 1/f curve at low frequency. b, Normalized photoresponse of the strain-tunable bP photoconductor measured as a function of modulation frequency. The device was measured at a Vd = 100 mV and excited by a 1,650-nm laser, showing a 3-dB frequency of 10 kHz. c, Specific detectivity (D*) as a function of wavelength at room temperature, for the bP device with 0.4% compressive and 1.0% tensile strain and for various commercially available photodetectors.

Extended Data Fig. 8 Setups for measuring photoluminescence and electroluminescence.

a, The IRPL setup. b, The infrared electroluminescence (IREL) setup. For both measurements, the emission from bP was collected by a reflective objective and sent to the external port of the FT-IR spectrometer, with fmod = 5 kHz, τLock-in = 300 μs and optical velocity = 0.0633 cm s−1. Ref., reference; Vac. chamber, vacuum chamber. The total interferogram from the HgCdTe (MCT) detector and the modulated interferogram from the voltage preamplifier were used to separate the photoluminescence/electroluminescence signal from the thermal background.

Extended Data Table 1 Strain-induced bandgap shift of semiconductors in the SWIR/MWIR range
Extended Data Table 2 Parameters used to calculate the theoretical quantum yield from the ABC model

Supplementary information

Supplementary Video 1

Dynamic emission spectrum tuning of bP-LED. Real-time measurement of emission spectrum from actively variable spectrum bP-LED. When tensile strain is applied by bending, the emission has a peak around 2.7 μm. When tensile strain is released and there is only compressive strain, the emission peak is shifted to 4.1 μm. The time period between the bending and release is 21 s. The spectrum is modulated without drift over 500 times of bending cycles and 6 hours of operation time.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, H., Uddin, S.Z., Lien, DH. et al. Actively variable-spectrum optoelectronics with black phosphorus. Nature 596, 232–237 (2021). https://doi.org/10.1038/s41586-021-03701-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03701-1

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing