Open Access
1 September 2017 Surface emitting ring quantum cascade lasers for chemical sensing
Rolf Szedlak, Jakob Hayden, Pedro Martín-Mateos, Martin Holzbauer, Andreas Harrer, Benedikt Schwarz, Borislav Hinkov, Donald MacFarland, Tobias Zederbauer, Hermann Detz, Aaron Maxwell Andrews, Werner Schrenk, Pablo Acedo, Bernhard Lendl, Gottfried Strasser
Author Affiliations +
Abstract
We review recent advances in chemical sensing applications based on surface emitting ring quantum cascade lasers (QCLs). Such lasers can be implemented in monolithically integrated on-chip laser/detector devices forming compact gas sensors, which are based on direct absorption spectroscopy according to the Beer–Lambert law. Furthermore, we present experimental results on radio frequency modulation up to 150 MHz of surface emitting ring QCLs. This technique provides detailed insight into the modulation characteristics of such lasers. The gained knowledge facilitates the utilization of ring QCLs in combination with spectroscopic techniques, such as heterodyne phase-sensitive dispersion spectroscopy for gas detection and analysis.

1.

Introduction

The midinfrared spectral region exhibits numerous ro-vibrational absorption lines of various substances. Therefore, light sources emitting in this wavelength range are desirable tools for sensing and chemical fingerprinting. Due to their compact size, tailorable emission characteristics,1,2 and high spectral brightness,3 quantum cascade lasers (QCLs)4 are commonly used for such absorption experiments.5 Surface emitting QCLs possess considerable advantages compared with their edge emitting counterparts. They facilitate on-chip testing and two-dimensional array integration. Due to their rather large emission aperture, these lasers can produce narrow beam profiles.6 The first surface emitting QCL7 was demonstrated utilizing a second-order distributed-feedback (DFB) grating. Subsequently, further approaches have been successfully implemented: in photonic crystals8,9 artificial periodic structures provide optical feedback and generate light beams in vertical direction. With random lasers10,11 almost diffraction limited broadband surface emission can be achieved, while losing tunability. Manipulating beam characteristics through wave engineering has been demonstrated with vertically emitting circular DFB gratings12 as well as graded photonic heterostructures.13

Surface emitting ring QCLs14 typically consist of a circular ring waveguide with a second-order DFB grating. The latter provides single-mode selection as well as vertical light emission while the circular ring geometry creates a collimated and rotationally symmetric far field. This desirable beam profile can be modified by well-directed manipulation of a regular DFB grating.15 Single-mode emission and continuous-wave operation at room temperature16,17 in combination with the collimated emission beam make ring QCLs suitable tools for chemical sensing applications.18,19 Due to their compact size and small footprint, these lasers are ideal light sources for the application in miniaturized sensing setups, such as substrate-integrated hollow waveguides.20

2.

Surface Emitting/Detecting On-Chip Sensor

Typical state-of-the-art applications make use of conventional sensing setups, which include a light source, a light–analyte interaction region, and an external separate detector, e.g., based on the mercury-cadmium-telluride (MCT) material system. Utilization of a quantum cascade detector (QCD)21,22 in combination with a separate QCL mounted on an individual chip demonstrates the capabilities of quantum cascade technology in spectroscopy.23 The development of bifunctional quantum cascade heterostructures24 enables the realization of monolithically integrated lasers (QCL) and detectors (QCD) fabricated on the same chip. Edge emitting and detecting bifunctional devices have successfully been used for sensing of liquids.25 However, the limited length of the interaction region between laser and detector makes gas sensing impractical. One possible solution is to guide the light on the chip and thereby increase the interaction length.26 Certainly, this approach is also limited in terms of absorption path length due to restrictions in on-chip mode guiding.

In contrast, utilization of surface emitting and detecting quantum cascade structures fabricated on the same chip in combination with an external light–analyte interaction region provides, in theory, arbitrarily long absorption paths and facilitates remote sensing experiments.27,28 Furthermore, the small footprint of such sensors enables the realization of compact sensing systems. Figure 1(a) shows a scanning electron microscope (SEM) image of a surface emitting ring QCL with a surface detecting disk QCD in its center.27 Vertical light emission and detection are provided by the second-order DFB grating (QCL) and the metal hole grating (QCD), respectively. A sketch of the corresponding sensing setup is shown in Fig. 2(a).

Fig. 1

SEM images of surface emitting and detecting quantum cascade sensors. (a) Ring QCL with a centered disk QCD.27 (b) Two concentric DFB ring quantum cascade laser/detector (QCLD) devices.28 Each ring can be used for light emission and detection.

OE_57_1_011005_f001.png

Fig. 2

(a) Sketch of the experimental setup for remote sensing based on a monolithically integrated surface emitting and detecting on-chip sensor. Light is emitted from the sensor chip, traverses the gas cell, and is back-reflected by a flat gold mirror.28 (b) Beam profile measured with a bolometer camera after two identical lenses. This pattern corresponds to the intensity distribution incident on the sensor chip.

OE_57_1_011005_f002.png

Light emitted from the ring QCL is collimated by a lens and modulated by a chopper in order to filter out the high-frequency electrical cross talk before it propagates through the gas cell. A flat gold mirror at the end of the cell reflects the light back toward the chip. After inverse propagation along its initial path, the light is focused onto the sensor chip where it is detected by the disk QCD. In this setup, an image of the laser source is projected back onto the sensor chip. Figure 2(b) shows the incident beam pattern, which was recorded in a mirrored setup with two identical lenses preceding a bolometer camera. The extent of the measured intensity pattern indicates that only a fraction of the emitted light is impinging on the QCD. An enhanced coupling between laser and detector can be achieved by adapting the detector geometry to the shape of the laser beam. Such an approach is depicted in Fig. 1(b), which shows two monolithically integrated concentric ring structures. The implemented second-order DFB gratings facilitate vertical light emission and detection, which enables mutual commutation of both rings, i.e., inner and outer ring can be used as laser and as detector, respectively.28 Both concepts, with the disk and ring detector, are based on conventional absorption spectroscopy according to the Beer–Lambert law.

3.

Modulation Characteristics and Dispersion Spectroscopy

In contrast, ring QCLs can also be used in combination with sophisticated spectroscopic techniques. Knowledge of the laser’s current-tuning characteristics29 is important for a variety of sensing techniques, such as wavelength- and frequency-modulation spectroscopy,30,31 intrapulse absorption spectroscopy,32 chirped laser dispersion spectroscopy,33 and heterodyne phase-sensitive dispersion spectroscopy (HPSDS).34,35 The latter is a molecular dispersion technique that probes a sample’s refractive index spectrum, rather than its absorption, via the phase of radio frequency-modulated laser light.36 Its sensitivity to the signal’s phase makes HPSDS immune to power fluctuations and therefore enables calibration-free sensor operation.

In the following, an HPSDS spectrometer is implemented and experimentally validated. Therefore, a surface emitting single-mode ring QCL [Fig. 3(a)] is sinusoidally modulated at radio frequencies and swept across a well-known absorption line. This technique allows a thorough investigation of the laser’s modulation characteristics. A detailed description of the quantum cascade heterostructure can be found in Ref. 37. In this study, characteristic laser tuning parameters are fitted to a known sample’s absorption profile using the model described in Ref. 38. Based on this characterization, HPSDS gas sensing experiments are performed.

Fig. 3

(a) SEM image of a surface emitting ring QCL with a second-order DFB grating for surface emission. (b) Schematic of the experimental setup used for the characterization of the laser’s modulation properties.

OE_57_1_011005_f003.png

The ring laser is operated at liquid nitrogen temperatures, and its emitted radiation is measured with an MCT detector after passing through a 10-cm-long gas cell filled with 0.4% CH4 in N2 at a pressure of 24 mbar. A sketch of the corresponding setup is shown in Fig. 3(b). Different modulation frequencies between 1 and 150 MHz are utilized in combination with four DC current values (65, 101, 150, and 200 mA). The corresponding laser submount temperatures (118, 110, 100, and 88 K) are reduced in order to stabilize the emission wavelength of the laser. All operating points are well above threshold, which is found to be at 27 mA (80 K) and at 43 mA (110 K). For each DC current, the laser is slowly tuned over the two absorption features at 1322.08/cm and 1322.15/cm. The tuning frequency of the current ramp is around 50 MHz with a peak-to-peak amplitude of 11.8 mA. Subsequently, the signal of the MCT is fed into a mixer and downmixed to 100 kHz in order to measure the amplitude and phase with a lock-in amplifier.35 Figure 4 shows the amplitude (a) and phase (b) of the downmixed signal at a modulation frequency of 100 MHz and a DC current of 200 mA. For all DC currents and modulation frequencies, only the larger feature at 1322.08/cm is fitted and used for the following analysis.

Fig. 4

(a) Amplitude and (b) FM–IM phase of an exemplary measurement with a modulation frequency of 100 MHz and a DC current of 200 mA. Experimental data (red) and fit (black) are shown.

OE_57_1_011005_f004.png

According to the notation in Ref. 38, Fig. 5(a) shows the ratio Δf/m=Δf/(ΔP/P0) between absolute frequency modulation (FM) and intensity modulation (IM) index. In Fig. 5(b), the FM–IM phase shift Θ is provided as a function of the modulation frequency. The region between 30 and 150 MHz is characterized by a rather flat behavior of Δf/m and Θ. However, below 20 MHz, the rise of the FM–IM ratio and phase indicates the onset of thermal tuning, which is a figure of merit in ring QCLs. Based on these tuning characteristics, HPSDS gas sensing measurements were performed using a different absorption line of CH4. The laser characterization confirmed that a low bias current and a high modulation frequency are beneficial for HPSDS. On the other hand, a large optical intensity is advantageous for a high signal-to-noise ratio. Operating the laser at a DC current of 92 mA (120 K) and 183.5 mA (100 K) with a modulation frequency of 150 MHz, different gas concentrations are investigated. Figure 6 shows the corresponding normalized phase in a concentration range from 1,000 to 12,500 ppm. The different linear ranges for the two bias current setpoints are clearly visible. This can be attributed to the different FM/IM indices for the two setpoints [compare Fig. 5(a)].35 The low bias current setpoint corresponds to a predominantly amplitude modulated case (small FM/IM index), yielding a large dynamic range, while a high bias current favors frequency modulation and gives larger phase shifts at the expense of limited linearity. The extracted 3σ limit of detection, defined as the noise-equivalent concentration, for DC currents of 92 and 183.5 mA is extrapolated from gas measurements with low peak absorption and amounts to 16 and 2 ppm, respectively. The noise floor σ is measured with no CH4 in the gas cell.

Fig. 5

(a) FM–IM ratio and (b) FM–IM phase as a function of the modulation frequency for four different DC currents.

OE_57_1_011005_f005.png

Fig. 6

Gas sensing results based on HPSDS. Normalized phase for different CH4 concentrations obtained for a DC current of 92 mA (black) and 183.5 mA (red), both at 150 MHz.

OE_57_1_011005_f006.png

4.

Conclusion

We present recent advances in the field of ring QCLs for chemical sensing applications. These light sources have been implemented in monolithically integrated on-chip laser/detector devices for gas sensing. Prototype devices featuring a disk and a ring detector have been realized. The latter enables mutual commutation between laser and detector. Furthermore, an introductive characterization of the radio frequency modulation characteristics up to 150 MHz of surface emitting ring QCLs is demonstrated. This knowledge paves the way for an implementation of these compact and versatile midinfrared light sources in gas sensing applications exploiting efficient spectroscopic techniques, such as HPSDS.

Disclosures

The authors declare no competing financial interest.

Acknowledgments

The authors acknowledge the support by the Austrian Science Fund (FWF) Project No. F 49 (Next-Lite). J.H. and B.L. acknowledge the financial support received from the Competence Center Austrian Smart Systems Integration Research Center as part of the Austrian Competence Centers for Excellent Technologies program. B.S. acknowledges the support from the FWF Project No. P 28914. H.D. acknowledges the support through an APART fellowship from the Austrian Academy of Sciences.

References

1. 

G. Wysocki et al., “Widely tunable mode-hop free external cavity quantum cascade laser for high resolution spectroscopic applications,” Appl. Phys. B, 81 (6), 769 –777 (2005). http://dx.doi.org/10.1007/s00340-005-1965-4 Google Scholar

2. 

B. G. Lee et al., “Widely tunable single-mode quantum cascade laser source for mid-infrared spectroscopy,” Appl. Phys. Lett., 91 (23), 231101 (2007). http://dx.doi.org/10.1063/1.2816909 APPLAB 0003-6951 Google Scholar

3. 

B. Hinkov et al., “Quantum cascade laser in a master oscillator power amplifier configuration with watt-level optical output power,” Opt. Express, 21 19180 –19186 (2013). http://dx.doi.org/10.1364/OE.21.019180 OPEXFF 1094-4087 Google Scholar

4. 

J. Faist et al., “Quantum cascade laser,” Science, 264 (5158), 553 –556 (1994). http://dx.doi.org/10.1126/science.264.5158.553 SCIEAS 0036-8075 Google Scholar

5. 

A. A. Kosterev and F. K. Tittel, “Chemical sensors based on quantum cascade lasers,” IEEE J. Quantum Electron., 38 582 –591 (2002). http://dx.doi.org/10.1109/JQE.2002.1005408 IEJQA7 0018-9197 Google Scholar

6. 

M. Schubert and F. Rana, “Analysis of terahertz surface emitting quantum-cascade lasers,” IEEE J. Quantum Electron., 42 257 –265 (2006). http://dx.doi.org/10.1109/JQE.2005.863138 IEJQA7 0018-9197 Google Scholar

7. 

D. Hofstetter et al., “Surface-emitting 10.1  μm quantum-cascade distributed feedback lasers,” Appl. Phys. Lett., 75 (24), 3769 –3771 (1999). http://dx.doi.org/10.1063/1.125450 APPLAB 0003-6951 Google Scholar

8. 

R. Colombelli et al., “Quantum cascade surface-emitting photonic crystal laser,” Science, 302 (5649), 1374 –1377 (2003). http://dx.doi.org/10.1126/science.1090561 SCIEAS 0036-8075 Google Scholar

9. 

M. S. Vitiello et al., “Photonic quasi-crystal terahertz lasers,” Nat. Commun., 5 5884 –5891 (2014). http://dx.doi.org/10.1038/ncomms6884 NCAOBW 2041-1723 Google Scholar

10. 

R. Degl’Innocenti et al., “Hyperuniform disordered terahertz quantum cascade laser,” Sci. Rep., 6 19325 –19331 (2016). http://dx.doi.org/10.1038/srep19325 SRCEC3 2045-2322 Google Scholar

11. 

S. Schönhuber et al., “Random lasers for broadband directional emission,” Optica, 3 1035 –1038 (2016). http://dx.doi.org/10.1364/OPTICA.3.001035 Google Scholar

12. 

G. Liang et al., “Single-mode surface-emitting concentric-circular-grating terahertz quantum cascade lasers,” Appl. Phys. Lett., 102 (3), 031119 (2013). http://dx.doi.org/10.1063/1.4789535 APPLAB 0003-6951 Google Scholar

13. 

G. Xu et al., “Efficient power extraction in surface-emitting semiconductor lasers using graded photonic heterostructures,” Nat. Commun., 3 952 –958 (2012). http://dx.doi.org/10.1038/ncomms1958 NCAOBW 2041-1723 Google Scholar

14. 

E. Mujagić et al., “Low divergence single-mode surface emitting quantum cascade ring lasers,” Appl. Phys. Lett., 93 (16), 161101 (2008). http://dx.doi.org/10.1063/1.3000630 APPLAB 0003-6951 Google Scholar

15. 

R. Szedlak et al., “The influence of whispering gallery modes on the far field of ring lasers,” Sci. Rep., 5 16668 –16675 (2015). http://dx.doi.org/10.1038/srep16668 SRCEC3 2045-2322 Google Scholar

16. 

Y. Bai et al., “High power, continuous wave, quantum cascade ring laser,” Appl. Phys. Lett., 99 (26), 261104 (2011). http://dx.doi.org/10.1063/1.3672049 APPLAB 0003-6951 Google Scholar

17. 

D. H. Wu and M. Razeghi, “High power, low divergent, substrate emitting quantum cascade ring laser in continuous wave operation,” APL Mater., 5 (3), 035505 (2017). http://dx.doi.org/10.1063/1.4978810 Google Scholar

18. 

M. Brandstetter et al., “Time-resolved spectral characterization of ring cavity surface emitting and ridge-type distributed feedback quantum cascade lasers by step-scan FT-IR spectroscopy,” Opt. Express, 22 2656 –2664 (2014). http://dx.doi.org/10.1364/OE.22.002656 OPEXFF 1094-4087 Google Scholar

19. 

H. Moser et al., “Application of a ring cavity surface emitting quantum cascade laser (RCSE-QCL) on the measurement of H2S in a CH4 matrix for process analytics,” Opt. Express, 24 6572 –6585 (2016). http://dx.doi.org/10.1364/OE.24.006572 OPEXFF 1094-4087 Google Scholar

20. 

E. Tütüncü et al., “Advanced gas sensors based on substrate-integrated hollow waveguides and dual-color ring quantum cascade lasers,” Analyst, 141 6202 –6207 (2016). http://dx.doi.org/10.1039/C6AN01130F ANLYAG 0365-4885 Google Scholar

21. 

D. Hofstetter, M. Beck and J. Faist, “Quantum-cascade-laser structures as photodetectors,” Appl. Phys. Lett., 81 (15), 2683 –2685 (2002). http://dx.doi.org/10.1063/1.1512954 APPLAB 0003-6951 Google Scholar

22. 

F. R. Giorgetta et al., “Quantum cascade detectors,” IEEE J. Quantum Electron., 45 1039 –1052 (2009). http://dx.doi.org/10.1109/JQE.2009.2017929 IEJQA7 0018-9197 Google Scholar

23. 

P. Jouy et al., “Mid-infrared spectroscopy for gases and liquids based on quantum cascade technologies,” Analyst, 139 2039 –2046 (2014). http://dx.doi.org/10.1039/C3AN01462B ANLYAG 0365-4885 Google Scholar

24. 

B. Schwarz et al., “A bi-functional quantum cascade device for same-frequency lasing and detection,” Appl. Phys. Lett., 101 (19), 191109 (2012). http://dx.doi.org/10.1063/1.4767128 APPLAB 0003-6951 Google Scholar

25. 

B. Schwarz et al., “Monolithically integrated mid-infrared lab-on-a-chip using plasmonics and quantum cascade structures,” Nat. Commun., 5 4085 –4091 (2014). http://dx.doi.org/10.1038/ncomms5085 NCAOBW 2041-1723 Google Scholar

26. 

E. J. Zhang et al., “Silicon photonic on-chip trace-gas spectroscopy of methane,” in Conf. on Lasers and Electro-Optics, (2016). Google Scholar

27. 

A. Harrer et al., “Mid-infrared surface transmitting and detecting quantum cascade device for gas-sensing,” Sci. Rep., 6 21795 –21800 (2016). http://dx.doi.org/10.1038/srep21795 SRCEC3 2045-2322 Google Scholar

28. 

R. Szedlak et al., “Remote sensing with commutable monolithic laser and detector,” ACS Photonics, 3 (10), 1794 –1798 (2016). http://dx.doi.org/10.1021/acsphotonics.6b00603 Google Scholar

29. 

B. Hinkov et al., “RF-modulation of mid-infrared distributed feedback quantum cascade lasers,” Opt. Express, 24 3294 –3312 (2016). http://dx.doi.org/10.1364/OE.24.003294 OPEXFF 1094-4087 Google Scholar

30. 

P. Kluczynski et al., “Wavelength modulation absorption spectrometry an extensive scrutiny of the generation of signals,” Spectrochim. Acta Part B, 56 (8), 1277 –1354 (2001). http://dx.doi.org/10.1016/S0584-8547(01)00248-8 Google Scholar

31. 

S. Borri et al., “Frequency modulation spectroscopy by means of quantum-cascade lasers,” Appl. Phys. B, 85 (2), 223 –229 (2006). http://dx.doi.org/10.1007/s00340-006-2343-6 Google Scholar

32. 

C. Reidl-Leuthner and B. Lendl, “Toward stand-off open-path measurements of NO and NO2 in the sub-parts per million meter range using quantum cascade lasers (QCLs) in the intra-pulse absorption mode,” Appl. Spectrosc., 67 1368 –1375 (2013). http://dx.doi.org/10.1366/13-07104 APSPA4 0003-7028 Google Scholar

33. 

A. Hangauer et al., “Chirped laser dispersion spectroscopy using a directly modulated quantum cascade laser,” Appl. Phys. Lett., 103 (19), 191107 (2013). http://dx.doi.org/10.1063/1.4829036 APPLAB 0003-6951 Google Scholar

34. 

P. Martín-Mateos and P. Acedo, “Heterodyne phase-sensitive detection for calibration-free molecular dispersion spectroscopy,” Opt. Express, 22 15143 –15153 (2014). http://dx.doi.org/10.1364/OE.22.015143 OPEXFF 1094-4087 Google Scholar

35. 

P. Martín-Mateos et al., “Heterodyne phase-sensitive dispersion spectroscopy in the mid-infrared with a quantum cascade laser,” Anal. Chem., 89 (11), 5916 –5922 (2017). http://dx.doi.org/10.1021/acs.analchem.7b00303 ANCHAM 0003-2700 Google Scholar

36. 

J. Hayden et al., “A quantitative comparison of dispersion- and absorption-spectroscopic gas sensing,” Proc. SPIE, 10110 101100Z (2017). http://dx.doi.org/10.1117/12.2252624 Google Scholar

37. 

Z. Liu et al., “Room-temperature continuous-wave quantum cascade lasers grown by MOCVD without lateral regrowth,” IEEE Photonics Technol. Lett., 18 1347 –1349 (2006). http://dx.doi.org/10.1109/LPT.2006.877006 IPTLEL 1041-1135 Google Scholar

38. 

A. Hangauer et al., “High frequency modulation capabilities and quasi single-sideband emission from a quantum cascade laser,” Opt. Express, 22 23439 –23455 (2014). http://dx.doi.org/10.1364/OE.22.023439 OPEXFF 1094-4087 Google Scholar

Biographies for the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Rolf Szedlak, Jakob Hayden, Pedro Martín-Mateos, Martin Holzbauer, Andreas Harrer, Benedikt Schwarz, Borislav Hinkov, Donald MacFarland, Tobias Zederbauer, Hermann Detz, Aaron Maxwell Andrews, Werner Schrenk, Pablo Acedo, Bernhard Lendl, and Gottfried Strasser "Surface emitting ring quantum cascade lasers for chemical sensing," Optical Engineering 57(1), 011005 (1 September 2017). https://doi.org/10.1117/1.OE.57.1.011005
Received: 2 June 2017; Accepted: 9 August 2017; Published: 1 September 2017
Lens.org Logo
CITATIONS
Cited by 8 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Quantum cascade lasers

Sensors

Biological and chemical sensing

Modulation

Frequency modulation

Spectroscopy

Absorption

Back to Top