4
Molecular dispersion spectroscopy based on FabryPerot quantum cascade lasers LUKASZ A. STERCZEWSKI, 1,2,JONAS WESTBERG, 1,AND GERARD WYSOCKI 1, * 1 Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA 2 Faculty of Electronics, Wroclaw University of Science and Technology, Wroclaw 50370, Poland *Corresponding author: [email protected] Received 7 November 2016; revised 9 December 2016; accepted 11 December 2016; posted 12 December 2016 (Doc. ID 280142); published 10 January 2017 Two FabryPerot quantum cascade lasers are used in a dif- ferential dual comb configuration to perform rapidly swept dispersion spectroscopy of low-pressure nitrous oxide with <1 ms acquisition time. Active feedback control of the laser injection current enables simultaneous wavelength modulation of both lasers at kilohertz rates. The system demonstrates similar performance in both absorption and dispersion spectroscopy modes and achieves a noise- equivalent absorption figure of merit in the low 10 -4 Hz p range. © 2017 Optical Society of America OCIS codes: (140.5965) Semiconductor lasers, quantum cascade; (300.6310) Spectroscopy, heterodyne; (120.2230) Fabry-Perot; (300.6380) Spectroscopy, modulation. https://doi.org/10.1364/OL.42.000243 Mid-IR FabryPerot quantum cascade laser (FP-QCL)based spectroscopy can potentially address a large spectroscopic appli- cation space by providing a unique combination of narrow instrumental linewidths and broad spectral coverage provided by coherent multimode light sources. Unlike many other spectroscopic techniques that are able to cover wide spectral bandwidths, a FP-QCLbased spectrometer does not use any moving parts, which enables ultrafast acquisition times (μs) that are ultimately limited by data transfer speeds and processing algorithms. The most common implementation of multimode spectroscopy is the dual-comb spectroscopy (DCS) technique [1], which is based on a pair of frequency comb lasers with slightly different optical mode spacing. By spectrally over- lapping the two frequency combs at the surface of a square law detector, an optical multiheterodyne mixing process occurs. This allows for the optical spectrum to be effectively downcon- verted to the radio-frequency domain, where reliable instru- mentation and well-established detection procedures are available for multiheterodyne spectroscopy (MHS) signal extraction. Most of the work on DCS has been performed in the near-IR due to the accessibility of stable frequency comb sources covering this spectral region [2]. However, access to the mid-IR part of the spectrum enables probing of the fundamen- tal rotational-vibrational transitions of many molecules that have important roles in various applications including environ- mental monitoring [3], industrial process control [4], and medical diagnostics [5]. The absorption cross sections in the mid-IR can be orders of magnitude greater than at near-IR, and therefore several different approaches to implement mid-IR MHS have recently been proposed. These include dif- ference frequency generation from coherent near-IR sources [6], optical parametric oscillators [7], microresonators [8], and multimode FP-QCLs [9]. FP-QCLs in particular offer an interesting potential for development of pure electronically controlled multiheterodyne light sources that can be integrated on a single semiconductor chip, which could offer cost-effective and high-resolution (<0.0001 cm -1 ) mid-IR spectrometers with broadband spectral coverage (>100 cm -1 ) and rapid response times (<1 μs). The coherence properties of these multimode light sources have been studied in both the mid-IR [9] and the terahertz [10], and frequency comb operation has been achieved by careful dispersion engineering of the de- vices [11,12]. A majority of contemporary QCL-based dual comb systems relies on measurements of the attenuation of the light as it propagates through the sample. Although this type of measure- ment is intuitively straightforward and relatively easy to per- form, there are a number of challenging applications that would benefit from measurements of the light dispersion (phase) within the probed medium [13]. A typical example is the measurement of strongly absorbing species (>10%), which normally requires a nonlinear correction of the exponential BeerLamberts law, while sample dispersion shows linear dependence on species concentration. Also, dispersion mea- surements are less sensitive to intensity fluctuations, and any amplitude drift and noise that can occur in MHS, such as RF noise in the detector electronics or fluctuations in hetero- dyne efficiency due to variations in the optical wavefront, can be largely suppressed. An intrinsic property of MHS is its ability to instantly access spectroscopic information associated with any or all available optical modes. This offers an un- matched opportunity of combining broadband and high- resolution measurements within the same system. This work is focused specifically on a proof-of-concept demonstration of high-resolution absorption and dispersion spectra of nitrous Letter Vol. 42, No. 2 / January 15 2017 / Optics Letters 243 0146-9592/17/020243-04 Journal © 2017 Optical Society of America

Molecular dispersion spectroscopy based on Fabry Perot ... · Molecular dispersion spectroscopy based on Fabry–Perot quantum cascade lasers LUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,†

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Molecular dispersion spectroscopy based on Fabry Perot ... · Molecular dispersion spectroscopy based on Fabry–Perot quantum cascade lasers LUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,†

Molecular dispersion spectroscopy based onFabry–Perot quantum cascade lasersLUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,† AND GERARD WYSOCKI1,*1Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA2Faculty of Electronics, Wroclaw University of Science and Technology, Wroclaw 50370, Poland*Corresponding author: [email protected]

Received 7 November 2016; revised 9 December 2016; accepted 11 December 2016; posted 12 December 2016 (Doc. ID 280142);published 10 January 2017

Two Fabry–Perot quantum cascade lasers are used in a dif-ferential dual comb configuration to perform rapidly sweptdispersion spectroscopy of low-pressure nitrous oxide with<1 ms acquisition time. Active feedback control of thelaser injection current enables simultaneous wavelengthmodulation of both lasers at kilohertz rates. The systemdemonstrates similar performance in both absorptionand dispersion spectroscopy modes and achieves a noise-equivalent absorption figure of merit in the low10−4∕

�������

Hzp

range. © 2017 Optical Society of America

OCIS codes: (140.5965) Semiconductor lasers, quantum cascade;

(300.6310) Spectroscopy, heterodyne; (120.2230) Fabry-Perot;

(300.6380) Spectroscopy, modulation.

https://doi.org/10.1364/OL.42.000243

Mid-IR Fabry–Perot quantum cascade laser (FP-QCL)–basedspectroscopy can potentially address a large spectroscopic appli-cation space by providing a unique combination of narrowinstrumental linewidths and broad spectral coverage providedby coherent multimode light sources. Unlike many otherspectroscopic techniques that are able to cover wide spectralbandwidths, a FP-QCL–based spectrometer does not useany moving parts, which enables ultrafast acquisition times(∼μs) that are ultimately limited by data transfer speeds andprocessing algorithms. The most common implementation ofmultimode spectroscopy is the dual-comb spectroscopy (DCS)technique [1], which is based on a pair of frequency comb laserswith slightly different optical mode spacing. By spectrally over-lapping the two frequency combs at the surface of a square lawdetector, an optical multiheterodyne mixing process occurs.This allows for the optical spectrum to be effectively downcon-verted to the radio-frequency domain, where reliable instru-mentation and well-established detection procedures areavailable for multiheterodyne spectroscopy (MHS) signalextraction. Most of the work on DCS has been performedin the near-IR due to the accessibility of stable frequency combsources covering this spectral region [2]. However, access to themid-IR part of the spectrum enables probing of the fundamen-tal rotational-vibrational transitions of many molecules that

have important roles in various applications including environ-mental monitoring [3], industrial process control [4], andmedical diagnostics [5]. The absorption cross sections in themid-IR can be orders of magnitude greater than at near-IR,and therefore several different approaches to implementmid-IR MHS have recently been proposed. These include dif-ference frequency generation from coherent near-IR sources[6], optical parametric oscillators [7], microresonators [8],and multimode FP-QCLs [9]. FP-QCLs in particular offeran interesting potential for development of pure electronicallycontrolled multiheterodyne light sources that can be integratedon a single semiconductor chip, which could offer cost-effectiveand high-resolution (<0.0001 cm−1) mid-IR spectrometerswith broadband spectral coverage (>100 cm−1) and rapidresponse times (<1 μs). The coherence properties of thesemultimode light sources have been studied in both the mid-IR[9] and the terahertz [10], and frequency comb operationhas been achieved by careful dispersion engineering of the de-vices [11,12].

A majority of contemporary QCL-based dual comb systemsrelies on measurements of the attenuation of the light as itpropagates through the sample. Although this type of measure-ment is intuitively straightforward and relatively easy to per-form, there are a number of challenging applications thatwould benefit from measurements of the light dispersion(phase) within the probed medium [13]. A typical example isthe measurement of strongly absorbing species (>10%), whichnormally requires a nonlinear correction of the exponentialBeer–Lambert’s law, while sample dispersion shows lineardependence on species concentration. Also, dispersion mea-surements are less sensitive to intensity fluctuations, and anyamplitude drift and noise that can occur in MHS, such asRF noise in the detector electronics or fluctuations in hetero-dyne efficiency due to variations in the optical wavefront, canbe largely suppressed. An intrinsic property of MHS is itsability to instantly access spectroscopic information associatedwith any or all available optical modes. This offers an un-matched opportunity of combining broadband and high-resolution measurements within the same system. This workis focused specifically on a proof-of-concept demonstrationof high-resolution absorption and dispersion spectra of nitrous

Letter Vol. 42, No. 2 / January 15 2017 / Optics Letters 243

0146-9592/17/020243-04 Journal © 2017 Optical Society of America

Page 2: Molecular dispersion spectroscopy based on Fabry Perot ... · Molecular dispersion spectroscopy based on Fabry–Perot quantum cascade lasers LUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,†

oxide (N2O) through analog mixing of two multiheterodynebeat notes to demonstrate the important functionalities avail-able with a mid-IR FP-QCL multiheterodyne spectrometer.These systems can operate in two modes of detection, conven-tional broadband MHS and single beat note high-resolutionswept MHS, which in combination with their rapid responsetimes provide a unique potential for a versatile, pure electroni-cally controlled mid-IR chemical sensing system. A briefdiscussion on the potential expansion of the basic principlesof the presented technique to retrieve spectroscopic dispersioninformation from all accessible multiheterodyne beat notes in aparallel manner is also provided.

A schematic of the experimental setup is shown in Fig. 1.The optical core configuration of the system is similar to thatreported in our earlier works [14,15], with important modifi-cations that include a reference detector and a phase compara-tor for dispersion measurements. The reference detector isused to realize a differential detection scheme, where two fastphotodetectors (VIGO systems, PV-4TE-10.6, 3 dB cutoffat >1.1 GHz) are used as optical heterodyne mixers andtwo multimode FP-QCLs are used as light sources. The lasersemit a broadband multimode optical spectrum in the1150–1210 cm−1 range with an optical mode spacing of ∼40.8and ∼40.64 GHz, for the signal (Sig.) and local oscillator (LO)laser, respectively. The light emitted from each FP-QCL is di-vided into two beams and each heterodyne mixer (photodetec-tor) receives light from both lasers adhering to the dual-detectorDCS scheme [9]. One beam derived from the signal laser passesthrough a 100 mm sample cell (PIKE Technologies) containingpure N2O at 10 Torr of absolute pressure. N2O exhibits amultitude of narrow absorption features within the lasers’ tun-ing range, which permits demonstration of the high-resolutioncapabilities of MHS. The reference detector receives light thatdid not interact with the sample. By frequency tuning the LOlaser through the injection current, the RF beat notes can beplaced at any arbitrary RF frequency. For convenience, thefrequency of the beat note coinciding with the targetN2O tran-sition was chosen to be equal to a quarter of the free spectralrange difference (ΔFSR) of the two lasers, which places thetarget beat note at a frequency of ΔFSR∕4 and ensures thatboth folded and non-folded beat notes are equidistant (sepa-rated by ΔFSR∕2) throughout the entire RF spectrum [1].This effectively doubles the number of accessible beat notes

within the available detector bandwidth (1.1 GHz in our case).The injection currents of both lasers are continuously scannedusing synchronized 1 kHz triangular waveforms from twoarbitrary waveform generators. In addition to the synchronizedscan, the lasers are frequency-locked to each other, whichassures stability of the two lasers during the frequency scanacross the absorption feature. Ramp rates exceeding 10 kHzintroduced a noticeable locking instability, thereby degradingthe spectroscopic results. For that reason, a ramping rate of1 kHz was chosen as a compromise between measurementstability and relatively short spectral scan time.

The detector signals were amplified by two low-noise RFamplifiers (HD Communications HD28110) and measuredby two real-time spectrum analyzers (Tektronix RSA6101A/5103A) that recorded frequency and amplitude time traceswithin a bandwidth of 20 MHz around the center frequencyof the targeted beat note. In addition, an RF splitter followed bya bandpass filter was introduced in the reference signal path toprovide an input to the frequency discriminator locking circuit.A schematic of the frequency discriminator system is shown inFig. 2. It is based on a 9th order Butterworth low-pass filter(∼2.3 dB∕MHz) with a cutoff frequency that correspondsto approximately ΔFSR∕4. Relative frequency offset betweenthe two lasers was stabilized by active feedback control ofthe injection current to one of the lasers. The error signal wasobtained via a frequency discriminator, which converts fre-quency variations of the first multiheterodyne RF beat note(used in this experiment) to a voltage signal used as an inputto a proportional-integral-differential (PID) controller (SRSSIM960, 100 kHz bandwidth) that actively controls the injec-tion current to one of the lasers. It should be noted that thelimited bandwidth of the PID-controller was not sufficientto reduce the linewidths of the beat notes, but was effectivein stabilization of the beat notes’ center frequencies.

In Fig. 3, the stability of the frequency discriminator lockingscheme is presented in both a time-series measurement of thebeat note center frequency and the corresponding Allandeviation plot. As can be seen in the figure, the unlocked con-figuration exhibits a significant frequency drift after ∼10 sec-onds, whereas the locked configuration remains stable duringthe entire duration of the measurement. This enabled rapidsimultaneous frequency tuning of both lasers and ensured thatthe beat notes stayed within their demodulation bandwidth

Fig. 1. Schematic of the differential dispersion spectroscopy setup. Two FP-QCLs are used in the dual comb configuration with two detectors.The RF beat notes from the detector signals are measured by two separate real-time spectrum analyzers. Two bandpass filters are used to isolate aparticular beat note from the signal detector and the corresponding beat note from the reference detector. The phase difference between these twobeat notes is measured by a phase detector and recorded by an oscilloscope. For details on the locking procedure see Fig. 2.

244 Vol. 42, No. 2 / January 15 2017 / Optics Letters Letter

Page 3: Molecular dispersion spectroscopy based on Fabry Perot ... · Molecular dispersion spectroscopy based on Fabry–Perot quantum cascade lasers LUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,†

throughout the measurements, which simplified the post-processing of the data.

High-resolution swept-frequency absorption and dispersionmeasurements of low-pressure N2O were performed by

extracting a single beat note from the RF spectrum. InFig. 4(a), the time-dependent power of a beat note attenuatedby interaction with the R35e transition of N2O is shown. Thedata were acquired at a 150 MS/s sampling frequency during a1 ms triangular current scan (∼1.5 mApp). This corresponds toan optical frequency scan of ∼1 GHz that could be performedat any arbitrarily selected center frequency. Scans with largeramplitudes introduced laser instabilities, and thus scans cover-ing a full-spectral range could not be realized with this FP-QCLpair. Two synchronized real-time spectrum analyzers were usedto simultaneously measure the RF beat notes from both detec-tors. Each of the data points shown in Fig. 4(a) represent an8 μs time-window averaged beat note power, chosen to maxi-mize averaging while not distorting the line shape during thefast scan. The beat note power has been normalized with re-spect to the reference beat note, which cancels the majorityof the common mode amplitude noise. However, amplitudenoise that is not in common mode, such as fringes from theabsorption cell or different RF noise characteristics in the de-tector signal paths, cannot be canceled by this procedure. Thenon-random noise features of the fit residual are clear evidenceof the aforementioned phenomena. The short-term detectionlimit for absorption measurements has been estimated based onthe standard deviation of the residuals obtained from spectralfits using line-by-line models based on the HITRAN database[16]. The residuals shown in Fig. 4(a) result in a bandwidth

Fig. 2. (a) Schematic of the frequency discriminator locking circuit.TheRFsignal fromthe referencedetector is split into twopaths; oneunder-goes a frequency independent attenuation (20 dB), whereas the other islow-pass filtered. The attenuated signals are used as inputs to an AnalogDevices RF gain and phase detector, which outputs a frequency dependentvoltage used to actively stabilize the beat note frequencies. RF IN,radio-frequency input; LNA, low noise amplifier; BPF, band-pass filter.(b) Filter characteristics of the ninth order Butterworth low-pass filteraround the desired beat note frequency. The left inset shows the filter cir-cuit, andthe right inset shows the filter characteristicsovera100MHzspan.

Fig. 3. (a) Time-series measurement of the detuning of the centerfrequency of the 40 MHz beat note in the unlocked (blue) and locked(red) configurations. (b) The corresponding Allan deviation plot clearlydemonstrates the improved relative frequency stability of the system.

Fig. 4. (a) Normalized beat note power for the beat note at40 MHz. The reference detector signal has been used to correct thebaseline. Each data point represents a time-window averaging of 8 μs.The data were acquired with a 1 kHz triangular ramp applied simul-taneously to both FP-QCLs. The solid line shows a HITRAN fit[16] based on a Voigt absorption line shape using the R35eN2O tran-sition at 1199.3576 cm−1. (b) The corresponding phase measurementobtained by detection of the differential phase between the two40 MHz beat notes measured by the signal and reference detectors.The solid line represents a HITRAN fit [16] based on a Voigtdispersion line shape using the same transition.

Letter Vol. 42, No. 2 / January 15 2017 / Optics Letters 245

Page 4: Molecular dispersion spectroscopy based on Fabry Perot ... · Molecular dispersion spectroscopy based on Fabry–Perot quantum cascade lasers LUKASZ A. STERCZEWSKI,1,2,† JONAS WESTBERG,1,†

normalized noise-equivalent absorption (NEA) of 2.8 × 10−4∕ffiffiffiffiffiffi

Hzp

. This is approximately an order of magnitude improve-ment of the NEAwith respect to our earlier work [15], which islargely due to the increased frequency scan rate of the system,permitted by the additional frequency-locking stability.

For the dispersion measurements both detector signals arebandpass filtered around the desired beat note frequency(ΔFSR∕4 in this case) and supplied to the inputs of a phasedetector (Analog Devices AD8302), which measures the phasedifference of the two beat notes. Since the beat note from thesignal channel is affected by the molecular interaction whileboth signal and reference beat notes share the same phase noisefrom the laser sources, the differential phase measurement ef-fectively cancels a large portion of any common mode phasenoise. As shown in Fig. 1, the output from the phase detectoris measured and recorded by a 100 MHz mixed signal oscillo-scope (2.5 GS/s). The recorded output from the phase detectoris shown in Fig. 4(b). Since this measurement is inherently dif-ferential, no further post-processing needs to be performed toextract the dispersion spectrum. The NEA of 3.0 × 10−4∕

ffiffiffiffiffiffi

Hzp

was calculated based on the dispersion spectrum residuals, us-ing the absorption measurement as a reference. Equivalently, aminimum detectable phase change of 6.6 × 10−6 rad∕

ffiffiffiffiffiffi

Hzp

canbe deduced from Fig. 4(b). It should be emphasized that thebaseline in the dispersion spectrum is typically less complexthan in the absorption spectra. While the fit in Fig. 4(a) useda second-order polynomial for the baseline, Fig. 4(b) was onlycorrected for a constant offset. The majority of the fringe-like residuals are induced by the windows of the gas cell, whichcan be addressed by a proper optical design that includesanti-reflection coatings, wedged windows, and angled opticalinterfaces.

In conclusion, a rapidly swept mid-IR QCL-based MHSsystem operating at 8.5 μm and capable of extracting high-resolution dispersion spectra recorded with one of the FP-QCLmodes has been demonstrated. The system offers a responsetime of less than 1 ms and allows for simultaneous acquisitionof absorption and dispersion spectra. The system is frequencystabilized using a cost-effective frequency discriminator circuitto provide relative optical frequency stability during the lasertuning at kilohertz rates. The assessment of dispersion spectraeffectively suppresses the influence of beat note amplitudenoise, typically limiting the sensitivity of conventional absorp-tion-based multiheterodyne mid-IR systems. In addition, itnegates the need of real-time power normalization of the de-tector signal, which is otherwise necessary to avoid inaccuracieswhen acquiring data subjected to transmission fluctuations.Such conditions may typically occur during the monitoringof transient chemical reactions in noisy environments, whichis often the case for combustion processes, where the rapid re-sponse time of a multiheterodyne dispersion system may poten-tially be a viable solution. Finally, the current proof of conceptis restricted to analog demodulation of a single-mode dispersion

spectrum and its extension to multimode acquisition wouldrequire implementation of parallel frequency-selective phasedetection. Alternatively, the basic mechanism of phase detec-tion can be implemented via digital signal processing of theentire RF spectrum using computational phase-retrieval andcoherent-averaging algorithms [17].

Funding. Defense Advanced Research Projects Agency(DARPA) (W31P4Q161001); National Science Foundation(NSF) (EEC-0540832).

Acknowledgment. The authors thank F. Capasso atHarvard University and his collaborators L. Diehl and M.Troccoli for providing the QCLs used in this study.

†These authors contributed equally to this work.

REFERENCES

1. S. Schiller, Opt. Lett. 27, 766 (2002).2. I. Coddington, N. Newbury, and W. Swann, Optica 3, 414 (2016).3. F. K. Tittel, D. Richter, and A. Fried, Solid-State Mid-Infrared Laser

Sources, D. I. T. Sorokina and D. K. L. Vodopyanov, eds., Topicsin Applied Physics (Springer, 2003), pp. 458–529.

4. D. J. Bamford, D. J. Cook, S. J. Sharpe, and A. D. Van Pelt, Appl. Opt.46, 3958 (2007).

5. Y. Wang, M. Nikodem, E. Zhang, F. Cikach, J. Barnes, S. Comhair,R. A. Dweik, C. Kao, and G. Wysocki, Sci. Rep. 5, 9096 (2015).

6. F. Keilmann, C. Gohle, and R. Holzwarth, Opt. Lett. 29, 1542 (2004).7. Z. Zhang, T. Gardiner, and D. T. Reid, Opt. Lett. 38, 3148 (2013).8. T. J. Kippenberg, R. Holzwarth, and S. A. Diddams, Science 332, 555

(2011).9. G. Villares, A. Hugi, S. Blaser, and J. Faist, Nat. Commun. 5, 5192

(2014).10. D. Burghoff, Y. Yang, D. J. Hayton, J.-R. Gao, J. L. Reno, and Q. Hu,

Opt. Express 23, 1190 (2015).11. D. Burghoff, T.-Y. Kao, N. Han, C. W. I. Chan, X. Cai, Y. Yang, D. J.

Hayton, J.-R. Gao, J. L. Reno, and Q. Hu, Nat. Photonics 8, 462(2014).

12. G. Villares, S. Riedi, J. Wolf, D. Kazakov, M. J. Süess, P. Jouy, M.Beck, and J. Faist, Optica 3, 252 (2016).

13. G. Wysocki and D. Weidmann, Opt. Express 18, 26123 (2010).14. Y. Wang, M. G. Soskind, W. Wang, and G. Wysocki, Appl. Phys. Lett.

104, 31114 (2014).15. A. Hangauer, J. Westberg, E. Zhang, and G. Wysocki, Opt. Express

24, 25298 (2016).16. L. S. Rothman, I. E. Gordon, Y. Babikov, A. Barbe, D. Chris Benner,

P. F. Bernath, M. Birk, L. Bizzocchi, V. Boudon, L. R. Brown, A.Campargue, K. Chance, E. A. Cohen, L. H. Coudert, V. M. Devi,B. J. Drouin, A. Fayt, J.-M. Flaud, R. R. Gamache, J. J. Harrison,J.-M. Hartmann, C. Hill, J. T. Hodges, D. Jacquemart, A. Jolly, J.Lamouroux, R. J. Le Roy, G. Li, D. A. Long, O. M. Lyulin, C. J.Mackie, S. T. Massie, S. Mikhailenko, H. S. P. Müller, O. V.Naumenko, A. V. Nikitin, J. Orphal, V. Perevalov, A. Perrin, E. R.Polovtseva, C. Richard, M. A. H. Smith, E. Starikova, K. Sung, S.Tashkun, J. Tennyson, G. C. Toon, V. G. Tyuterev, and G.Wagner, J. Quant. Spectrosc. Radiat. Transfer 130, 4 (2013).

17. D. Burghoff, Y. Yang, and Q. Hu, Sci. Adv. 2, e1601227 (2016).

246 Vol. 42, No. 2 / January 15 2017 / Optics Letters Letter