Next Article in Journal
Synergistic Mechanism of Rare-Earth Modification TiO2 and Photodegradation on Benzohydroxamic Acid
Next Article in Special Issue
TDLAS Monitoring of Carbon Dioxide with Temperature Compensation in Power Plant Exhausts
Previous Article in Journal
A Two-Stage Method for Parameter Identification of a Nonlinear System in a Microbial Batch Process
Previous Article in Special Issue
Real-Time Vision through Haze Based on Polarization Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mid-Infrared Tunable Laser-Based Broadband Fingerprint Absorption Spectroscopy for Trace Gas Sensing: A Review

1
State Key Laboratory of Precision Measuring Technology and Instruments, Tianjin University, Tianjin 300072, China
2
Key Laboratory of Micro Opto-electro Mechanical System Technology, Tianjin University, Ministry of Education, Tianjin 300072, China
3
Key Laboratory of Advanced Electrical Engineering and Energy Technology, Tianjin Polytechnic University, Tianjin 300387, China
4
School of Mechanical Engineering, Hebei University of Technology, Tianjin 300130, China
*
Author to whom correspondence should be addressed.
Submission received: 30 November 2018 / Revised: 9 January 2019 / Accepted: 11 January 2019 / Published: 18 January 2019
(This article belongs to the Special Issue State-of-the-art Laser Gas Sensing Technologies)

Abstract

:
The vast majority of gaseous chemical substances exhibit fundamental rovibrational absorption bands in the mid-infrared spectral region (2.5–25 μm), and the absorption of light by these fundamental bands provides a nearly universal means for their detection. A main feature of optical techniques is the non-intrusive in situ detection of trace gases. We reviewed primarily mid-infrared tunable laser-based broadband absorption spectroscopy for trace gas detection, focusing on 2008–2018. The scope of this paper is to discuss recent developments of system configuration, tunable lasers, detectors, broadband spectroscopic techniques, and their applications for sensitive, selective, and quantitative trace gas detection.

1. Introduction

Laser-based trace gas sensing is becoming more popular in a wide variety of areas including urban and industrial emission measurement [1,2], environmental and pollution monitoring [3,4], chemical analysis and industrial process control [5,6], medical diagnostics [7,8], homeland security [9], and scientific research [10,11]. With the increase in global environmental, ecological, and energy issues, laser-based trace gas detection technology has attracted unprecedented attention. Laser gas sensing is based on the analysis of characteristic spectra of molecules, mostly known as tunable diode laser absorption spectroscopy (TDLAS) or tunable laser absorption spectroscopy (TLAS). Laser absorption spectrometers (LAS), also known laser gas analyzers (LGA), enjoy the merits of non-contact, fast response time, high sensitivity and selectivity, the potential to be calibration-free, low maintenance requirements, and a long life cycle. LAS is particularly suitable for in situ, online analysis, and real-time detection.
Laser gas sensing has undergone tremendous progress along with the advancement in tunable semiconductor lasers in the last few decades. Direct absorption spectroscopy (DAS) is the most common technique for simple optical configuration, signal processing, and potential absolute measurement. DAS often suffers from low sensitivity (absorbance ~10−3) due to the interference from 1/f noise in the system and laser power fluctuation. There are basically two ways to improve the sensitivity in this situation: (1) reduce the noise in the signal; or (2) increase the absorption. The former can be achieved by using modulation technique, e.g., wavelength modulation spectroscopy (WMS) and frequency modulation spectroscopy (FMS), with a typical sensitivity of absorbance ~10−5. The latter can be obtained by placing the gas inside a cavity in which the light passes through multiple times to increase the interaction length, e.g., multiple-pass or long path absorption cells, and cavity enhanced absorption spectroscopy (CEAS) [12,13]. Both ways of reducing noise and increasing absorption can be applied to a same system, e.g., cavity enhanced wavelength modulation spectrometry [13] and noise-immune cavity-enhanced optical heterodyne molecular spectroscopy (NICE-OHMS) [14,15].
FMS is a method of optical heterodyne spectroscopy capable of rapid measurement of the absorption or dispersion associated with narrow spectral features. The absorption or dispersion is measured by detecting the heterodyne beat signal that occurs when the FMS optical spectrum of the probe wave is distorted by the spectral feature of interest. Recently, dispersion spectroscopy, namely chirped laser dispersion spectroscopy (CLaDS) [16] or heterodyne phase sensitive dispersion spectroscopy (HPSDS) [17], has attracted attention for its immunity to optical intensity changes and superb linearity in the measurement of concentration.
CEAS and its new versions, e.g., cavity ring-down spectroscopy (CRDS) [18], broadband cavity ring-down spectroscopy [19], phase-shift cavity ring-down spectroscopy [20], integrated cavity output spectroscopy (ICOS) [21], and continuous wave cavity enhanced absorption spectrometry (cw-CEAS) [22], provide much larger pathlength enhancement by using a resonant cavity, and thus have highly sensitive absorbance ~10−7–10−9.
Practically, the simplest and most promising method to enhance the signal of trace gas detection is to perform the detection at wavelengths where the transitions have larger line strengths, e.g., using fundamental rovibrational bands or electronic transitions. The fundamental rovibrational bands of a vast majority of gaseous chemical substances, located at the mid-infrared spectral region (MIR, 2.5–25 μm), are due to the transitions of molecular rovibrational energy states. In general, these bands have stronger line strengths than the overtone and combination bands typically used in the visible and near-IR regions. The MIR spectrum depends on the physical properties of the molecule such as the number and type of atoms, the bond angles, and the bond strength. Thus, the MIR spectrum is uniquely characterized by highly specific spectroscopic features and is considered the molecular signature, which allows both the identification and quantification of the molecular species, especially suitable for larger molecules, e.g., volatile organic compounds (VOCs) [23].
VOCs are gaseous organic chemicals at the conditions of normal temperature and pressure (NTP, 293.15 K and 101.325 kPa). There are several hundred types of VOCs, some of which are dangerous to human health or cause harm to the environment. VOCs monitoring has attracted attention for long time. Commonly used spectroscopic techniques for VOCs detection are Fourier-transform infrared spectroscopy (FTIR) and differential optical absorption spectroscopy (DOAS) [24]; however, their sensitivity, selectivity, and fragile optical setup are not always sufficient for harsh applications.
Recently, newly commercialized MIR detectors and lasers, especially quantum cascade lasers (QCLs) [25] and interband cascade lasers (ICLs) [26], have stimulated the development of high-performance, compact, and rugged gas sensors. Traditionally, TLAS use a discrete narrow absorption lines of small molecules for gas sensing. For larger molecules, however, so many lines overlapping with each other results in the spectral features being broad and smooth, except for occasional spikes [23,27]. These spectral features are distinct from those of the discrete narrow absorption lines with a Lorentzian, Gaussian, or Voigt profile. Detection of trace gas with broadband absorption is much more difficult than with an isolated narrow spectral line. Extra effort should be made to cope with the challenges of the broadband of larger molecules.
In this paper, we primarily review tunable laser-based broadband absorption spectroscopy for trace gas detection in 2008–2018. After a brief overview of the principle (Section 2), we discuss the system configuration, including MIR tunable lasers, detectors, and optical configuration in Section 3. We discuss broadband spectroscopic techniques concerning derivative spectroscopy (Section 4), WMS (Section 5), and optical frequency comb spectroscopy (Section 6). Section 7 is a collection of MIR gas sensing applications. Section 8 gives conclusions and prospects.

2. Principle

Quantitative spectral analysis is based on the Beer-Lambert law, which gives the relationship between the incident and the transmitted radiation through a gas cell filled with a molecular gas sample:
I ( ν ) = I 0 ( ν ) × e x p { σ ( ν ) × L × C } ,
where I0 and I are the incident and transmitted radiant powers, respectively; σ is absorption cross section of the molecule in cm2/molecule; L is absorption pathlength in cm; C is the density of the molecule in molecule/cm3. Usually, the absorption cross section σ is also used to describe the absorption intensity. The line strength is retrieved by spectrally integrating the absorption line shape and applying the ideal gas law:
S ( T ) = K B T A X i L P   r i s o ,
where KB, T (K), and P (Pa) are the Boltzmann constant, gas temperature, and total pressure of the gas sample, respectively; Xi is the amount fraction of i species; A (cm−1) is integral absorbance; riso is a correction factor for isotopic fractionation of the gas sample.
WMS is commonly used to improve the sensitivity of gas sensing. The WMS theory and signal model have been detailed previously [28], and so are only briefly reviewed here. A periodic sawtooth ramp ridden by a high-frequency sinusoidal is applied to the laser injection current, thus the laser wavenumber ν ( t ) = ν c + ν a cos ω t is scanned across the transition of gas to be detected, where νc and να, are the laser center wavenumber and modulation depth, respectively; ω is the radian frequency. In case of ideal conditions, ignoring all kinds of interference, the modulated absorption signal is detected by a photodiode and then processed using a lock-in amplifier to demodulate the signal at the harmonics (1f, 2f, 3f, etc.). The second harmonic component (WMS-2f) is commonly used for calculating the concentration of target gas. In the case of optically thin ( σ ( ν ) L C 0.05 ), the ideal 2f signal is modeled as:
A i d e a l   2 f = 2 I 0 C L π 0 π α ( ν c + ν a c o s θ ) c o s 2 θ d θ I 0 C L ,
where α is the absorption coefficient and θ = ω t is the phase angle. When the incident laser intensity I0 and optical path L are constant, the amplitude of WMS-2f signal is proportional to the gas concentration. Practically, apart from the 2f signal descript in Equation (3), the detected signal consists of random noise and the derivation of optical fringes [29,30]. The optical fringes appear as unpleasant spectral features that are usually mixed with the target absorption, and constitute one of the major obstacles in the gas detection. In a well-designed and -fabricated system, the optical fringes should be reduced, and only small residual fringes remain with sinusoidal waveforms, while random noise is seen as small time-varying wiggles superimposed on the true underlying signal, with small standard deviation. Thus, the detected signal could be described as:
A detected   2 f = e n + a j ( t ) × cos ( ω j ( t ) × t ) + A ideal   2 f ,
where αj(t) and ωj(t) are the instantaneous amplitude and frequency of jth fringe component, respectively; Aideal 2f is the WMS-2f signal modeled by Equation (3). The profiles of second harmonic of absorption, fringes, and noise will inherit the featuresof their origination. These profile differences among WMS-2f, harmonic of optical fringes, and noise will be a novel breakthrough point to distinguish and eliminate the interference from the signal (details in Section 5.3).

3. System Configuration

A typical LAS consists of a laser, a photodetector, and an optical configuration for light interaction with gas. For WMS-based LAS, there are, additionally, a laser modulator and a signal demodulator, the latter usually by a lock-in amplifier (LIA).
The laser is the LAS’s key component; it usually needs to be continuously tunable mode-hop-free, reliable, single frequency with narrow linewidth (typically <1 MHz), and to have low noise intensity. Historically, lead-salt diode lasers have been developed in a MIR gas sensor. However, these lasers require cooling to liquid nitrogen temperatures and present problems for mode hops and multi-mode operation. Recently, great progress in laser technology has brought many types of excellent lasers, e.g., QCLs and ICLs.
High-sensitive and low-noise detectors are essential for trace gas detection. The most popular commercial infrared detector is a mercury-cadmium-telluride (MCT, or HgCdTe) photoconductive semiconductor-based detector. The MCT detector enjoys a very wide spectral response (2 to 25 µm) and higher speed of detection. Its main limitation is that it needs cooling to reduce noise due to the thermally excited current carriers. Alternatively, newly developed quantum heterostructure detectors could play a vital role in future infrared detection [31].
The optical configuration provides interaction between light and gas’ the interaction length directly relates with the sensitivity of gas detection. Thus, a long interaction length is desired to achieve high sensitivity. Multiple-pass cells (MPCs) and open long path are commonly used in LAS to measure low-concentration components or to observe weak spectra in gas. The requirements of compactness, small sample volume, and fast response time stimulated the development of a new type of gas cell. Recently, a hollow waveguide (HWG)-based gas cell has been found to boast the advantages of small sample volume and fast response time [8,32], whereas substrate-integrated hollow waveguides (iHWG) are compact integrated sensors [33]. On the other hand, the need for non-fixed open-path gas detection, e.g., leak detection, aroused the development of standoff remote sensing without a retroreflector [34,35,36].

3.1. Mid-Infrared Tunable Lasers

MIR tunable laser technology has undergone tremendous development in the past decade, which involves QCL, ICL, difference frequency generator (DFG) [37], optical parametric oscillator (OPO) [38], fluoride fiber lasers [39], hollow-core fiber gas laser [40], VCSEL [41], and II–VI chalcogenides-based MIR lasers [42]. As there have been many excellent reviews of MIR light sources [25,43,44,45], here we present only a brief overview of the newly developed, highly reliable, and most widely used in spectroscopy, focusing on QCL, ICL, VCSEL, and optical frequency comb (OFC).

3.1.1. Quantum Cascade Lasers

QCL, first demonstrated by Faist et al. in 1994 [46], emits by intersubband transitions between energy levels inside superlattice quantum wells rather than the material bandgap energies in conventional lasers. The most attractive distributed feedback (DFB) QCLs were commercialized in 2004. Now, many commercial providers offer cw- and RT-operated QCLs in different configurations ranging from Fabry-Perot devices, to DFB resonators, to external cavities-based (EC-) QCLs, as well as high-power devices (with watts) in the infrared to terahertz spectral region. QCLs are attractive for infrared countermeasure, metrology, high-resolution spectroscopy, and chemical sensing applications [7,47,48,49,50,51,52,53].
QCLs enjoy a broad gain profile of hundreds of wavenumbers and a narrow linewidth about 0.5 MHz by providing monochromatic feedback, i.e., DFB and EC [54]. The natural linewidth can be as low as a few hundred hertz. The emission wavelength of DFB-QCLs can be tuned only a couple of wavenumbers, whereas EC-QCLs can provide hundreds of wavenumbers of coverage with drawbacks of slow mechanical speed, instability, alignment of multiple optical components, and high price. Broadband QCLs have been also built by an array of DFB QCLs with closely spaced emission wavelengths and fabricated monolithically with wavelength coverage of several micrometers [55]. QCL arrays enjoy broad tuning and high spectral resolution, which opens up a wide range of new possibilities for fast, compact, and mechanically robust solutions with high customizability [56].
Overall, QCLs have been greatly improved, but many challenges remain, such as intervalley scattering, heat removal from the core region, and interface scattering, which limit the performance of QCLs, especially at short wavelengths [57]. Although InP-based QCLs emitting at wavelengths of 3–4 μm have been demonstrated recently [58,59], they are still not commercially available.

3.1.2. Interband Cascade Lasers

ICL was presented by Yang in 1995 [60]. RT-cw operation of ICL was first demonstrated in 2008 [61]. Only a few years ago, DFB-ICLs became commercially available with an optical power of milliwatts and a spectral tuning range of a few wavenumbers by Nanoplus GmbH [26]. Like QCLs, ICLs employ the concept of bandstructure engineering to achieve an optimized laser design and reuse injected electrons to emit multiple photons. However, ICLs’ photons are generated by interband transitions rather than the intersubband transitions used in QCLs, which allows ICLs to achieve lower input powers than is possible with QCLs.
DFB-ICLs provide single mode, narrow linewidth (~0.7 MHz) [62], low power consumption (hundreds of milliwatts), a compact system solution, and RT-cw emission in the 3–6 μm range, which fill the MIR gap perfectly. DFB-ICLs have been used for the ppb-level detection of many important gases, especially hydrocarbon species [63,64,65,66,67,68,69], which leads to many important applications in various areas, for example, clinical diagnostics [70,71], combustion probing [72], environmental monitoring [73,74], and remote sensing.

3.1.3. Mid-Infrared Vertical-Cavity Surface-Emitting Lasers

MIR VCSELs have been paid great attention in the past decade for their advantages of low power consumption, low beam divergence, narrow and single-fundamental-mode, high wavelength tunability, and on-wafer testing capability [45]. The buried tunnel junction (BTJ) concept yields high-performance VCSELs in the wavelength ranges of 1.3–2.6 μm [75] and 2.3–3.0 μm [76], respectively. An interband cascade VCSEL has achieved lasing to λ~3.4 μm in pulsed mode at temperatures up to 70 °C [77], and a 4 μm VCSEL has been developed by using a single-stage active region with eight type-II quantum wells combined with BTJ technology [78]. VCSELs operate with very low threshold currents of several mA and very low power consumption of milliwatts. They provide a single mode by distributed Bragg reflector with a moderate linewidth of tens of MHz. VCSELs are particularly suited for compact and battery-powered sensors.

3.1.4. Mid-Infrared Optical Frequency Comb

OFCs consist of a series of discrete, narrow, stable, equally spaced spectral lines that have a fixed phase relationship between them. These combs can span a broadband of frequency range that have found important applications in areas such as metrology, spectroscopy, and optical communications [79]. To date, OFC sources have extended from the ultraviolet, visible, infrared, and terahertz spectral regions [80]. We only focus on the MIR-OFC, including mode-locked, difference frequency generators (DFG), optical parametric oscillation (OPO), and direct modulation OFC.
Mode-Locked Laser. The most popular way of generating a frequency comb is with a mode-locked laser. Mode-locked lasers produce a series of optical pulses separated in time by the round-trip time of the laser cavity. The spectrum of such a pulse train approximates a series of Dirac delta functions separated by the repetition rate of the laser. In the past decade, there have been hundreds of demonstrations of generating OFC with various lasers including Ti: sapphire solid-state lasers [81], Er: fiber lasers [82], Kerr-lens mode-locked lasers [83], QCLs [84], and ICLs [85] with repetition rates typically between to MHz to 10 GHz. The issues with mode-locked OFC are lower output power, still experimental, and not commercially available.
Difference Frequency Generators. DFG is the most versatile method to generate mode-hop-free tunable broad laser by a nonlinear optical process. The two necessary conditions to achieve MIR OFC output are high-coherence pump lasers and a suitable order nonlinear crystal. There are more than a dozen demonstrated crystals, among which LiNbO3 and ZnGeP2 are the most popular. DFG-based OFC have more power and stability and do not require an oscillating cavity or a high threshold of the optical parametric process [86], whereas the drawbacks are high cost and a relatively low conversion efficiency [37,86,87].
Optical parametric oscillator. An OPO is another parametric nonlinear optical process used to provide a high output power and versatile sources of coherent radiation for spectral regions inaccessible to lasers. OPO has been shown providing high power OFCs based on Yb: fiber laser covering several micrometers [88,89]. Researchers have not only shown that degenerate synchronously OPOs are efficient tools to transfer near-infrared (NIR) frequency combs to the mid-infrared; also, they can utilize pump lasers in NIR and expand the spectrum to the mid-infrared [38,90]. OPO offer high output power with broad spectral coverage, but require a free-space resonator, external pumping sources, and many optical components [91]. This makes OPO sources bulky, vulnerable to any external disturbances, and very impractical in field applications.
Direct Frequency-Modulation Combs. Direct FM combs have been experimentally demonstrated in semiconductor lasers, such as QCLs [91,92,93,94,95,96,97], ICLs [85], quantum dot (QD) [98], and quantum dash lasers [99]. These lasers are passively mode-locked with cw or quasi-cw output. Direct FM combs enjoy broadband, wide repetition frequency from kHz to THz [100] with line width of kHz to MHz level. Direct FM comb offers the possibility of a portable, chip-scale device with low power consumption, which are desired in spectroscopic trace gas sensing.

3.2. Infrared Detector

There has been exciting progress in MCT junction technology, which could design and fabricate a high-performance MCT photovoltaic detector for operation in RT and in situ applications [101]. The commercialized detector benefits from the use of optimized material, device architecture, concentrators of radiation, enhanced absorption, and shields against thermal radiation, and can achieve directivity as high as 1011 cm·√Hz·W−1 in RT conditions [102]. Alternatively, more progress has been noted in quantum engineering-based detectors.

3.2.1. Quantum Heterostructure Detector

Quantum heterostructure-based infrared detectors, including quantum cascade (QC), quantum well (QW), quantum dot, and combinations of both QDs and QWs in a dot-in-a-well (DWELL) strategy detectors, have high detectivity and low dark currents [31,103,104]. QC detectors have low noise due to their low interference of background radiation. QW detectors are characterized by a narrow spectral band and are easily fabricated. However, the disadvantages of QW detectors include low operating temperatures and the requirement of a scattering filter to adjust the incident light angle [105]. QD detectors have higher operating temperatures and are capable of absorbing normally incident photons. However, they have much lower quantum efficiency due to the lower absorption and capture probabilities of incident light [104,106]. DWELL devices have shown promise in terms of detectivity and spectral range, though their operating temperatures remained fairly low.
Most quantum heterostructure detectors are still at the experimental stage, with different fabrication technologies and materials being used. Few commercial products have been reported, which leads to the restriction of such detectors’ application in MIR detection. However, recent advances in this field such as new device-chip hybridization [107], antimonide-based membrane synthesis integration, and strain engineering [108] may be scalable methods for the fabrication of commercially available heterostructure detectors.

3.2.2. Infrared Avalanche Photodiodes

An avalanche photodiode (APD) provides an internal multiplication necessary to achieving high avalanche gain at low bias with low noise and high bandwidth [109]. Presently, MCT APD is the most promising for trace gases in standoff remote sensing with high sensitivity. The demands of night vision and LIDAR systems stimulate the development of MCT APD arrays [110]. Overall, commercialized high-sensitive and bandwidth MCT APD would play a vital role in many areas.

3.3. New Types of Gas Cells

Recently, new types of gas cells have been developed to achieve long optical path in a smaller volume, including modified MPCs [111,112,113,114,115,116], circular multi-reflection (CMR) cells [117,118,119,120,121,122,123,124], and HWG [32,63,69,125,126,127,128]. Guo and Sun reviewed the progress in modified MPCs and CMR cells and compared them in terms of optical pathlength (OPL), volume, and path-to-volume ratio (PVR) [116]. HWG-based gas cells have been discussed in detail in books and reviews [129,130,131]. We compared the main features among the modified MPCs, CMR cells and HWG, as shown in Table 1.
There are three categories of HWG employed in spectroscopic gas sensing, namely Ag/AgI-coated HWG (Ag/AgI-HWG), photonic bandgaps HWG (PBG-HWG), and iHWG [131], and the comparison of them is as shown in Table 1. HWGs work as both an optical waveguide and gas transmission cell that provide an extended OPL yielding high sensitivity measurements [69]. It is worth noting that filling PBG-HWG with analyte gas for sensing is difficult owing to the considerable back-pressure building up in the hollow structure. HWGs are ideal candidates for gas cells due to their high PVR and ability to transmit light at a fairly wide wavelength range, which means building a MIR sensor platform is feasible.

3.4. Fully Integrated Sensors

Profiting from the development of MIR lasers, detectors, and gas cells, as described above, the laser spectrometer can be potentially integrated into a miniaturized and compact system for gas sensing, as shown in Figure 1, maintaining or even enhancing the achievable sensitivity [132]. In this subsection, we introduce the progress and applications of the fully integrated MIR laser spectrometer, focusing on QCL or ICL-coupled HWG gas sensors.
An integrated sensor consists of a compact MIR laser, a new gas cell, and a detector, as shown in Figure 1. The configuration provides an attractive solution for miniaturized and practical applications. The laser should be a QCL, ICL, EC-QCL, or VECSEL. The miniaturized gas cells, i.e., HWG or iHWG, cover the wavelength range of 3.0–11.0 μm, and provide a response time as fast as seconds [32] and a sample volume of sub-milliliter [131]. iHWG-based cells provide excellent modularity and mechanical stability, with effective OPL of hundreds of millimeters and larger losses per unit length than Ag/AgI-HWG-based cells [32], which could achieve an effective OPL of several meters. The sensor could operate in the DAS, WMS, or intrapulse modulation regime as well [33,133,134,135,136].

3.5. Open Path Detection without Retroreflectors

The requirements of non-fixed open-path gas detection, e.g., the atmospheric environmental monitoring, leak detection, security early warning, etc., promoted the standoff remote sensing. We focus on the open-path-averaged gas concentrations by the backscatter light from a remote hard target or topographic target. Other open-path systems that are deployed as point samplers or long-path with retroreflectors are beyond the scope of this review.
In order to realize standoff gas detection, a number of laser-based techniques are available, such as TDLAS [34,35,137,138], PAS [139], differential absorption lidar (DIAL) [140], CLaDS [141], and more recently active coherent laser spectrometers (ACLaS) [36]. The basic architectures of these techniques are shown in Figure 2. The OPL is often variable and unknown; therefore, performance is often quoted in a similar fashion to that of open path gas detectors, using the pathlength-integrated unit of ppm·m [142]. The detection limits for such systems commonly range from sub ppm·m to several hundred ppm·m [35,36,141], typically over distances of open path from several meters to hundreds of meters or even kilometers. The performance of this system is typically limited by the level of received laser power, which is dependent on the incident laser power, distance between receiver and scattering target, type of scattering materials, and size of the receiver aperture [142,143,144]. High-power and broadband tuning lasers, such as OPO and EC-QCLs, are desired to achieve a higher signal-to-noise ratio (SNR) [145,146,147]. However, high-sensitive and low-noise detectors are particularly important for applications in public areas where eye safety should be considered.
Recently, standoff trace detection has achieved tremendous progress and been applied in leaks [35], environmental monitoring [141], explosives [148,149], combustion [147], unmanned aerial vehicles (UAV) [150,151], and many other promising applications as well.

4. Detection Methods: Derivative Spectroscopy

MIR spectroscopy is attractive for the strong fingerprint signature, and the commonly used DAS often suffers from low selectivity for the spectra overlapping of broadband absorption. Derivative spectroscopy uses first or higher derivatives of absorbance with respect to wavelength for qualitative analysis and for quantification with higher selectivity [152]. The derivatization of zero-order spectrum can lead to separation of overlapped signals or elimination of background caused by other compounds in a sample [153,154].
Derivative spectra could be obtained by the Savitzky-Golay (SG) smoothing/differentiation procedure, which is widely implemented in instrumental software or in packages for spectral data processing. The resolution enhancement in the second derivative spectrum depends on the data spacing in original spectra, absorption peak profile, parameters of SG (i.e., window size and polynomial order). To maximize the separation of the peaks in a second derivative spectrum, the original spectra should be recorded at high resolution and using appropriate parameters [155]. Other methods used to calculate the derivative spectra include numerical differentiation [156] and continuous wavelet transform [157]. The latter has been proven to be efficient in the analysis of overlapping spectra and is advantageous for providing higher SNR and flexibility in searching for absorption peaks.
Derivative spectroscopy can be used in various spectral region including UV [158], visible [159], NIR [160,161], and MIR. The method has a close dependence on instrumental parameters, like speed of scan, the linewidth, and SNR. The derivatization can amplify the noise signals in the resulting curves, which normally leads to a higher SNR. Another disadvantage is the non-robust character of the selected parameters of the elaborated methods. The selected parameters of this method are applicable only for the studied system and every change in composition requires re-optimization and the selection of new parameters of derivatization [153]. Without a homogeneous protocol of optimization, the parameters of the method vary, and most researchers did not describe the parameter selection in their published articles.

5. Detection Methods: Modulation Spectroscopy for Wideband Absorption

WMS has been demonstrated to have high sensitivity for sensing gas with an isolated spectral line. When the modulation index is small, WMS is approximately expressed as derivative spectroscopy. For the MIR spectral region, however, the fingerprint spectra are often broad, serried, crowded, and even overlap within the coverage of a tunable laser. To ensure the detection sensitivity and selectivity, the essential procedures include optimizing modulation index, varied modulation amplitude, removing fringes and noise interference, and multicomponent spectral fitting.

5.1. Optimizing the Modulation Index

Since the recorded harmonic signals used in WMS are heavily dependent on the spectral line profile and modulation index adopted in the WMS system, the situation for sensing a larger molecular gas with a broadband spectrum is quite different. Moreover, the similarity of signal waveform between a broadband spectrum and the intrinsic optical fringes interference will seriously complicate the signal processing [30] and deteriorate the sensing sensitivity and precision.
The modulation index always plays a pivotal role in WMS-based measurement. A modulation index of 2.2 is recognized as the optimum to achieve the maximum SNR with isolated Gaussian or Lorentzian line profile. For broadband spectrum, however, the reordered harmonic signal would broaden to overlap with the adjacent spectrum, interference, and optical fringes with the so-called modulation index optimum. The overlapping may deteriorate and even disable the WMS measurement. So the modulation index determination should balance the spectra discrimination and the SNR in WMS with broadband spectrum.
We investigated the determination of modulation index for the broadband absorption, taking the carbon disulfide (CS2) spectrum around 2177.6 cm−1 as an example. To evaluate the spectra discrimination of the WMS-2f signal, we particularly defined a parameter, SD [27]:
S D = h C D 0.5 × ( h A + h B ) ,
where A and B are the adjacent valleys of neighboring WMS-2f signals, C is the middle point of the line connecting A and B, point D at the signal curve is vertical to point C, hA and hB are the amplitude of absorption valley A and B, respectively. hCD is the height difference between point C and D. The parameter SD is a constant from 0 to 1, which represents the WMS-2f completely overlapping and parting, respectively, as shown in the insert panel of Figure 3b.
The parameter SD and normalized amplitude of WMS-2f signals under different modulation index was plotted in Figure 3c, which reverse with the modulation index. To balance the SD and normalized amplitude of WMS-2f signals, the modulation index around 1.0 should be a good compromise for optimizing WMS with the band spectrum of CS2 detection.

5.2. Varied Modulation Amplitude

A modulation amplitude setting always confronts multi-spectrum with different widths, which creates a significant dilemma. Any single modulation amplitude cannot cover a large spectral width difference, e.g., a difference of the half width at half maximum is more than 50%. A practical way to achieve this is to use a varied modulation amplitude for multi-spectrum detection, namely WMS with varied modulation amplitude (WMS-VMA) [162].
The WMS-VMA was realized mainly by a homemade digital lock-in amplifier (DLIA), which performs the modulation, demodulation of WMS-1f and WMS-2f, and generation of reference signal by an integrated field-programmable gate array-based circuit. The DLIA generates an arbitrary waveform signal by direct digital frequency synthesis. The sine waveform with varied amplitude is prepared in advance and stored in the DLIA’s random access memory, and then read out to control the sequence using the frequency controller integrated in the DLIA. The method has been verified by a multiparameter optical sensor with typical broadband spectroscopy [162].

5.3. Removing Fringes and Noise Interference

LAS always suffers interference from electric noise and optical fringes; the latter are caused by multiple reflections upon optical interface, i.e., the so-called “etalon effect.” The interference causes low sensitivity and precision in the spectrometers. Since the optical fringes are small and sine-like in a well-designed and -fabricated system, the signal profiles of molecular absorption always exhibit distinct differences to those of optical fringes and electric noise. The WMS-2f signal profile would inherit their difference with an optimized modulation index. So, the molecular absorption can be distinguished from optical fringes and electric noise by the signal profile.
An empirical mode decomposition (EMD) algorithm has been successfully used in WMS to remove optical fringes [27] and noise [163] by characterization of the signal profile. The procedure for employing EMD to decompose and reconstruct the WMS-2f signal is described in Figure 4a. The detected 2f signal was decomposed into intrinsic mode functions (IMF), which have a specific physical source and are meaningful. The simplest criterion, whether IMF from molecular absorption or interference is the correlation coefficient (R) between an IMF and the reference 2f signal, is shown in Figure 4b,c.
The components with lower R most likely come from fringes or noise, while components with larger R with reference absorption must be the absorption. All these components with higher R can be added together to reconstruct a WMS-2f free from the interference of fringes and noise. Generally, this method can improve the sensor sensitivity by about 30% [164].

5.4. Multicomponent Spectral Fitting

Multicomponent spectral fitting is mainly used to eliminate spectral interference, which commonly occurs in MIR detection. Furthermore, benefitting from the use of multivariate regression and nonlinear least square fitting, it can calculate the multi-component concentration in a gas mixture, i.e., achieving multiple component detection simultaneously using a single DFB ICL [164,165]. The Levenberg-Marquardt (LM) algorithm, also known as the damped least-squares method, was applied for the data fitting. Reference WMS-2f signals of all the components were obtained beforehand. The concentrations of all components in the mixture comply with the constraint condition of non-negative parameters. In each iteration of the fitting routine, the WMS-2f signal of the mixture was simulated with the updated parameters. Once the routine converged, the best fitting parameters were determined as the concentrations of the components.
Multicomponent spectral detection could benefit from the redundancy of the multiple spectra, not only in the magnitude of the absorption but also in the line shape related to temperature and pressure broadening. To make full use of the information buried in the detected spectral lines, we presented an improved multicomponent spectral fitting algorithm for the sensor. We also applied the method of normalizedWMS-2f by 1f for the sensor immunity of laser energy fluctuation [162].

6. Detection Methods: Optical Frequency Comb Spectroscopy

Frequency combs enjoy high spectral resolution and broad spectral coverage that make them a unique spectroscopic tool for precision spectroscopy and for multi-species detection [80,166]. Direct frequency comb spectroscopy (DFCS) employs optical frequency combs to probe spectral features in a parallel fashion [167]. We briefly review the DFCS including frequency comb-based Fourier transform spectroscopy (FC-FTS), cavity-enhanced direct frequency comb spectroscopy (CE-DFCS), and virtual imaging phased array spectroscopy (VIPAS).

6.1. Frequency Comb Fourier Transform Spectroscopy

FC-FTS is the measurement of the Fourier transform of the interferogram based on frequency combs source, which offers excellent spectral brightness and spatial coherence. There are two implementations of FC-FTS: Michelson interferometer-based Fourier transform spectroscopy and dual-comb spectroscopy (DCS) [168]. Each presents its own distinct advantages, but both rely on the same physical principle.
Michelson interferometer-based FC-FTS. In FC-FTS, OFC works as the light source for Fourier transform spectrometers (FTS). The high spectral brightness, together with the spatial and temporal coherence of the combs, enable acquisition times orders of magnitude shorter than in conventional FTIR spectroscopy. Thus, FC-FTS could be promising in standoff chemical sensing of transient, non-repeatable phenomenal like combustion, plasmas, and explosions [169]. Presently, the applications of FC-FTS is still limited by spectral width, lower comb intensities, mechanically scanned mirrors to record the interferogram, placing limits on the acquisition of spectra [170]. The interferometer records the interference pattern between two combs as slightly different because the light reflected by the moving mirror of the interferometer is Doppler shifted [171,172].
Dual comb spectroscopy. DCS uses two frequency combs of slightly differing line spacing, one for reference and the other for sample detection [168]. From each pair of optical lines, one from each comb, a radio frequency beat note is generated on a detector. In this way, optical frequencies are converted into radio frequencies such that the amplitude and phase changes caused by the interaction of one of the combs with a sample can be detected. DCS has an advantage similar to a Michelson interferometer but without moving parts and employing a single point detector, in which case a minimum detectable absorption ~1 × 10−8 cm−1 has been demonstrated [173].
Despite an intriguing potential for the measurement of molecular spectra spanning tens of nanometers within tens of microseconds at Doppler-limited resolution, the development of dual-comb spectroscopy is hindered by the demanding stability requirements of the laser combs [168,174,175]. For an ideal, the interference sampled waveform can be Fourier transformed to display the signal spectrum. In reality, the main difficulty comes from the time and phase fluctuations of the frequency comb. However, we experimentally demonstrate that the means of real-time dual-comb spectroscopy can be used to overcome this problem [176]. The true value of DCS for sensitive molecular detection lies in the MIR [177,178], and compact design will be obtained using the semiconductor laser comb technology in the future [179,180,181].

6.2. Cavity-Enhanced Direct Frequency Comb Spectroscopy

CE-DFCS combines broad spectral bandwidth, high spectral resolution, precise frequency calibration, and ultrahigh detection sensitivity all in one experimental platform based on an optical frequency comb interacting with a high-finesse optical cavity [166,182]. Michael et al. demonstrate a minimum detectable absorption of 8 × 10−10 cm−1, a spectral resolution of 800 MHz, and 200 nm of spectral coverage [183]. Moreover, combined with VIPA technology, the minimum detectable concentration is 1.7 × 1011 cm−3 [184]. Whereas, a great deal of CE-DFCS applications are in the visible and NIR spectral region, MIR CE-DFCS is attracting a lot of attention [182,185,186]. It is noteworthy that the mode spacing between the cavity and comb should be well matched, or the operation will always too be complicated [186,187].

6.3. Virtual Imaging Phased Array Spectroscopy

VIPAS provides alternative approaches that circumvent this problem by directly measuring the power or phase of individual comb teeth that have interacted with the molecular gas [188]. In this case, a novel high-resolution crossed spectral disperser is employed to project the various frequency comb modes onto a two-dimensional digital camera. The distinguishing feature of the present approach is the use of a side-entrance etalon called a VIPA disperser in the visible spectral range. The multiple reflections within the VIPA etalon interfere such that the exiting beam has different frequencies emerging at different angles. Sensing with a broadband comb directly interrogates an absorbing sample, after which the spectrum is dispersed in two dimensions and sensed with detector arrays. Thus, the spectrometer transforms the one-dimensional comb into something more reminiscent of a two-dimensional ‘brush’.
With the VIPA method, Diddams et al. [189] resolved 2200 comb modes covering a 6.5 THz span with resolution 1.2 GHz, while Gohle et al. [190] resolved 4000 modes covering a 4 THz span with resolution 1 GHz. However, due to the limitations in optical coating and array detector technology, VIPA is available in the visible and NIR ranges. Proof-of-principle demonstrations have been carried out in the MIR wavelength region spectral resolution 600 MHz (0.02 cm−1) [191], and resolution 1 GHz (0.03 cm−1) [192]. MID APD arrays may change this situation in the near future.

7. Summary of MIR Gas Sensing

MIR trace gas sensing in the molecular fingerprint region developed rapidly in the near decade mostly due to the commercialization of MIR tunable lasers, i.e., DFB-QCL and DFB-ICL, which could be proven by rough statistics on the number of studies published annually on TOPIC: (Mid-Infrared Lasers) AND TOPIC: (Sensing) in the Web of Science, as shown in Figure 5. Though many prominent works on MIR trace gas sensing were performed before 2008, NO was detected through the ν1 band near 1875 cm−1 with limit of detection (LOD) of ppb level by DAS [193] and WMS [194], respectively, for industrial processes and vehicle emissions monitoring. We only summarize gas detection based on MIR absorption spectroscopy with tunable lasers in the near decade for the last 10 years; see Table 2 for single component detection and Table 3 for multi-component detection.

8. Conclusions and Future Prospects

MIR spectral trace gas sensing is particularly attractive for its unique and strong fingerprint absorption. Higher sensitivity has been achieved by solving all the challenges of the spectral feature, i.e., broad, serried, crowding, and even overlapping in the MIR region. Benefiting from the methods mentioned above, multicomponent simultaneous detection is an expected achievement.
The invention and commercialization of high-performance MIR lasers, i.e., DFB-QCL and DFB-ICL, promoted the development and application of MIR tunable laser-based trace gas sensors, especially in miniaturized and portable applications. Though more than 100 types of gas have been detected by tunable laser-based sensors, there are huge demands for higher detection sensitivity, or, in extreme conditions or scientific exploration, more other types of gas need to be detected. Thus, we believe greater progress will be achieved in the next decade, which may include:
(1) Compact integrated gas sensors. Compact sensors with lower power consumption or battery power could benefit from VCSEL, DFB-ICL, and iHWG. Even a sensor system on a chip will become possible using integrated optics with the fabrication of miniaturized devices integrating the electronics and optics. These sensors could play a vital role in portable and wearable applications that could be applied for breath analysis, diagnostics, metabolomics, environmental safety detection, and related applications.
(2) Multicomponent gas sensors. Multicomponent sensors will achieve more progress, benefiting from the wider wavelength coverage by integrated laser arrays, OFC, or EC-QCL. More species could be detected simultaneously by a particularly devised broadband laser, which could expand their applications in scientific research, including combustion diagnosis, chemical reaction process dynamics, exhaled breath analysis, and metabolomics.
(3) Standoff remote sensing. The techniques of open-path standoff detection by backscattered MIR light provide a promising method of prompt and flexible assessment of atmospheric environmental, leaks, explosive, and security in handheld devices or UAV. The detection sensitivity could be substantially improved by newly developed high-performance MIR detectors and progress in high-power DFB-QCL.
(4) Ultra-sensitive sensing. With the development of a mid-infrared laser source and high-performance detector, combined with cavity enhancement technology and noise immunity technology, ultra-high detection sensitivity becomes possible.

Author Contributions

This review article was jointly written and proof-read by all authors. Z.D. proposed the idea, drafted the outline and structure, and contributed to the principle, detection methods as well as conclusion. S.Z. collected references, composed the whole manuscript, and contributed to the summary. J.L. contributed to the system configuration and detection methods. N.G. contributed to the system configuration. K.T. contributed to the detection methods.

Funding

This research was funded by the National Natural Science Foundation of China (61505142), the Tianjin Natural Science Foundation (16JCQNJC02100), the Science & Technology Development Fund of the Tianjin Education Commission for Higher Education (2017KJ085), and the Natural Science Foundation of Hebei Province (F2014202065).

Acknowledgments

This work was supported by the open project of Key Laboratory of Micro Opto-electro Mechanical System Technology, Tianjin University, Ministry of Education.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jane, H.; Ralph, P.T. Optical gas sensing: A review. Meas. Sci. Technol. 2013, 24. [Google Scholar] [CrossRef]
  2. Miller, D.J.; Sun, K.; Tao, L.; Khan, M.A.; Zondlo, M.A. Open-path, quantum cascade-laser-based sensor for high-resolution atmospheric ammonia measurements. Atmos. Meas. Tech. 2014, 7, 81–93. [Google Scholar] [CrossRef] [Green Version]
  3. Lassen, M.; Lamard, L.; Balslev-Harder, D.; Peremans, A.; Petersen, J.C. Mid-infrared photoacoustic spectroscopy for atmospheric NO2 measurements. In Proceedings of the SPIE OPTO, San Francisco, CA, USA, 30 January–1 February 2018; p. 7. [Google Scholar]
  4. Huszár, H.; Pogány, A.; Bozóki, Z.; Mohácsi, Á.; Horváth, L.; Szabó, G. Ammonia monitoring at ppb level using photoacoustic spectroscopy for environmental application. Sens. Actuator B-Chem. 2008, 134, 1027–1033. [Google Scholar] [CrossRef]
  5. Lackner, M. Tunable diode laser absorption spectroscopy (TDLAS) in the process industries—A review. Rev. Chem. Eng. 2007, 23, 65–147. [Google Scholar] [CrossRef]
  6. Haas, J.; Mizaikoff, B. Advances in Mid-Infrared Spectroscopy for Chemical Analysis. Annu. Rev. Anal. Chem. 2016, 9, 45–68. [Google Scholar] [CrossRef] [PubMed]
  7. Ghorbani, R.; Schmidt, F.M. Real-time breath gas analysis of CO and CO2 using an EC-QCL. Appl. Phys. B-Lasers Opt. 2017, 123. [Google Scholar] [CrossRef]
  8. Liu, L.; Xiong, B.; Yan, Y.; Li, J.; Du, Z. Hollow Waveguide-Enhanced Mid-Infrared Sensor for Real-Time Exhaled Methane Detection. IEEE Photonics Technol. Lett. 2016, 28, 1613–1616. [Google Scholar] [CrossRef]
  9. Bauer, C.; Sharma, A.K.; Willer, U.; Burgmeier, J.; Braunschweig, B.; Schade, W.; Blaser, S.; Hvozdara, L.; Müller, A.; Holl, G. Potentials and limits of mid-infrared laser spectroscopy for the detection of explosives. Appl. Phys. B-Lasers Opt. 2008, 92, 327–333. [Google Scholar] [CrossRef]
  10. Röpcke, J.; Welzel, S.; Lang, N.; Hempel, F.; Gatilova, L.; Guaitella, O.; Rousseau, A.; Davies, P.B. Diagnostic studies of molecular plasmas using mid-infrared semiconductor lasers. Appl. Phys. B-Lasers Opt. 2008, 92, 335–341. [Google Scholar] [CrossRef]
  11. Bolshov, M.A.; Kuritsyn, Y.A.; Romanovskii, Y.V. Tunable diode laser spectroscopy as a technique for combustion diagnostics. Spectroc. Acta Part B-Atom. Spectrosc. 2015, 106, 45–66. [Google Scholar] [CrossRef]
  12. O’Keefe, A.; Deacon, D.A.G. Cavity ring-down optical spectrometer for absorption measurements using pulsed laser sources. Rev. Sci. Instrum. 1988, 59, 2544–2551. [Google Scholar] [CrossRef]
  13. Zybin, A.; Kuritsyn, Y.A.; Mironenko, V.R.; Niemax, K. Cavity enhanced wavelength modulation spectrometry for application in chemical analysis. Appl. Phys. B-Lasers Opt. 2004, 78, 103–109. [Google Scholar] [CrossRef]
  14. Foltynowicz, A.; Schmidt, F.M.; Ma, W.; Axner, O. Noise-immune cavity-enhanced optical heterodyne molecular spectroscopy: Current status and future potential. Appl. Phys. B-Lasers Opt. 2008, 92, 313. [Google Scholar] [CrossRef]
  15. Talicska, C.N.; Porambo, M.W.; Perry, A.J.; McCall, B.J. Mid-infrared concentration-modulated noise-immune cavity-enhanced optical heterodyne molecular spectroscopy of a continuous supersonic expansion discharge source. Rev. Sci. Instrum. 2016, 87, 063111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Nikodem, M.; Krzempek, K.; Karwat, R.; Dudzik, G.; Abramski, K.; Wysocki, G. Chirped laser dispersion spectroscopy with differential frequency generation source. Opt. Lett. 2014, 39, 4420–4423. [Google Scholar] [CrossRef] [PubMed]
  17. Martín-Mateos, P.; Acedo, P. Heterodyne phase-sensitive detection for calibration-free molecular dispersion spectroscopy. Opt. Express 2014, 22, 15143–15153. [Google Scholar] [CrossRef] [PubMed]
  18. Zhou, S.; Han, Y.; Li, B. Simultaneous detection of ethanol, ether and acetone by mid-infrared cavity ring-down spectroscopy at 3.8 μm. Appl. Phys. B-Lasers Opt. 2016, 122. [Google Scholar] [CrossRef]
  19. Takehiro, H.; Takayuki, O.; Masafumi, I.; Norihiko, N.; Masaru, H. Optical-Fiber-Type Broadband Cavity Ring-Down Spectroscopy Using Wavelength-Tunable Ultrashort Pulsed Light. Jpn. J. Appl. Phys 2013, 52, 040201. [Google Scholar] [CrossRef]
  20. Grilli, R.; Ciaffoni, L.; Orr-Ewing, A.J. Phase-shift cavity ring-down spectroscopy using mid-IR light from a difference frequency generation PPLN waveguide. Opt. Lett. 2010, 35, 1383–1385. [Google Scholar] [CrossRef]
  21. Moyer, E.J.; Sayres, D.S.; Engel, G.S.; St. Clair, J.M.; Keutsch, F.N.; Allen, N.T.; Kroll, J.H.; Anderson, J.G. Design considerations in high-sensitivity off-axis integrated cavity output spectroscopy. Appl. Phys. B-Lasers Opt. 2008, 92, 467–474. [Google Scholar] [CrossRef]
  22. Van Helden, J.H.; Lang, N.; Macherius, U.; Zimmermann, H.; Röpcke, J. Sensitive trace gas detection with cavity enhanced absorption spectroscopy using a continuous wave external-cavity quantum cascade laser. Appl. Phys. Lett. 2013, 103, 131114. [Google Scholar] [CrossRef]
  23. Sigrist, M.W. Air monitoring by spectroscopic techniques. In Proteomics; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1994; pp. 335–338. [Google Scholar]
  24. Du, Z.; Zhai, Y.; Li, J.; Hu, B. Techniques of On-Line Monitoring Volatile Organic Compounds in Ambient Air with Optical Spectroscopy. Spectrosc. Spectr. Anal. 2009, 29, 3199–3203. [Google Scholar] [CrossRef]
  25. Wu, D.; Razeghi, M.; Lu, Q.Y.; Mcclintock, R.; Slivken, S.; Zhou, W. Recent progress of quantum cascade laser research from 3 to 12 μm at the Center for Quantum Devices. Appl. Opt. 2017, 56, H30–H44. [Google Scholar] [CrossRef]
  26. Vurgaftman, I.; Weih, R.; Kamp, M.; Meyer, J.R.; Canedy, C.L.; Kim, C.S.; Kim, M.; Bewley, W.W.; Merritt, C.D.; Abell, J. Interband cascade lasers. J. Phys. D-Appl. Phys. 2015, 48, 123001. [Google Scholar] [CrossRef] [Green Version]
  27. Du, Z.; Li, J.; Cao, X.; Gao, H.; Ma, Y. High-sensitive carbon disulfide sensor using wavelength modulation spectroscopy in the mid-infrared fingerprint region. Sens. Actuator B-Chem. 2017, 247, 384–391. [Google Scholar] [CrossRef]
  28. Reid, J.; Labrie, D. Second-harmonic detection with tunable diode lasers—Comparison of experiment and theory. Appl. Phys. B-Lasers Opt. 1981, 26, 203–210. [Google Scholar] [CrossRef]
  29. Werle, P. Accuracy and precision of laser spectrometers for trace gas sensing in the presence of optical fringes and atmospheric turbulence. Appl. Phys. B-Lasers Opt. 2011, 102, 313–329. [Google Scholar] [CrossRef]
  30. Xiong, B.; Du, Z.; Li, J. Modulation index optimization for optical fringe suppression in wavelength modulation spectroscopy. Rev. Sci. Instrum. 2015, 86, 113104. [Google Scholar] [CrossRef]
  31. Downs, C.; Vandervelde, E.T. Progress in Infrared Photodetectors Since 2000. Sensors 2013, 13, 5054–5098. [Google Scholar] [CrossRef] [Green Version]
  32. Francis, D.; Hodgkinson, J.; Livingstone, B.; Black, P.; Tatam, R.P. Low-volume, fast response-time hollow silica waveguide gas cells for mid-IR spectroscopy. Appl. Opt. 2016, 55, 6797–6806. [Google Scholar] [CrossRef]
  33. Wilk, A.; Carter, J.C.; Chrisp, M.; Manuel, A.M.; Mirkarimi, P.; Alameda, J.B.; Mizaikoff, B. Substrate-Integrated Hollow Waveguides: A New Level of Integration in Mid-Infrared Gas Sensing. Anal. Chem. 2013, 85, 11205–11210. [Google Scholar] [CrossRef] [PubMed]
  34. Frish, M.B.; Wainner, R.T.; Laderer, M.C.; Green, B.D.; Allen, M.G. Standoff and Miniature Chemical Vapor Detectors Based on Tunable Diode Laser Absorption Spectroscopy. IEEE Sens. J. 2010, 10, 639–646. [Google Scholar] [CrossRef]
  35. Diaz, A.; Thomas, B.; Castillo, P.; Gross, B.; Moshary, F. Active standoff detection of CH4 and N2O leaks using hard-target backscattered light using an open-path quantum cascade laser sensor. Appl. Phys. B-Lasers Opt. 2016, 122. [Google Scholar] [CrossRef]
  36. Macleod, N.A.; Rose, R.; Weidmann, D. Middle infrared active coherent laser spectrometer for standoff detection of chemicals. Opt. Lett. 2013, 38, 3708–3711. [Google Scholar] [CrossRef] [PubMed]
  37. Lee, K.F.; Hensley, C.J.; Schunemann, P.G.; Fermann, M.E. Midinfrared frequency comb by difference frequency of erbium and thulium fiber lasers in orientation-patterned gallium phosphide. Opt. Express 2017, 25, 17411–17416. [Google Scholar] [CrossRef] [PubMed]
  38. Vodopyanov, K.L.; Sorokin, E.; Sorokina, I.T.; Schunemann, P.G. Mid-IR frequency comb source spanning 4.4–5.4 μm based on subharmonic GaAs optical parametric oscillator. Opt. Lett. 2011, 36, 2275–2277. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, J.; Wu, M.; Huang, B.; Tang, P.; Zhao, C.; Shen, D.; Fan, D.; Turitsyn, S.K. Widely Wavelength-Tunable Mid-Infrared Fluoride Fiber Lasers. IEEE J. Sel. Top. Quantum Electron. 2018, 24, 1–7. [Google Scholar] [CrossRef] [Green Version]
  40. Xu, M.; Yu, F.; Knight, J. Mid-infrared 1 W hollow-core fiber gas laser source. Opt. Lett. 2017, 42, 4055–4058. [Google Scholar] [CrossRef]
  41. Jayaraman, V.; Segal, S.; Lascola, K.; Burgner, C.; Towner, F.; Cazabat, A.; Cole, G.D.; Follman, D.; Heu, P.; Deutsch, C. Room temperature continuous wave mid-infrared VCSEL operating at 3.35um. In Proceedings of the SPIE OPTO, San Francisco, CA, USA, 19 February 2018; p. 7. [Google Scholar]
  42. Mirov, S.B.; Fedorov, V.V.; Martyshkin, D.; Moskalev, I.S.; Mirov, M.; Vasilyev, S. Progress in Mid-IR Lasers Based on Cr and Fe-Doped II–VI Chalcogenides. IEEE J. Sel. Top. Quantum Electron. 2015, 21, 292–310. [Google Scholar] [CrossRef]
  43. Schunemann, P.G.; Zawilski, K.T.; Pomeranz, L.A.; Creeden, D.J.; Budni, P.A. Advances in nonlinear optical crystals for mid-infrared coherent sources. J. Opt. Soc. Am. B 2016, 33, D36–D43. [Google Scholar] [CrossRef]
  44. Jung, D.; Bank, S.; Lee, M.L.; Wasserman, D. Next-generation mid-infrared sources. J. Opt. 2017, 19, 123001. [Google Scholar] [CrossRef]
  45. Tournié, E.; Baranov, A.N. Chapter 5—Mid-Infrared Semiconductor Lasers: A Review. In Semiconductors and Semimetals; Coleman, J.J., Bryce, A.C., Jagadish, C., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; Volume 86, pp. 183–226. [Google Scholar]
  46. Faist, J.; Capasso, F.; Sivco, D.L.; Sirtori, C.; Hutchinson, A.L.; Cho, A.Y. Quantum Cascade Laser. Science 1994, 264, 553. [Google Scholar] [CrossRef] [PubMed]
  47. Curl, R.F.; Capasso, F.; Gmachl, C.; Kosterev, A.A.; McManus, B.; Lewicki, R.; Pusharsky, M.; Wysocki, G.; Tittel, F.K. Quantum cascade lasers in chemical physics. Chem. Phys. Lett. 2010, 487, 1–18. [Google Scholar] [CrossRef]
  48. Westberg, J.; Sterczewski, L.A.; Wysocki, G. Mid-infrared multiheterodyne spectroscopy with phase-locked quantum cascade lasers. Appl. Phys. Lett. 2017, 110, 141108. [Google Scholar] [CrossRef] [Green Version]
  49. Gong, L.; Lewicki, R.; Griffin, R.J.; Flynn, J.H.; Lefer, B.L.; Tittel, F.K. Atmospheric ammonia measurements in Houston, TX using an external-cavity quantum cascade laser-based sensor. Atmos. Chem. Phys. 2011, 11, 9721–9733. [Google Scholar] [CrossRef] [Green Version]
  50. Cui, X.; Dong, F.; Zhang, Z.; Sun, P.; Xia, H.; Fertein, E.; Chen, W. Simultaneous detection of ambient methane, nitrous oxide, and water vapor using an external-cavity quantum cascade laser. Atmos. Environ. 2018, 189, 125–132. [Google Scholar] [CrossRef]
  51. Spearrin, R.M.; Goldenstein, C.S.; Schultz, I.A.; Jeffries, J.B.; Hanson, R.K. Simultaneous sensing of temperature, CO, and CO2 in a scramjet combustor using quantum cascade laser absorption spectroscopy. Appl. Phys. B-Lasers Opt. 2014, 117, 689–698. [Google Scholar] [CrossRef]
  52. Nadeem, F.; Mandon, J.; Khodabakhsh, A.; Cristescu, S.M.; Harren, F.J.M. Sensitive Spectroscopy of Acetone Using a Widely Tunable External-Cavity Quantum Cascade Laser. Sensors 2018, 18, 2050. [Google Scholar] [CrossRef]
  53. Maity, A.; Pal, M.; Maithani, S.; Banik, G.D.; Pradhan, M. Wavelength modulation spectroscopy coupled with an external-cavity quantum cascade laser operating between 7.5 and 8 µm. Laser Phys. Lett. 2018, 15. [Google Scholar] [CrossRef]
  54. Hugi, A.; Maulini, R.; Faist, J. External cavity quantum cascade laser. Semicond. Sci. Technol. 2010, 2535, 83001–83014. [Google Scholar] [CrossRef]
  55. Zhou, W.; Wu, D.; McClintock, R.; Slivken, S.; Razeghi, M. High performance monolithic, broadly tunable mid-infrared quantum cascade lasers. Optica 2017, 4, 1228–1231. [Google Scholar] [CrossRef]
  56. Rauter, P.; Capasso, F. Multi-wavelength quantum cascade laser arrays. Laser Photonics Rev. 2015, 9, 452–477. [Google Scholar] [CrossRef] [Green Version]
  57. Razeghi, M.; Bandyopadhyay, N.; Bai, Y.; Lu, Q.; Slivken, S. Recent advances in mid infrared (3–5 µm) Quantum Cascade Lasers. Opt. Mater. Express 2013, 3, 1872–1884. [Google Scholar] [CrossRef]
  58. Bandyopadhyay, N.; Bai, Y.; Tsao, S.; Nida, S.; Slivken, S.; Razeghi, M. Room temperature continuous wave operation of λ ~ 3–3.2 μm quantum cascade lasers. Appl. Phys. Lett. 2012, 101, 241110. [Google Scholar] [CrossRef]
  59. Wolf, J.M.; Bismuto, A.; Beck, M.; Faist, J. Distributed-feedback quantum cascade laser emitting at 3.2 μm. Opt. Express 2014, 22, 2111–2118. [Google Scholar] [CrossRef]
  60. Yang, R.Q. Infrared laser based on intersubband transitions in quantum wells. Superlattices Microstruct. 1995, 17, 77. [Google Scholar] [CrossRef]
  61. Kim, M.; Canedy, C.L.; Bewley, W.W.; Kim, C.S.; Lindle, J.R.; Abell, J.; Vurgaftman, I.; Meyer, J.R. Interband cascade laser emitting at λ=3.75μm in continuous wave above room temperature. Appl. Phys. Lett. 2008, 92, 191110. [Google Scholar] [CrossRef]
  62. Du, Z.; Luo, G.; An, Y.; Li, J. Dynamic spectral characteristics measurement of DFB interband cascade laser under injection current tuning. Appl. Phys. Lett. 2016, 109, 011903. [Google Scholar] [CrossRef]
  63. Jose Gomes da Silva, I.; Tutuncu, E.; Nagele, M.; Fuchs, P.; Fischer, M.; Raimundo, I.M.; Mizaikoff, B. Sensing hydrocarbons with interband cascade lasers and substrate-integrated hollow waveguides. Analyst 2016, 141, 4432–4437. [Google Scholar] [CrossRef]
  64. Li, J.; Du, Z.; Zhang, Z.; Song, L.; Guo, Q. Hollow waveguide-enhanced mid-infrared sensor for fast and sensitive ethylene detection. Sens. Rev. 2017, 37, 82–87. [Google Scholar] [CrossRef]
  65. Li, C.; Zheng, C.; Dong, L.; Ye, W.; Tittel, F.K.; Wang, Y. Ppb-level mid-infrared ethane detection based on three measurement schemes using a 3.34-μm continuous-wave interband cascade laser. Appl. Phys. B-Lasers Opt. 2016, 122. [Google Scholar] [CrossRef]
  66. Li, C.; Dong, L.; Zheng, C.; Tittel, F.K. Compact TDLAS based optical sensor for ppb-level ethane detection by use of a 3.34 μm room-temperature CW interband cascade laser. Sens. Actuator B-Chem. 2016, 232, 188–194. [Google Scholar] [CrossRef] [Green Version]
  67. He, Q.; Zheng, C.; Lou, M.; Ye, W.; Wang, Y.; Tittel, F.K. Dual-feedback mid-infrared cavity-enhanced absorption spectroscopy for H2CO detection using a radio-frequency electrically-modulated interband cascade laser. Opt. Express 2018, 26, 15436–15444. [Google Scholar] [CrossRef] [PubMed]
  68. Du, Z.; Zhen, W.; Zhang, Z.; Li, J.; Gao, N. Detection of methyl mercaptan with a 3393-nm distributed feedback interband cascade laser. Appl. Phys. B-Lasers Opt. 2016, 122. [Google Scholar] [CrossRef]
  69. Li, J.; Luo, G.; Du, Z.; Ma, Y. Hollow waveguide enhanced dimethyl sulfide sensor based on a 3.3 μm interband cascade laser. Sens. Actuator B-Chem. 2018, 255, 3550–3557. [Google Scholar] [CrossRef]
  70. Risby, T.H.; Tittel, F.K. Current status of midinfrared quantum and interband cascade lasers for clinical breath analysis. Opt. Eng. 2010, 49, 11. [Google Scholar] [CrossRef]
  71. Ghorbani, R.; Schmidt, F.M. ICL-based TDLAS sensor for real-time breath gas analysis of carbon monoxide isotopes. Opt. Express 2017, 25, 12743–12752. [Google Scholar] [CrossRef]
  72. Girard, J.J.; Spearrin, R.M.; Goldenstein, C.S.; Hanson, R.K. Compact optical probe for flame temperature and carbon dioxide using interband cascade laser absorption near 4.2 µm. Combust. Flame 2017, 178, 158–167. [Google Scholar] [CrossRef]
  73. Sonnenfroh, D.M. Interband cascade laser–based sensor for ambient CH4. Opt. Eng. 2010, 49, 111118. [Google Scholar] [CrossRef]
  74. Edlinger, M.V.; Scheuermann, J.; Weih, R.; Zimmermann, C.; Nähle, L.; Fischer, M.; Koeth, J.; Höfling, S.; Kamp, M. Monomode Interband Cascade Lasers at 5.2 μm for Nitric Oxide Sensing. IEEE Photonics Technol. Lett. 2014, 26, 480–482. [Google Scholar] [CrossRef]
  75. Sprengel, S.; Andrejew, A.; Federer, F.; Veerabathran, G.K.; Boehm, G.; Amann, M.C. Continuous wave vertical cavity surface emitting lasers at 2.5 μm with InP-based type-II quantum wells. Appl. Phys. Lett 2015, 106, 151102. [Google Scholar] [CrossRef]
  76. Andrejew, A.; Sprengel, S.; Amann, M.-C. GaSb-based vertical-cavity surface-emitting lasers with an emission wavelength at 3 μm. Opt. Lett. 2016, 41, 2799–2802. [Google Scholar] [CrossRef] [PubMed]
  77. Bewley, W.W.; Canedy, C.L.; Kim, C.S.; Merritt, C.D.; Warren, M.V.; Vurgaftman, I.; Meyer, J.R.; Kim, M. Room-temperature mid-infrared interband cascade vertical-cavity surface-emitting lasers. Appl. Phys. Lett. 2016, 109, 151108. [Google Scholar] [CrossRef]
  78. Veerabathran, G.K.; Sprengel, S.; Andrejew, A.; Amann, M.C. Room-temperature vertical-cavity surface-emitting lasers at 4 μm with GaSb-based type-II quantum wells. Appl. Phys. Lett. 2017, 110, 071104. [Google Scholar] [CrossRef]
  79. Cossel, K.C.; Waxman, E.M.; Finneran, I.A.; Blake, G.A.; Ye, J.; Newbury, N.R. Gas-phase broadband spectroscopy using active sources: Progress, status, and applications [Invited]. J. Opt. Soc. Am. B 2017, 34, 104–129. [Google Scholar] [CrossRef] [PubMed]
  80. Schliesser, A.; Picqué, N.; Hänsch, T.W. Mid-infrared frequency combs. Nat. Photonics 2012, 6, 440–449. [Google Scholar] [CrossRef] [Green Version]
  81. Galli, I.; Cappelli, F.; Cancio, P.; Giusfredi, G.; Mazzotti, D.; Bartalini, S.; De Natale, P. High-coherence mid-infrared frequency comb. Opt. Express 2013, 21, 28877–28885. [Google Scholar] [CrossRef]
  82. Droste, S.; Ycas, G.; Washburn Brian, R.; Coddington, I.; Newbury Nathan, R. Optical Frequency Comb Generation based on Erbium Fiber Lasers. Nanophotonics 2016, 5, 196. [Google Scholar] [CrossRef]
  83. Cizmeciyan, M.N.; Cankaya, H.; Kurt, A.; Sennaroglu, A. Kerr-lens mode-locked femtosecond Cr2+:ZnSe laser at 2420 nm. Opt. Lett. 2009, 34, 3056–3058. [Google Scholar] [CrossRef]
  84. Burghoff, D.; Kao, T.-Y.; Han, N.; Chan, C.W.I.; Cai, X.; Yang, Y.; Hayton, D.J.; Gao, J.-R.; Reno, J.L.; Hu, Q. Terahertz laser frequency combs. Nat. Photonics 2014, 8, 462. [Google Scholar] [CrossRef]
  85. Bagheri, M.; Frez, C.; Sterczewski, L.A.; Gruidin, I.; Fradet, M.; Vurgaftman, I.; Canedy, C.L.; Bewley, W.W.; Merritt, C.D.; Kim, C.S.; et al. Passively mode-locked interband cascade optical frequency combs. Sci. Rep. 2018, 8, 3322. [Google Scholar] [CrossRef] [PubMed]
  86. Sotor, J.; Martynkien, T.; Schunemann, P.G.; Mergo, P.; Rutkowski, L.; Soboń, G. All-fiber mid-infrared source tunable from 6 to 9 μm based on difference frequency generation in OP-GaP crystal. Opt. Express 2018, 26, 11756–11763. [Google Scholar] [CrossRef] [PubMed]
  87. Zhu, F.; Hundertmark, H.; Kolomenskii, A.A.; Strohaber, J.; Holzwarth, R.; Schuessler, H.A. High-power mid-infrared frequency comb source based on a femtosecond Er:fiber oscillator. Opt. Lett. 2013, 38, 2360–2362. [Google Scholar] [CrossRef] [PubMed]
  88. Jin, Y.; Cristescu, S.M.; Harren, F.J.M.; Mandon, J. Two-crystal mid-infrared optical parametric oscillator for absorption and dispersion dual-comb spectroscopy. Opt. Lett. 2014, 39, 3270–3273. [Google Scholar] [CrossRef]
  89. Liu, P.; Wang, S.; He, P.; Zhang, Z. Dual-channel operation in a synchronously pumped optical parametric oscillator for the generation of broadband mid-infrared coherent light sources. Opt. Lett. 2018, 43, 2217–2220. [Google Scholar] [CrossRef] [PubMed]
  90. Smolski, V.O.; Yang, H.; Gorelov, S.D.; Schunemann, P.G.; Vodopyanov, K.L. Coherence properties of a 2.6–7.5 μm frequency comb produced as a subharmonic of a Tm-fiber laser. Opt. Lett. 2016, 41, 1388–1391. [Google Scholar] [CrossRef] [PubMed]
  91. Lu, Q.Y.; Razeghi, M.; Slivken, S.; Bandyopadhyay, N.; Bai, Y.; Zhou, W.J.; Chen, M.; Heydari, D.; Haddadi, A.; McClintock, R.; et al. High power frequency comb based on mid-infrared quantum cascade laser at λ ~ 9 μm. Appl. Phys. Lett. 2015, 106, 051105. [Google Scholar] [CrossRef]
  92. Faist, J.; Villares, G.; Scalari, G.; Rösch, M.; Bonzon, C.; Hugi, A.; Beck, M. Quantum Cascade Laser Frequency Combs. Nanophotonics 2016, 5, 272. [Google Scholar] [CrossRef]
  93. Dong, M.; Cundiff, S.T.; Winful, H.G. Physics of frequency-modulated comb generation in quantum-well diode lasers. Phys. Rev. A 2018, 97, 053822. [Google Scholar] [CrossRef]
  94. Hugi, A.; Villares, G.; Blaser, S.; Liu, H.C.; Faist, J. Mid-infrared frequency comb based on a quantum cascade laser. Nature 2012, 492, 229. [Google Scholar] [CrossRef]
  95. Jouy, P.; Wolf, J.M.; Bidaux, Y.; Allmendinger, P.; Mangold, M.; Beck, M.; Faist, J. Dual comb operation of λ ~ 8.2 μm quantum cascade laser frequency comb with 1 W optical power. Appl. Phys. Lett. 2017, 111, 141102. [Google Scholar] [CrossRef]
  96. Villares, G.; Hugi, A.; Blaser, S.; Faist, J. Dual-comb spectroscopy based on quantum-cascade-laser frequency combs. Nat. Commun. 2014, 5, 5192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kazakov, D.; Piccardo, M.; Wang, Y.; Chevalier, P.; Mansuripur, T.S.; Xie, F.; Zah, C.-e.; Lascola, K.; Belyanin, A.; Capasso, F. Self-starting harmonic frequency comb generation in a quantum cascade laser. Nat. Photonics 2017, 11, 789–792. [Google Scholar] [CrossRef] [Green Version]
  98. Liu, L.; Zhang, X.; Xu, T.; Dai, Z.; Dai, S.; Liu, T. Simple and seamless broadband optical frequency comb generation using an InAs/InP quantum dot laser. Opt. Lett. 2017, 42, 1173–1176. [Google Scholar] [CrossRef] [PubMed]
  99. Panapakkam, V.; Anthur, A.; Vujicic, V.; Gaimard, Q.; Merghem, K.; Aubin, G.; Lelarge, F.; Viktorov, E.; Barry, L.P.; Ramdane, A. Asymmetric corner frequency in the 1/f FM-noise PSD of optical frequency combs generated by quantum-dash mode-locked lasers. Appl. Phys. Lett. 2016, 109, 181102. [Google Scholar] [CrossRef]
  100. Forrer, A.; Rösch, M.; Singleton, M.; Beck, M.; Faist, J.; Scalari, G. Coexisting frequency combs spaced by an octave in a monolithic quantum cascade laser. Opt. Express 2018, 26, 23167–23177. [Google Scholar] [CrossRef] [PubMed]
  101. Martyniuk, P.; Rogalski, A. Performance limits of the mid-wave InAsSb/AlAsSb nBn HOT infrared detector. Opt. Quantum Electron. 2014, 46, 581–591. [Google Scholar] [CrossRef]
  102. Piotrowski, A.; Piotrowski, J. Uncooled Infrared Detectors in Poland, History and Recent Progress. In Proceedings of the 26th European Conference on Solid-State Transducers (Eurosensors), Krakow, Poland, 9–12 September 2012; pp. 1506–1512. [Google Scholar]
  103. Deng, Z.; Jeong, K.S.; Guyot-Sionnest, P. Colloidal Quantum Dots Intraband Photodetectors. ACS Nano 2014, 8, 11707–11714. [Google Scholar] [CrossRef]
  104. El-Tokhy, M.S.; Mahmoud, I.I.; Konber, H.A. Comparative study between different quantum infrared photodetectors. Opt. Quantum Electron. 2009, 41, 933–956. [Google Scholar] [CrossRef]
  105. Gueriaux, V.; l’Isle, N.B.d.; Berurier, A.; Huet, O.; Manissadjian, A.; Facoetti, H.; Marcadet, X.; Carras, M.; Trinité, V.; Nedelcu, A. Quantum well infrared photodetectors: Present and future. Opt. Eng. 2011, 50, 20. [Google Scholar] [CrossRef]
  106. Asano, T.; Hu, C.; Zhang, Y.; Liu, M.; Campbell, J.C.; Madhukar, A. Design Consideration and Demonstration of Resonant-Cavity-Enhanced Quantum Dot Infrared Photodetectors in Mid-Infrared Wavelength Regime (3–5 μm). IEEE J. Quantum Electron. 2010, 46, 1484–1491. [Google Scholar] [CrossRef]
  107. Dianat, P. A Scalable Low-Cost Manufacturing to Hybridize Infrared Detectors with Si Read-Out Circuits. In Proceedings of the 2018 IEEE Research and Applications of Photonics In Defense Conference (RAPID), Miramar Beach, FL, USA, 22–24 August 2018; pp. 1–3. [Google Scholar]
  108. Zamiri, M.; Anwar, F.; Klein, B.A.; Rasoulof, A.; Dawson, N.M.; Schuler-Sandy, T.; Deneke, C.F.; Ferreira, S.O.; Cavallo, F.; Krishna, S. Antimonide-based membranes synthesis integration and strain engineering. Proc. Natl. Acad. Sci. USA 2017, 114, E1–E8. [Google Scholar] [CrossRef] [PubMed]
  109. Singh, A.; Pal, R. Infrared Avalanche Photodiode Detectors. Def. Sci. J. 2017, 67, 159. [Google Scholar] [CrossRef] [Green Version]
  110. Singh, A.; Srivastav, V.; Pal, R. HgCdTe avalanche photodiodes: A review. Opt. Laser Technol. 2011, 43, 1358–1370. [Google Scholar] [CrossRef]
  111. Shen, C.; Zhang, Y.; Ni, J. Compact cylindrical multipass cell for laser absorption spectroscopy. Chin. Opt. Lett. 2013, 11, 091201. [Google Scholar] [CrossRef]
  112. Mohamed, T.; Zhu, F.; Chen, S.; Strohaber, J.; Kolomenskii, A.A.; Bengali, A.A.; Schuessler, H.A. Multipass cell based on confocal mirrors for sensitive broadband laser spectroscopy in the near infrared. Appl. Opt. 2013, 52, 7145–7151. [Google Scholar] [CrossRef]
  113. Kühnreich, B.; Höh, M.; Wagner, S.; Ebert, V. Direct single-mode fibre-coupled miniature White cell for laser absorption spectroscopy. Rev. Sci. Instrum. 2016, 87, 023111. [Google Scholar] [CrossRef] [PubMed]
  114. Ren, W.; Jiang, W.; Tittel, F.K. Single-QCL-based absorption sensor for simultaneous trace-gas detection of CH4 and N2O. Appl. Phys. B-Lasers Opt. 2014, 117, 245–251. [Google Scholar] [CrossRef]
  115. Liu, K.; Wang, L.; Tan, T.; Wang, G.; Zhang, W.; Chen, W.; Gao, X. Highly sensitive detection of methane by near-infrared laser absorption spectroscopy using a compact dense-pattern multipass cell. Sens. Actuator B-Chem. 2015, 220, 1000–1005. [Google Scholar] [CrossRef] [Green Version]
  116. Guo, Y.; Sun, L. Compact optical multipass matrix system design based on slicer mirrors. Appl. Opt. 2018, 57, 1174–1181. [Google Scholar] [CrossRef]
  117. Ofner, J.; Krüger, H.-U.; Zetzsch, C. Circular multireflection cell for optical spectroscopy. Appl. Opt. 2010, 49, 5001–5004. [Google Scholar] [CrossRef] [PubMed]
  118. Manninen, A.; Tuzson, B.; Looser, H.; Bonetti, Y.; Emmenegger, L. Versatile multipass cell for laser spectroscopic trace gas analysis. Appl. Phys. B-Lasers Opt. 2012, 109, 461–466. [Google Scholar] [CrossRef]
  119. Tuzson, B.; Mangold, M.; Looser, H.; Manninen, A.; Emmenegger, L. Compact multipass optical cell for laser spectroscopy. Opt. Lett. 2013, 38, 257–259. [Google Scholar] [CrossRef] [PubMed]
  120. Jouy, P.; Mangold, M.; Tuzson, B.; Emmenegger, L.; Chang, Y.C.; Hvozdara, L.; Herzig, H.P.; Wägli, P.; Homsy, A.; Rooij, N.F.D. Mid-infrared spectroscopy for gases and liquids based on quantum cascade technologies. Analyst 2014, 139, 2039–2046. [Google Scholar] [CrossRef]
  121. Knox, D.A.; King, A.K.; McNaghten, E.D.; Brooks, S.J.; Martin, P.A.; Pimblott, S.M. Novel utilisation of a circular multi-reflection cell applied to materials ageing experiments. Appl. Phys. B-Lasers Opt. 2015, 119, 55–64. [Google Scholar] [CrossRef]
  122. Mangold, M.; Tuzson, B.; Hundt, M.; Jágerská, J.; Looser, H.; Emmenegger, L. Circular paraboloid reflection cell for laser spectroscopic trace gas analysis. J. Opt. Soc. Am. A 2016, 33, 913–919. [Google Scholar] [CrossRef]
  123. Graf, M.; Looser, H.; Emmenegger, L.; Tuzson, B. Beam folding analysis and optimization of mask-enhanced toroidal multipass cells. Opt. Lett. 2017, 42, 3137–3140. [Google Scholar] [CrossRef]
  124. Graf, M.; Emmenegger, L.; Tuzson, B. Compact, circular, and optically stable multipass cell for mobile laser absorption spectroscopy. Opt. Lett. 2018, 43, 2434–2437. [Google Scholar] [CrossRef]
  125. Xiong, B.; Du, Z.; Liu, L.; Zhang, Z.; Li, J.; Cai, Q. Hollow-waveguide-based carbon dioxide sensor for capnography. Chin. Opt. Lett. 2015, 13, 111201–111204. [Google Scholar] [CrossRef]
  126. Tütüncü, E.; Nägele, M.; Fuchs, P.; Fischer, M.; Mizaikoff, B. iHWG-ICL: Methane Sensing with Substrate-Integrated Hollow Waveguides Directly Coupled to Interband Cascade Lasers. ACS Sens. 2016, 1, 847–851. [Google Scholar] [CrossRef]
  127. Gayraud, N.; Kornaszewski, Ł.W.; Stone, J.M.; Knight, J.C.; Reid, D.T.; Hand, D.P.; MacPherson, W.N. Mid-infrared gas sensing using a photonic bandgap fiber. Appl. Opt. 2008, 47, 1269–1277. [Google Scholar] [CrossRef] [PubMed]
  128. Jin, W.; Xuan, H.F.; Ho, H.L. Sensing with hollow-core photonic bandgap fibers. Meas. Sci. Technol. 2010, 21, 094014. [Google Scholar] [CrossRef]
  129. Mizaikoff, B. Mid-IR fiber-optic sensors. Anal. Chem. 2003, 75, 258A–267A. [Google Scholar] [CrossRef] [PubMed]
  130. Charlton, C.M.; Thompson, B.T.; Mizaikoff, B. Hollow Waveguide Infrared Spectroscopy and Sensing. In Frontiers in Chemical Sensors; Springer: Berlin/Heidelberg, Germany, 2005; pp. 133–167. [Google Scholar]
  131. Jin-Yi, L.; Zhen-Hui, D.; Wang, R.X.; Xu, Y.; Song, L.M.; Guo, Q.H. Applications of Hollow Waveguide in Spectroscopic Gas Sensing. Spectrosc. Spectr. Anal. 2017, 37, 2259–2266. [Google Scholar] [CrossRef]
  132. Liu, N.; Sun, J.; Deng, H.; Ding, J.; Zhang, L.; Li, J. Recent progress on gas sensor based on quantum cascade lasers and hollow fiber waveguides. In Proceedings of the Fourth International Conference on Optical and Photonics Engineering, Chengdu, China, 26–30 September 2016; p. 5. [Google Scholar]
  133. Wörle, K.; Seichter, F.; Wilk, A.; Armacost, C.; Day, T.; Godejohann, M.; Wachter, U.; Vogt, J.; Radermacher, P.; Mizaikoff, B. Breath Analysis with Broadly Tunable Quantum Cascade Lasers. Anal. Chem. 2013, 85, 2697–2702. [Google Scholar] [CrossRef] [PubMed]
  134. Tütüncü, E.; Kokoric, V.; Szedlak, R.; Macfarland, D.; Zederbauer, T.; Detz, H.; Andrews, A.M.; Schrenk, W.; Strasser, G.; Mizaikoff, B. Advanced gas sensors based on substrate-integrated hollow waveguides and dual-color ring quantum cascade lasers. Analyst 2016, 141, 6202–6207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Tütüncü, E.; Kokoric, V.; Wilk, A.; Seichter, F.; Schmid, M.; Hunt, W.E.; Manuel, A.M.; Mirkarimi, P.; Alameda, J.B.; Carter, J.C.; et al. Fiber-Coupled Substrate-Integrated Hollow Waveguides: An Innovative Approach to Mid-infrared Remote Gas Sensors. ACS Sens. 2017, 2, 1287–1293. [Google Scholar] [CrossRef]
  136. Kokoric, V.; Theisen, J.; Wilk, A.; Penisson, C.; Bernard, G.; Mizaikoff, B.; Gabriel, J.-C.P. Determining the Partial Pressure of Volatile Components via Substrate-Integrated Hollow Waveguide Infrared Spectroscopy with Integrated Microfluidics. Anal. Chem. 2018, 90, 4445–4451. [Google Scholar] [CrossRef]
  137. Van Well, B.; Murray, S.; Hodgkinson, J.; Pride, R.; Strzoda, R.; Gibson, G.; Padgett, M. An open-path, hand-held laser system for the detection of methane gas. J. Opt. A Pure Appl. Opt. 2005, 7, S420–S424. [Google Scholar] [CrossRef]
  138. Sun, J.; Ding, J.; Liu, N.; Yang, G.; Li, J. Detection of multiple chemicals based on external cavity quantum cascade laser spectroscopy. Spectrochim. Acta Part A Mol. Spectrosc. 2018, 191, 532–538. [Google Scholar] [CrossRef]
  139. Chen, X.; Cheng, L.; Guo, D.; Kostov, Y.; Choa, F.S. Quantum cascade laser based standoff photoacoustic chemical detection. Opt. Express 2011, 19, 20251–20257. [Google Scholar] [CrossRef] [PubMed]
  140. Puiu, A.; Fiorani, L.; Rosa, O.; Borelli, R.; Pistilli, M.; Palucci, A. Lidar/DIAL detection of acetone at 3.3 μm by a tunable OPO laser system. Laser Phys. 2014, 24, 085606. [Google Scholar] [CrossRef]
  141. Daghestani, N.S.; Brownsword, R.; Weidmann, D. Analysis and demonstration of atmospheric methane monitoring by mid-infrared open-path chirped laser dispersion spectroscopy. Opt. Express 2014, 22 (Suppl. 7), A1731–A1743. [Google Scholar] [CrossRef]
  142. Jaworski, P.; Stachowiak, D.; Nikodem, M. Standoff detection of gases using infrared laser spectroscopy. In Proceedings of the SPIE Photonics Europe, Brussels, Belgium, 3–7 April 2016; p. 6. [Google Scholar]
  143. Wang, Z.; Sanders, S.T. Toward single-ended absorption spectroscopy probes based on backscattering from rough surfaces: H2O vapor measurements near 1350 nm. Appl. Phys. B-Lasers Opt. 2015, 121, 187–192. [Google Scholar] [CrossRef]
  144. Andresen, B.F.; Lwin, M.; Fulop, G.F.; Corrigan, P.; Gross, B.; Norton, P.R.; Moshary, F.; Ahmed, S. Mid-infrared backscattering measurements of building materials using a quantum cascade laser. In Proceedings of the Conference on Infrared Technology and Applications XXXVI, Orlando, FL, USA, 5–9 April 2010; p. 766043. [Google Scholar]
  145. Lambert-Girard, S.; Allard, M.; Piché, M.; Babin, F. Broadband and tunable optical parametric generator for remote detection of gas molecules in the short and mid-infrared. Appl. Opt. 2015, 54, 2594–2605. [Google Scholar] [CrossRef] [PubMed]
  146. Brumfield, B.E.; Phillips, M.C. Standoff detection of turbulent chemical mixture plumes using a swept external cavity quantum cascade laser. Opt. Eng. 2017, 57. [Google Scholar] [CrossRef]
  147. Peng, W.Y.; Goldenstein, C.S.; Mitchell Spearrin, R.; Jeffries, J.B.; Hanson, R.K. Single-ended mid-infrared laser-absorption sensor for simultaneous in situ measurements of H2O, CO2, CO, and temperature in combustion flows. Appl. Opt. 2016, 55, 9347–9359. [Google Scholar] [CrossRef] [PubMed]
  148. Wallin, S.; Pettersson, A.; Östmark, H.; Hobro, A. Laser-based standoff detection of explosives: A critical review. Anal. Bioanal. Chem. 2009, 395, 259–274. [Google Scholar] [CrossRef]
  149. Gaudio, P. Laser Based Standoff Techniques: A Review on Old and New Perspective for Chemical Detection and Identification. In Cyber and Chemical, Biological, Radiological, Nuclear, Explosives Challenges: Threats and Counter Efforts; Martellini, M., Malizia, A., Eds.; Springer International Publishing: Cham, Germany, 2017; pp. 155–177. [Google Scholar] [CrossRef]
  150. Golston, L.M.; Tao, L.; Brosy, C.; Schäfer, K.; Wolf, B.; McSpiritt, J.; Buchholz, B.; Caulton, D.R.; Pan, D.; Zondlo, M.A.; et al. Lightweight mid-infrared methane sensor for unmanned aerial systems. Appl. Phys. B-Lasers Opt. 2017, 123. [Google Scholar] [CrossRef] [Green Version]
  151. Roşca, S.; Suomalainen, J.; Bartholomeus, H.; Herold, M. Comparing terrestrial laser scanning and unmanned aerial vehicle structure from motion to assess top of canopy structure in tropical forests. Interface Focus 2018, 8, 20170038. [Google Scholar] [CrossRef] [Green Version]
  152. Joanna, K. Basic Principles and Analytical Application of Derivative Spectrophotometry. In Macro to Nano Spectroscopy; Uddin, J., Ed.; IntechOpen: Rijeka, Croatia, 2012; pp. 253–268. [Google Scholar]
  153. Patel, K.N.; Patel, J.K.; Rajput, G.C.; Rajgor, N.B. Derivative spectrometry method for chemical analysis: A review. Der Pharm. Lett. 2010, 2, 139–150. [Google Scholar]
  154. Bosch Ojeda, C.; Sanchez Rojas, F. Recent applications in derivative ultraviolet/visible absorption spectrophotometry: 2009–2011: A review. Microchem. J. 2013, 106, 1–16. [Google Scholar] [CrossRef]
  155. Czarnecki, M.A. Resolution Enhancement in Second-Derivative Spectra. Appl. Spectrosc. 2015, 69, 67–74. [Google Scholar] [CrossRef] [PubMed]
  156. Xi-yang, L.; Nan, G.; Zhen-hui, D.; Jin-yi, L.; Chao, C.; Zong-hua, Z. Infrared Spectroscopy Quantitative Detection Method Based on Second Order Derivative Spectrum and Characteristic Absorption Window. Spectrosc. Spectr. Anal. 2017, 37, 1765–1770. [Google Scholar] [CrossRef]
  157. Elzanfaly, E.S.; Hassan, S.A.; Salem, M.Y.; El-Zeany, B.A. Continuous Wavelet Transform, a powerful alternative to Derivative Spectrophotometry in analysis of binary and ternary mixtures: A comparative study. Spectroc. Acta Part A-Mol. Biomol. Spectrosc. 2015, 151, 945–955. [Google Scholar] [CrossRef] [PubMed]
  158. Lee, W.; Choi, S.; Kim, T.K.; Kang, J.; Park, K.S.; Kim, Y.-H. Determination of Insulator-to-Semiconductor Transition in Sol-Gel Oxide Semiconductors Using Derivative Spectroscopy. Materials 2016, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  159. El-Sayed, A.-A.Y.; El-Salem, N.A. Recent Developments of Derivative Spectrophotometry and Their Analytical Applications. Anal. Sci. 2005, 21, 595–614. [Google Scholar] [CrossRef] [Green Version]
  160. Hamid, D.; Frederic, L.; Brian, W.P.; Fabien, C. Application of spectral derivative data in visible and near-infrared spectroscopy. Phys. Med. Biol. 2010, 55, 3381. [Google Scholar] [CrossRef]
  161. Deng, H.; Sun, J.; Li, P.; Liu, Y.; Yu, B.; Li, J. Sensitive detection of acetylene by second derivative spectra with tunable diode laser absorption spectroscopy. Opt. Appl. 2016, 46, 353–363. [Google Scholar] [CrossRef]
  162. Du, Z.; Yan, Y.; Li, J.; Zhang, S.; Yang, X.; Xiao, Y. In situ, multiparameter optical sensor for monitoring the selective catalytic reduction process of diesel engines. Sens. Actuator B-Chem. 2018, 267, 255–264. [Google Scholar] [CrossRef]
  163. Meng, Y.; Liu, T.; Liu, K.; Jiang, J.; Wang, R.; Wang, T.; Hu, H. A Modified Empirical Mode Decomposition Algorithm in TDLAS for Gas Detection. IEEE Photonics J. 2014, 6, 1–7. [Google Scholar] [CrossRef]
  164. Wang, S.; Du, Z.; Yuan, L.; Ma, Y.; Wang, X.; Han, R.; Meng, S. Measurement of Atmospheric Dimethyl Sulfide with a Distributed Feedback Interband Cascade Laser. Sensors 2018, 18, 3216. [Google Scholar] [CrossRef] [PubMed]
  165. Du, Z.; Wan, J.; Li, J.; Luo, G.; Gao, H.; Ma, Y. Detection of Atmospheric Methyl Mercaptan Using Wavelength Modulation Spectroscopy with Multicomponent Spectral Fitting. Sensors 2017, 17, 379. [Google Scholar] [CrossRef] [PubMed]
  166. Foltynowicz, A.; Masłowski, P.; Ban, T.; Adler, F.; Cossel, K.C.; Briles, T.C.; Ye, J. Optical frequency comb spectroscopy. Faraday Discuss. 2011, 150. [Google Scholar] [CrossRef]
  167. Thorpe, M.J.; Moll, K.D.; Jones, R.J.; Safdi, B.; Ye, J. Broadband Cavity Ringdown Spectroscopy for Sensitive and Rapid Molecular Detection. Science 2006, 311, 1595–1599. [Google Scholar] [CrossRef] [PubMed]
  168. Coddington, I.; Swann, W.C.; Newbury, N.R. Coherent Multiheterodyne Spectroscopy Using Stabilized Optical Frequency Combs. Phys. Rev. Lett. 2008, 100, 013902. [Google Scholar] [CrossRef] [PubMed]
  169. Lee, J.; Kim, Y.-J.; Lee, K.; Lee, S.; Kim, S.-W. Time-of-flight measurement with femtosecond light pulses. Nat. Photonics 2010, 4, 716. [Google Scholar] [CrossRef]
  170. Mandon, J.; Guelachvili, G.; Picqué, N. Fourier transform spectroscopy with a laser frequency comb. Nat. Photonics 2009, 3, 99. [Google Scholar] [CrossRef]
  171. Foltynowicz, A.; Masłowski, P.; Fleisher, A.J.; Bjork, B.J.; Ye, J. Cavity-enhanced optical frequency comb spectroscopy in the mid-infrared application to trace detection of hydrogen peroxide. Appl. Phys. B-Lasers Opt. 2012, 110, 163–175. [Google Scholar] [CrossRef] [Green Version]
  172. Khodabakhsh, A.; Ramaiah-Badarla, V.; Rutkowski, L.; Johansson, A.C.; Lee, K.F.; Jiang, J.; Mohr, C.; Fermann, M.E.; Foltynowicz, A. Fourier transform and Vernier spectroscopy using an optical frequency comb at 3–5.4 μm. Opt. Lett. 2016, 41, 2541–2544. [Google Scholar] [CrossRef] [PubMed]
  173. Bernhardt, B.; Ozawa, A.; Jacquet, P.; Jacquey, M.; Kobayashi, Y.; Udem, T.; Holzwarth, R.; Guelachvili, G.; Hänsch, T.W.; Picqué, N. Cavity-enhanced dual-comb spectroscopy. Nat. Photonics 2009, 4, 55. [Google Scholar] [CrossRef]
  174. Bernhardt, B.; Sorokin, E.; Jacquet, P.; Thon, R.; Becker, T.; Sorokina, I.T.; Picqué, N.; Hänsch, T.W. Mid-infrared dual-comb spectroscopy with 2.4 μm Cr2+:ZnSe femtosecond lasers. Appl. Phys. B-Lasers Opt. 2010, 100, 3–8. [Google Scholar] [CrossRef] [Green Version]
  175. Rieker, G.B.; Giorgetta, F.R.; Swann, W.C.; Kofler, J.; Zolot, A.M.; Sinclair, L.C.; Baumann, E.; Cromer, C.; Petron, G.; Sweeney, C.; et al. Frequency-comb-based remote sensing of greenhouse gases over kilometer air paths. Optica 2014, 1, 290–298. [Google Scholar] [CrossRef] [Green Version]
  176. Ideguchi, T.; Poisson, A.; Guelachvili, G.; Picqué, N.; Hänsch, T.W. Adaptive real-time dual-comb spectroscopy. Nat. Commun. 2014, 5, 3375. [Google Scholar] [CrossRef]
  177. Muraviev, A.V.; Smolski, V.O.; Loparo, Z.E.; Vodopyanov, K.L. Massively parallel sensing of trace molecules and their isotopologues with broadband subharmonic mid-infrared frequency combs. Nat. Photonics 2018, 12, 209–214. [Google Scholar] [CrossRef]
  178. Ycas, G.; Giorgetta, F.R.; Baumann, E.; Coddington, I.; Herman, D.; Diddams, S.A.; Newbury, N.R. High-coherence mid-infrared dual-comb spectroscopy spanning 2.6 to 5.2 μm. Nat. Photonics 2018, 12, 202–208. [Google Scholar] [CrossRef] [Green Version]
  179. Link, S.M.; Klenner, A.; Mangold, M.; Zaugg, C.A.; Golling, M.; Tilma, B.W.; Keller, U. Dual-comb modelocked laser. Opt. Express 2015, 23, 5521–5531. [Google Scholar] [CrossRef]
  180. Link, S.M.; Klenner, A.; Keller, U. Dual-comb modelocked lasers: Semiconductor saturable absorber mirror decouples noise stabilization. Opt. Express 2016, 24, 1889–1902. [Google Scholar] [CrossRef] [PubMed]
  181. Link, S.M.; Maas, D.J.H.C.; Waldburger, D.; Keller, U. Dual-comb spectroscopy of water vapor with a free-running semiconductor disk laser. Science 2017, 356, 1164. [Google Scholar] [CrossRef] [PubMed]
  182. Thorpe, M.J.; Ye, J. Cavity-enhanced direct frequency comb spectroscopy. Appl. Phys. B-Lasers Opt. 2008, 91, 397–414. [Google Scholar] [CrossRef] [Green Version]
  183. Thorpe, M.J.; Balslev-Clausen, D.; Kirchner, M.S.; Ye, J. Cavity-enhanced optical frequency comb spectroscopy: Application to human breath analysis. Opt. Express 2008, 16, 2387–2397. [Google Scholar] [CrossRef] [PubMed]
  184. Thorpe, M.J.; Adler, F.; Cossel, K.C.; de Miranda, M.H.G.; Ye, J. Tomography of a supersonically cooled molecular jet using cavity-enhanced direct frequency comb spectroscopy. Chem. Phys. Lett. 2009, 468, 1–8. [Google Scholar] [CrossRef]
  185. Spaun, B.; Changala, P.B.; Patterson, D.; Bjork, B.J.; Heckl, O.H.; Doyle, J.M.; Ye, J. Continuous probing of cold complex molecules with infrared frequency comb spectroscopy. Nature 2016, 533, 517. [Google Scholar] [CrossRef]
  186. Changala, P.B.; Spaun, B.; Patterson, D.; Doyle, J.M.; Ye, J. Sensitivity and resolution in frequency comb spectroscopy of buffer gas cooled polyatomic molecules. Appl. Phys. B-Lasers Opt. 2016, 122, 292. [Google Scholar] [CrossRef]
  187. Masłowski, P.; Cossel, K.C.; Foltynowicz, A.; Ye, J. Cavity-Enhanced Direct Frequency Comb Spectroscopy. In Cavity-Enhanced Spectroscopy and Sensing; Gagliardi, G., Loock, H.-P., Eds.; Springer: Berlin/Heidelberg, Germnay, 2014; pp. 271–321. [Google Scholar] [CrossRef]
  188. Xiao, S.; Weiner, A.M. 2-D wavelength demultiplexer with potential for ≥ 1000 channels in the C-band. Opt. Express 2004, 12, 2895–2902. [Google Scholar] [CrossRef]
  189. Diddams, S.A.; Hollberg, L.; Mbele, V. Molecular fingerprinting with the resolved modes of a femtosecond laser frequency comb. Nature 2007, 445, 627. [Google Scholar] [CrossRef] [PubMed]
  190. Gohle, C.; Stein, B.; Schliesser, A.; Udem, T.; Hänsch, T.W. Frequency Comb Vernier Spectroscopy for Broadband, High-Resolution, High-Sensitivity Absorption and Dispersion Spectra. Phys. Rev. Lett. 2007, 99, 263902. [Google Scholar] [CrossRef]
  191. Nugent-Glandorf, L.; Neely, T.; Adler, F.; Fleisher, A.J.; Cossel, K.C.; Bjork, B.; Dinneen, T.; Ye, J.; Diddams, S.A. Mid-infrared virtually imaged phased array spectrometer for rapid and broadband trace gas detection. Opt. Lett. 2012, 37, 3285–3287. [Google Scholar] [CrossRef]
  192. Fleisher, A.J.; Bjork, B.J.; Bui, T.Q.; Cossel, K.C.; Okumura, M.; Ye, J. Mid-Infrared Time-Resolved Frequency Comb Spectroscopy of Transient Free Radicals. J. Phys. Chem. Lett. 2014, 5, 2241–2246. [Google Scholar] [CrossRef] [Green Version]
  193. Nelson, D.D.; Shorter, J.H.; McManus, J.B.; Zahniser, M.S. Sub-part-per-billion detection of nitric oxide in air using a thermoelectrically cooled mid-infrared quantum cascade laser spectrometer. Appl. Phys. B-Lasers Opt. 2002, 75, 343–350. [Google Scholar] [CrossRef]
  194. Weber, W.H.; Remillard, J.T.; Chase, R.E.; Richert, J.F.; Capasso, F.; Gmachl, C.; Hutchinson, A.L.; Sivco, D.L.; Baillargeon, J.N.; Cho, A.Y. Using a Wavelength-Modulated Quantum Cascade Laser to Measure NO Concentrations in the Parts-per-Billion Range for Vehicle Emissions Certification. Appl. Spectrosc. 2002, 56, 706–714. [Google Scholar] [CrossRef]
  195. Martín-Mateos, P.; Jerez, B.; de Dios, C.; Acedo, P. Mid-infrared heterodyne phase-sensitive dispersion spectroscopy using difference frequency generation. Appl. Phys. B-Lasers Opt. 2018, 124. [Google Scholar] [CrossRef]
  196. Song, F.; Zheng, C.; Yan, W.; Ye, W.; Wang, Y.; Tittel, F.K. Interband cascade laser based mid-infrared methane sensor system using a novel electrical-domain self-adaptive direct laser absorption spectroscopy (SA-DLAS). Opt. Express 2017, 25, 31876–31888. [Google Scholar] [CrossRef] [PubMed]
  197. Zheng, C.; Ye, W.; Sanchez, N.P.; Li, C.; Dong, L.; Wang, Y.; Griffin, R.J.; Tittel, F.K. Development and field deployment of a mid-infrared methane sensor without pressure control using interband cascade laser absorption spectroscopy. Sens. Actuator B-Chem. 2017, 244, 365–372. [Google Scholar] [CrossRef]
  198. Dong, L.; Li, C.; Sanchez, N.P.; Gluszek, A.K.; Griffin, R.J.; Tittel, F.K. Compact CH4 sensor system based on a continuous-wave, low power consumption, room temperature interband cascade laser. Appl. Phys. Lett. 2016, 108. [Google Scholar] [CrossRef]
  199. Song, F.; Zheng, C.; Yu, D.; Zhou, Y.; Yan, W.; Ye, W.; Zhang, Y.; Wang, Y.; Tittel, F.K. Interband cascade laser-based ppbv-level mid-infrared methane detection using two digital lock-in amplifier schemes. Appl. Phys. B-Lasers Opt. 2018, 124. [Google Scholar] [CrossRef]
  200. Manfred, K.M.; Ritchie, G.A.D.; Lang, N.; Röpcke, J.; van Helden, J.H. Optical feedback cavity-enhanced absorption spectroscopy with a 3.24 μm interband cascade laser. Appl. Phys. Lett. 2015, 106. [Google Scholar] [CrossRef]
  201. Triki, M.; Nguyen Ba, T.; Vicet, A. Compact sensor for methane detection in the mid infrared region based on Quartz Enhanced Photoacoustic Spectroscopy. Infrared Phys. Technol. 2015, 69, 74–80. [Google Scholar] [CrossRef]
  202. Zhu, F.; Bicer, A.; Askar, R.; Bounds, J.; Kolomenskii, A.A.; Kelessides, V.; Amani, M.; Schuessler, H.A. Mid-infrared dual frequency comb spectroscopy based on fiber lasers for the detection of methane in ambient air. Laser Phys. Lett. 2015, 12. [Google Scholar] [CrossRef]
  203. Khodabakhsh, A.; Rutkowski, L.; Morville, J.; Johansson, A.C.; Soboń, G.; Foltynowicz, A. Continuous-Filtering Vernier Spectroscopy at 3.3 μm Using a Femtosecond Optical Parametric Oscillator. In Proceedings of the Conference on Lasers and Electro-Optics, San Jose, CA, USA, 14–19 May 2017; p. SW1L.5. [Google Scholar]
  204. Vainio, M.; Merimaa, M.; Halonen, L. Frequency-comb-referenced molecular spectroscopy in the mid-infrared region. Opt. Lett. 2011, 36, 4122–4124. [Google Scholar] [CrossRef]
  205. Silander, I.; Hausmaninger, T.; Ma, W.; Harren, F.J.; Axner, O. Doppler-broadened mid-infrared noise-immune cavity-enhanced optical heterodyne molecular spectrometry based on an optical parametric oscillator for trace gas detection. Opt. Lett. 2015, 40, 439–442. [Google Scholar] [CrossRef] [PubMed]
  206. Li, C.; Dong, L.; Zheng, C.; Lin, J.; Wang, Y.; Tittel, K.F. Ppbv-Level Ethane Detection Using Quartz-Enhanced Photoacoustic Spectroscopy with a Continuous-Wave, Room Temperature Interband Cascade Laser. Sensors 2018, 18, 723. [Google Scholar] [CrossRef] [PubMed]
  207. Razeghi, M.; Jahjah, M.; Lewicki, R.; Tittel, F.K.; Krzempek, K.; Stefanski, P.; So, S.; Thomazy, D. CW DFB RT diode laser based sensor for trace-gas detection of ethane using novel compact multipass gas absorption cell. In Proceedings of the Quantum Sensing and Nanophotonic Devices X, San Francisco, CA, USA, 3–7 February 2013. [Google Scholar]
  208. Krzempek, K.; Lewicki, R.; Nähle, L.; Fischer, M.; Koeth, J.; Belahsene, S.; Rouillard, Y.; Worschech, L.; Tittel, F.K. Continuous wave, distributed feedback diode laser based sensor for trace-gas detection of ethane. Appl. Phys. B-Lasers Opt. 2012, 106, 251–255. [Google Scholar] [CrossRef]
  209. Zhao, Z.-S.; Liao, Y.; Grattan, K.T.; Wang, W.; Jiang, D.; Zhang, W.; Wei, L.; Feng, Q.; Li, P.; Wang, Y.; et al. Research on propane leak detection system and device based on mid infrared laser. In Proceedings of the AOPC 2017: Fiber Optic Sensing and Optical Communications, Beijing, China, 4–6 June 2017. [Google Scholar]
  210. Adler, F.; Masłowski, P.; Foltynowicz, A.; Cossel, K.C.; Briles, T.C.; Hartl, I.; Ye, J. Mid-infrared Fourier transform spectroscopy with a broadband frequency comb. Opt. Express 2010, 18, 21861–21872. [Google Scholar] [CrossRef] [PubMed]
  211. Belyanin, A.A.; Smowton, P.M.; Lewicki, R.; Witinski, M.; Li, B.; Wysocki, G. Spectroscopic benzene detection using a broadband monolithic DFB-QCL array. In Proceedings of the Novel In-Plane Semiconductor Lasers XV, San Francisco, CA, USA, 15–18 February 2016. [Google Scholar]
  212. Lassen, M.; Harder, D.B.; Brusch, A.; Nielsen, O.S.; Heikens, D.; Persijn, S.; Petersen, J.C. Photo-acoustic sensor for detection of oil contamination in compressed air systems. Opt. Express 2017, 25, 1806–1814. [Google Scholar] [CrossRef] [PubMed]
  213. Lundqvist, S.; Kluczynski, P.; Weih, R.; von Edlinger, M.; Nähle, L.; Fischer, M.; Bauer, A.; Höfling, S.; Koeth, J. Sensing of formaldehyde using a distributed feedback interband cascade laser emitting around 3493nm. Appl. Opt. 2012, 51, 6009–6013. [Google Scholar] [CrossRef] [PubMed]
  214. Tanaka, K.; Miyamura, K.; Akishima, K.; Tonokura, K.; Konno, M. Sensitive measurements of trace gas of formaldehyde using a mid-infrared laser spectrometer with a compact multi-pass cell. Infrared Phys. Technol. 2016, 79, 1–5. [Google Scholar] [CrossRef]
  215. Dong, L.; Yu, Y.; Li, C.; So, S.; Tittel, F.K. Ppb-level formaldehyde detection using a CW room-temperature interband cascade laser and a miniature dense pattern multipass gas cell. Opt. Express 2015, 23, 19821–19830. [Google Scholar] [CrossRef] [Green Version]
  216. Ren, W.; Luo, L.; Tittel, F.K. Sensitive detection of formaldehyde using an interband cascade laser near 3.6 μm. Sens. Actuator B-Chem. 2015, 221, 1062–1068. [Google Scholar] [CrossRef]
  217. Tuzson, B.; Jagerska, J.; Looser, H.; Graf, M.; Felder, F.; Fill, M.; Tappy, L.; Emmenegger, L. Highly Selective Volatile Organic Compounds Breath Analysis Using a Broadly-Tunable Vertical-External-Cavity Surface-Emitting Laser. Anal. Chem. 2017, 89, 6377–6383. [Google Scholar] [CrossRef]
  218. Borri, S.; Patimisco, P.; Galli, I.; Mazzotti, D.; Giusfredi, G.; Akikusa, N.; Yamanishi, M.; Scamarcio, G.; De Natale, P.; Spagnolo, V. Intracavity quartz-enhanced photoacoustic sensor. Appl. Phys. Lett. 2014, 104, 091114. [Google Scholar] [CrossRef]
  219. Liu, X.; Zhang, G.; Huang, Y.; Wang, Y.; Qi, F. Two-dimensional temperature and carbon dioxide concentration profiles in atmospheric laminar diffusion flames measured by mid-infrared direct absorption spectroscopy at 4.2 μm. Appl. Phys. B-Lasers Opt. 2018, 124. [Google Scholar] [CrossRef]
  220. Nwaboh, J.A.; Qu, Z.; Werhahn, O.; Ebert, V. Interband cascade laser-based optical transfer standard for atmospheric carbon monoxide measurements. Appl. Opt. 2017, 56, E84–E93. [Google Scholar] [CrossRef] [PubMed]
  221. Dang, J.; Yu, H.; Zheng, C.; Wang, L.; Sui, Y.; Wang, Y. Development a low-cost carbon monoxide sensor using homemade CW-DFB QCL and board-level electronics. Opt. Laser Technol. 2018, 101, 57–67. [Google Scholar] [CrossRef]
  222. Lee, D.D.; Bendana, F.A.; Schumaker, S.A.; Spearrin, R.M. Wavelength modulation spectroscopy near 5 μm for carbon monoxide sensing in a high-pressure kerosene-fueled liquid rocket combustor. Appl. Phys. B-Lasers Opt. 2018, 124. [Google Scholar] [CrossRef]
  223. Diemel, O.; Pareja, J.; Dreizler, A.; Wagner, S. An interband cascade laser-based in situ absorption sensor for nitric oxide in combustion exhaust gases. Appl. Phys. B-Lasers Opt. 2017, 123. [Google Scholar] [CrossRef]
  224. Shi, C.; Wang, D.; Wang, Z.; Ma, L.; Wang, Q.; Xu, K.; Chen, S.-C.; Ren, W. A Mid-Infrared Fiber-Coupled QEPAS Nitric Oxide Sensor for Real-Time Engine Exhaust Monitoring. IEEE Sens. J. 2017, 17, 7418–7424. [Google Scholar] [CrossRef]
  225. Upadhyay, A.; Wilson, D.; Lengden, M.; Chakraborty, A.L.; Stewart, G.; Johnstone, W. Calibration-Free WMS Using a cw-DFB-QCL, a VCSEL, and an Edge-Emitting DFB Laser With In-Situ Real-Time Laser Parameter Characterization. IEEE Photonics J. 2017, 9, 1–17. [Google Scholar] [CrossRef] [Green Version]
  226. Almodovar, C.A.; Spearrin, R.M.; Hanson, R.K. Two-color laser absorption near 5 μm for temperature and nitric oxide sensing in high-temperature gases. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 572–581. [Google Scholar] [CrossRef]
  227. Wojtas, J.; Gluszek, A.; Hudzikowski, A.; Tittel, F.K. Mid-Infrared Trace Gas Sensor Technology Based on Intracavity Quartz-Enhanced Photoacoustic Spectroscopy. Sensors 2017, 17, 513. [Google Scholar] [CrossRef]
  228. Sur, R.; Peng, W.Y.; Strand, C.; Mitchell Spearrin, R.; Jeffries, J.B.; Hanson, R.K.; Bekal, A.; Halder, P.; Poonacha, S.P.; Vartak, S.; et al. Mid-infrared laser absorption spectroscopy of NO2 at elevated temperatures. J. Quant. Spectrosc. Radiat. Transf. 2017, 187, 364–374. [Google Scholar] [CrossRef]
  229. Sun, J.; Deng, H.; Liu, N.; Wang, H.; Yu, B.; Li, J. Mid-infrared gas absorption sensor based on a broadband external cavity quantum cascade laser. Rev. Sci. Instrum. 2016, 87, 123101. [Google Scholar] [CrossRef] [PubMed]
  230. Gambetta, A.; Cassinerio, M.; Coluccelli, N.; Fasci, E.; Castrillo, A.; Gianfrani, L.; Gatti, D.; Marangoni, M.; Laporta, P.; Galzerano, G. Direct phase-locking of a 8.6-mum quantum cascade laser to a mid-IR optical frequency comb: Application to precision spectroscopy of N2O. Opt. Lett. 2015, 40, 304–307. [Google Scholar] [CrossRef] [PubMed]
  231. Nikodem, M.; Wysocki, G. Chirped laser dispersion spectroscopy for remote open-path trace-gas sensing. Sensors 2012, 12, 16466–16481. [Google Scholar] [CrossRef] [PubMed]
  232. Razeghi, M.; Lewicki, R.; Sudharsanan, R.; Kosterev, A.A.; Thomazy, D.M.; Brown, G.J.; Risby, T.H.; Solga, S.; Schwartz, T.B.; Tittel, F.K. Real time ammonia detection in exhaled human breath using a distributed feedback quantum cascade laser based sensor. In Proceedings of the Quantum Sensing and Nanophotonic Devices VIII, San Francisco, CA, USA, 23–27 January 2011. [Google Scholar]
  233. Li, J.; Yang, S.; Wang, R.; Du, Z.; Wei, Y. Ammonia detection using hollow waveguide enhanced laser absorption spectroscopy based on a 9.56 μm quantum cascade laser. In Proceedings of the Applied Optics and Photonics China (AOPC2017), Beijing, China, 4–6 January 2017; p. 8. [Google Scholar]
  234. Plunkett, S.; Parrish, M.E.; Shafer, K.H.; Shorter, J.H.; Nelson, D.D.; Zahniser, M.S. Hydrazine detection limits in the cigarette smoke matrix using infrared tunable diode laser absorption spectroscopy. Spectroc. Acta Part A-Mol. Biomol. Spectrosc. 2002, 58, 2505–2517. [Google Scholar] [CrossRef]
  235. Taslakov, M.; Simeonov, V.; Froidevaux, M.; van den Bergh, H. Open-path ozone detection by quantum-cascade laser. Appl. Phys. B-Lasers Opt. 2005, 82, 501–506. [Google Scholar] [CrossRef]
  236. Ren, W.; Jiang, W.; Sanchez, N.P.; Patimisco, P.; Spagnolo, V.; Zah, C.-E.; Xie, F.; Hughes, L.C.; Griffin, R.J.; Tittel, F.K. Hydrogen peroxide detection with quartz-enhanced photoacoustic spectroscopy using a distributed-feedback quantum cascade laser. Appl. Phys. Lett. 2014, 104, 041117. [Google Scholar] [CrossRef] [Green Version]
  237. Siciliani de Cumis, M.; Viciani, S.; Borri, S.; Patimisco, P.; Sampaolo, A.; Scamarcio, G.; De Natale, P.; D’Amato, F.; Spagnolo, V. Widely-tunable mid-infrared fiber-coupled quartz-enhanced photoacoustic sensor for environmental monitoring. Opt. Express 2014, 22, 28222–28231. [Google Scholar] [CrossRef]
  238. Spagnolo, V.; Patimisco, P.; Borri, S.; Scamarcio, G.; Bernacki, B.E.; Kriesel, J. Part-per-trillion level SF6 detection using a quartz enhanced photoacoustic spectroscopy-based sensor with single-mode fiber-coupled quantum cascade laser excitation. Opt. Lett. 2012, 37, 4461–4463. [Google Scholar] [CrossRef]
  239. Spagnolo, V.; Patimisco, P.; Borri, S.; Scamarcio, G.; Bernacki, B.E.; Kriesel, J. Mid-infrared fiber-coupled QCL-QEPAS sensor. Appl. Phys. B-Lasers Opt. 2013, 112, 25–33. [Google Scholar] [CrossRef]
  240. Waclawek, J.P.; Moser, H.; Lendl, B. Compact quantum cascade laser based quartz-enhanced photoacoustic spectroscopy sensor system for detection of carbon disulfide. Opt. Express 2016, 24, 6559–6571. [Google Scholar] [CrossRef] [PubMed]
  241. Cao, X.; Li, J.; Gao, H.; Du, Z.; Ma, Y. Simultaneous Determination of Carbon Disulfide, Carbon Monoxide, and Dinitrogen Oxide by Differential Absorption Spectroscopy Using a Distributed Feedback Quantum Cascade Laser. Anal. Lett. 2017, 50, 2342–2350. [Google Scholar] [CrossRef]
  242. Wysocki, G.; McCurdy, M.; So, S.; Weidmann, D.; Roller, C.; Curl, R.F.; Tittel, F.K. Pulsed quantum-cascade laser-based sensor for trace-gas detection of carbonyl sulfide. Appl. Opt. 2004, 43, 6040–6046. [Google Scholar] [CrossRef] [PubMed]
  243. Du, Z.; Li, J.; Gao, H.; Luo, G.; Cao, X.; Ma, Y. Ultrahigh-resolution spectroscopy for methyl mercaptan at the ν 2 -band by a distributed feedback interband cascade laser. J. Quant. Spectrosc. Radiat. Transf. 2017, 196, 123–129. [Google Scholar] [CrossRef]
  244. Dong, L.; Tittel, F.K.; Li, C.; Sanchez, N.P.; Wu, H.; Zheng, C.; Yu, Y.; Sampaolo, A.; Griffin, R.J. Compact TDLAS based sensor design using interband cascade lasers for mid-IR trace gas sensing. Opt. Express 2016, 24, A528–A535. [Google Scholar] [CrossRef] [PubMed]
  245. Zheng, C.; Ye, W.; Sanchez, N.P.; Gluszek, A.K.; Hudzikowski, A.J.; Li, C.; Dong, L.; Griffin, R.J.; Tittel, F.K. Infrared Dual-Gas CH4/C2H6Sensor Using Two Continuous-Wave Interband Cascade Lasers. IEEE Photonics Technol. Lett. 2016, 28, 2351–2354. [Google Scholar] [CrossRef]
  246. Ye, W.; Li, C.; Zheng, C.; Sanchez, N.P.; Gluszek, A.K.; Hudzikowski, A.J.; Dong, L.; Griffin, R.J.; Tittel, F.K. Mid-infrared dual-gas sensor for simultaneous detection of methane and ethane using a single continuous-wave interband cascade laser. Opt. Express 2016, 24, 16973–16985. [Google Scholar] [CrossRef]
  247. Razeghi, M.; Ye, W.; Zheng, C.; Tittel, F.K.; Sanchez, N.P.; Gluszek, A.K.; Hudzikowski, A.J.; Lou, M.; Dong, L.; Griffin, R.J. A compact mid-infrared dual-gas CH4/C2H6 sensor using a single interband cascade laser and custom electronics. In Proceedings of the Quantum Sensing and Nano Electronics and Photonics XIV, San Francisco, CA, USA, 29 January–2 February 2017; p. 1011134. [Google Scholar]
  248. Krzempek, K.; Abramski, K.M.; Nikodem, M. All-fiber mid-infrared difference frequency generation source and its application to molecular dispersion spectroscopy. Laser Phys. Lett. 2017, 14. [Google Scholar] [CrossRef]
  249. Sampaolo, A.; Csutak, S.; Patimisco, P.; Giglio, M.; Menduni, G.; Passaro, V.; Tittel, F.K.; Deffenbaugh, M.; Spagnolo, V. Methane, ethane and propane detection using a compact quartz enhanced photoacoustic sensor and a single interband cascade laser. Sens. Actuator B-Chem. 2019, 282, 952–960. [Google Scholar] [CrossRef]
  250. Schiller, C.L.; Bozem, H.; Gurk, C.; Parchatka, U.; Königstedt, R.; Harris, G.W.; Lelieveld, J.; Fischer, H. Applications of quantum cascade lasers for sensitive trace gas measurements of CO, CH4, N2O and HCHO. Appl. Phys. B-Lasers Opt. 2008, 92, 419–430. [Google Scholar] [CrossRef]
  251. Haakestad, M.W.; Lamour, T.P.; Leindecker, N.; Marandi, A.; Vodopyanov, K.L. Intracavity trace molecular detection with a broadband mid-IR frequency comb source. J. Opt. Soc. Am. B 2013, 30, 631–640. [Google Scholar] [CrossRef]
  252. Dong, L.; Cao, Y.; Sanchez, N.P.; Griffin, R.J.; Tittel, F.K. Mid-infrared detection of atmospheric CH4, N2O and H2O based on a single continuous wave quantum cascade laser. In Proceedings of the CLEO 2015, San Jose, CA, USA, 10–15 May 2015; p. AF2J.3. [Google Scholar]
  253. Wei, C.; Pineda, D.I.; Paxton, L.; Egolfopoulos, F.N.; Spearrin, R.M. Mid-infrared laser absorption tomography for quantitative 2D thermochemistry measurements in premixed jet flames. Appl. Phys. B-Lasers Opt. 2018, 124. [Google Scholar] [CrossRef]
  254. Shorter, J.H.; Nelson, D.D.; Barry McManus, J.; Zahniser, M.S.; Milton, D.K. Multicomponent Breath Analysis With Infrared Absorption Using Room-Temperature Quantum Cascade Lasers. IEEE Sens. J. 2009, 10, 76–84. [Google Scholar] [CrossRef] [PubMed]
  255. Sun, K.; Tao, L.; Miller, D.J.; Khan, M.A.; Zondlo, M.A. Inline multi-harmonic calibration method for open-path atmospheric ammonia measurements. Appl. Phys. B-Lasers Opt. 2012, 110, 213–222. [Google Scholar] [CrossRef]
  256. Ma, Y.; Lewicki, R.; Razeghi, M.; Tittel, F.K. QEPAS based ppb-level detection of CO and N2O using a high power CW DFB-QCL. Opt. Express 2013, 21, 1008–1019. [Google Scholar] [CrossRef]
  257. Reidl-Leuthner, C.; Ofner, J.; Tomischko, W.; Lohninger, H.; Lendl, B. Simultaneous open-path determination of road side mono-nitrogen oxides employing mid-IR laser spectroscopy. Atmos. Environ. 2015, 112, 189–195. [Google Scholar] [CrossRef]
  258. Razeghi, M.; Yu, Y.; Sanchez, N.P.; Lou, M.; Zheng, C.; Wu, H.; Głuszek, A.K.; Hudzikowski, A.J.; Griffin, R.J.; Tittel, F.K. CW DFB-QCL- and EC-QCL-based sensor for simultaneous NO and NO2 measurements via frequency modulation multiplexing using multi-pass absorption spectroscopy. In Proceedings of the Quantum Sensing and Nano Electronics and Photonics XIV, San Francisco, CA, USA, 29 January–2 February 2017. [Google Scholar]
  259. Yu, Y.; Sanchez, N.P.; Yi, F.; Zheng, C.; Ye, W.; Wu, H.; Griffin, R.J.; Tittel, F.K. Dual quantum cascade laser-based sensor for simultaneous NO and NO2 detection using a wavelength modulation-division multiplexing technique. Appl. Phys. B-Lasers Opt. 2017, 123, 164. [Google Scholar] [CrossRef]
  260. Jagerska, J.; Jouy, P.; Tuzson, B.; Looser, H.; Mangold, M.; Soltic, P.; Hugi, A.; Bronnimann, R.; Faist, J.; Emmenegger, L. Simultaneous measurement of NO and NO2 by dual-wavelength quantum cascade laser spectroscopy. Opt. Express 2015, 23, 1512–1522. [Google Scholar] [CrossRef] [PubMed]
  261. Jágerská, J.; Jouy, P.; Hugi, A.; Tuzson, B.; Looser, H.; Mangold, M.; Beck, M.; Emmenegger, L.; Faist, J. Dual-wavelength quantum cascade laser for trace gas spectroscopy. Appl. Phys. Lett. 2014, 105. [Google Scholar] [CrossRef]
  262. Chen, X.; Yang, C.-G.; Hu, M.; Shen, J.-K.; Niu, E.-C.; Xu, Z.-Y.; Fan, X.-L.; Wei, M.; Yao, L.; He, Y.-B.; et al. Highly-sensitive NO, NO 2, and NH 3 measurements with an open-multipass cell based on mid-infrared wavelength modulation spectroscopy. Chin. Phys. B 2018, 27. [Google Scholar] [CrossRef]
  263. Chen, X.; Yang, C.; Hu, M.; Xu, Z.; Fan, X.; Wei, M.; Yao, L.; He, Y.; Kan, R. High sensitivity measurement of NO, NO2 and NH3 using MIR-QCL and time division multiplexing WMS technology. In Proceedings of the Hyperspectral Remote Sensing Applications and Environmental Monitoring and Safety Testing Technology, Beijing, China, 9–11 May 2016. [Google Scholar]
  264. Lee, B.H.; Wood, E.C.; Zahniser, M.S.; McManus, J.B.; Nelson, D.D.; Herndon, S.C.; Santoni, G.W.; Wofsy, S.C.; Munger, J.W. Simultaneous measurements of atmospheric HONO and NO2 via absorption spectroscopy using tunable mid-infrared continuous-wave quantum cascade lasers. Appl. Phys. B-Lasers Opt. 2010, 102, 417–423. [Google Scholar] [CrossRef]
  265. Rawlins, W.T.; Hensley, J.M.; Sonnenfroh, D.M.; Oakes, D.B.; Allen, M.G. Quantum cascade laser sensor for SO2 and SO3 for application to combustor exhaust streams. Appl. Opt. 2005, 44, 6635–6643. [Google Scholar] [CrossRef] [PubMed]
  266. Young, C.; Kim, S.S.; Luzinova, Y.; Weida, M.; Arnone, D.; Takeuchi, E.; Day, T.; Mizaikoff, B. External cavity widely tunable quantum cascade laser based hollow waveguide gas sensors for multianalyte detection. Sens. Actuator B-Chem. 2009, 140, 24–28. [Google Scholar] [CrossRef]
Figure 1. Overview of the recently emerging gas cell-based laser gas sensing principles. A gas cell with small volume and easy integration includes: Ag/AgI-HWG-Ag/AgI-coated hollow waveguides, PBG-HWG-photonic bandgaps hollow waveguides, iHWG-substrate-integrated hollow waveguides and CMRC-circular multi-reflection cell. The laser source includes: ICL—interband cascade laser, QCL—quantum cascade laser, EC-QCL—external cavity coupled QCL. The detector includes: Pyroel. —pyroelectric detector, DTGS—deuterated triglycine sulfate detector, Thermo. —thermopile detector, MCT—mercury cadmium telluride semiconductor detector, QCD—quantum cascade detector.
Figure 1. Overview of the recently emerging gas cell-based laser gas sensing principles. A gas cell with small volume and easy integration includes: Ag/AgI-HWG-Ag/AgI-coated hollow waveguides, PBG-HWG-photonic bandgaps hollow waveguides, iHWG-substrate-integrated hollow waveguides and CMRC-circular multi-reflection cell. The laser source includes: ICL—interband cascade laser, QCL—quantum cascade laser, EC-QCL—external cavity coupled QCL. The detector includes: Pyroel. —pyroelectric detector, DTGS—deuterated triglycine sulfate detector, Thermo. —thermopile detector, MCT—mercury cadmium telluride semiconductor detector, QCD—quantum cascade detector.
Applsci 09 00338 g001
Figure 2. Diagrams of the basic architectures of standoff gas detection techniques with non-cooperators: (a) TDLAS [35], (b) DIAL [149], (c) PAS [139], (d) ACLaS [36].
Figure 2. Diagrams of the basic architectures of standoff gas detection techniques with non-cooperators: (a) TDLAS [35], (b) DIAL [149], (c) PAS [139], (d) ACLaS [36].
Applsci 09 00338 g002
Figure 3. (a) broadband spectra of CS2 around 2176 cm−1 with 30.5 ppm × 5 m; (b) simulated WMS-2f signals of Figure 1a with modulation index 0.2 to 2.2, the insert panel describing the definition of spectrum discrimination (SD); (c) the SD and normalized WMS-2f signals amplitude under various modulation index [27].
Figure 3. (a) broadband spectra of CS2 around 2176 cm−1 with 30.5 ppm × 5 m; (b) simulated WMS-2f signals of Figure 1a with modulation index 0.2 to 2.2, the insert panel describing the definition of spectrum discrimination (SD); (c) the SD and normalized WMS-2f signals amplitude under various modulation index [27].
Applsci 09 00338 g003
Figure 4. Schematic of removing optical fringes and noise with the empirical mode decomposition (EMD) algorithm. (a) flow chart of the algorithm; (b) WMS-2f signal: recorded in the upper panel, the reconstructed and simulated in lower panel; (c) the intrinsic mode functions (IMF) decomposed from the recorded WMS-2f [27].
Figure 4. Schematic of removing optical fringes and noise with the empirical mode decomposition (EMD) algorithm. (a) flow chart of the algorithm; (b) WMS-2f signal: recorded in the upper panel, the reconstructed and simulated in lower panel; (c) the intrinsic mode functions (IMF) decomposed from the recorded WMS-2f [27].
Applsci 09 00338 g004
Figure 5. Census on mid-infrared lasers AND sensing (TOPIC) in the Web of Science.
Figure 5. Census on mid-infrared lasers AND sensing (TOPIC) in the Web of Science.
Applsci 09 00338 g005
Table 1. Comparison of different types of gas cells.
Table 1. Comparison of different types of gas cells.
TypeRef.OPL/mVolume/LPVR/m/LADsDADs
Modified MPCs2 mirrors[111]220.5540(c) (d) (e)②⑤
6 mirrors[112]3141.25251.2(b) (c) (e)①②⑤
3 mirrors[113]1.460.334.4(c) (d)①②⑤
2 mirrors[114]57.60.225256(a) (b) (d) (e)②⑤
2 mirrors[115]26.40.2894.3(b) (c) (e)②⑤
6 mirrors[116]32.40.4867.5(c) (d) (e)①②⑤
CMR cells1 mirrors[117]1.050.07813.5(c) (d) (e)③④
6 mirrors[118]3.10.024129.2(a) (c) (d) (e)③④
1 mirrors[119]2.160.0454(a) (c) (d) (e)③④
1 mirrors[120]4.080.04102(a) (c) (d) (e)③④
6 mirrors[121]0.690.01353.1(a) (c) (d) (e)③④
1 mirrors[122]12.24//(a) (c) (d) (e)③④
1 mirrors[123]9.90.05198(a) (c) (d) (e)
65 mirrors[124]100.1471.4(a) (c) (d) (e)③④
HWGsAg/AgI-HWG[32,69]50.0041250(a) (d) (e)
iHWG[126]0.250.001250(a) (d) (e)
iHWG[63]0.0750.0003250(a) (d) (e)
PBF-HWG[127]15 × 10−6200,000(a) (d) (e)④⑤
PBF-HWG[128]0.084.5 × 10−7177,778(a) (d) (e)④⑤
Advantages indexDisadvantage index
(a) easy configurations, fewer than three components
(b) ultra-long pathlength (>50 m)
(c) keeping focal properties, low-aberrations
(d) compact structure
(e) suitable for MIR
① multiple components (≥3 optical elements) and requiring adjustment
② large volume (>0.1 L)
③ hard to align, sensitive to vibrations or input conditions
④ spot diffusion introduced by aberrations
⑤ slow gas exchanging
Table 2. Summary of single component detection by MIR spectroscopy (2008–2018).
Table 2. Summary of single component detection by MIR spectroscopy (2008–2018).
SpeciesBandsWavelength/nmLaser TypeTechniquesLOD 1/ppbRefsApplications
CH4ν13392ICLWMS-2f[email protected] s[8]Exhaled breath analysis
3451.9DFGHPSDS 2250[195]Technique research
ν33291ICLSA-DAS
DAS
6.3@240 s;
[email protected] s
[196,197]Atmospheric
DAS1.4@60 s[198]
WMS13.07@2 s[199]
3300DAS15@60 s[73]
3240OF-CEAS 33@2 s[200]Technique research
3366DAS3.8 × 104[63]Industrial emission and process control
3260QW-DFB-DL 4PAS;WMS1.5 × 104@12 s[201]Environmental
3200DFGFCS60@80 ms[202]Atmospheric
3250DROPO 5FCS4@15 ms[203]
3390OPOFCS/[204]Technique research
3270OPONICE-OHMS 60.09@20 s[205]Ultrasensitive detection research
GaSb laserWMS13[150]Atmospheric
ν47791QCLCLaDS 760 ppb@100 s[141]
C2H6ν103330ICLWMS1.5@23 s[206]Technique research
3340WMS-2f;
WMS-2f/1f
DAS;
1.2@4 s;
1.0@4 s;
7.92@1 s
[65,66]Atmospheric
3360LDWMS0.13@1 s[207]Environmental
LDWMS-2f0.24@1 s[208]Technique research
C2H4ν13266ICLWMS-2f53@24 s[64]Industrial emission
C2H2ν4 +ν57263EC-QCLWMS-2f/1f3@110 s[53]Exhaled breath analysis
C3H8ν23370.4ICLWMS-2f460@1 s[209]Leakage monitoring
C5H8/3333.3OPOFCS7@30 s[210]Technique research
C6H6ν149640QCLDAS12@200 s[211]Atmospheric
C10H22/3380ICLPAS0.3[212]Industrial process
H2COν43493ICLDAS103[213]Workplace monitoring
3356WMS-2f73@40 s[214]Combustion emission
ν13599ICLDAS;
WMS
0.26@300 s;
0.069@90 s
[215]Technique research
WMS-2f1.5@140 s[216]Atmospheric
DF-RFM 825@1 s[67]Technique research
CH3OHν93390OPOFCS40@30 s[210]Technique research
C2H5OH/3367OPOFCS40@30 s[210]Technique research
CH3COCH3ν93380VECSEL 9DAS13@300 s[217]Breath VOCs detection
3389.8OPOFCS9.1@30 s[210]Technique research
ν58000ECQCLWMS-2f15@10 s[52]Spoilage monitoring of agricultural products
ν13300OPODIAL 101.2 × 105[140]Atmospheric
CO2ν34330QCLI-QEPAS 11300 ppt@4 s[218]Technique research
4200ICLDAS5 × 104[72]Combustion diagnose
4172/[219]
COν14691.2ICLWMS[email protected] s[71]Exhaled breath analysis
4600DAS500@14 s[220]Atmospheric
4764QCLWMS26@1 s[221]Indoor air
4980WMS/[222]Combustion diagnose
NOν15184ICLDAS3 × 104@10 ms[223]Combustion emission
5200QCLQEPAS 12120[224]Engine exhaust monitoring
5250WMS-2f/[225]Technique research
5030DAS/[226]Gas sensing in high temperature
5263I-QEPAS;
WMS-2f
4.8@30 ms[227]Environmental
NO2ν36250QCLWMS360 (600 K);
760 (800 K)
[228]Gas sensing in high temperature
ν1 + ν33250–3550OPOPAS14@170 s[3]Environmental pollution
N2Oν37782ECQCLDAS7.36 × 103[229]Toxic industrial chemical detection
ν28600QCL/DFGFCS0.3[230]Atmospheric
ν14530QCLCLaDS1.2 × 103[231]Atmospheric
NH3ν210,400ECQCLPAS1[49]Atmospheric
10,340QCLQEPAS;
2f-WMS
6@1 s[232]Exhaled breath analysis
9560DAS17.3@3 s[233]Atmospheric
9060WMS0.3[2]Atmospheric
ν12958.5OPOFCS25@30 s[210]Exhaled breath analysis
N2H4ν1210,363LDDAS400[234]Chemical analysis
O3ν39697QCLDIAS 13300[235]Atmospheric
H2Oν26700QCLOA-ICOS 14280[21]Chemical analysis
ν32666.7OPOFCS5.3@30 s[210]Technique research
H2O2ν67730QCLQEPAS12@100 s[236]Breath diagnosis
ν33760OPO FCS8[171]Exhaled breath analysis
H2Sν27900QCLQEPAS450@3 s
330@30 s
[237]Environmental pollution
SF6ν310,540QCLQEPAS0.05@1 s[238,239]Technique research
CS2ν1 + ν34590QCLQEPAS28@1 s[240]Industrial process
DOAS;
WMS
10.5;
60@240 s
[27,241]Atmospheric
OCSν34860QCLDAS[email protected] s[242]Exhaled breath analysis
CH3SHν23393ICLDAS[email protected] s[68]Industrial emission
3392WMS7.1@295 s[165,243]Atmospheric
CH3SCH3ν183370ICLWMS-2f/1f2.8@125 s[69]Environmental
3337ICLWMS9.6@164 s[164]Atmospheric
1 LOD: Limit of Detection; 2 HPSDS: Heterodyne phase sensitive dispersion spectroscopy; 3 OF-CEAS: Optical feedback cavity-enhanced absorption spectroscopy; 4 QW-DFB-DL: Quantum wells distributed feedback diode laser; 5 DROPO: Doubly resonant optical parametric oscillator; 6 NICE-OHMS: Noise-immune cavity-enhanced optical heterodyne molecular spectrometry; 7 CLaDS: Chirped laser dispersion spectroscopy; 8 DF-RFM: Dual-feedback RF modulation; 9 VECSEL: Vertical-external cavity surface-emitting laser; 10 DIAL: Differential absorption lidar; 11 I-QEPAS: Intracavity Quartz-Enhanced Photoacoustic Spectroscopy; 12 QEPAS: Quartz-enhanced photoacoustic spectroscopy; 13 DIAS: Differential absorption spectroscopy; 14 OA-ICOS: Off-axis integrated cavity output spectroscopy;
Table 3. Summary of information on multi-component simultaneous detection in the last 10 years.
Table 3. Summary of information on multi-component simultaneous detection in the last 10 years.
SpeciesBandsWavelength/nmLaser TypeTechniquesLOD/ppbRefsApplications
CH4/C2H6ν3/ν103291ICLDAS5@1 s[244]Atmospheric
33378@1 s
CH4/C2H6ν3/ν103291ICLDAS2.7@1 s[245]Atmospheric
3337WMS-2f[email protected] s
CH4/C2H6ν3/ν103337ICLWMS[email protected] s[246,247]Atmospheric
[email protected] s
CH4/C2H6ν3/ν103404DFGCLaDS360@1 s;
60@100 s
[248]Technique research
3335.5/
CH4/C2H6/C3H8ν3/ν10/ν13345ICLQEPAS90@1 s[249]Oil and gas industry monitoring
7@1 s
3 × 103@1 s
CH4/CO/H2COν4/ν1/ν27880/4633
5683
QCLWMS-2f0.5 for H2 CO@2 s[250]Atmospheric
CH4/N2Oν4/ν37700QCLDOAS3 × 104[35]Remote gas leakage detection
3.3 × 103
CH4/N2Oν4/ν37800QCLWMS-2f5.9@1 s[114]Atmospheric
2.6@1 s
CH4/H2CO/C2H4/C2H2/COν3/ν1/ν11/ν3/ν12500–5000OPOFCS1.7[251]Technique research
310
320
110
270
CH4/NOν3/ν13000–5400DROPOFCS20[172]Technique research
15
CH4/N2O/H2Oν4/ν3/ν27710QCLDAS23@1 s[252]Atmospheric
6.5@1 s
6.2 × 104@1 s
CH4/N2O/H2Oν4/ν3/ν28000ECQCLWMS-2f4.8@1 s[50]Atmospheric
0.9@1 s
3.1 × 104@1 s
CH3OH/C2H5OHν810,100ECQCLDAS130@1 s[146]Atmospheric VOCs
1.2 × 103@1 s
C2H5OH/(C2H5)2O/CH3COCH3/3800QCLCRDS157[18]Atmospheric VOCs
60
280
CO2/COν3/ν14730EC-QCLWMS-2f6.5 × 105[7]Exhaled breath analysis
9
CO2/COν3/ν14250QCLWMS106[51]Combustion diagnose
4860
CO2/COν3/ν14193ICLDAS/[253]Combustion diagnose
4979QCL
CO2/N2Oν3/ν14466QCLDAS2.7; 0.2[118]Technique research
WMS4.3@1 s;/
H2O/CO2/COν3/ν3/ν12551LDWMS-2f/1f1.4 × 107[147]Combustion diagnose
4176ICL6 × 106
4865QCL4 × 106
NO/CO/
N2O
ν15263QCLDAS0.5@1 s[254]Exhaled breath analysis
45660.8@1 s
CO/N2Oν14500QCLWMS0.36[255]Atmospheric
0.15
CO/N2Oν14610QCLQEPAS0.09@5 s[256]Atmospheric
0.05@5 s
NO/NO2ν1/ν35263QCLDAS597.3@1 s[257]Atmospheric
6135438.3@1 s
NO/NO2ν1/ν35263QCLFMS4@1 s[258]Technique research
61349@1 s
NO/NO2ν1/ν35263QCLWMDM-2f 10.75@100 s[259]Atmospheric
61340.9@200 s
NO/NO2ν1/ν35263QCLDAS1.5@100 s[260]Vehicle exhaust emission
62500.5@100 s
NO/NO2ν1/ν35250QCLDAS1.5@100 s[261]Environmental monitoring
62500.5@100 s
NO/NO2/NH3ν1/ν1/ν35263QCLWMS0.2@100 s;
0.96@1 s
[262,263]Industrial emission
62500.12@100 s; 0.94@1 s
90630.1@100 s;
0.86@1 s
NH3/C4H10ν2/ν248500QCLMHS 2/[48]Technique research
NO2/
HONO
ν1/6234QCLTILDAS 30.03[264]Atmospheric pollution
60240.3
SO2/SO3ν37500QCLDAS(1–2) × 103[265]Industrial emission
7160
C2H5Cl/CH2Cl2/CHCl3//ν47949EC-QCLDAS4[266]Atmospheric VOCs
7 × 103
11
1 WMDM-2f: Wavelength modulation-division multiplexing-2f; 2 MHS: Multiheterodyne spectroscopy; 3 TILDAS: Tunable infrared laser differential absorption spectroscopy.

Share and Cite

MDPI and ACS Style

Du, Z.; Zhang, S.; Li, J.; Gao, N.; Tong, K. Mid-Infrared Tunable Laser-Based Broadband Fingerprint Absorption Spectroscopy for Trace Gas Sensing: A Review. Appl. Sci. 2019, 9, 338. https://0-doi-org.brum.beds.ac.uk/10.3390/app9020338

AMA Style

Du Z, Zhang S, Li J, Gao N, Tong K. Mid-Infrared Tunable Laser-Based Broadband Fingerprint Absorption Spectroscopy for Trace Gas Sensing: A Review. Applied Sciences. 2019; 9(2):338. https://0-doi-org.brum.beds.ac.uk/10.3390/app9020338

Chicago/Turabian Style

Du, Zhenhui, Shuai Zhang, Jinyi Li, Nan Gao, and Kebin Tong. 2019. "Mid-Infrared Tunable Laser-Based Broadband Fingerprint Absorption Spectroscopy for Trace Gas Sensing: A Review" Applied Sciences 9, no. 2: 338. https://0-doi-org.brum.beds.ac.uk/10.3390/app9020338

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop