Next Article in Journal
The Indirect Role of Gluteus Medius Muscle in Knee Joint Stability during Unilateral Vertical Jump and Landing on Unstable Surface in Young Trained Males
Next Article in Special Issue
NH3 and N2O Real World Emissions Measurement from a CNG Heavy Duty Vehicle Using On-Board Measurement Systems
Previous Article in Journal
Thermal Performance Enhancement Using Absorber Tube with Inner Helical Axial Fins in a Parabolic Trough Solar Collector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fourier Transform Infrared (FTIR) Spectroscopy for Measurements of Vehicle Exhaust Emissions: A Review

by
Barouch Giechaskiel
* and
Michaël Clairotte
European Commission—Joint Research Centre (JRC), 21027 Ispra, Italy
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(16), 7416; https://doi.org/10.3390/app11167416
Submission received: 24 July 2021 / Revised: 9 August 2021 / Accepted: 10 August 2021 / Published: 12 August 2021
(This article belongs to the Special Issue The Effect of Vehicle Emissions on Secondary Aerosol and Air Quality)

Abstract

:

Featured Application

The review showed that sampling with FTIR from the tailpipe of vehicles for the determination of various gaseous pollutants is a possible alternative to currently regulated techniques.

Abstract

Pollution from vehicles is a serious concern for the environment and human health. Vehicle emission regulations worldwide have limits for pollutants such as hydrocarbons, CO, and NOx. The measurements are typically conducted at engine dynamometers (heavy-duty engines) sampling from the tailpipe or at chassis dynamometers (light-duty vehicles) sampling from the dilution tunnel. The latest regulations focused on the actual emissions of the vehicles on the road. Greenhouse gases (GHG) (such as CO2, CH4, N2O), and NH3 have also been the subject of some regulations. One instrument that can measure many gaseous compounds simultaneously is the Fourier transform infrared (FTIR) spectrometer. In this review the studies that assessed FTIRs since the 1980s are summarized. Studies with calibration gases or vehicle exhaust gas in comparison with well-established techniques were included. The main conclusion is that FTIRs, even when used at the tailpipe and not at the dilution tunnel, provide comparable results with other well-established techniques for CO2, CO, NOx, while for hydrocarbons, higher deviations were noticed. The introduction of FTIRs in the regulation needs a careful description of the technical requirements, especially interference tests. Although the limited results of prototype portable FTIRs for on-road measurement are promising, their performance at the wide range of environmental conditions (temperature, pressure, vibrations) needs further studies.

1. Introduction

Vehicle emissions are regulated since the 1970s [1]. The measurements are conducted on chassis dynamometers (light-duty vehicles) or in engine test cells (heavy-duty engines). The instruments described in the regulations are sampling from the full dilution tunnel, where the whole exhaust gas is diluted, or directly from the tailpipe (undiluted exhaust). The control of the regulated pollutants (e.g., CO, NOx) with advanced aftertreatment devices [2] has led in some cases to increased emission of non-regulated pollutants (e.g., N2O, NH3). The measurement techniques for regulated pollutants are well-defined in the regulation (e.g., non-dispersive infrared (NDIR) for CO and CO2). For non-regulated pollutants, only recently, a Global Technical Regulation for light-duty vehicles (GTR 15) prescribes possible measurement techniques. One method that can measure many compounds is FTIR (Fourier transform infrared) spectroscopy (= study of the interaction between light with matter) [3]. Many compounds absorb infrared energy at an intrinsic wave number (or wavelength) proportionally to their concentration. In an FTIR spectrometer, some of the infrared (IR) radiation is absorbed by the sample, and some of it is passed through (transmitted). The resulting molecular absorption and transmission response can be used to identify the components of the sample and their concentration. FTIR, compared to other IR techniques, can measure many components in real-time due to the use of an interferometer that allows the collection of a broad range of wavelengths. By contrast, NDIR analyzers measure one compound due to the use of an optical filter that allows the selection of a narrow wavelength area, specific to the compound of interest.
FTIR is used in geology, chemistry, materials, medicine, and biology research fields on solid, liquid, and gaseous samples [4]. FTIR has been used in a wide range of air pollution-related studies in both ambient air and environmental chambers [5,6]. Already in the late 1970s, an FTIR system was installed in a van for air pollution measurements for the Environmental Protection Agency (EPA) of the United States of America (USA) [7,8]. FTIR has also been applied for aerosol analysis [9,10], for example, to assess the hygroscopicity of ambient particles [11] or analyze surface functional groups of particles [12]. Another application was the measurement of stack emissions from various processes, such as incinerators [13]. Other researchers have used open path FTIR to record CO, CO2, and N2O [14] or NH3 [15] emissions from high traffic roadside sites or even emissions in aircraft exhaust [16]. Extractive in cell FTIR has been used in many applications, e.g., trains [17]. Extractive measurement with an activated charcoal tube was used to measure volatile organic compounds (VOC) of a heavy-duty engine [18]. Applications in atmospheric and environmental studies were reviewed elsewhere [19,20].
The application of FTIR spectroscopy to vehicles’ exhaust analysis began in the early 1980s [21,22,23]. It received a lot of attention for the study of alternative fuels due to its ability to discriminate oxidized species in tailpipe gases [24,25,26,27]. Other studies focused on TWC (three-way catalysts) [28,29,30] and low ambient temperatures [27,29]. Later FTIR was used to measure emissions from NOx reduction systems [31] due to its ability to simultaneously measure separately the oxides of nitrogen [27,32,33].
However, the technique did not spread out in the industry due to several practical problems, such as complexity of calibration (due to cross sensitivity of, e.g., CO2 and H2O), slow response, and poor concentration accuracy when compared to the regulated analytical techniques [31,34]. The evolution of computers and algorithms made it possible to have an instrument that can measure and analyze the data in real time. Since then, and especially in the last decade, the use of FTIR spectrometers has widespread.
It has been used for the measurement of gas concentrations for various studies, e.g., soot oxidation [35], or SCR (selective catalytic reduction for NOx) [36,37] and catalyst evaluation [38,39,40] with synthetic gases. It has also been used in engine test beds to assess ethanol [41,42], biodiesel [43] such as Jatropha [44], dimethyl ether (DME) [45], or hydro-treated vegetable oil (HVO) [46], homogeneous charge compression ignition (HCCI) engines [47], gasoline compression ignition engine [48], post injection effect on emissions [49], NH3 sensors [50], or even modeling of emissions [51]. FTIR instruments have also been used on chassis dynamometers: Small gasoline engines [52] or even diesel trucks [53,54]. For example, for exhaust gas recirculation (EGR) [55], alternative fuels [56,57], reactive nitrogen compounds [58,59], impact of low temperature on non-regulated pollutants [60], and retrofit evaluation [61,62] of diesel vehicles. Similarly, chassis dynamometer studies with gasoline vehicles [63] focused on unregulated emissions [64,65,66,67,68], NH3 [69,70,71,72,73], effect of exhaust gas reforming on emissions [74], low temperature [60,75,76], alternative fuels [65,77], and hybrids [78,79,80]. Motorcycles’ non-regulated pollutants emissions have also been assessed with FTIR [81,82,83,84,85].
The on-road application started in 2000 [86]. Since then other researchers measured emissions on the road [87,88,89], greenhouse gases (GHG) [90], nitrogen species [91], cold start emissions [92,93,94,95,96] of gasoline vehicles and the impact of ambient temperature [97]. A few also studied compressed natural gas (CNG) [98], diesel fueled vehicles [99,100,101] and their non-regulated pollutants [102].
Currently in the European Union (EU), FTIR is allowed for tailpipe (undiluted) NH3 measurements for heavy duty engines (Commission Regulation (EU) 582/2011). The same applies globally with UNECE (United Nations Economic Commission for Europe) Regulation 49. It is also prescribed in the UNECE light-duty vehicles GTR 15 (Global Technical Regulation) for ethanol, formaldehyde, acetaldehyde and N2O from the dilution tunnel. The future Euro 7 regulation on light-duty and heavy-duty vehicles moves in a direction that these pollutants should be tested on the road [103]. Thus, using the FTIR for simultaneous analysis of various pollutants sampling from the tailpipe (undiluted exhaust) is an attractive option [104,105] that needs to be assessed.
The reviews on the topic are limited and 20 years old [31]. Furthermore, assessing FTIRs for regulatory purposes has not been discussed before. The objective of this paper is to review studies that have evaluated FTIR systems for automotive exhaust gas applications. Special emphasis is given to portable applications and low emission levels to assess the suitability for future regulations.

2. Materials and Methods

2.1. Regulatory Background

In the legislation of the EU, light-duty vehicles include passenger cars and light commercial vehicles, while heavy-duty vehicles include trucks and buses.
Originally emission regulations for light-duty vehicles were introduced in the 1970s with Directive 70/220/EEC (only CO and HC) [1,106]. In the years after, other pollutants were added, and various reductions of the emission limits were applied. In 2007 regulation 715/2007 defined Euro 5 and Euro 6 standards. The pollutants currently regulated are CO, HC (for positive ignition (PI) engines only), NOx, HC + NOx (only for compression ignition (CI) engines), particulate matter (PM), and particle number (PN) (only for CI and PI direct injection engines). CO2 has limits only as fleet-average. The measurements were conducted mostly from bags that collect a sample from a dilution tunnel with constant volume sampling (CVS), where the whole exhaust gas was diluted.
In 2017, with Euro 6d-temp, additional on-road testing was added to the laboratory type approval and in-service conformity testing. Limits were defined only for PN and NOx, while CO and CO2 had to be measured. The limits were the laboratory type approval limits taking into account the additional measurement uncertainty of the on-board portable emissions measurement systems (PEMS) with temporary conformity factors [107,108]. Since 2020, with Euro 6d the revised conformity factors were applicable. Regulation (EU) 2018/858 introduced a new EU type-approval framework (from September 2020), with an effective market surveillance system to control the conformity of vehicles already in circulation (new In-Service Conformity process from September 2019).
Heavy-duty standards were originally introduced with Directive 88/77/EEC (applicable from 1992). Current Euro VI emission standards were introduced by Regulation 595/2009 followed by a number of amendments that specified technical details. Limits are applicable for CO, non-methane hydrocarbons (NMHC), CH4 (for gas engines) NOx, PM, PN, and NH3. The measurements can be conducted from bags, directly from the full dilution tunnel, from a proportional partial flow system (PDFS), or directly from the tailpipe. Euro VI regulation introduced in-use testing with field measurement using PEMS. Conformity factors are applicable that take into account the additional measurement uncertainty of the PEMS.
At a global level, GTR 15 (global technical regulation) for light-duty vehicles includes additional pollutants (only methodology, not limits), and the EU is highly likely to adopt most of them in future regulations. None of these pollutants, however, are prescribed as candidates for on-board testing. Table 1 summarizes the pollutants, their principle of measurement and their inclusion or not in the current EU regulations. Some of them were recently introduced in other countries (e.g., N2O in China 6, aldehydes in Brazil, Korea, USA) [109].
Based on Table 1, FTIR is permitted to be used for the measurement of NH3, N2O, C2H5OH, CH2O, and CH3CHO. However, with the current regulations, FTIR would have to be connected to the tailpipe for the measurement of NH3, but at the dilution tunnel for the rest of the pollutants. As FTIR can determine simultaneously many of the regulated pollutants (CO2, CO, NO, NO2, CH4), and has the capabilities to be used at the tailpipe (e.g., for NH3), it is an attractive candidate for on-road testing.

2.2. FTIR Description

FTIR can be used to measure substances (solids, liquids, gases, powders, polymers, organics, etc.) that absorb in the mid-infrared (i.e., approximately 400–4000 cm−1). As a basic principle of the interaction matter–radiation, absorption can occur at intrinsic energy levels, specific for each molecule. When mid-infrared radiations cross a molecule, some specific frequencies (or wavelengths) will be absorbed if they correspond to the transition of vibrational levels of the molecule. Consequently, FTIR can be used for qualitative and quantitative analysis. However, it cannot detect noble gases, such as helium (He) and argon (Ar), or diatomic gases, such as oxygen (O2), nitrogen (N2), and hydrogen (H2), as their vibration does not create a dipole moment, thus do not have absorbance bands in the infrared region of the electromagnetic spectrum. Others molecules absorb very little radiation and are, therefore, not detectable at low concentrations (e.g., H2S > 200 ppm [110]).
The heart of every FTIR instrument is an optical device called an interferometer (Figure 1) [3]. The oldest and most common type is the Michelson interferometer. The infrared source is usually a heated ceramic (at ca. 1200 °C). A collimating mirror collects light from the source and makes its rays parallel. A beamsplitter (in KBr) transmits approximately half of the light incident upon it and reflects the remaining half. A fraction of the light transmitted travels to a fixed mirror, while the other fraction travels to a moving mirror (see Figure 1). The lights are reflected by the two mirrors back to the beamsplitter, where they are recombined into a single light beam. This light beam interacts with the sample (exhaust gas) in a gas cell and finally strikes the detector. A multireflection cell is used to obtain a long optical path length with the minimum possible volume of the cell [111].
Because the path that one beam travels is a fixed length and the other is constantly changing as its mirror moves, the signal which exits the interferometer is the result of these two beams “interfering” with each other. The resulting signal is called an interferogram (i.e., a plot of light intensity versus optical path difference). The interferograms measured are then Fourier transformed to yield a spectrum (i.e., a plot intensity versus frequency/wavenumber).
There is also a laser (not shown in the figure) whose light follows the infrared beam. This laser light is used to measure the optical path difference of the interferometer. The spectral resolution (in cm−1) depends on the inverse of the optical path difference. This gives the wavelength accuracy (or Connes’ advantage) compared to dispersive instruments where the scale depends on the mechanical movement of diffraction gratings. Thus, the FTIR can give very accurate frequencies in the spectrum—this enables processing techniques such as spectral subtraction. It was shown that for automotive applications, a resolution of 0.5 cm−1 is the best compromise to obtain suitable fineness of the spectra (required to build robust calibration method) without compromising signal to noise ratio [8]. At lower resolutions (e.g., 1 cm−1), water’s absorbance bands may create interference problems that affect the detection limits of many compounds [112]. The accuracy and detection limits of the individual compounds measured are dependent on the intensity of the respective absorbance bands and their interference, but also on the detector [113,114]. For most compounds, the detection limit is well below 1 ppm for 1 s resolution [55,115]. The resolution of 0.5 cm−1 is achievable only with a photonic detector (e.g., Mercury Cadmium Telluride—MCT) that needs to be cooled down with liquid N2. Peltier cooled MCT or heating detector (e.g., Deuterated TriGlycide Sulfate—DTG)) are less sensitive and reactive compared to photonic detectors. However, they might be more suitable for portable systems. Another advantage of the FTIR is the multiplex advantage (or Fellgett advantage) because all wavelengths are collected simultaneously and thus a spectrum can be obtained very quickly. In contrast, with “dispersive spectroscopy” a monochromatic light beam shines at a sample, the absorbed light is measured, and this is repeated for each different wavelength. The throughput advantage (or the Jacquinot advantage) arises because, unlike dispersive spectrometers, FTIR spectrometers have no slits, which attenuate the infrared light, resulting in a higher signal-to-noise ratio.
Other advantages of the FTIR systems are: measurement of many pollutants simultaneously, real-time operation, possibility to measure undiluted exhaust gas, the fast response time (at least as fast as the conventional analyzers) for transient engine operation, low minimum detection limits, large dynamic range for dilute and direct sampling, high accuracy with negligible cross-sensitivity [31]. On the other hand, the interference primarily from H2O and CO2 can be high, and the signal-to-noise ratio can be affected by external vibrations. Figure 2a plots absorbance of CO and different nitrogen compounds (lower part), and CO2 and H2O (upper part) in arbitrary scales for better visualization. The example is based on the reference library acquired with a HR-FTIR MKS Multigas analyzer 2030 (Andover, MA, USA) at 191 °C, 0.95 atm and with a 5.11 m optical path measurement cell. Such reference spectra can be found from instrument manufacturers or from the Hitran (Cambridge, MA, USA) database [116,117]. It is evident that there are areas with overlap (e.g., NO2, NH3 and H2O). CO2 measurement is relatively straightforward, but the selection of the bands needs to be conducted considering H2O interference, and between CO absorbance peaks. CO requires the careful selection of infrared absorbance bands to stay within the desired absorbance range and in an area between H2O and potential N2O absorption region. NO and NO2 absorb only in regions with very high levels of water interference, as well as CO and CO2 interferences. Thus, for these compounds, the selection of the absorption bands needs to be conducted between, e.g., H2O absorption peaks, in the area between ca. 1850 and 1940 cm−1 for NO, and between 1570 and 1650 cm−1 for NO2. Figure 2b provides the example of NH3 calibration method, for which a rather straightforward selection of the absorption bands is possible within an area where no interference is expected (between 900 and 970 cm−1). Figure 2c provides the example of the N2O, for which the absorption bands need to be selected in an area located outside the CO2 absorption range, and between the CO absorption peaks. More detailed discussion on spectral regions for the analysis and interferences can be found in the literature [28,47,104,118,119]. Estimations of cross-sensitivities simulating exhaust gas gave an influence of <<1% for most compounds (NO2, CO, NH3) [114], but >1% for acetaldehyde (CH3CHO), acetone (CH3COCH3), C6H6, and C7H8, because their absorption spectrums are not sharp and strong, and/or the region of possible quantification is small due to the co-existing components. The actual interference might be different if there are interfering compounds not included in the software, if some wavelengths are saturated due to high concentrations, and/or the resolution is not enough to resolve the species, e.g., water, NO, NO2, and NH3 [120].
Principally, transmission-FTIR spectroscopy follows the Lambert–Beer law. Consequently, for a given optical path length, a linear relationship between the compound concentration and the absorbance mediated by this compound can be assumed. However, the output voltage of the detector is not a linear function of the incident radiation power [121] and non-linearity can be seen for components such as CO2, H2O, CO, and NOx with a wide range of variations [8]. Thus, for some components, the calibration might not be a linear function, but quadratic, cubic, etc. Because the mirror travels only a finite distance, the interferogram is truncated, and the Fourier transform results in a spectrum with broader features and spurious oscillations at the wings of the features. An apodization function is applied to reduce the magnitude of these oscillations but further broadens the spectral features. This broadening causes the instrument to have a nonlinear response to changes in absorption of the various gases being measured [122]. This can be modeled and corrected.
FTIR requires no daily calibration per se. This is possible because the daily background/zero scan is compared point per point to the measurement spectrum and compensates for any instrument drift in the final absorbance spectrum. Since the same number of molecules always absorb the same fraction of incident energy (independent of the total amount of energy), the calibration factor remains the same for a given compound and wavenumber and as a result there is no span drift or calibration drift [115]. However, since background and measurement spectrums are acquired successively, the FTIR needs to be very stable, in terms of infrared source, detectors, but also in terms of composition of the gas included in the instrument. Such stability is insured by keeping the instrument always on (heated), and by purging the optical system with dry gaseous N2.
The sampling lines and the sampling cell for exhaust gas application need to be heated above 100 °C (usually 191 °C) to avoid water condensation, which would lead to an underestimation of the hygroscopic compounds (e.g., NH3). In addition, the sampling line located upstream is usually equipped with a heated filter that prevents the deposition of particles on the mirrors’ surface of the multipath gas cell, and thus, the modification of the optical path length.

2.3. Processing of the Interferogram

The interferogram produced by the interferometer is converted mathematically (Fourier transform) into an intensity versus wavenumber plot known as a single-beam spectrum (Figure 3). The single-beam spectrum of the gas exhaust sample is ratioed (logarithmically) against the single-beam background spectrum (produced by passing nitrogen gas through the cell) to produce the absorbance spectrum [123]. Other processing of raw data (apodization) may be used to smoothen discontinuities at the beginning and end of the spectra, reduce, or eliminate noise [3,124].
Then, to utilize the complete information of the complex spectra and to handle the large data set, multivariate analysis is often used (i.e., data analytical methods that deal with more than one variable at a time). Some combinations of variables (wavenumbers) of a given data set are highly correlated with each other. Principal component analysis (PCA) is one such widely used dimensionality reduction technique to extract the informative region of the spectra for every individual species. Unlike classification or clustering, regression is used in the quantification of particular dependent variables (expected pollutants). Some of the commonly used multivariant regression methods are classical least squares (CLS), multiple linear regression (MLR), principal component regression (PCR), partial least squares (PLS) [125]. Such calibration methods can be appropriate when the gas composition of the mixture to be analyzed (here, the exhaust gas) is not known. In that case, the compound of interest can be added to the mixture (matrix) at different concentration levels (spiking) in order to build a suitable multivariate regression model. This model might, however, highly depends on the complexity and composition of the gas mixture (matrix) used to build it.
Another approach is to compare the spectra of the compound of interest to the foreseen interfering compounds (e.g., water or CO2) in order to: (i) identify the wavelength region where not too many interferences are expected, and (ii) identify the other wavelength regions where interferences are expected. The first region is then used to build the regression model, sometime in a straightforward way (see Figure 2b), or less straightforward way (see Figure 2c). The second set of wavelength regions will be used to foresee the possible interference brought by the compound of interest when building a regression model for other compounds. For illustration, Figure 2c shows how the N2O model is built in a wavelength area where the possible interference of CO absorption was identified. In this case, the region used to build the regression model was selected to avoid the interference of CO2 and H2O (see Figure 2a). Typically, specific wavelengths where no other species absorb are used to predict the volume concentration of the compound, assuming a linear or quadratic relationship between concentration and absorption [28,47,118]. For each compound, standard gas cylinders of several concentration levels are used to calibrate the model. Such an approach has the advantage of being more robust and less dependent on the composition of the mixture. However, this approach requires to be exhaustive in consideration of the expected interferences, thus having a library including the pure spectra of all the compounds expected to be found in vehicle exhaust.
Once the regression model is built, analysis of the spectral residual is also a crucial step in order to identify potential interfering compounds in the models created. It is important to highlight that the reference spectra must be recorded at the same conditions, meaning the same instrument (same optical system, including optical path), and same temperature and pressure as used for the exhaust measurements [126,127]. The processing of the data is not standardized, but FTIR data can be re-processed differently with a different method to reveal the concentration data of other components. It is important to mention that FTIR is a non-destructive measurement technique. Thus, it could be possible to direct the analyzed gaseous sample toward another measuring instrument for complementary analysis. It should be, however, recalled that the sampling gas is heated (e.g., 191 °C) and filtered.

3. Results

The following sections summarize the results of the studies that assessed FTIRs against calibration gases or reference instruments. The sections present shortly the principle of operation of the reference instruments (details can be found elsewhere, e.g., [128,129,130]) and any advantages and disadvantages over FTIR. Then the results of the studies that have assessed FTIRs are summarized. The details of the studies can be found in the Appendix A. Each table gives first the results with calibration gases. Then, the comparison of the FTIR and reference analyzers is divided into cases that both instruments were at the same location (dilution tunnel or tailpipe), or different (FTIR at tailpipe, reference at dilution tunnel). In the last case, the differences can be affected significantly by the uncertainty of the exhaust flow determination.
To put the assessed values into perspective, Table 2 presents Euro 6/VI limits for light-duty vehicles and heavy-duty engines, an approximation of the mass emissions of 1 ppm concentration of various pollutants. Exhaust flows of 2 kg/km and 10 kg/kWh were assumed as the highest values for large engine displacement light-duty vehicles and heavy-duty engines [103]. The density ratios u of the relevant regulations were used.

3.1. Carbon Monoxide (CO) and Carbon Dioxide (CO2)

CO and CO2 are typically measured with NDIR (Non-Dispersive Infrared) absorption type detectors. NDIR is one of the IR methods, which is not spectroscopic analysis but uses non-dispersive infrared light. In NDIR, the region of wavelengths used for analyzing the target component is selected by using an optical filter (multilayer interference filter) or a gas filter (a cell enclosing interfering gases). Other gases that show absorption in the same wavelength region will contribute to the measurement result of the target component. For example, when analyzing CO in the engine exhaust gas, CO2 and H2O also included in engine exhaust gas are likely to interfere with the measured CO concentration [131]. The CO NDIR analyzer may require a sample conditioning column containing CaSO4, or indicating silica gel to remove water vapor, and containing ascarite to remove carbon dioxide from the CO analysis stream (USA EPA Title 40, Chapter I, Subchapter C, Part 86, Subpart B, §86.111-94). Many analyzers use coolers. The regulation requires interference checks of the CO analyzer with CO2 and H2O at concentrations expected during the tests. The CO response has to be within 2% or ±50 ppm (whichever is larger). For analyzers measuring from the dilution tunnel, the allowed interference is between −1 ppm to 3 ppm. Regarding CO2 measurements of diluted exhaust, the regulation requires that the CO2 concentration in the dilute exhaust sample bag is <3% for gasoline and diesel engines, <2.2% for LPG engines, and <1.5% for natural gas and biomethane engines. When the concertation of water is <3% H2O the error is <1% [132].
For tailpipe sampling, in order to minimize the interference for non-diluted exhaust, additional detectors sensitive only to the interfering component can be added [133], or the H2O can be removed (e.g., by a cooler) [134]. Errors after corrections with additional H2O detector, the main interfering component in the exhaust gas, are <2% for H2O concentration up to 12% [133]. Removal of the water needs a dry-to-wet correction, which has a low error, as long as no condensation takes place at the sampling lines, e.g., at cold start. In such cases errors of up to 10% have been reported during the cold start period [135].
Table A1 summarizes the studies assessing FTIR CO2 readings with calibration cylinders or NDIR analyzers. Table A2 summarizes for CO. Regarding CO2, in most cases, the differences were within ±5% (slope 0.95–1.05), with only a few exceptions that the differences were around 10%. There were no particular differences between gasoline or diesel vehicles. The correlation coefficient was high (>0.95). The only exception (R2 = 0.70, FTIR at tailpipe vs. dilution tunnel) was when a calculated from the intake air exhaust flow was used instead of measured with exhaust flow meter. Higher differences (−21%) were reported when the response of the FTIR was slow [34].
Regarding CO, the mean differences were within ± 5% of the reference analyzers, but the scatter was very high in some cases, exceeding 50%. The slopes were within 0.8 and 1.2 with no indications of higher slopes when the FTIR connected at the tailpipe was compared to the dilution tunnel or bags. The worse performance of the FTIR CO compared to the CO2 might have to do with the higher interference of water and CO2, and the higher concentration range of CO compared to the CO2.

3.2. Nitrogen Monoxide (NO) and Nitrogen Dioxide (NO2)

NO and NOx (sum of NO and NO2) are typically measured with chemiluminescence analyzers (CLA) developed in early 1970s [120]. NO reacts with O3 and the excited NO2 molecules spontaneously return to normal state emitting light (chemiluminescence) [136]. The sensitivity for NO is influenced by the existence of CO2 or H2O in the sample gas, because the excited NO2 can react with CO2 and H2O (quenching effect) [128]. The effect is <1%, in particular for systems with H2O sensors that can compensate for H2O concentrations up to 14% [137]. The light can also be filtered before being measured by the photomultiplier to minimize interference [130].
The CLA can be used to measure NOx. In this case, a carbon or active metal-based furnace converts NO2 into NO (NOx converter) [120,136]. Regulations require a conversion efficiency of >95%. However, NH3 can also be converted to NO in the converter. The conversion efficiency depends on the converter material, the temperature and the concentrations of NO and NH3. With a 5:1 ratio of NH3 to NO, an error of 30% was estimated for NOx [120]. Nevertheless, laboratory CLA today has a cross-sensitivity of <1 ppm for 1000 ppm NH3. NH3 could also cause a loss of conversion efficiency of the NOx converter over time. On the other hand, NOx can be converted to N2 or NH4NO3 (ammonium nitrate), resulting in a lower concentration [120]. Ammonium nitrate formation is very rapid even at room temperature if NH3 and HNO3 are present. HNO3, in turn, is readily formed when NO2 dissolves in liquid water. Ammonia scrubbers can be used to avoid this problem with CLA [138]. The disadvantages over FTIR are the extra expenses required for an ozonizer, NO2 to NO converter with converter efficiencies >95% (which need to be checked every month according to EU Regulation), possible NO2 losses in the chiller, influence due to quenching from water and CO2 [136,139]. FTIR was also found to provide accurate measurements of NOx in the presence of NH3 [120].
Table A3 summarizes the studies that the FTIR NOx reading was compared to calibration cylinders or CLA. The comparisons with calibration gases were within ±5%, while with reference CLA within ±10%, with a few exceptions. For example, a study had 30% difference due to different exhaust flows used for the FTIR and the PEMS [110]. Higher differences were reported when the response of the FTIR was slow [34]. Comparison of tailpipe FTIR with dilution tunnel bags were also within ±10%, with only a few exceptions. In some cases, the laboratory analyzers were suspected [98].
The comparisons of FTIR and CLA for NO2 did not always give acceptable differences, with differences exceeding 10% in particular at different sampling locations [140]. One explanation might be due to the fact that NO2 is not directly measured but calculated from the difference between NOx and NO. Wrong CLA NO2 measurements can also originate from different time delays in the analyzer for NO and NOx [141], and as mentioned before, converter-related side reactions. Partly lower NO2 concentrations peaks due to reaction of NO2 with soot contamination have also been reported [141]. Finally, despite the advantages of FTIR spectroscopy, the fundamental interference of NO2 measurement by water vapor remains. It should be added that NO2 (and NH3) are considered “sticky” substances and the sampling setup is very important. NO2 can be underestimated if it adsorbs on surfaces in the sampling system, dissolves in the condensed water [142], or forms a salt in the presence of NH3 and remains in the sampling filter. Chiller penetration is defined in the USA regulation for analyzers that do not use NO2 to NO converter upstream of the chiller. Heated CLA systems are used when high NO2 concentrations are expected in order to avoid the loss of water-soluble NO2 inside the dehumidifier. The dry-to-wet correction introduces an uncertainty, which is significant during cold start [135]. Lower NO2 concentrations peaks over test cycle as reaction of NO2 with soot contamination have been reported [141].
All reference instruments presented in the table applied the CLA principle. Only in one case the reference instrument was a PEMS applying the NDUV principle [102]. The NDUV measurement method is based on the absorption of ultraviolet (UV) radiation from a broadband UV source to quantify the amount of NOx in a given sample [139,143]. The NO and NO2 signal occurs during the very large water wavelength in addition to other smaller contributing signals such as sulfur compounds [144]. For this reason, the sample is typically dried. The portable FTIR and the NDUV analyzer agreed well (slope 1.03, R2 0.97) [102]. This was expected as NDUV and CLA usually agree. For example, a study found the two principles within 5% for 20–130 ppm concentrations [145].

3.3. Nitrous Oxide (N2O)

Nitrous oxide (N2O) can be measured using FTIR, NDIR, QCL-IR, and GC with ECD. One study [102] compared two FTIRs measuring the tailpipe exhaust of diesel vehicle and found a slope of 0.96 (R2 = 0.95) for concentrations up to 60 ppm. Another study compared the FTIR reading from bags with exhaust gas injected with N2O and found differences within 3% of the expected values [146], indicating no particular interference from H2O or other exhaust gas components.

3.4. Ammonia (NH3)

Ammonia can be measured using FTIR, QCL or LDS from the tailpipe.
A Quantum Cascade Laser (QCL) element emits an intermittent mid-infrared laser beam in a sample cell at a very specific wave number by applying a pulsed current [128,147,148]. During a pulse, the element temperature varies, and, consequently, the actual wave number continuously shifts within a narrow range. Compounds that absorb energy in this oscillation range will decrease the light intensity and will be detected at the relevant wave numbers.
The measurement principle of a Laser Diode Spectrometer (LDS) is based on single line molecular absorption spectroscopy [149]. A diode laser is used for emitting a near-infrared (compared to QCL mid-infrared) light beam through the sample gas, and the light is detected by a receiver. The laser diode output is tuned for a gas-specific absorption line. Cross-sensitivities of the method is insignificant as, spectrally, the laser light is much narrower than the gas absorption line.
Table A4 summarizes the studies that FTIR was assessed in measuring NH3. The calibration gases readings were within 5% of the specified values. In general, the agreement between FTIRs was good. Steady-state engine tests with NH3 dosing showed the importance of the FTIR response time and the sampling line conditioning. The differences in dynamic response indicated buffering of NH3 (known to “stick” on walls) within the sampling line before breakthrough to the analyzer cell [150].
The comparison of FTIRs with QCLs was also good on average. In one case, examining real-time graphs, higher and sharper peaks were obtained with the QCL due to the 10-Hz frequency [151]. Comparison of an FTIR with chemical ionization mass spectrometry (CI-MS) gave good agreement [152], and comparisons with an in-situ (cross-duct, without sampling line) LDS was also in good agreement [149].
There were only a few studies comparing tailpipe with dilution tunnel [152,153]. However, as pointed out by these studies, severe losses of ammonia can occur in the dilution tunnel resulting in lower emissions.
The sampling location and the sampling conditions are very important for NH3 measurements. As it was mentioned in the NOx section, the formation of ammonium nitrate is also possible in the presence of HNO3. Low sample line temperature can result in condensation inside the heated line, which can lead to loss of NH3. On the other hand, upstream of the SCR in the presence of HNCO and H2O, a high sample line temperature can produce NH3 [150]. For such cases, a temperature of 113 °C was recommended. Downstream of the SCR, no such phenomena were observed.

3.5. Hydrocarbons (HCs) and Methane (CH4)

Total hydrocarbons (THC) and methane (CH4) are typically measured with Flame Ionization Detection (FID). In FID, the sample gas is introduced into a hydrogen flame, where some of the HCs in the sample gas are ionized [154,155]. Due to the ions, an electric current is generated at the applied electric potential, is nearly proportional to the amount of carbon atoms. For this reason, the HCs concentration is called “THC” and the unit is “ppmC”. A heated FID should be used in applications such as diesel exhaust that contains significant quantities of high boiling point HCs to avoid condensation and loss of heavier HCs [128]. The FID detector temperature should be set to 113 °C for alcohol-fueled vehicles (instead of 190 °C, which is used for diesel-fueled vehicles) (USA EPA Title 40, Chapter I, Subchapter C, Part 86, Subpart B, §86.111–94). This is based on the higher water vapor of methanol and the fact that methanol can undergo decomposition reactions if the oven is too hot [156].
The FID sensitivity for each HC is represented by a “response factor” that indicates the relative sensitivity compared to propane as the calibration gas. Generally, the FID has lower sensitivity to oxygenated HC components such as alcohols, aldehydes [157,158,159,160] and no sensitivity to formaldehyde. For regulatory purposes, FIDs are calibrated to have a response factor of 1 for propane. Toluene and propylene have to be within 0.9 and 1.1. The FID has negligible interference from inorganic components such as CO, CO2, H2O, and NO, except for O2 in the sample gas. The O2 concentration in the sample can affect not only FID sensitivity for HCs but also the zero point. Some THC analyzers have a compensation function for O2 interference using an additional O2 analyzer. The regulation allows a maximum of 1.5% oxygen interference.
For measuring CH4 in the engine exhaust gas with the FID analyzer, sample conditioning parts, i.e., Gas Chromatography (GC) or Non-Methane Cutter (NMC), are placed before the detector to extract or separate CH4 from the other HC components. The GC-FID-based CH4 analyzer is only suitable for batch measurement rather than continuous measurements. A Non-Methane Cutter (NMC) is a catalyst that selectively oxidizes HCs and only minimally CH4. The regulation requires the determination of the methane conversion efficiency, which ideally should be 0%, and the ethane conversion as an approximation of the conversion efficiency of the non-methane hydrocarbons, which should be >98%. If the CH4 response is <1.05, it may be omitted from the calculations of the emissions.
Non-Methane Hydrocarbons (NMHCs) are calculated as the difference of THC and CH4. (details in regulation, it is not a simple subtraction as conversion efficiencies need to be considered). In EU for high ethanol content fuels a different density is used to calculate THCs and NMHCs (0.934 g/L E85 vs. 0.646 g/L E10). In USA, non-methane organic gases (NMOG) include NMHCs, alcohols and carbonyls. Therefore, to measure NMOG (gasoline with ethanol > 25%), one must separately measure the principle alcohols, aldehydes, and ketones, subtract their inaccurate contribution to the FID signal, and add in their correct concentrations. Ethanol is traditionally measured by impinger collection and GC. Alternatively, the ethanol levels in diluted emissions sampled into Tedlar bags can be read directly via a photoacoustic non-dispersive infrared method [161], thus avoiding the need for off-line impinge analysis. Aldehyde measurement is via DNPH cartridge collection and off-line high-pressure liquid chromatography. Thus, ethanol and acetaldehyde are measured via batch methods, thus precluding time-resolved NMOG data.
FTIR spectroscopy provides an attractive alternative because all of the species contributing to NMOG are infrared active. For FTIR, in general, the analysis of infrared spectra works best for components with strong and sharp absorbance bands [113]. It is progressively more difficult for FTIR to speciate hydrocarbons as the number of carbons increases for two reasons [162]: (i) the individual molecular rotation-vibration lines coalesce into broader bandshapes. (ii) they progressively overlap with each other.
Table A5 summarizes the comparisons of FTIR with FID systems for THC. There is a high scatter of the results. The differences are typically within 15%, but underestimation by a factor of 2 has also been reported [110,163]. Another study found FTIR measuring 34% higher than the FID from the dilution tunnel [164]. The explanation was the reduced sensitivity of FID to methanol and formaldehyde. It should be emphasized that this is a comparison of two technics, which are both estimating THC. None of them are specifically quantifying every single HC. FID is non-specific, and FTIR is non-exhaustive.
Table A6 summarizes the FTIR studies with CH4. The differences were around 5% when measuring calibration gases and 10% compared to FIDs with NMC when measuring CH4 from the tailpipe. The difference to other methods (GS) was within 5%. A few older studies found higher differences (up to 18%) at levels of <10 ppm CH4. To put results into perspective of those older studies, when the FTIR was used at the tailpipe and then to measure the CH4 from the bags, the differences remained the same (slope 0.88 and R2 0.96) [165], indicating uncertainties with the FTIR. Similarly, a study that measured CH4 from cylinders found a 3.5 ppm offset due to interference from other gases [166]. The same study found more than double emissions compared to chemical ionization mass spectrometry (CI-MS).
Table A7 summarizes the FTIR studies with NMHC. The differences were around 5%, with some cases having >30% differences. For these cases, an uncertainty analysis revealed that the most likely reason for the differences was the reference method. The reference method uncertainty was around 80–100 ppm (almost 50–100%), while for FTIR was around 25 ppm (10–15%) [162]. Another study, even assuming ethane and ethylene to have an error of 10% and the other components to have an error of 50%, found that the FTIR had a lower absolute error than FID for NMHC [167].
A promising proposal to measure real-time NMHC and NMOG with FTIR is to use a group of HCs, as a surrogate to NHMC and then combined it with ethanol, acetaldehyde and formaldehyde to calculate NMOG [162]. That study also showed that typically the top five organic compounds account for about 60% of the HC emissions with CH4 the most abundant, and the top 10 compounds account for 80% for fuels ranging from gasoline to 85% ethanol/gasoline blends. The same study concluded that FTIR can measure NMHC and NMOG within 5% to the regulatory method [162].
Only a few studies compared ethanol emissions from FTIR and other methods. In one study, the differences were <10% for emission levels ranging from 100–1000 mg/km [160]. Another study with three FTIRs found on average −7% (0% to −12%) compared to PTR-Qi-ToF-MS during the cold start phase (emissions 150 mg/km) [168].
Regarding methanol many studies in the 1990s gave good results: slopes 0.72–1.10 (FTIR vs. impinger + CS) [26,112,169,170,171]. One study compared alcohols and found an agreement of 22% (FTIR vs. impinger + CS) for emission levels up to 250 mg/km and fuels up to E100 [172].

3.6. Formaldehyde (CH2O) and Acetaldehyde (CH3CHO)

The accepted techniques for measuring formaldehyde and acetaldehyde are DNPH + HPLC, FTIR, and PTR + MS; all from the dilution tunnel. The most common and cheaper method is using HPLC (high-performance liquid chromatography) to separate and quantify the carbonyls after extraction of 2,4-DNPH (dinitrophenylhydrazine) cartridges [173,174], while PTR-MS (proton transfer reaction—mass spectrometry) [168] is the least common. The DNPH-HPLC method also has its uncertainties, as the quantification by HPLC is not always straightforward due to the chromatogram peak resolution. NO2, NO, and CO can have an impact on the quantification of formaldehyde and acetaldehyde due to the consumption of the DNPH during sampling [175].
Table A8 summarizes the studies that assessed the FTIR measuring formaldehyde. Sampling from the dilution tunnel showed in general good agreement with the DNPH + HPLC methodology (slopes close to 1), with one exception where the FTIR was on average 20–25% higher. From the tailpipe, the differences were larger (±30%).
Table A9 summarizes the studies that assessed the FTIR measuring acetaldehyde. The scatter of the results is high. One study found more than double concentrations of acetaldehyde with the FTIR than with the DNPH + HPLC method. It seems that the reaction of acetaldehyde took place in the exhaust pipe [176]. Another study with a flexi-fuel vehicle found 30% higher concentrations with the FTIR (compared to DNPH-HPLC), but at the same study, the FTIR was 30% lower than the PTR-MS method [168]. The emission levels were around 0.4 mg/km. Other studies found differences within 10%. A study that checked the FTIR formaldehyde concentrations according to analyte spiking EPA method 301 on a coal-fired burner found differences on the order of 10% [177]. The same study found 5% differences for acetaldehyde.

3.7. Other Compounds

A few only studies assessed alkanes. For example, differences from propane (C3H8) gas cylinders were 3.3% at 3200 ppm [95], and 13% at 2 ppm or lower concentrations. One ppm interference from CO2 and/or H2O was found at 15 ppm levels [178]. Ethane was measured 5–10% lower [166].
Isocyanic acid (HCNO) was assessed in one study for SCR (selective catalytic reduction for NOx) applications with a heavy-duty diesel engine, and all three FTIRs were within a few ppm for concentrations up to 15 ppm [150].
Aromatic hydrocarbons (benzene, toluene) from FTIR were compared to the analysis of bags after Tenax TA adsorption and GC-MS analysis. Benzene was from −6.5% to +1.5% from bags for various fuels (gasoline, ethanol, methanol) [179]. The emission levels were around 2.5 mg/km, and the peaks at the cold start were around 20–30 ppm. The same study found differences of −3.5% to 2.4% (spike 40–60 ppm, emissions 7.5 mg/km) for toluene.
A very good correlation has been found for propylene, ethylene, acetylene, and benzene [180]. Results from other gases such as ethane, ethyne, 1,3-butadiene [166], acetylene, propene [113,171,181,182] can be found in the literature.

4. Discussion

FTIR analysis is seeing increased use in engine exhaust measurements. Since the first prototype instruments in the 1980s, many laboratories use FTIRs as a standard technique for engine development (see introduction). This review summarized the differences between FTIR and other methods for various gaseous components of engine exhaust. The differences were in most cases at acceptable levels (5–10%), but in some cases and for some compounds, higher differences were noticed. The main question is whether FTIR can be used for on-road regulatory purposes. To answer this question, the following topics need some analysis:
  • Can FTIR measure undiluted exhaust accurately enough?
  • Can FTIR be used on-board?
  • How the FTIR accuracy can be ensured for regulatory purposes?

4.1. Tailpipe Applications

Interestingly, GTR 15 allows the use of FTIR for ethanol, formaldehyde, acetaldehyde, and N2O only from the dilution tunnel. This is because there is no exhaust gas measurement to determine the emissions from the tailpipe. Only NH3 has to be measured from the tailpipe. At the moment, there is a limit only for heavy-duty engines (in ppm) in the EU regulation (not in GTR). Such specifications would need two FTIRs for the measurement of non-regulated pollutants (e.g., one for NH3 at the tailpipe and one at the dilution tunnel for the other pollutants). Permitting measurement of all pollutants from the tailpipe would simplify the setup. Furthermore, FTIR could be used instead of other analyzers (NDIR for CO and CO2), CLD (for NOx), and possibly for hydrocarbons (instead of FID). Indeed, the use of FTIR at the tailpipe is a commonly accepted technique for research and development (see “Introduction”).
Figure 4 gives an overview of the “Results” section summarizing the studies that FTIRs was compared with reference values: (i) calibration gases (“Cylinder”), (ii) reference instrument measuring in parallel with the FTIR at the dilution tunnel or at the tailpipe (“Parallel”), (iii) reference instrument at a different location (FTIR at the tailpipe versus reference at the dilution tunnel) (TP vs. CVS). Only cases where at least two studies were available were taken into account. It should be mentioned that the mean values of slopes or already averaged values do not give the complete scatter of the tests. Furthermore, a different number of tests in each case make any comparisons between different compounds doubtful. On the other hand, it has to be reminded that the results summarize 40 years of experience with a wide range of instruments manufacturers (and users).
The mean differences from the reference values were ±2.5% for CO2, CO, NOx, and NH3, without any particular deviation when FTIR and reference instrument were at different locations. The variability (one standard deviation) of the means was 5% for CO2 and 10% for the other three gases. Again no particularly higher variability of the “TP vs. CVS” cases. For HCs (THC, NMHC) the mean differences were 5–30% and the variability 30%. Smaller mean differences and variability (10%) was calculated for CH4. The “TP vs. CVS” cases had similar means and variability. CH2O had closer to CH4 behavior, while CH3CHO was closer to NH3, but the number of tests was very limited to draw any conclusions. The results are reasonable, considering that the uncertainty of the different equipment to which FTIR was compared was not the same. As it was discussed in the respective sections, the reference instruments for THCs, NMHCs, and carbonyls have higher uncertainty than those for, e.g., CO2 and CO. On the one hand, the THC measurement with FID is not specific, as different HCs can have different response factors in the flame; and on the other hand, THC estimate with the FTIR might be not exhaustive, as some HCs can be not quantified if they are not initially included in the calibration method. A proposal to bring closer the two methods was to use a group of HCs as a surrogate to NHMC and then combined it with ethanol, acetaldehyde, and formaldehyde to calculate NMOG [162]. Another way of research may be to apply chemometric tools, such as principal components regression (PCR) and partial least squares regression (PLSR), on the FTIR multivariate data to predict the THC estimated with FID. Either way, a higher uncertainty margin would be necessary when using FTIR to assess compliance to THCs standards.
The analysis was repeated considering only the studies of the last ten years (i.e., 2011 and afterwards), in order to see whether there were any trends of improvements (i.e., closer agreement with the references). For THCs, NMHC, CH4, and acetaldehyde there were no studies or only one study was available for each case, thus no conclusion could be drawn. For acetaldehyde, practically the same studies remained, thus there was no meaning for any comparison. For CO2, CO, NOx, and NH3, the mean differences and variabilities remained the same or slightly improved (in particular for CO), but without any statistically significant difference (only 2–6 studies available per case).

4.2. FTIR and Interferences

The higher differences for some components when measuring exhaust gas, compared to calibration gases, can be attributed to analytical and sampling interferences. Analytical interference (also called background or spectral interference) occurs when two or more compounds have overlapping absorbance bands in their infrared spectra. To minimize such interferences, appropriate resolution, selection of wave lengths, an appropriate library of expected components, and post-processing of the spectra are necessary [120].
Sampling system interferences are interferences that prohibit or prevent delivery of the target compounds to the FTIR gas cell (e.g., moisture condensation, reactive gases). Regulations, for example, require a heated sampling line (191 °C) when sampling undiluted exhaust in order to avoid the wall adsorption and/or dissolution of hydrophilic compounds (e.g., NH3, NO2, aldehydes, or ethanol) in condensed water. A study noticed the delay in oxygenated species reaching the tailpipe during cold start because of their condensation onto cold exhaust system surfaces and dissolution into condensed water [162]. FTIR systems might have differences in real-time operation. It was shown that an FTIR with a slower gas replacement rate and lower sampling frequency was not able to detect some of the rapid concentration fluctuations, e.g., for CO [183]. Another study noticed that during decelerations, the NH3 concentration did not drop to near zero as it would be expected during fuel cut-offs (NH3 formation is strongly inhibited by O2) [184]. Partly the lower response time of the instrument could explain this. However, it was suggested that an important reason was the outgassing of NH3 from metal surfaces, which act as temporary NH3 storage reservoirs [184]. A dedicated study on NH3 found that response attenuation rates were due to mixing and diffusion during transport as well as NH3 wall storage. Mixing/diffusion effects caused attenuation with a mean time constant of around 1.6 s. Wall storage attenuation had a mean time constant of 72 s [185]. The stored NH3 on the sampling lines was around 11 mg. It was concluded that, in practical terms, shorter lines at a higher temperature, with flow rates > 10 L/min proved the best for transient response testing [185].

4.3. On-Board Applications

In the previous paragraphs, it was shown that tailpipe application is possible, paying attention to analytical and sampling interferences. Are FTIRs robust enough for on-board applications? Portable systems were already available in the 1980s. However, portable does not necessarily mean suitable for on-board measurements. The main concerns are:
  • Size, weight, power consumption.
  • Effect of environmental conditions (temperature, altitude, vibrations).
  • Safety (liquid N2 for cooling, other gases on-board, e.g., N2 for purging).
The importance of size, weight, and power consumption is different for light-duty and heavy-duty applications. The size and weight are very important for light-duty applications, especially for small city cars. Commercial portable FTIRs are split into units that can fit in the vehicle cabin and/or on the hook. The weight without accessories (e.g., pumps, batteries, and heated lines) is around 50 kg, but including accessories is around 100 kg, which is slightly heavier compared to commercial PEMS (portable emissions measurement systems). Even though in the 1980s the need for power generators of 10 kW was reported [186], today’s portable systems are <0.5 kW (after warming up in the laboratory). These values are still higher or comparable to PEMS based on other principles. An important point for energy consumption is the location of the sampling pump. If it is located upstream of the FTIR, it needs to be heated, thus the energy consumption will be higher compared to a downstream location. Furthermore, at the downstream location, the pressure can be lower than atmospheric pressure, which is an advantage for on-board measurement because it can be maintained and fixed easier. The liquid N2 on-board is a concern. One solution is keeping the liquid N2 in a sealed container, with only one small tube connected, venting the evaporating N2 to the atmosphere or to the rest of the exhaust gases from the FTIR pump.
The effect of the environmental conditions should be well characterized. FTIRs are sensitive to vibrations: the better the resolution is, the longer the displacement of the moving mirror in the interferometer, and the higher the effect of the vibrations on the optical system. Vibration tests in the late 1990s concluded that FTIRs were best isolated by simply placing them on the rear seat of the vehicle [86]. Today FTIR suppliers claim that their systems are vibrations robust. For example, wire rope isolators can be used [187]. Static single mirror solutions have also been presented [188]. Experiments with NOx PEMS (based on CLA or NDUV) showed that sudden temperature changes resulted in zero drift [189]. FTIR results are also sensitive to temperature. The recent CEN standard on PEMS performance prescribes appropriate testing procedures to properly assess the influence of temperature, pressure, and vibrations [190]. Long-term stability and robustness due to vibrations, contamination of optics, etc., should also be assessed.
As with all portable systems, comparison with laboratory versions or other well-established techniques are necessary to increase their confidence in them. Sometimes portable systems might not have the appropriate spectral resolution, response time, or detection limits [104]. For example, in the past, drying of the sample has been used [86], slow response (30 s) [78], or the low optical resolution of 4 cm−1 [110] might have negatively affected the accuracy of the results. It should also be emphasized that the studies in Appendix A were with prototype portable systems, thus further studies with commercial systems are needed.

4.4. Regulatory Requirements

It is clear that regulations cannot prescribe in detail all technical aspects of FTIRs. Some basic and important parameters can be described, but appropriate tests are necessary to confirm the instrument’s internal hardware and software performances. Table 3 summarizes the technical specifications for FTIRs for NH3 measurements in the current regulations. Most of them are based on the experience of the users and the capabilities of the instruments. Nevertheless, some specifications could be further restricted or better be controlled for future low emissions vehicles. Such recommendations can be based on recent (2019–2020) standards and methodologies for FTIRs [191,192,193,194].
For example, SAE J2992 [192] includes a few more requirements (e.g., repeatability and noise). The detection limit, typically determined with zero gas [191,192], could also be determined using interfering gases [193]. Similarly, in addition to the accuracy test, defined as the deviation of the analyzer reading from the reference value, an interference test could be added. SAE J2992 requires interference testing with a gas containing CO2, CO, NOx, and N2O (mix or separately). For most gases (e.g., NH3, N2O, etc.), the maximum total permitted interference is 1%. A spectral residual test is also recommended [191]. USA regulations require appropriate analytical procedures for the interpretation of infrared spectra [193,194].
In the EU RDE regulation, a maximum zero drift of 5 ppm and span drift of 2% is allowed for NOx. This is much stricter than the 2% of full scale currently prescribed for FTIR. Actually, the drift of FTIR should be negligible as it was discussed before, not only for NH3, but for all compounds. Thus these drift requirements could be even stricter.
In USA EPA regulations, FTIR analyzer may be used to measure CH4, C2H6, NMHC, and non-methane-non-ethane hydrocarbon (NMNEHC) for continuous sampling for natural gas engines (Title 40/I/U/1065/C/§1065.266). The FTIR analyzer must have combined interferences that are within ±2% (recommended ±1%) with CH4, NMHC, or NMNEHC concentrations expected at the standard (Title 40/I/U/1065/D/§1065.366). Such uncertainties are at the same levels permitted for oxygen interference for FID analyzers.
Furthermore, for regulatory purposes, more quality checks should be included. SAE J2992 has a separate chapter of tests performed prior to and after an emissions cycle test (leak, zero, span, pre- and post-drift checks) [192]. As in CEN/TS 17337, adjustment factors for zero and span, or even zero and span drift could be allowed [191].

4.5. Measurement Uncertainty

One important aspect for future regulations is the uncertainty topic. At the moment, the uncertainty for PEMS analyzers is based on a simple (single point—worst case) model. Assuming a 2 ppm accuracy (including interferences), a 2.5 mg/km uncertainty is calculated for a 3 L engine for the analyzer [107] (see also Table 2). Combining the uncertainty of the exhaust flow, the distance, the trip dynamics, this value could be doubled. This value can be 5 times higher for heavy-duty engines. The values are half of the proposed future NH3 limits [103]. This means that the accuracy requirement instead of “±3% of the reading or ±2 ppm, whichever is larger”, should change, for example, to “± 2% of the reading for concentrations >50 ppm, and ±1 ppm for lower concentrations”. This change would significantly reduce the uncertainty if mass limits are set (and not concentration in ppm). CEN/TS 17337, applicable to stationary sources emissions provides a detailed analysis [191]. A big step was the publication of CEN PEMS performance standard, where a second by second calculation is provided for the calculation of the final uncertainty [190]. Related to the uncertainty is also the traceability topic. The regulated gases have reached high levels of accuracy and traceability. However, this is not the case for non-regulated gases. For example, formaldehyde is tricky to calibrate because it tends to polymerize against the cylinder walls of the gas container. An old study found transfer efficiency of formaldehyde of 95.5% (±13.5%) for all vehicle types (gasoline, diesel, methanol) [24]. At the time of writing, the uncertainty of the formaldehyde calibration gas is much higher than the 2% uncertainty of the regulated gases. In another study, acetaldehyde standard (25 ppm in N2) could not be used because an impurity was detected in the gas bottle [168].

5. Conclusions

This review summarized the studies that assessed FTIRs performance on the measurement of vehicle exhaust emissions. The mean differences compared to regulated or other methods were around ±2.5% for CO2, CO, NOx, and NH3 with a variability (one standard deviation) of 5% for CO2 and 10% for CO, NOx, and NH3. For CH4, acetaldehyde, and formaldehyde, the mean differences were ±10% (variability 10–20%), but for total hydrocarbons, much higher differences were noticed. The differences were similar regardless of the sampling location of the FTIR (dilution tunnel or tailpipe). Assessment of prototype portable FTIRs on the road confirmed these findings also on-board, but for a narrow range of environmental and driving conditions. Based on these results, FTIRs may be an alternative for on-road testing. However, more studies with commercial portable systems are necessary to cover a wider range of environmental and driving conditions. The introduction of FTIRs in the regulation will require strict technical and performance requirements and procedures based on recently developed standards.

Author Contributions

Conceptualization, B.G.; formal analysis, B.G.; writing—original draft preparation, B.G.; writing—review and editing, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are summarized in the tables.

Conflicts of Interest

The authors declare no conflict of interest. The opinions expressed in this manuscript are those of the authors and should in no way be considered to represent an official opinion of the European Commission. Mention of trade names or commercial products does not constitute endorsement or recommendation by the European Commission or the authors.

Appendix A

The following tables summarize the studies that evaluated FTIRs comparing them with reference values. The following terminology applies:
  • Year: Year of the study (published). The studies are given in chronological order; separately for laboratory and portable systems.
  • Instrument: Manufacturer and model of the FTIR (as given in the study). In italics in case the FTIR was portable (and the letter “P” is added).
  • Difference: Difference of FTIR from the reference value. If many points were available, the mean and the range (in brackets) is given.
  • Slope: Linear regression analysis slope.
  • R2: Coefficient of determination of linear regression.
  • Range: Concentration or emissions range (from plots or tables of the studies).
  • Points: Number of points that the mean or the slope and R2 were calculated.
  • Comment: Particularities of the tests and additional information: CNG = compressed natural gas; D = diesel; E = ethanol; FFV = flex-fuel vehicle; G = gasoline; M = methanol; x = number of cars; y = years.
  • Ref: Citation of study.
For each study, the following subdivisions were made whenever data were available:
  • FTIR measuring cylinder (calibration gas).
  • FTIR and reference, both sampling from the dilution tunnel.
  • FTIR and reference, both sampling from the tailpipe.
  • FTIR sampling from the tailpipe, reference from the dilution tunnel.
Table A1. Comparisons of FTIR with CO2 analyzers (all NDIR).
Table A1. Comparisons of FTIR with CO2 analyzers (all NDIR).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR measuring cylinder
2010MKS MultiGas 2030−10.5%--12%1mix with CO, NO, C3H8[95]
2010AVL Sesam-1.01-1–14%10 [195]
2019AVL Sesam<1%--1500 and 15,000 ppm2 [196]
2005P: Temet Gasmet CR2000−4.4%--14.5%1 [110]
2005P: Temet Gasmet CR2000−5.4%--10.0%1mix with CO[110]
2015P: Spectrum 2, PerkinElmer0.4% (−9%...4%)1.021.000–20%17 [181]
Both FTIR and reference connected at the dilution tunnel
1990Nicolet Sesam−0.6%--300–370 g/mimeansG, Dx2[165]
1990Mattson Horiba-0.920.990.4–2.0%25G, fuels M85[112]
1993Mattson REA13.7% (0.8%...27%)1.050.980.2–1%3802.5 y (G, CNG)[170]
1993Nicolet REGA3.0% (−13%...18%)0.930.970.2–1%2682.5 y (G, CNG)[170]
1994Nicolet REGA−2.8% (−9%...7.5%)0.960.890.4–0.6%12G[146]
1994Nicolet REGA 70002.1%--0.4%1G, fuel M85[171]
Both FTIR and reference connected at the tailpipe
1990Nicolet Sesam−1.3%--300–370 g/mimeansG, Dx2[165]
2010AVL Sesam (MKS)−5% 600–1200 g/kWh25G, fuels[123]
2019AVL Sesam±2%--4–13%9G, D, CNG, moped, moto.[135]
2005P: Temet Gasmet CR20003.1%--12.8%1engine[110]
2005P: Temet Gasmet CR20002−12.6%--400 g/km1G[110]
2006P: Temet Gasmet CR2000-1.020.968–15%cycleG[163]
2020P: Bruker Matrix MG52-0.990.99<13%cycleD[102]
FTIR at the tailpipe and reference at the dilution tunnel
1985Prototype-0.980.9430–340 g/mi80G[164]
1998VW Sesam-1.010.96300–550 g/mi65Gx10[113]
2013MEXA-6000FT10% (±1%)--200 g/km6Gx2 (E0–10–20, M15–30)[179]
2017Gasmet CR2000 1.060.97140–250 g/km20Dx2, Gx2, CNG[197]
2018AVL Sesam±4%--110–190 g/km11D[61]
2020AVL Sesam-1.010.90200–440 g/km9D[198]
2021AVL Sesam0.2% (±0.8%)--140–650 g/km37G[199]
2006P: Temet Gasmet CR2000−2.1% (−5%...1%)--180–230 g/km11G[163]
2018P: Nicolet Antaris IGS6.4% (−19%...61%)0.960.7085–300 g/km18D, CNG[98]
1 P: portable (in italics). 2 Comparison on the road with PEMS.
Table A2. Comparisons of FTIR with CO analyzers (all NDIR).
Table A2. Comparisons of FTIR with CO analyzers (all NDIR).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR measuring cylinder
2010MKS MultiGas 2030−0.8%--8%1 [95]
2019AVL Sesam<1%--1000 ppm4 [196]
2000P: Nicolet Protégé 460<0.5%--0–10 ppm3 [86]
2000P: Nicolet-1.021.000–19 ppm5–10 [178]
2000P: Nicolet−3.6%--19 ppm2Mix (CO2, H2O)[178]
2005P: Temet Gasmet CR20008.7%--4900 ppm1 [110]
2015P: Spectrum 2, PerkinElmer−1.2% (−30%...15%)0.991.000–5000 ppm15 [181]
Both FTIR and reference connected at the dilution tunnel
1990Nicolet Sesam−4.2%--1–10 g/mimeansG, Dx2[165]
1990Mattson Horiba-0.800.870–2500 ppm25G, fuels M85[112]
1991Bio-Rad Digilab FTS-60-0.820.960–400 ppm20Many G, fuels[169]
1993Mattson REA4% (−51%...68%)0.940.980–0.8%3902.5 y (G, CNG)[170]
1993Nicolet REGA0% (−87%..109%)1.200.860–0.8%2682.5 y (G, CNG)[170]
1994Nicolet REGA−0.8% (−6%...7%)0.961.004–265 ppm12G[146]
1994Nicolet REGA 70000.4%--178 ppm1G, fuel M85[171]
2000P: Nicolet Protégé 4603.5%--5–26 ppm2G[86]
Both FTIR and reference connected at the tailpipe
1990Nicolet Sesam4.4%--1–10 g/mimeansG, Dx2[165]
2010AVL Sesam (MKS)−5%-0.9920–45 g/kWh25G, fuels[123]
2010AVL Sesam−2.5% (0%…9%)--5–170 g/kWh66 snowmobiles[195]
2005P: Temet Gasmet CR2000−12%--2500 ppm1engine[110]
2005P: Temet Gasmet CR200021.3%--10 g/km1G[110]
2006P: Temet Gasmet CR2000-0.870.990–6000 ppmcycleG[163]
2020P: Bruker Matrix MG52-1.000.980–2000 ppmcycleD[102]
FTIR at the tailpipe and reference at the dilution tunnel
1985Prototype-1.040.980–37 g/mi80G[164]
1998VW Sesam-1.080.990–6 g/mi65Gx10[113]
2013MEXA-6000FT5.8% (−4%...8%)--0.27–0.37 g/km6Gx2 (E0–10–20, M15–30)[179]
2017Gasmet CR2000 0.950.9750–3000 mg/km15Dx2, Gx2, CNG[197]
2006P: Temet Gasmet CR2000−1.5% (−11%.4%)0.950.980.5–1.1 g/km11G[163]
2007P: Nicolet-0.900.900.1–1.1 g/mi15Gx4[92]
2018P: Nicolet Antaris IGS−4.7% (−75%...69%)1.010.9210–400 mg/km18D, CNG[98]
1 P: portable (in italics). 2 Comparison on the road with PEMS.
Table A3. Comparisons of FTIR with NOx analyzers (all CLD unless otherwise specified).
Table A3. Comparisons of FTIR with NOx analyzers (all CLD unless otherwise specified).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR measuring cylinder
2010MKS MultiGas 20303.5%--4000 ppm1 [95]
2010AVL Sesam-1.00-0–4000 ppm14 [195]
2000P: Nicolet-0.951.000–3 ppm5–10 [178]
2000P: Nicolet−2.1% and 1.1%--3 ppm2Mix (CO2, H2O)[178]
2000P: Nicolet Protégé 460±10%--1–10 ppm5 [86]
2005P: Temet Gasmet CR20000.3%--1451 ppm1 [110]
Both FTIR and reference connected at the dilution tunnel
1990Nicolet Sesam−5.2%--0.5–1 g/mimeansG, Dx2[165]
1990Mattson Horiba-0.950.940–50 ppm25G, fuels M85[112]
1991Bio-Rad Digilab FTS-60-0.930.890–40 ppm20Many G, fuels[169]
1993Mattson REA4% (−40%...48%)1.040.780–100 ppm3682.5 y (G, CNG)[170]
1993Nicolet REGA32% (−31%..86%)1.360.970–100 ppm2172.5 y (G, CNG)[170]
1994Nicolet REGA −6.6% (−11%.1%)0.961.004–14 ppm12G[146]
1994Nicolet REGA 70001.0%--3 ppm1G, fuel M85[171]
2015AVL Sesam1.3%...10.5%--40–80 ppm2Dx2[140]
2000P: Nicolet Protégé 4600.5%--0.5–3 ppm2G[86]
Both FTIR and reference connected at the tailpipe
1990Nicolet Sesam8.0%--0.5–1 g/mimeansG, Dx2[165]
1996Nicolet REGA 7000−10%--0–4000 ppmcycleG[30]
2010AVL Sesam (MKS)5%-1.008–20 g/kWh25G, fuels[123]
2010Nicolet Magna 5605.4% (3%..9%)--20–100 ppm4CNG + NO2 injection[142]
2016Not disclosed<5%--300–1000 ppmsteadyD[200]
2018AVL Sesam (MKS)-1.031.000–600 ppm600various[141]
2018AVL Sesam4% (±9%)--50–1200 mg/km22D[61]
2000P: Nicolet± 5%--0–70 ppmcycleG[178]
2005P: Temet Gasmet CR200013%--70 ppm1engine[110]
2005P: Temet Gasmet CR20002−30%--1.9 g/km1G[110]
2006P: Temet Gasmet CR2000-1.050.940–800 ppmcycleG[163]
2017P: MIDAC I-series2-1.090.990–400 ppmcycleD[201]
2020P: Bruker Matrix MG52-1.030.970–1000 ppmcycleD[102]
FTIR at the tailpipe and reference at the dilution tunnel
1985Prototype-0.980.970–3.2 g/mi80G[164]
1998VW Sesam-0.980.990–0.8 g/mi65Gx10[113]
2013MEXA-6000FT7.1% (5%...10%)--0.5 g/km6Gx2 (E0–10–20, M15–30)[179]
2015AVL Sesam−8.8%...3.0%--190–310 ppm4D[140]
2018AVL Sesam11% (±10%)--50–1200 mg/km22D[61]
2020AVL Sesam-0.940.790–350 mg/km9D[198]
2017Gasmet CR2000-0.790.9910–1200 mg/km12Dx2, Gx2, CNG[197]
2006P: Temet Gasmet CR20003.6% (2.7%...4%)--0.2–0.3 g/km11G[163]
2007P: Nicolet-1.020.965–40 mg/mi8Gx4[92]
2018P: Nicolet Antaris IGS32% (−22%...117%)1.210.925–650 mg/km18D, CNG[98]
1 P: portable (in italics). 2 Comparison on the road with PEMS.
Table A4. Comparisons of FTIR with NH3 analyzers (FTIR, QCL, LDS).
Table A4. Comparisons of FTIR with NH3 analyzers (FTIR, QCL, LDS).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR measuring cylinder
2009AVL Sesam± 5%--20 ppm1Daily checks[202]
2012AVL Sesam−0.2%--100 ppm1 [203]
2016Horiba<2.5%--100 ppm1 [150]
2016Not disclosed<5%--0–350 ppm3D engine NH3 injection[200]
Both FTIRs connected at the tailpipe
2016Horiba<2.5%--50 ppm1D[150]
2017P: MIDAC I-series-1.030.800–300 ppmcycleG[201]
2020P: Bruker Matrix MG5-0.820.960–100 ppmcycleD[102]
Both FTIR and QCL connected at the tailpipe
2011Horiba0–5%--0–350 ppmmanyD and NH3 injection[147]
2015MKS 2030-HS−6%1.091.000–25 ppm4Dx2, G, FFV[151]
2020Not disclosed-0.980.9925–100 ppmcycleD[204]
Both FTIR and LDS or CI-MS at the tailpipe
2018Gasmet DX4000 2−9% (−19%...1%)--0–25 ppm5D[149]
2004Nicolet Avatar 370 3-1.010.9925–100 ppmcycleG[152]
1 P: portable (in italics). 2 vs. LDS. 3 vs. CI-MS.
Table A5. Comparisons of FTIR with THC analyzers (FID).
Table A5. Comparisons of FTIR with THC analyzers (FID).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
Both FTIRs connected at the dilution tunnel
1985Prototype-1.340.970–3.7 g/mi80G[164]
1990Nicolet Sesam1.9% (−2%...1%)--0.2–0.9 g/mimeansG, Dx2[165]
1990Mattson Horiba-1.130.990–1200 ppm25G (fuels, M85)[112]
Both FTIR and reference (FID) connected at the tailpipe
1990Nicolet Sesam17.6% (1%...38%)--0.2–0.9 g/mimeansG, Dx2[165]
1996Nicolet REGA 7000−15%--<10,000 ppmcyclesG[30]
2005P: Temet Gasmet CR2000−66%, −63%--600 ppm engine G, G[110]
2006P: Temet Gasmet CR2000−55%0.410.990–1600 ppmcycleG[163]
FTIR at the tailpipe and reference at the dilution tunnel
1994Nicolet REGA 70001.8% (−4%...4%)--50 ppm5CNG[171]
1998VW Sesam 1.090.97<0.4 g/mi65Gx10[113]
2000P: Nicolet5%--550 mg/mi1G[178]
2006P: Temet Gasmet CR2000−55% (−5% cal.)--160 mg/km10G[163]
2018P: Nicolet Antaris IGS46% (−63%...280%)0.630.690–300 mg/km16D, CNG[98]
1P: portable (in italics).
Table A6. Comparisons of FTIR with CH4 analyzers (FID), unless specified otherwise.
Table A6. Comparisons of FTIR with CH4 analyzers (FID), unless specified otherwise.
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR measuring cylinder
1992Nicolet REGA 700015%--15–17 ppm2Mix [166]
2005P: Temet Gasmet CR2000−5.1%--1509 ppm1 [110]
2015P: Spectrum 2, Perkin Elmer1.0% (−2.4%...7.1%)1.021.000–50%17 [181]
Both FTIRs connected at the dilution tunnel (or tailpipe if specified)
1991Bio-Rad Digilab FTS-60-0.890.960–10 ppm20Many G, fuels[169]
1994Nicolet REGA 7000−17.9%--3 ppm1G (M85)[171]
2016Thermo Fisher Antaris IGS2.4% (0.8%...4.2%)--1500–3000 ppm12D-CNG[167]
FTIR at the tailpipe and reference at the dilution tunnel
1998VW Sesam-1.090.990–0.07 g/mi65Gx10[113]
2020AVL Sesam-0.960.990–40 mg/km9D[198]
FTIR versus FTIR or GS
1994Nicolet REGA 7000 25%--43 ppm1CNG[171]
1994Sesam II 35% (−29%...20%)1.110.9940–250 ppm12G[180]
2017Gasmet CR2000 2-0.970.953–105 mg/km17Dx2, Gx2, CNG[197]
2020P: Bruker Matrix MG53-1.070.930–6000 ppmcycleD[102]
1 P: portable (in italics). 2 vs. GS. 3 vs. FTIR
Table A7. Comparisons of FTIR with NMHC analyzers (FID).
Table A7. Comparisons of FTIR with NMHC analyzers (FID).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIRs measuring cylinder
2000P: Nicolet Protégé 4600–9%--1–10 ppm3 [86]
Both FTIRs connected at the dilution tunnel (or tailpipe if specified)
1991Bio-Rad Digilab FTS-60-1.030.960–120 ppm20Many G, fuels[169]
2000P: Nicolet Protégé 460−14%...−5%--2 and 8 ppm2G[86]
2016Thermo Fisher Antaris IGS97% (31%...292%)--50–350 ppm12D-CNG[167]
FTIR at the tailpipe and reference at the dilution tunnel
2007P: Nicolet9.5% (−32%...52%)1.040.993–51 mg/mi8Gx4[92]
2017AVL Sesam5%0.95–1.05-0–2500 ppmmanyG, FFV (E10, E50, E85)[162]
1P: portable (in italics).
Table A8. Comparisons of FTIR with formaldehyde (CH2O) methodology (DNPH + HPLC).
Table A8. Comparisons of FTIR with formaldehyde (CH2O) methodology (DNPH + HPLC).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR connected at the dilution tunnel
1986Prototype-1.010.990.3–8.5 ppm78G (methanol)[24]
1990Mattson Horiba-1.030.990–80 ppm25G, fuels M85[112]
1992Nicolet REGA 700017%1.140.990–45 mg/mi6G[166]
1993Mattson REA18% (−60%...98%)1.250.910–4 ppm2062.5 y (G, CNG)[170]
1993Nicolet REGA34% (−33%...96%)1.040.950–4 ppm2292.5 y (G, CNG)[170]
1994Nicolet REGA 70000.9%--10 ppm1G, fuel M85[171]
1994Sesam II−13% (−21%...−6%)0.880.882–16 ppm6G[180]
FTIR at the tailpipe and reference at the dilution tunnel
1990Nicolet Sesam-0.850.943–72 mg/mi15G, Dx2[165]
1991Bio-Rad Digilab FTS-60-0.840.970–8 ppm20Many G, fuels[169]
1998VW Sesam-0.680.570–4 mg/mi65Gx10[113]
1998VW Sesam15%--2 mg/mi2G[113]
2006MEXA 4000 FT-1.010.910–6 ppm26D[176]
2013MEXA 6000 FT−0.5% (−2.2%...2.4%)--2 mg/km (<60 ppm)6Gx2 (E0–10–20, M15–30)[179]
2016AVL Sesam−5%...27%--0–4 mg/km3Motorcycle (flexi)[205]
2017Gasmet CR2000-1.320.980–20 mg/km20Dx2, Gx2, CNG[197]
2017MKS, Sesam, MEXA8% (−4%...32%)-0.82–0.940–8 mg/km8FFV[168]
Table A9. Comparisons of FTIR with acetaldehyde (CH3CHO) methodology (DNPH + HPLC).
Table A9. Comparisons of FTIR with acetaldehyde (CH3CHO) methodology (DNPH + HPLC).
YearInstrument 1DifferenceSlopeR2RangePointsCommentRef.
FTIR connected at the dilution tunnel
2010Gasmet CR2000−20%, +5%--10–35 mg/km2FFV (E5, E85)[160]
FTIR at the tailpipe and reference at the dilution tunnel
2006MEXA 4000 FT-2.450.790–2 ppm27D[176]
2013MEXA 6000 FT2.2% (−1%...5%)--1–2 mg/km (<100 ppm)3Gx2 (E0, E10, E20)[179]
2016AVL Sesam−1%...5%--0–35 mg/km3Motorcycle (flex-fuel)[205]
2017Gasmet CR2000-1.080.900–27 mg/km22Dx2, Gx2, CNG[197]
2017MKS, Sesam, MEXA5% (0%...9%)--0–43 mg/km8FFV[168]

References

  1. Berg, W. Legislation for the reduction of exhaust gas emissions. In Traffic and Environment; Gruden, D., Ed.; Springer: Berlin/Heidelberg, Germany, 2003; Volume 3T, pp. 175–253. ISBN 978-3-540-00050-1. [Google Scholar]
  2. Keenan, M. Exhaust Emissions Control: 60 Years of Innovation and Development; SAE Technical Paper 2017-24-0120: Warrendale, PA, USA, 2017. [Google Scholar] [CrossRef]
  3. Smith, B.C. Fundamentals of Fourier Transform Infrared Spectroscopy, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2011; ISBN 978-1-4200-6929-7. [Google Scholar]
  4. Guerrero-Pérez, M.O.; Patience, G.S. Experimental Methods in Chemical Engineering: Fourier Transform Infrared Spectroscopy—FTIR. Can. J. Chem. Eng. 2020, 98, 25–33. [Google Scholar] [CrossRef]
  5. Doyle, G.J.; Tuazon, E.C.; Graham, R.A.; Mischke, T.M.; Winer, A.M.; Pitts, J.N. Simultaneous Concentrations of Ammonia and Nitric Acid in a Polluted Atmosphere and Their Equilibrium Relationship to Particulate Ammonium Nitrate. Environ. Sci. Technol. 1979, 13, 1416–1419. [Google Scholar] [CrossRef]
  6. Tuazon, E.C.; Winer, A.M.; Pitts, J.N. Trace Pollutant Concentrations in a Multiday Smog Episode in the California South Coast Air Basin by Long Path Length Fourier Transform Infrared Spectroscopy. Environ. Sci. Technol. 1981, 15, 1232–1237. [Google Scholar] [CrossRef] [PubMed]
  7. Herget, W.F.; Brasher, J.D. Remote Measurement of Gaseous Pollutant Concentrations Using a Mobile Fourier Transform Interferometer System. Appl. Opt. 1979, 18, 3404–3420. [Google Scholar] [CrossRef]
  8. Herget, W.F.; Lowry, S.R. Auto Exhaust Gas Analysis by FTIR Spectroscopy. In Proceedings of the Optics, Electro-Optics, and Laser Applications in Science and Engineering, Los Angeles, CA, USA, 20–25 January 1991; Proc. SPIE 1433 Measurement of Atmospheric Gases. pp. 275–289. [Google Scholar] [CrossRef]
  9. Allen, D.T.; Palen, E. Recent Advances in Aerosol Analysis by Infrared Spectroscopy. J. Aerosol Sci. 1989, 20, 441–455. [Google Scholar] [CrossRef]
  10. Takahama, S.; Dillner, A.M.; Weakley, A.T.; Reggente, M.; Bürki, C.; Lbadaoui-Darvas, M.; Debus, B.; Kuzmiakova, A.; Wexler, A.S. Atmospheric Particulate Matter Characterization by Fourier Transform Infrared Spectroscopy: A Review of Statistical Calibration Strategies for Carbonaceous Aerosol Quantification in US Measurement Networks. Atmos. Meas. Tech. 2019, 12, 525–567. [Google Scholar] [CrossRef] [Green Version]
  11. Lu, P.-D.; He, T.; Zhang, Y.-H. Relative Humidity Anneal Effect on Hygroscopicity of Aerosol Particles Studied by Rapid-Scan FTIR-ATR Spectroscopy. Geophys. Res. Lett. 2008, 35. [Google Scholar] [CrossRef]
  12. Liu, J.; Wang, L.; Sun, P.; Wang, P.; Li, Y.; Ma, H.; Wu, P.; Liu, Z. Effects of Iron-Based Fuel Borne Catalyst Addition on Microstructure, Element Composition and Oxidation Activity of Diesel Exhaust Particles. Fuel 2020, 270, 117597. [Google Scholar] [CrossRef]
  13. Mao, Z.; Demirgian, J.; Mathew, A.; Hyre, R. Use of Fourier Transform Infrared Spectrometry as a Continuous Emission Monitor. Waste Manag. 1995, 15, 567–577. [Google Scholar] [CrossRef]
  14. Bradley, K.S.; Brooks, K.B.; Hubbard, L.K.; Popp, P.J.; Stedman, D.H. Motor Vehicle Fleet Emissions by OP-FTIR. Environ. Sci. Technol. 2000, 34, 897–899. [Google Scholar] [CrossRef]
  15. Phillips, F.A.; Naylor, T.; Forehead, H.; Griffith, D.W.T.; Kirkwood, J.; Paton-Walsh, C. Vehicle Ammonia Emissions Measured in an Urban Environment in Sydney, Australia, Using Open Path Fourier Transform Infra-Red Spectroscopy. Atmosphere 2019, 10, 208. [Google Scholar] [CrossRef] [Green Version]
  16. Schäfer, K.; Heland, J.; Lister, D.H.; Wilson, C.W.; Howes, R.J.; Falk, R.S.; Lindermeir, E.; Birk, M.; Wagner, G.; Haschberger, P.; et al. Nonintrusive Optical Measurements of Aircraft Engine Exhaust Emissions and Comparison with Standard Intrusive Techniques. Appl. Opt. 2000, 39, 441–455. [Google Scholar] [CrossRef] [PubMed]
  17. Vojtisek-Lom, M.; Jirků, J.; Pechout, M. Real-World Exhaust Emissions of Diesel Locomotives and Motorized Railcars during Scheduled Passenger Train Runs on Czech Railroads. Atmosphere 2020, 11, 582. [Google Scholar] [CrossRef]
  18. Lechner, B.; Paar, H.; Sturm, P.J. Measurement of VOCs in Vehicle Exhaust by Extractive FTIR Spectroscopy. In Proceedings of the Europto Remote Sensing, SPIE, Barcelona, Spain, 25–29 September 2000; Fujisada, H., Lurie, J.B., Ropertz, A., Weber, K., Eds.; Proc. SPIE 4169. pp. 432–439. [Google Scholar] [CrossRef]
  19. Bacsik, Z.; Mink, J.; Keresztury, G. FTIR Spectroscopy of the Atmosphere Part 2. Applications. Appl. Spectrosc. Rev. 2005, 40, 327–390. [Google Scholar] [CrossRef]
  20. Simonescu, M.C. Application of FTIR spectroscopy in environmental studies. In Advanced Aspects of Spectroscopy; Farrukh, M.A., Ed.; InTech: London, UK, 2012; ISBN 978-953-51-0715-6. [Google Scholar]
  21. Butler, J.W.; Maker, P.D.; Korniski, T.J.; Haack, L.P. On-Line Characterization of Vehicle Emissions by FT-IR and Mass Spectrometry; SAE Technical Paper 810429: Warrendale, PA, USA, 1981. [Google Scholar] [CrossRef]
  22. Staab, J.; Klingenberg, H.; Schürmann, D. Strategy for the Development of a New Multicomponent Exhaust Emissions Measurement Technique; SAE Technical Paper 830437: Warrendale, PA, USA, 1983. [Google Scholar] [CrossRef]
  23. Staab, J.; Klingenberg, H.; Pflüger, H.; Herget, W.F.; Tromp, M.I. First Experiences in Testing a New Multicomponent Exhaust Gas Sampling and Analyzing System; SAE Technical Paper 851659: Warrendale, PA, USA, 1985. [Google Scholar] [CrossRef]
  24. Haack, L.P.; LaCourse, D.L.; Korniski, T.J. Comparison of Fourier Transform Infrared Spectrometry and 2,4-Dinitrophenylhydrazine-Impinger Techniques for the Measurement of Formaldehyde in Vehicle Exhaust. Anal. Chem. 1986, 58, 68–72. [Google Scholar] [CrossRef]
  25. Adachi, M.; Yamagishi, Y.; Inoue, K.; Ishida, K. Automotive Emission Analyses Using FTIR Spectrophotometer; SAE Technical Paper 920723: Warrendale, PA, USA, 1992. [Google Scholar] [CrossRef]
  26. Zweidinger, R.A.; Crews, W.S.; Braddock, J.N. Comparison of Three Analytical Methods for the Determination of Methanol in Vehicle Evaporative and Exhaust Samples; SAE Technical Paper 930377: Warrendale, PA, USA, 1993. [Google Scholar] [CrossRef]
  27. Laurikko, J.; Aakko, P. The Effect of Ambient Temperature on the Emissions of Some Nitrogen Compounds:A Comparative Study on Low-, Medium- and High-Mileage Three-Way Catalyst Vehicles; SAE Technical Paper 950933: Warrendale, PA, USA, 1995. [Google Scholar] [CrossRef]
  28. Bianchi, D.; Gass, J.L.; Bouly, C.; Maret, D. Determination of Efficiency of Exhaust Gas Catalyst by FTIR Spectroscopy; SAE Technical Paper 910839: Warrendale, PA, USA, 1991. [Google Scholar] [CrossRef]
  29. Laurikko, J.; Nylund, N.-O. Regulated and Unregulated Emissions from Catalyst Vehicles at Low Ambient Temperatures; SAE Technical Paper 930946: Warrendale, PA, USA, 1993. [Google Scholar] [CrossRef]
  30. Lee, G.R.; McDonald, C.R.; Widdicombe, K.A.; van den Brink, P.J. The Role of Methane on Catalyst Conversion of NOx: Study Based on FTIR; SAE Technical Paper 961155: Warrendale, PA, USA, 1996. [Google Scholar] [CrossRef]
  31. Adachi, M. Emission Measurement Techniques for Advanced Powertrains. Meas. Sci. Technol. 2000, 11, R113–R129. [Google Scholar] [CrossRef]
  32. Babu, P.R.; Nagalingam, B.; Mehta, P.S. Hydrocarbon Modeling for Two-Stroke SI Engine; SAE Technical Paper 940403: Warrendale, PA, USA, 1994. [Google Scholar] [CrossRef]
  33. Ambs, J.L.; McClure, B.T. The Influence of Oxidation Catalysts on NO2 in Diesel Exhaust; SAE Technical Paper 932494: Warrendale, PA, USA, 1993. [Google Scholar] [CrossRef]
  34. Arai, M. Continuous Analysis of Vehicle Exhaust Gas by Fourier-Transform Infrared Spectroscopy. Anal. Sci. 1993, 9, 77–82. [Google Scholar] [CrossRef] [Green Version]
  35. Soltani, S.; Andersson, R.; Andersson, B. The Effect of Exhaust Gas Composition on the Kinetics of Soot Oxidation and Diesel Particulate Filter Regeneration. Fuel 2018, 220, 453–463. [Google Scholar] [CrossRef]
  36. Stanciulescu, M.; Charland, J.-P.; Kelly, J.F. Effect of Primary Amine Hydrocarbon Chain Length for the Selective Catalytic Reduction of NOx from Diesel Engine Exhaust. Fuel 2010, 89, 2292–2298. [Google Scholar] [CrossRef]
  37. Zhao, W.; Liu, Y.; Wei, H.; Zhang, R.; Luo, G.; Hou, H.; Chen, S.; Zhang, R. NO Removal by Plasma-Enhanced NH3-SCR Using Methane as an Assistant Reduction Agent at Low Temperature. Appl. Sci. 2019, 9, 2751. [Google Scholar] [CrossRef] [Green Version]
  38. Davies, C.; Thompson, K.; Cooper, A.; Golunski, S.; Taylor, S.H.; Bogarra Macias, M.; Doustdar, O.; Tsolakis, A. Simultaneous Removal of NOx and Soot Particulate from Diesel Exhaust by In-Situ Catalytic Generation and Utilisation of N2O. Appl. Catal. B Environ. 2018, 239, 10–15. [Google Scholar] [CrossRef]
  39. Wang, C.; Zhang, C.; Zhao, Y.; Yan, X.; Cao, P. Poisoning Effect of SO2 on Honeycomb Cordierite-Based Mn–Ce/Al2O3 Catalysts for NO Reduction with NH3 at Low Temperature. Appl. Sci. 2018, 8, 95. [Google Scholar] [CrossRef] [Green Version]
  40. Auvinen, P.; Nevalainen, P.; Suvanto, M.; Oliva, F.; Llamas, X.; Barciela, B.; Sippula, O.; Kinnunen, N.M. A Detailed Study on Regeneration of SO2 Poisoned Exhaust Gas after-Treatment Catalysts: In Pursuance of High Durability and Low Methane, NH3 and N2O Emissions of Heavy-Duty Vehicles. Fuel 2021, 291, 120223. [Google Scholar] [CrossRef]
  41. de Melo, T.C.C.; Machado, G.B.; Belchior, C.R.; Colaço, M.J.; Barros, J.E.; de Oliveira, E.J.; de Oliveira, D.G. Hydrous Ethanol–Gasoline Blends—Combustion and Emission Investigations on a Flex-Fuel Engine. Fuel 2012, 97, 796–804. [Google Scholar] [CrossRef]
  42. Gailis, M.; Pirs, V.; Jansons, M.; Birzietis, G.; Dukulis, I. An Experimental Investigation on Aldehyde and Methane Emissions from Hydrous Ethanol and Gasoline Fueled SI Engine; SAE Technical Paper 2020-01-2047: Warrendale, PA, USA, 2020. [Google Scholar] [CrossRef]
  43. Liu, S.; Chen, W.; Zhu, Z.; Jiang, S.; Ren, T.; Guo, H. A Review of the Developed New Model Biodiesels and Their Effects on Engine Combustion and Emissions. Appl. Sci. 2018, 8, 2303. [Google Scholar] [CrossRef] [Green Version]
  44. Tan, P.; Hu, Z.; Lou, D.; Li, Z. Exhaust Emissions from a Light-Duty Diesel Engine with Jatropha Biodiesel Fuel. Energy 2012, 39, 356–362. [Google Scholar] [CrossRef]
  45. Yu-sheng, Z.; Chun-lan, M.; Hai-ying, S.; Shao-ren, Z. Study on Formaldehyde Emission in a DME-Fueled Direct-Injection Diesel Engine; SAE Technical Paper 2007-01-1909: Warrendale, PA, USA, 2007. [Google Scholar] [CrossRef]
  46. Hunicz, J.; Matijošius, J.; Rimkus, A.; Kilikevičius, A.; Kordos, P.; Mikulski, M. Efficient Hydrotreated Vegetable Oil Combustion under Partially Premixed Conditions with Heavy Exhaust Gas Recirculation. Fuel 2020, 268, 117350. [Google Scholar] [CrossRef]
  47. Lemel, M.; Hultqvist, A.; Vressner, A.; Nordgren, H.; Persson, H.; Johansson, B. Quantification of the Formaldehyde Emissions from Different HCCI Engines Running on a Range of Fuels; SAE Technical Paper 2005-01-3724: Warrendale, PA, USA, 2005. [Google Scholar] [CrossRef] [Green Version]
  48. Jiang, C.; Huang, G.; Liu, G.; Qian, Y.; Lu, X. Optimizing Gasoline Compression Ignition Engine Performance and Emissions: Combined Effects of Exhaust Gas Recirculation and Fuel Octane Number. Appl. Therm. Eng. 2019, 153, 669–677. [Google Scholar] [CrossRef]
  49. Wu, Y.; Wang, P.; Farhan, S.M.; Yi, J.; Lei, L. Effect of Post-Injection on Combustion and Exhaust Emissions in DI Diesel Engine. Fuel 2019, 258, 116131. [Google Scholar] [CrossRef]
  50. Brosha, E.L.; Prikhodko, V.Y.; Kreller, C.R.; Pihl, J.A.; Curran, S.; Parks, J.E.; Mukundan, R. Response Characteristics of Stable Mixed-Potential NH3 Sensors in Diesel Engine Exhaust. Emiss. Control Sci. Technol. 2017, 3, 112–121. [Google Scholar] [CrossRef]
  51. Zarante, P.H.B.; Sodré, J.R. Comparison of Aldehyde Emissions Simulation with FTIR Measurements in the Exhaust of a Spark Ignition Engine Fueled by Ethanol. Heat Mass Transf. 2018, 54, 2079–2087. [Google Scholar] [CrossRef] [Green Version]
  52. Zardini, A.A.; Suarez-Bertoa, R.; Forni, F.; Montigny, F.; Otura-Garcia, M.; Carriero, M.; Astorga, C. Reducing the Exhaust Emissions of Unregulated Pollutants from Small Gasoline Engines with Alkylate Fuel and Low-Ash Lube Oil. Environ. Res. 2019, 170, 203–214. [Google Scholar] [CrossRef]
  53. Yamada, H.; Misawa, K.; Suzuki, D.; Tanaka, K.; Matsumoto, J.; Fujii, M.; Tanaka, K. Detailed Analysis of Diesel Vehicle Exhaust Emissions: Nitrogen Oxides, Hydrocarbons and Particulate Size Distributions. Proc. Combust. Inst. 2011, 33, 2895–2902. [Google Scholar] [CrossRef]
  54. Clairotte, M.; Suarez-Bertoa, R.; Zardini, A.A.; Giechaskiel, B.; Pavlovic, J.; Valverde, V.; Ciuffo, B.; Astorga, C. Exhaust Emission Factors of Greenhouse Gases (GHGs) from European Road Vehicles. Environ. Sci. Eur. 2020, 32, 125. [Google Scholar] [CrossRef]
  55. Nevius, T.; Rooney, R.; Zummer, R. Speciation of Nitrogen Oxides in a Light Duty Diesel Engine during an EGR System Failure; SAE Technical Paper 2012-01-0876: Warrendale, PA, USA, 2012. [Google Scholar] [CrossRef]
  56. Aakko-Saksa, P.; Koponen, P.; Roslund, P.; Laurikko, J.; Nylund, N.-O.; Karjalainen, P.; Rönkkö, T.; Timonen, H. Comprehensive Emission Characterisation of Exhaust from Alternative Fuelled Cars. Atmos. Environ. 2020, 236, 117643. [Google Scholar] [CrossRef]
  57. Suarez-Bertoa, R.; Kousoulidou, M.; Clairotte, M.; Giechaskiel, B.; Nuottimäki, J.; Sarjovaara, T.; Lonza, L. Impact of HVO Blends on Modern Diesel Passenger Cars Emissions during Real World Operation. Fuel 2019, 235, 1427–1435. [Google Scholar] [CrossRef]
  58. Bielaczyc, P.; Szczotka, A.; Woodburn, J. An Overview of Emissions of Reactive Nitrogen Compounds from Modern Light Duty Vehicles Featuring SI Engines. Combust. Engines 2014, 159, 48–53. [Google Scholar] [CrossRef]
  59. Andersson, J.; May, J.; Favre, C.; Bosteels, D.; de Vries, S.; Heaney, M.; Keenan, M.; Mansell, J. On-Road and Chassis Dynamometer Evaluations of Emissions from Two Euro 6 Diesel Vehicles. SAE Int. J. Fuels Lubr. 2014, 7, 919–934. [Google Scholar] [CrossRef] [Green Version]
  60. Suarez-Bertoa, R.; Astorga, C. Impact of Cold Temperature on Euro 6 Passenger Car Emissions. Environ. Pollut. 2018, 234, 318–329. [Google Scholar] [CrossRef] [PubMed]
  61. Giechaskiel, B.; Suarez-Bertoa, R.; Lähde, T.; Clairotte, M.; Carriero, M.; Bonnel, P.; Maggiore, M. Evaluation of NOx Emissions of a Retrofitted Euro 5 Passenger Car for the Horizon Prize “Engine Retrofit”. Environ. Res. 2018, 166, 298–309. [Google Scholar] [CrossRef]
  62. Giechaskiel, B.; Suarez-Bertoa, R.; Lahde, T.; Clairotte, M.; Carriero, M.; Bonnel, P.; Maggiore, M. Emissions of a Euro 6b Diesel Passenger Car Retrofitted with a Solid Ammonia Reduction System. Atmosphere 2019, 10, 180. [Google Scholar] [CrossRef] [Green Version]
  63. Hirschberger, R. Automotive Emissions Analysis with Spectroscopic Techniques. In Encyclopedia of Analytical Chemistry: Applications, Theory and Instrumentation, 1st ed.; Meyers, R., Ed.; Wiley: New York, NY, USA, 2006; ISBN 978-0-471-97670-7. [Google Scholar]
  64. Aakko-Saksa, P.; Rantanen-Kolehmainen, L.; Koponen, P.; Engman, A.; Kihlman, J. Biogasoline Options—Possibilities for Achieving High Bio-Share and Compatibility with Conventional Cars. SAE Int. J. Fuels Lubr. 2011, 4, 298–317. [Google Scholar] [CrossRef]
  65. Aakko-Saksa, P.T.; Rantanen-Kolehmainen, L.; Skyttä, E. Ethanol, Isobutanol, and Biohydrocarbons as Gasoline Components in Relation to Gaseous Emissions and Particulate Matter. Environ. Sci. Technol. 2014, 48, 10489–10496. [Google Scholar] [CrossRef]
  66. Suarez-Bertoa, R.; Zardini, A.A.; Keuken, H.; Astorga, C. Impact of Ethanol Containing Gasoline Blends on Emissions from a Flex-Fuel Vehicle Tested over the Worldwide Harmonized Light Duty Test Cycle (WLTC). Fuel 2015, 143, 173–182. [Google Scholar] [CrossRef]
  67. Suarez-Bertoa, R.; Lähde, T.; Pavlovic, J.; Valverde, V.; Clairotte, M.; Giechaskiel, B. Laboratory and On-Road Evaluation of a GPF-Equipped Gasoline Vehicle. Catalysts 2019, 9, 678. [Google Scholar] [CrossRef] [Green Version]
  68. Mehsein, K.; Norsic, C.; Chaillou, C.; Nicolle, A. Minimizing Secondary Pollutant Formation through Identification of Most Influential Volatile Emissions in Gasoline Exhausts: Impact of the Vehicle Powertrain Technology. Atmos. Environ. 2020, 226, 117394. [Google Scholar] [CrossRef]
  69. Durbin, T.D.; Wilson, R.D.; Norbeck, J.M.; Miller, J.W.; Huai, T.; Rhee, S.H. Estimates of the Emission Rates of Ammonia from Light-Duty Vehicles Using Standard Chassis Dynamometer Test Cycles. Atmos. Environ. 2002, 36, 1475–1482. [Google Scholar] [CrossRef] [Green Version]
  70. Woodburn, J.; Bielaczyc, P.; Szczotka, A. Chassis Dynamometer Testing of Ammonia Emissions from Light-Duty Si Vehicles in the Context of Emissions of Reactive Nitrogen Compounds; SAE Technical Paper 2013-01-1346: Warrendale, PA, USA, 2013. [Google Scholar] [CrossRef]
  71. Suarez-Bertoa, R.; Zardini, A.A.; Astorga, C. Ammonia Exhaust Emissions from Spark Ignition Vehicles over the New European Driving Cycle. Atmos. Environ. 2014, 97, 43–53. [Google Scholar] [CrossRef]
  72. Wang, X.; Ge, Y.; Gong, H.; Yang, Z.; Tan, J.; Hao, L.; Su, S. Ammonia Emissions from China-6 Compliant Gasoline Vehicles Tested over the WLTC. Atmos. Environ. 2019, 199, 136–142. [Google Scholar] [CrossRef]
  73. Liu, Y.; Wang, H.; Li, N.; Tan, J.; Chen, D. Research on Ammonia Emissions from Three-Way Catalytic Converters Based on Small Sample Test and Vehicle Test. Sci. Total Environ. 2021, 795, 148926. [Google Scholar] [CrossRef]
  74. Bogarra, M.; Herreros, J.M.; Tsolakis, A.; York, A.P.E.; Millington, P.J. Study of Particulate Matter and Gaseous Emissions in Gasoline Direct Injection Engine Using On-Board Exhaust Gas Fuel Reforming. Appl. Energy 2016, 180, 245–255. [Google Scholar] [CrossRef] [Green Version]
  75. Clairotte, M.; Adam, T.W.; Zardini, A.A.; Manfredi, U.; Martini, G.; Krasenbrink, A.; Vicet, A.; Tournié, E.; Astorga, C. Effects of Low Temperature on the Cold Start Gaseous Emissions from Light Duty Vehicles Fuelled by Ethanol-Blended Gasoline. Appl. Energy 2013, 102, 44–54. [Google Scholar] [CrossRef]
  76. Suarez-Bertoa, R.; Astorga, C. Isocyanic Acid and Ammonia in Vehicle Emissions. Transp. Res. Part D Transp. Environ. 2016, 49, 259–270. [Google Scholar] [CrossRef]
  77. Huai, T.; Durbin, T.; Rhee, S.H.; Norbeck, J.M. Investigation of Emission Rates of Ammonia, Nitrous Oxide and Other Exhaust Compounds from Alternative Fuel Vehicles Using a Chassis Dynamometer. Int. J. Automot. Technol. 2003, 4, 9–19. [Google Scholar]
  78. Reyes, F.; Grutter, M.; Jazcilevich, A.; González-Oropeza, R. Tecnical Note: Analysis of Non-Regulated Vehicular Emissions by Extractive FTIR Spectrometry: Tests on a Hybrid Car in Mexico City. Atmos. Chem. Phys. 2006, 6, 5339–5346. [Google Scholar] [CrossRef] [Green Version]
  79. Suarez-Bertoa, R.; Astorga, C. Unregulated Emissions from Light-Duty Hybrid Electric Vehicles. Atmos. Environ. 2016, 136, 134–143. [Google Scholar] [CrossRef]
  80. Suarez-Bertoa, R.; Pavlovic, J.; Trentadue, G.; Otura-Garcia, M.; Tansini, A.; Ciuffo, B.; Astorga, C. Effect of Low Ambient Temperature on Emissions and Electric Range of Plug-in Hybrid Electric Vehicles. ACS Omega 2019, 4, 3159–3168. [Google Scholar] [CrossRef]
  81. Daemme, L.C.; Penteado, R.d.A.; Furlan, C.; Errera, M.; Zotin, F.M.Z. An Investigation on Aldehyde and Ammonia Emissions from a 4-Stroke Gasoline-Fueled Motorcycle. Ammonia Emission Reduction by Using a SCR Catalyst; SAE Technical Paper 2013-36-0181: Warrendale, PA, USA, 2013. [Google Scholar] [CrossRef]
  82. Daemme, L.C.; Penteado, R.; Zotin, F.; Errera, M. Regulated and Unregulated Emissions from a Flex Fuel Motorcycle Fuelled with Various Gasoline/Ethanol Blends; SAE Technical Paper 2014-32-0032: Warrendale, PA, USA, 2014. [Google Scholar] [CrossRef]
  83. Daemme, L.C.; Penteado, R.; Zotin, F.; Corrêa, S.M.; Errera, M.R.; Forcetto, A. The Effect of Fuel Sulfur Content on Ammonia, Aldehyde and Regulated Emissions Emitted from a Euro Iii Motorcycle; SAE Technical Paper 2016-36-0158: Warrendale, PA, USA, 2016. [Google Scholar] [CrossRef]
  84. Clairotte, M.; Adam, T.W.; Chirico, R.; Giechaskiel, B.; Manfredi, U.; Elsasser, M.; Sklorz, M.; DeCarlo, P.F.; Heringa, M.F.; Zimmermann, R.; et al. Online Characterization of Regulated and Unregulated Gaseous and Particulate Exhaust Emissions from Two-Stroke Mopeds: A Chemometric Approach. Anal. Chim. Acta 2012, 717, 28–38. [Google Scholar] [CrossRef]
  85. Zardini, A.A.; Platt, S.M.; Clairotte, M.; El Haddad, I.; Temime-Roussel, B.; Marchand, N.; Ježek, I.; Drinovec, L.; Močnik, G.; Slowik, J.G.; et al. Effects of Alkylate Fuel on Exhaust Emissions and Secondary Aerosol Formation of a 2-Stroke and a 4-Stroke Scooter. Atmos. Environ. 2014, 94, 307–315. [Google Scholar] [CrossRef]
  86. Jetter, J.; Maeshiro, S.; Hatcho, S.; Klebba, R. Development of an On-Board Analyzer for Use on Advanced Low Emission Vehicles; SAE Technical Paper 2000-01-1140: Warrendale, PA, USA, 2000. [Google Scholar] [CrossRef]
  87. Li, H.; Andrews, G.E.; Khan, A.A.; Savvidis, D.; Daham, B.; Bell, M.; Tate, J.; Ropkins, K. Analysis of Driving Parameters and Emissions for Real World Urban Driving Cycles Using an On-Board Measurement Method for a Euro 2 SI Car; SAE Technical Paper 2007-01-2066: Warrendale, PA, USA, 2007. [Google Scholar] [CrossRef]
  88. Li, H.; Andrews, G.E.; Savvidis, D.; Daham, B.; Ropkins, K.; Bell, M.; Tate, J. Comparisons of the Exhaust Emissions for Different Generations of SI Cars under Real World Urban Driving Conditions; SAE Technical Paper 2008-01-0754: Warrendale, PA, USA, 2008. [Google Scholar] [CrossRef]
  89. Khalfan, A.; Andrews, G.; Li, H. Real World Driving: Emissions in Highly Congested Traffic; SAE Technical Paper 2017-01-2388: Warrendale, PA, USA, 2017. [Google Scholar] [CrossRef]
  90. Li, H.; Khalfan, A.; Andrews, G. Determination of GHG Emissions, Fuel Consumption and Thermal Efficiency for Real World Urban Driving Using a SI Probe Car. SAE Int. J. Engines 2014, 7, 1370–1381. [Google Scholar] [CrossRef]
  91. Khalfan, A.; Li, H.; Andrews, G. Speciation of Nitrogen Compounds in the Tailpipe Emissions from a SI Car under Real World Driving Conditions. SAE Int. J. Engines 2014, 7, 1961–1983. [Google Scholar] [CrossRef] [Green Version]
  92. Collins, J.F.; Shepherd, P.; Durbin, T.D.; Lents, J.; Norbeck, J.; Barth, M. Measurements of In-Use Emissions from Modern Vehicles Using an on-Board Measurement System. Environ. Sci. Technol. 2007, 41, 6554–6561. [Google Scholar] [CrossRef]
  93. Li, H.; Andrews, G.E.; Savvidis, D.; Daham, B.; Ropkins, K.; Bell, M.; Tate, J.E. Characterization of Regulated and Unregulated Cold Start Emissions for Different Real World Urban Driving Cycles Using a SI Passenger Car; SAE Technical Paper 2008-01-1648: Warrendale, PA, USA, 2008. [Google Scholar] [CrossRef]
  94. Li, H.; Andrews, G.E.; Savvidis, D.; Ropkins, K.; Tate, J.; Bell, M. Investigation of Regulated and Non-Regulated Cold Start Emissions Using a Euro 3 SI Car as a Probe Vehicle under Real World Urban Driving Conditions; SAE Technical Paper 2008-01-2428: Warrendale, PA, USA, 2008. [Google Scholar] [CrossRef]
  95. Sentoff, K.M.; Robinson, M.K.; Holmén, B.A. Second-by-Second Characterization of Cold-Start Gas-Phase and Air Toxic Emissions from a Light-Duty Vehicle. Transp. Res. Rec. 2010, 2158, 95–104. [Google Scholar] [CrossRef]
  96. Khalfan, A.; Li, H.; Andrews, G. Cold Start SI Passenger Car Emissions from Real World Urban Congested Traffic; SAE Technical Paper 2015-01-1064: Warrendale, PA, USA, 2015. [Google Scholar] [CrossRef]
  97. Li, H.; Andrews, G.E.; Savvidis, D. Influence of Cold Start and Ambient Temperatures on Greenhouse Gas (GHG) Emissions, Global Warming Potential (GWP) and Fuel Economy for SI Car Real World Driving. SAE Int. J. Fuels Lubr. 2010, 3, 133–148. [Google Scholar] [CrossRef]
  98. Vojtíšek-Lom, M.; Beránek, V.; Klír, V.; Jindra, P.; Pechout, M.; Voříšek, T. On-Road and Laboratory Emissions of NO, NO2, NH3, N2O and CH4 from Late-Model EU Light Utility Vehicles: Comparison of Diesel and CNG. Sci. Total Environ. 2018, 616–617, 774–784. [Google Scholar] [CrossRef]
  99. Hadavi, S.A.; Li, H.; Przybyla, G.; Jarrett, R.; Andrews, G. Comparison of Gaseous Emissions for B100 and Diesel Fuels for Real World Urban and Extra Urban Driving. SAE Int. J. Fuels Lubr. 2012, 5, 1132–1154. [Google Scholar] [CrossRef]
  100. Hadavi, S.; Andrews, G.E.; Li, H.; Przybyla, G.; Vazirian, M. Diesel Cold Start into Congested Real World Traffic: Comparison of Diesel and B100 for Ozone Forming Potential; SAE Technical Paper 2013-01-1145: Warrendale, PA, USA, 2013. [Google Scholar] [CrossRef]
  101. Przybyla, G.; Hadavi, S.; Li, H.; Andrews, G.E. Real World Diesel Engine Greenhouse Gas Emissions for Diesel Fuel and B100; SAE Technical Paper 2013-01-1514: Warrendale, PA, USA, 2013. [Google Scholar] [CrossRef]
  102. Suarez-Bertoa, R.; Pechout, M.; Vojtíšek, M.; Astorga, C. Regulated and Non-Regulated Emissions from Euro 6 Diesel, Gasoline and CNG Vehicles under Real-World Driving Conditions. Atmosphere 2020, 11, 204. [Google Scholar] [CrossRef] [Green Version]
  103. Samaras, Z.; Andersson, J.; Aakko-Saksa, P.; Cuelenaere, R.; Mellios, G. Additional Technical Issues for Euro 7 (LDV). In Proceedings of the AGVES Webinar, Brussels, Belgium, 27 April 2021. [Google Scholar]
  104. Vojtíšek, M.; Pechout, M. Portable, on-Board FTIR Spectrometers: A Universal Tool for Real-World Monitoring of Greenhouse Gases, Reactive Nitrogen Compounds, and Other Gaseous Pollutants? In Proceedings of the 24th International Transport and Air Pollution Conference Webinar, Graz, Austria, 30 March–1 April 2021. [Google Scholar]
  105. Ligterink, N.; Vojtíšek, M.; Stelwagen, U.; von Goethem, S. On-Road Emission Measurements beyond Type Approval PEMS. In Proceedings of the 24th International Transport and Air Pollution Conference Webinar, Graz, Austria, 30 March–1 April 2021. [Google Scholar]
  106. Giechaskiel, B.; Joshi, A.; Ntziachristos, L.; Dilara, P. European Regulatory Framework and Particulate Matter Emissions of Gasoline Light-Duty Vehicles: A Review. Catalysts 2019, 9, 586. [Google Scholar] [CrossRef] [Green Version]
  107. Giechaskiel, B.; Clairotte, M.; Valverde-Morales, V.; Bonnel, P.; Kregar, Z.; Franco, V.; Dilara, P. Framework for the Assessment of PEMS (Portable Emissions Measurement Systems) Uncertainty. Environ. Res. 2018, 166, 251–260. [Google Scholar] [CrossRef] [PubMed]
  108. Giechaskiel, B.; Bonnel, P.; Perujo, A.; Dilara, P. Solid Particle Number (SPN) Portable Emissions Measurement Systems (PEMS) in the European Legislation: A Review. Int. J. Environ. Res. Public Health 2019, 16, 4819. [Google Scholar] [CrossRef] [Green Version]
  109. DieselNet Emission Standards 2021. Available online: https://dieselnet.com/standards (accessed on 11 August 2021).
  110. Daham, B.; Andrews, G.E.; Li, H.; Ballesteros, R.; Bell, M.C.; Tate, J.; Ropkins, K. Application of a Portable FTIR for Measuring On-Road Emissions; SAE Technical Paper 2005-01-0676: Warrendale, PA, USA, 2005. [Google Scholar] [CrossRef] [Green Version]
  111. White, J.U. Long Optical Paths of Large Aperture. J. Opt. Soc. Am. 1942, 32, 285. [Google Scholar] [CrossRef]
  112. Kawarabayashi, S.; Yamagishi, Y.; Kachi, H.; Kobayashi, S. Measurement of Vehicle Emissions with Fourier Transform Infrared (FTIR) Method. Int. Symp. COMODIA 1990, 90, 365–369. [Google Scholar]
  113. Baronick, J.; Heller, B.; Lach, G.; Luf, H. Modal Measurement of Raw Exhaust Volume and Mass Emissions by SESAM; SAE Technical Paper 980047: Warrendale, PA, USA, 1998. [Google Scholar] [CrossRef]
  114. Adachi, M.; Inoue, K.; Yamagishi, Y.; Ishida, K. Discussion of Operating Parameters and Analysis Capability for a Fourier Transform Infrared Emission Analyzer; SAE Technical Paper 971018: Warrendale, PA, USA, 1997. [Google Scholar] [CrossRef]
  115. Herget, W.F.; Staab, J.; Klingenberg, H.; Riedel, W.J. Progress in the Prototype Development of a New Multicomponent Exhaust Gas Sampling and Analyzing System; SAE Technical Paper 840470: Warrendale, PA, USA, 1984. [Google Scholar] [CrossRef]
  116. Rothman, L.S.; Gordon, I.E. The HITRAN molecular database. In Proceedings of the AIP Conference Proceedings, Gaithersburg, MD, USA, 30 September–4 October 2012. [Google Scholar] [CrossRef] [Green Version]
  117. Gordon, I.E.; Rothman, L.S.; Hill, C.; Kochanov, R.V.; Tan, Y.; Bernath, P.F.; Birk, M.; Boudon, V.; Campargue, A.; Chance, K.V.; et al. The HITRAN2016 Molecular Spectroscopic Database. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 3–69. [Google Scholar] [CrossRef]
  118. Heland, J.; Schäfer, K. Determination of Major Combustion Products in Aircraft Exhausts by FTIR Emission Spectroscopy. Atmos. Environ. 1998, 32, 3067–3072. [Google Scholar] [CrossRef]
  119. Zhao, Y.; Pan, Y.; Rutherford, J.; Mitloehner, F.M. Estimation of the Interference in Multi-Gas Measurements Using Infrared Photoacoustic Analyzers. Atmosphere 2012, 3, 246–265. [Google Scholar] [CrossRef] [Green Version]
  120. Hoard, J.; Snow, R.; Xu, L.; Gierczak, C.; Hammerle, R.; Montreuil, C.; Farooq, S.I. NOx Measurement Errors in Ammonia-Containing Exhaust; SAE Technical Paper 2007-01-0330: Warrendale, PA, USA, 2007. [Google Scholar] [CrossRef]
  121. Keens, A.; Simon, A. Correction of the Nonlinear Response of Infrared Detectors. In Proceedings of the 9th International Conference on Fourier Transform Spectroscopy, Calgary, AB, Canada, 23–27 August 1993; Proc. SPIE 2089. pp. 222–223. [Google Scholar] [CrossRef]
  122. Russwurm, G.M.; Phillips, W. Uncertainties in FTIR Data Due to a Nonlinear Response. In Proceedings of the Photonics East (ISAM, VVDC, IEMB), Boston, MA, USA, 1–6 November 1998; Proc. SPIE 3534, Environmental Monioting and Remediation Technologies. pp. 176–186. [Google Scholar] [CrossRef]
  123. Wallner, T.; Frazee, R. Study of Regulated and Non-Regulated Emissions from Combustion of Gasoline, Alcohol Fuels and Their Blends in a DI-SI Engine; SAE Technical Paper 2010-01-1571: Warrendale, PA, USA, 2010. [Google Scholar] [CrossRef]
  124. Gautam, R.; Vanga, S.; Ariese, F.; Umapathy, S. Review of Multidimensional Data Processing Approaches for Raman and Infrared Spectroscopy. EPJ Tech. Instrum. 2015, 2, 1–38. [Google Scholar] [CrossRef] [Green Version]
  125. Bacsik, Z.; Mink, J.; Keresztury, G. FTIR Spectroscopy of the Atmosphere. I. Principles and Methods. Appl. Spectrosc. Rev. 2004, 39, 295–363. [Google Scholar] [CrossRef]
  126. Gierczak, C.A.; Andino, J.M.; Butler, J.W.; Heiser, G.A.; Jesion, G.; Korniski, T.J. FTIR: Fundamentals and Applications in the Analysis of Dilute Vehicle Exhaust. In Proceedings of the Optics, Elctro-Optics, and Laser Applications in Science and Engineering, Los Angeles, CA, USA, 20–25 January 1991; Proc. SPIE 1433, Measurement of Atmospheric Gases. pp. 315–328. [Google Scholar] [CrossRef]
  127. Maricq, M. Exhaust emissions. In Encyclopedia of Automotive Engineering; Crolla, D., Foster, D.E., Kobayashi, T., Vaughan, N., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2014; pp. 1–14. ISBN 978-1-118-35417-9. [Google Scholar]
  128. Adachi, M.; Nakamura, H. Engine Emissions Measurement Handbook: HORIBA Automotive Test Systems; SAE International: Warrendale, PA, USA, 2014; ISBN 978-0-7680-8012-4. [Google Scholar]
  129. Majewski, W.A.; Khair, M.K. Diesel Emissions and Their Control; SAE International: Warrendale, PA, USA, 2006; ISBN 978-0-7680-0674-2. [Google Scholar]
  130. Kasab, J.J.; Strzelec, A. Automotive Emissions Regulations and Exhaust Aftertreatment Systems; SAE International: Warrendale, PA, USA, 2020; ISBN 978-0-7680-9955-3. [Google Scholar]
  131. Sun, Y.-W.; Liu, W.-Q.; Zeng, Y.; Wang, S.-M.; Huang, S.-H.; Xie, P.-H.; Yu, X.-M. Water Vapor Interference Correction in a Non Dispersive Infrared Multi-Gas Analyzer. Chin. Phys. Lett. 2011, 28, 073302. [Google Scholar] [CrossRef]
  132. Inoue, K.; Ishihara, M.; Akashi, K.; Adachi, M.; Ishida, K. Numerical Analysis of Mass Emission Measurement Systems for Low Emission Vehicles; SAE Technical Paper 1999-01-0150: Warrendale, PA, USA, 1999. [Google Scholar] [CrossRef]
  133. Nakamura, H.; Kihara, N.; Adachi, M.; Ishida, K. Development of a Wet-Based NDIR and Its Application to on-Board Emission Measurement System; SAE Technical Paper 2002-01-0612: Warrendale, PA, USA, 2002. [Google Scholar] [CrossRef]
  134. Dinh, T.-V.; Choi, I.-Y.; Son, Y.-S.; Kim, J.-C. A Review on Non-Dispersive Infrared Gas Sensors: Improvement of Sensor Detection Limit and Interference Correction. Sens. Actuators B Chem. 2016, 231, 529–538. [Google Scholar] [CrossRef]
  135. Giechaskiel, B.; Zardini, A.A.; Clairotte, M. Exhaust Gas Condensation during Engine Cold Start and Application of the Dry-Wet Correction Factor. Appl. Sci. 2019, 9, 2263. [Google Scholar] [CrossRef] [Green Version]
  136. Gluck, S.; Glenn, C.; Logan, T.; Vu, B.; Walsh, M.; Williams, P. Evaluation of NOx Flue Gas Analyzers for Accuracy and Their Applicability for Low-Concentration Measurements. J. Air Waste Manag. Assoc. 2003, 53, 749–758. [Google Scholar] [CrossRef]
  137. Akard, M.; Nakamura, H.; Aoki, S.; Kihara, N.; Adachi, M. Performance Results and Design Considerations for a New In-Use Testing Instrument; SAE Technical Paper 2005-01-3606: Warrendale, PA, USA, 2005. [Google Scholar] [CrossRef]
  138. Shah, S.D.; Mauti, A.; Richert, J.F.O.; Loos, M.J.; Chase, R.E. Measuring NOx in the Presence of Ammonia; SAE Technical Paper 2007-01-0331: Warrendale, PA, USA, 2007. [Google Scholar] [CrossRef]
  139. Baronick, J.D.; Heller, B.; Lach, G.; Bogner, F.; Gerspach, U.; Schimpl, H.; Gruber, D.; Fabinski, W.; Moede, M.; Zöchbauer, M. Evaluation of an UV Analyzer for NOx Vehicle Emission Measurement; SAE Technical Paper 2001-01-0213: Warrendale, PA, USA, 2001. [Google Scholar] [CrossRef]
  140. Czerwinski, J.; Comte, P.; Güdel, M.; Mayer, A.; Lemaire, J.; Reutimann, F.; Berger, A.H. Investigations of NO2 in Legal Test Procedure for Diesel Passenger Cars; SAE Technical Paper 2015-24-2510: Warrendale, PA, USA, 2015. [Google Scholar] [CrossRef]
  141. Arlitt, B. Korrelationen NOx: NO, NO2 Und NOx Analyse Mit FTIR Und CLD; AVL List GmbH: Graz, Austria, 2018. [Google Scholar]
  142. Olsen, D.B.; Kohls, M.; Arney, G. Impact of Oxidation Catalysts on Exhaust NO2/NOx Ratio from Lean-Burn Natural Gas Engines. J. Air Waste Manag. Assoc. 2010, 60, 867–874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Meinel, H.; Just, T. Measurement of NOx Exhaust Emissions by a New NDUV Analyzer. In Proceedings of the 14th Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, Washington, DC, USA, 26 January 1976. [Google Scholar]
  144. Cao, T.; Durbin, T.D.; Cocker, D.R.; Wanker, R.; Schimpl, T.; Pointner, V.; Oberguggenberger, K.; Johnson, K.C. A Comprehensive Evaluation of a Gaseous Portable Emissions Measurement System with a Mobile Reference Laboratory. Emiss. Control Sci. Technol. 2016, 2, 173–180. [Google Scholar] [CrossRef] [Green Version]
  145. Varella, R.; Giechaskiel, B.; Sousa, L.; Duarte, G. Comparison of Portable Emissions Measurement Systems (PEMS) with Laboratory Grade Equipment. Appl. Sci. 2018, 8, 1633. [Google Scholar] [CrossRef] [Green Version]
  146. Ballantyne, V.F.; Howes, P.; Stephanson, L. Nitrous Oxide Emissions from Light Duty Vehicles; SAE Technical Paper 940304: Warrendale, PA, USA, 1994. [Google Scholar] [CrossRef]
  147. Rahman, M.; Hara, K.; Nakatani, S.; Tanaka, Y. Development of a Fast Response Nitrogen Compounds Analyzer Using Quantum Cascade Laser for Wide-Range Measurement; SAE Technical Paper 2011-26-0044: Warrendale, PA, USA, 2011. [Google Scholar] [CrossRef]
  148. Hara, K.; Nakatani, S.; Rahman, M.; Nakamura, H.; Tanaka, Y.; Ukon, J. Development of Nitrogen Components Analyzer Utilizing Quantum Cascade Laser; SAE Technical Paper 2009-01-2743: Warrendale, PA, USA, 2009. [Google Scholar] [CrossRef]
  149. Murtonen, T.; Vesala, H.; Koponen, P.; Pettinen, R.; Kajolinna, T.; Antson, O. NH3 Sensor Measurements in Different Engine Applications; SAE Technical Paper 2018-01-1814: Warrendale, PA, USA, 2018. [Google Scholar] [CrossRef]
  150. Seykens, X.; van den Tillaart, E.; Lilova, V.; Nakatani, S. NH3 Measurements for Advanced SCR Applications; SAE Technical Paper 2016-01-0975: Warrendale, PA, USA, 2016. [Google Scholar] [CrossRef]
  151. Suarez-Bertoa, R.; Zardini, A.A.; Lilova, V.; Meyer, D.; Nakatani, S.; Hibel, F.; Ewers, J.; Clairotte, M.; Hill, L.; Astorga, C. Intercomparison of Real-Time Tailpipe Ammonia Measurements from Vehicles Tested over the New World-Harmonized Light-Duty Vehicle Test Cycle (WLTC). Environ. Sci. Pollut. Res. 2015, 22, 7450–7460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Mohn, J.; Forss, A.M.; Bruhlmann, S.; Zeyer, K.; Luscher, R.; Emmenegger, L.; Novak, P.; Heeb, N. Time-Resolved Ammonia Measurement in Vehicle Exhaust. Int. J. Environ. Pollut. 2004, 22, 342–356. [Google Scholar] [CrossRef]
  153. Pisano, J.T.; Sauer, C.; Durbin, T.; Mackay, G. Measurement of Low Concentration NH3 in Diesel Exhaust Using Tunable Diode Laser Adsorption Spectroscopy (TDLAS); SAE Technical Paper 2009-01-1519: Warrendale, PA, USA, 2009. [Google Scholar] [CrossRef]
  154. McWilliam, I.G.; Dewar, R.A. Flame Ionization Detector for Gas Chromatography. Nature 1958, 181, 760. [Google Scholar] [CrossRef]
  155. Cheng, W.K.; Summers, T.; Collings, N. The Fast-Response Flame Ionization Detector. Prog. Energy Combust. Sci. 1998, 24, 89–124. [Google Scholar] [CrossRef]
  156. Palo, D.R.; Dagle, R.A.; Holladay, J.D. Methanol Steam Reforming for Hydrogen Production. Chem. Rev. 2007, 107, 3992–4021. [Google Scholar] [CrossRef]
  157. Dietz, W.A. Response Factors for Gas Chromatographic Analyses. J. Chromatogr. Sci. 1967, 5, 68–71. [Google Scholar] [CrossRef]
  158. Jorgensen, A.D.; Picel, K.C.; Stamoudis, V.C. Prediction of Gas Chromatography Flame Ionization Detector Response Factors from Molecular Structures. Anal. Chem. 1990, 62, 683–689. [Google Scholar] [CrossRef]
  159. Wallner, T. Correlation between Speciated Hydrocarbon Emissions and Flame Ionization Detector Response for Gasoline/Alcohol Blends. J. Eng. Gas Turbines Power 2011, 133, 082801. [Google Scholar] [CrossRef]
  160. Sandstroem-Dahl, C.; Erlandsson, L.; Gasste, J.; Lindgren, M. Measurement Methodologies for Hydrocarbons, Ethanol and Aldehyde Emissions from Ethanol Fuelled Vehicles. SAE Int. J. Fuels Lubr. 2010, 3, 453–466. [Google Scholar] [CrossRef]
  161. Loo, J.F.; Parker, D.T. Evaluation of a Photoacoustic Gas Analyzer for Ethanol Vehicle Emissions Measurement; SAE Technical Paper 2000-01-0794: Warrendale, PA, USA, 2000. [Google Scholar] [CrossRef]
  162. Gierczak, C.A.; Kralik, L.L.; Mauti, A.; Harwell, A.L.; Maricq, M.M. Measuring NMHC and NMOG Emissions from Motor Vehicles via FTIR Spectroscopy. Atmos. Environ. 2017, 150, 425–433. [Google Scholar] [CrossRef]
  163. Li, H.; Ropkins, K.; Andrews, G.E.; Daham, B.; Bell, M.; Tate, J.; Hawley, G. Evaluation of a FTIR Emission Measurement System for Legislated Emissions Using a SI Car; SAE Technical Paper 2006-01-3368: Warrendale, PA, USA, 2006. [Google Scholar] [CrossRef] [Green Version]
  164. Butler, J.W.; Maker, P.D.; Korniski, T.J.; Haack, L.P.; McKelvy, F.E.; Colvin, A.D. A System for On-Line Measurement of Multicomponent Emissions and Engine Operating Parameters; SAE Technical Paper 851657: Warrendale, PA, USA, 1985. [Google Scholar] [CrossRef]
  165. Heller, B.; Klingenberg, H.; Lach, G.; Winckler, J. Performance of a New System for Emission Sampling and Measurement (SESAM); SAE Technical Paper 900275: Warrendale, PA, USA, 1990. [Google Scholar] [CrossRef]
  166. Shore, P.R.; deVries, R.S. On-Line Hydrocarbon Speciation Using FTIR and CI-MS; SAE Technical Paper 922246: Warrendale, PA, USA, 1992. [Google Scholar] [CrossRef]
  167. Wright, N.; Osborne, D.; Music, N. Comparison of Hydrocarbon Measurement with FTIR and FID in a Dual Fuel Locomotive Engine; SAE Technical Paper 2016-01-0978: Warrendale, PA, USA, 2016. [Google Scholar] [CrossRef]
  168. Suarez-Bertoa, R.; Clairotte, M.; Arlitt, B.; Nakatani, S.; Hill, L.; Winkler, K.; Kaarsberg, C.; Knauf, T.; Zijlmans, R.; Boertien, H.; et al. Intercomparison of Ethanol, Formaldehyde and Acetaldehyde Measurements from a Flex-Fuel Vehicle Exhaust during the WLTC. Fuel 2017, 203, 330–340. [Google Scholar] [CrossRef]
  169. Rieger, P.L.; Maddox, C.E. Analysis of Exhaust from Clean-Fuel Vehicles Using FTIR Spectroscopy. In Proceedings of the Optics, Electro-Optics, and Laser Applications in Science and Engineering, Los Angeles, CA, USA, 20–25 January 1991; Proc. SPIE 1433, Measurement of Atmopheric Gases. pp. 290–301. [Google Scholar] [CrossRef]
  170. McArver, A.Q. Comparison of Data from Two FTIR Based Automotive Emissions Analysis Systems with Those Obtained from Conventional Automotive Emissions Instrumentation; SAE Technical Paper 932660: Warrendale, PA, USA, 1993. [Google Scholar] [CrossRef]
  171. Roberts, J.P.; Lowry, S.R. Modal Measurements of Some Important NMOG Species by FT-IR Spectroscopy; SAE Technical Paper 940739: Warrendale, PA, USA, 1994. [Google Scholar] [CrossRef]
  172. Silva, K.C.C.; Daemme, L.C.; Macedo, V.d.C.; Penteado, R.d.A.; Ostapuik, I.F.; Corrêa, S.M. Measurement of Legislated Emissions, Unburned Alcohol and Potential Formation of Ozone from a Light Flex-Fuel Vehicle; SAE Technical Paper 2014-36-0260: Warrendale, PA, USA, 2014. [Google Scholar] [CrossRef]
  173. Oberdorfer, P.E. The Determination of Aldehydes in Automobile Exhaust Gas; SAE Technical Paper 670123: Warrendale, PA, USA, 1967. [Google Scholar] [CrossRef]
  174. Wagner, T.; Wyszyński, M.L. Aldehydes and Ketones in Engine Exhaust Emissions—A Review. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 1996, 210, 109–122. [Google Scholar] [CrossRef]
  175. Williams, J.; Li, H.; Ross, A.B.; Hargreaves, S.P. Quantification of the Influence of NO2, NO and CO Gases on the Determination of Formaldehyde and Acetaldehyde Using the DNPH Method as Applied to Polluted Environments. Atmos. Environ. 2019, 218, 117019. [Google Scholar] [CrossRef]
  176. Shiotani, H.; Goto, S.; Kinoshita, K.; Nikolic, D. Characteristics of Aldehydes and VOCs Emission from Off-Road Engines; SAE Technical Paper 2006-32-0023: Warrendale, PA, USA, 2006. [Google Scholar] [CrossRef]
  177. Shareef, G.S.; Ogle, L.; Ferry, K.; LaCosse, J.; Johnson, K.; McCarthy, J. Measurement of Aldehydes in Internal Combustion Engine Exhaust. In Proceedings of the SPE/EPA Exploration and Production Environmental Conference, Dallas, TX, USA, 3 March 1997; p. SPE-37878-MS. [Google Scholar]
  178. Truex, T.J.; Collins, J.F.; Jetter, J.J.; Knight, B.; Hayashi, T.; Kishi, N.; Suzuki, N. Measurement of Ambient Roadway and Vehicle Exhaust Emissions—An Assessment of Instrument Capability and Initial on-Road Test Results with an Advanced Low Emission Vehicle; SAE Technical Paper 2000-01-1142: Warrendale, PA, USA, 2000. [Google Scholar] [CrossRef]
  179. Zhang, F.; Wang, J.; Tian, D.; Wang, J.-X.; Shuai, S.-J. Research on Unregulated Emissions from an Alcohols-Gasoline Blend Vehicle Using FTIR, HPLC and GC-MS Measuring Methods. SAE Int. J. Engines 2013, 6, 1126–1137. [Google Scholar] [CrossRef]
  180. Aakko, P.; Lappi, M.; Laurikko, J.; Nylund, N.-O. Experiences in Analyzing the Exhaust Gas Emissions of Vehicles Using Fast, on-Line FTIR. In Proceedings of the Nordic Seminar on Gas Analysis in Combustion, Tampere, Finland, 4–5 October 1994. [Google Scholar]
  181. Liang, Y.; Tang, X.; Zhang, X.; Tian, F.; Sun, Y.; Dong, H. Portable Gas Analyzer Based on Fourier Transform Infrared Spectrometer for Patrolling and Examining Gas Exhaust. J. Spectrosc. 2015, 2015, 136516. [Google Scholar] [CrossRef] [Green Version]
  182. Dasrath, D.; Frazee, R.; Hwang, J.; Northrop, W. Comparison and Optimization of Fourier Transform Infrared Spectroscopy and Gas Chromatography-Mass Spectroscopy for Speciating Unburned Hydrocarbons from Diesel Low Temperature Combustion; SAE Technical Paper 2017-01-0992: Warrendale, PA, USA, 2017. [Google Scholar] [CrossRef]
  183. Lamas, J.E.; Hara, K.; Mori, Y. Optimization of Automotive Exhaust Sampling Parameters for Evaluation of after-Treatment Systems Using FTIR Exhaust Gas Analyzers; SAE Technical Paper 2019-01-0746: Warrendale, PA, USA, 2019. [Google Scholar] [CrossRef]
  184. Woodburn, J. Emissions of Reactive Nitrogen Compounds (RNCs) from Two Vehicles with Turbo-Charged Spark Ignition Engines over Cold Start Driving Cycles. Combust. Engines 2021. [Google Scholar] [CrossRef]
  185. Hoard, J.; Venkataramanan, N.; Marshik, B.; Murphy, W. NH3 Storage in Sample Lines; SAE Technical Paper 2014-01-1586: Warrendale, PA, USA, 2014. [Google Scholar] [CrossRef]
  186. Herget, W.F.; Brasher, J.D. Remote Fourier Transform Infrared Air Pollution Studies. Opt. Eng 1980, 19. [Google Scholar] [CrossRef]
  187. Balaji, P.S.; Rahman, M.E.; Moussa, L.; Lau, H.H. Wire Rope Isolators for Vibration Isolation of Equipment and Structures—A Review. IOP Conf. Ser. Mater. Sci. Eng. 2015, 78, 012001. [Google Scholar] [CrossRef] [Green Version]
  188. Köhler, M.; Schardt, M.; Rauscher, M.; Koch, A. Gas Measurement Using Static Fourier Transform Infrared Spectrometers. Sensors 2017, 17, 2612. [Google Scholar] [CrossRef] [Green Version]
  189. Valverde, V.; Giechaskiel, B.; Carriero, M. Real Driving Emissions: 2018–2019 Assessment of Portable Emissions Measurement Systems (PEMS) Measurement Uncertainty; Publications Office: Luxembourg, 2020. [Google Scholar]
  190. CEN/TS 17507:2021. Road Vehicles—Portable Emission Measuring Systems (PEMS)—Performance Assessment; BSI: London, UK, 2021. [Google Scholar]
  191. CEN/TS 17337:2019. Stationary Source Emissions. Determination of Mass Concentration of Multiple Gaseous Species. Fourier Transform Infrared Spectroscopy; BSI: London, UK, 2019; ISBN 978-0-580-51513-2. [Google Scholar]
  192. SAE J2992. FTIR Gas Analyzer Performance Evaluation/Qualification for Automotive Testing; SAE International: Warrendale, PA, USA, 2020. [Google Scholar]
  193. ASTM D6348 D22 Committee. Test Method for Determination of Gaseous Compounds by Extractive Direct Interface Fourier Transform Infrared (FTIR) Spectroscopy; ASTM International: West Conshohocken, PA, USA, 2020. [Google Scholar]
  194. EPA. Method 320: Measurement of Vapor Phase Organic and Inorganic Emissions by Extractive Fourier Transform Infrared (FTIR) Spectroscopy; EPA: Washington, DC, USA, 2020.
  195. Miers, S.A.; Green, C.; Meldrum, J.; Chmielewski, M. Measurement of Regulated and Unregulated Exhaust Emissions from Snowmobiles in the 2009 SAE Clean Snowmobile Challenge. SAE Int. J. Fuels Lubr. 2010, 3, 1112–1121. [Google Scholar] [CrossRef]
  196. Kumar, A.; Arlitt, B.; Jacksier, T. Effects of Isotopic Calibration Gases on IR Quantification Analyzer Techniques to Measure CO and CO2 in Engine Emissions Testing; SAE Technical Paper 2019-01-0076: Warrendale, PA, USA, 2019. [Google Scholar] [CrossRef]
  197. Aakko-Saksa, P.; Roslund, P.; Koponen, P. Development and Validation of Comprehensive Emission Measurement Methods for Alternative Fuels at VTT; VTT: Helsinki, Finland, 2017; p. 102. [Google Scholar]
  198. Valverde, V.; Giechaskiel, B. Assessment of Gaseous and Particulate Emissions of a Euro 6d-Temp Diesel Vehicle Driven >1300 Km Including Six Diesel Particulate Filter Regenerations. Atmosphere 2020, 11, 645. [Google Scholar] [CrossRef]
  199. Giechaskiel, B.; Valverde, V.; Kontses, A.; Suarez-Bertoa, R.; Selleri, T.; Melas, A.; Otura, M.; Ferrarese, C.; Martini, G.; Balazs, A.; et al. Effect of Extreme Temperatures and Driving Conditions on Gaseous Pollutants of a Euro 6d-Temp Gasoline Vehicle. Atmosphere 2021, 12, 1011. [Google Scholar] [CrossRef]
  200. Li, N.; El-Hamalawi, A.; Barrett, R.; Wheatley, A.; Robinson, J. Nitrogen Species Measurement Investigation Using Two Different FTIR. In Proceedings of the 18th IASTEM International Conference; International Academy of Science, Technology, Engnineering and Management, Berlin, Germany, 29 March 2016; pp. 24–29. [Google Scholar]
  201. Suarez-Bertoa, R.; Mendoza-Villafuerte, P.; Riccobono, F.; Vojtisek, M.; Pechout, M.; Perujo, A.; Astorga, C. On-Road Measurement of NH3 Emissions from Gasoline and Diesel Passenger Cars during Real World Driving Conditions. Atmos. Environ. 2017, 166, 488–497. [Google Scholar] [CrossRef]
  202. Livingston, C.; Rieger, P.; Winer, A. Ammonia Emissions from a Representative In-Use Fleet of Light and Medium-Duty Vehicles in the California South Coast Air Basin. Atmos. Environ. 2009, 43, 3326–3333. [Google Scholar] [CrossRef]
  203. Penteado, R.; Daemme, L.; Melo, T. An Experimental Investigation on Regulated and Unregulated Emissions from Four 4-Stroke Gasoline-Powered Motorcycle; SAE Technical Paper 2012-01-1098: Warrendale, PA, USA, 2012. [Google Scholar] [CrossRef]
  204. Li, N.; El-Hamalawi, A.; Barrett, R.; Baxter, J. Ammonia Measurement Investigation Using Quantum Cascade Laser and Two Different Fourier Transform Infrared Spectroscopy Methods; SAE Technical Paper 2020-01-0365: Warrendale, PA, USA, 2020. [Google Scholar] [CrossRef]
  205. Daemme, L.C.; Penteado, R.; Corrêa, S.M.; Zotin, F.; Errera, M.R. Emissions of Criteria and Non-Criteria Pollutants by a Flex-Fuel Motorcycle. J. Braz. Chem. Soc. 2016, 27, 2192–2202. [Google Scholar] [CrossRef]
Figure 1. Principle of operation of FTIR (Fourier transform infrared) spectroscopy. IR = infrared.
Figure 1. Principle of operation of FTIR (Fourier transform infrared) spectroscopy. IR = infrared.
Applsci 11 07416 g001
Figure 2. Example of absorbance versus wavenumbers: (a) for various species with concentrations of 398 ppm (CO), 97 ppm (NO), 136 ppm (NO2), 93 ppm (N2O), 93 ppm (NH3), 0.8% (CO2), 2.03% (H2O). For better visualization, the spectra of H2O and CO2 were reduced by a factor of 10 and displayed in an inverted axis. For the same reason, NH3 and HCN spectra were increased by a factor of 2 and 4, respectively; (b) NH3, area B of panel (a); (c) N2O, area C of panel (a).
Figure 2. Example of absorbance versus wavenumbers: (a) for various species with concentrations of 398 ppm (CO), 97 ppm (NO), 136 ppm (NO2), 93 ppm (N2O), 93 ppm (NH3), 0.8% (CO2), 2.03% (H2O). For better visualization, the spectra of H2O and CO2 were reduced by a factor of 10 and displayed in an inverted axis. For the same reason, NH3 and HCN spectra were increased by a factor of 2 and 4, respectively; (b) NH3, area B of panel (a); (c) N2O, area C of panel (a).
Applsci 11 07416 g002aApplsci 11 07416 g002b
Figure 3. Processing of an interferogram.
Figure 3. Processing of an interferogram.
Applsci 11 07416 g003
Figure 4. Overview of FTIR assessment studies. For each component, the mean deviations from the reference instruments were calculated based on the studies of Appendix A. “Cylinder” refers to calibration gases. “Parallel” means FTIR and reference instruments were measuring both from the dilution tunnel or the tailpipe. “TP vs. CVS” refers to cases where the FTIR was measuring from the tailpipe, while the reference instrument from the dilution tunnel. Error bars show one standard deviation of at least two studies. Numbers give the number of studies for the calculation of the means.
Figure 4. Overview of FTIR assessment studies. For each component, the mean deviations from the reference instruments were calculated based on the studies of Appendix A. “Cylinder” refers to calibration gases. “Parallel” means FTIR and reference instruments were measuring both from the dilution tunnel or the tailpipe. “TP vs. CVS” refers to cases where the FTIR was measuring from the tailpipe, while the reference instrument from the dilution tunnel. Error bars show one standard deviation of at least two studies. Numbers give the number of studies for the calculation of the means.
Applsci 11 07416 g004
Table 1. Pollutants, principle of measurement, and the sampling option for light-duty (LD) vehicles and heavy-duty (HD) engines in the European Union (EU) regulation and the global technical regulation (GTR 15) for LD vehicles. PEMS (portable emissions measurement systems) refer to the EU regulation for both LD and HD vehicles. Measurements can be conducted directly from the dilution tunnel with constant volume sampling (CVS), the bags, the proportional partial flow dilution system (PFDS), or the tailpipe (TP), depending on the regulation.
Table 1. Pollutants, principle of measurement, and the sampling option for light-duty (LD) vehicles and heavy-duty (HD) engines in the European Union (EU) regulation and the global technical regulation (GTR 15) for LD vehicles. PEMS (portable emissions measurement systems) refer to the EU regulation for both LD and HD vehicles. Measurements can be conducted directly from the dilution tunnel with constant volume sampling (CVS), the bags, the proportional partial flow dilution system (PFDS), or the tailpipe (TP), depending on the regulation.
CompoundMeasurement PrincipleGTR15 (LD)EU LDEU HDPEMS
PNVPR + PNCCVSCVS+PFDSTP
PMGravimetric (filter)CVSCVS+PFDS
CO2NDIRbagsbags+CVS, +TPTP
CONDIRbagsbags+CVS, +TPTP 3
THCFID or HFID (for diesel)bagsbags+CVS, +TPTP 3
CH4NMC or GC combined with FIDbagsbags+CVS, +TPTP 3
NMHCCalculated (THC–CH4)(bags)(bags)calculatedTP 3
NOxCLA or NDUVbagsbags+CVS, +TPTP
NOCLA or NDUV (only bags)CVS or bags
NO2NDUV, QCL (or CLA) or (NOx,bags–NOCVS)CVS
N2OGC with ECD, QCL-IR, NDIR, FTIR 1CVS or bags
NH3LDS or QCL or FTIR 2TP TP 4
C2H5OHImpinger + GC, FTIR, PAS, PTR-MS, dir. GCCVS
CH2ODNPH + HPLC, FTIR, PTR + MSCVS
CH3CHODNPH + HPLC, FTIR, PTR + MSCVS
1 interferences < 0.1 ppm; 2 with interference < 2 ppm at max CO2 and H2O; 3 only for heavy-duty vehicles; 4 LDS or FTIR for HD engines.C2H5OH = ethanol; CH3CHO = acetaldehyde; CH4 = methane; CLA = chemiluminescence analyzer; CO = carbon monoxide; CO2 = carbon dioxide; DNPH = 2,4-Dinitrophenylhydrazine; ECD = electron-capture detection; FID = flame ionization detection; FTIR = Fourier transform infrared spectrometry; GC = gas chromatography; CH2O = formaldehyde; HFID = heated flame ionization detection; HPLC = high-performance liquid chromatography; IR = infrared; LDS = laser diode spectrometer; N2O = nitrous oxide; NDIR = non-dispersive infrared spectrometry; NDUV = non-dispersive ultra-violet spectrometry; NH3 = ammonia; NMC = non-methane cutter; NMHC = non methane hydrocarbons; NO = nitrogen oxide; NO2 = nitrogen dioxide; NOx = nitrogen oxides; PAS = photoacoustic spectrometry; PM = particulate matter; PN = particle number; PNC = particle number counter; PTR-MS = proton transfer reaction—mass spectrometry; QCL = quantum cascade laser; THC = total hydrocarbons; VPR = volatile particle remover.
Table 2. Current Euro 6/VI limits in the laboratory (min from compression ignition and positive ignition engines) for light-duty (LD) vehicles and heavy-duty (HD) engines. The uncertainty of 1 ppm is also expressed in mass considering exhaust flow of 2 kg/m3 for LD and 10 kg/m3 (HD) large engine displacement vehicles.
Table 2. Current Euro 6/VI limits in the laboratory (min from compression ignition and positive ignition engines) for light-duty (LD) vehicles and heavy-duty (HD) engines. The uncertainty of 1 ppm is also expressed in mass considering exhaust flow of 2 kg/m3 for LD and 10 kg/m3 (HD) large engine displacement vehicles.
PollutantCONOxHCNH3N2O
Euro 6 limits (min of LD) [mg/km]50060100--
1 ppm uncertainty (LD) [mg/km]1.93.21.01.23.0
Euro VI limits (min of HD) [mg/kWh]400046066010 ppm-
1 ppm uncertainty (HD) [mg/kWh]9.715.94.85.915.2
Table 3. Example of technical requirements for FTIR for NH3 measurements in UNECE Reg. 49.
Table 3. Example of technical requirements for FTIR for NH3 measurements in UNECE Reg. 49.
SpecificationRequirement
Sampling lineStainless steel or PTFE, as short as possible, heated at 190 °C (±10 °C)
Spectral resolution0.5 cm−1
Linearityoffset ≤ 0.5% max, slope 0.99–1.01, SEE ≤ 1% max, R2 ≥ 0.998
Detection limit<2 ppm under all conditions of testing
Accuracy±3% of the reading or ±2 ppm, whichever is larger
Zero and span drift<2% of full scale
Rise time≤5 s
Response time≤20 s
PTFE = polytetrafluoroethylene; SEE = standard error of estimate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Giechaskiel, B.; Clairotte, M. Fourier Transform Infrared (FTIR) Spectroscopy for Measurements of Vehicle Exhaust Emissions: A Review. Appl. Sci. 2021, 11, 7416. https://doi.org/10.3390/app11167416

AMA Style

Giechaskiel B, Clairotte M. Fourier Transform Infrared (FTIR) Spectroscopy for Measurements of Vehicle Exhaust Emissions: A Review. Applied Sciences. 2021; 11(16):7416. https://doi.org/10.3390/app11167416

Chicago/Turabian Style

Giechaskiel, Barouch, and Michaël Clairotte. 2021. "Fourier Transform Infrared (FTIR) Spectroscopy for Measurements of Vehicle Exhaust Emissions: A Review" Applied Sciences 11, no. 16: 7416. https://doi.org/10.3390/app11167416

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