Next Article in Journal
Extended Environmental Multimedia Modeling System (EEMMS) with Analytic Hierarchy Process for Dual Evaluation of Energy Consumption and Pollutants in Solid Waste
Previous Article in Journal
The Unhappy Effects of the Antidepressant Fluoxetine on the Freshwater Microalga Raphidocelis subcapitata
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigating the Spectral Characteristics of High-Temperature Gases in Low-Carbon Chemical Pool Fires and Developing a Spectral Model

1
Guangxi Key Laboratory of Environmental Pollution Control Theory and Technology, Guilin University of Technology, Guilin 541006, China
2
College of Environmental Science and Engineering, Guilin University of Technology, Guilin 541004, China
*
Author to whom correspondence should be addressed.
Toxics 2025, 13(10), 877; https://doi.org/10.3390/toxics13100877 (registering DOI)
Submission received: 28 August 2025 / Revised: 2 October 2025 / Accepted: 9 October 2025 / Published: 14 October 2025

Abstract

Low-carbon chemical fires pose significant hazards, and remote sensing of high-temperature gas emissions from these fires is a critical method for identifying and assessing their environmental impact. Analyzing the spectral characteristics of gases produced by low-carbon chemical pool fires and developing spectral radiation models can establish a foundation for remote pollution monitoring. However, such studies remain scarce. Using a custom-built high-temperature gas spectroscopy platform, this study extracts spectral features of gases emitted by low-carbon chemical pool fires. We investigate spectral interference mechanisms among combustion products and develop a high-precision spectral radiation model to support remote fire pollution monitoring. Experimental results reveal distinct spectral bands for key gases: CO2 peaks near 2.7 μm and 4.35 μm, SO2 at 4.05 μm, 7.5 μm, and 9.0 μm, NO at 5.5 μm, and NO2 at 3.6 μm and 6.3 μm. The proposed spectral radiation model accurately simulates the position and shape of spectral peaks. For carbon disulfide and acetonitrile combustion products, the model achieves prediction accuracies of 83.4–96.9% and 79.2–95.3%, respectively.

Graphical Abstract

1. Introduction

Natural and human-made disasters persistently threaten ecological balance and societal stability. Among these, fires pose a severe and underestimated risk. Annual statistics reveal staggering human casualties and economic losses from fires, particularly in the chemical industry, where their impact is amplified [1,2].
Low-carbon chemicals are defined as substances produced from greenhouse gases (e.g., CO2, N2O, HFCs, PFCs, SF6) or through processes that substantially reduce emissions. Their development is critical for advancing low-carbon economies and is now a priority for industries aiming to minimize their carbon footprints. However, flammable low-carbon chemicals present significant hazards during production, transportation, and storage, with risks of explosions and potential misuse in attacks. Recent incidents have caused severe economic, human, and environmental damage, underscoring their societal risks [3,4]. Aerospace remote sensing offers a rapid, large-scale solution for monitoring pollution from low-carbon chemical fires. Identifying spectral signatures of these fires is essential for improving remote detection capabilities.
The spectral characteristics of chemical combustion serve as critical indicators for detecting and identifying pollution from chemical fires. Flame spectra primarily originate from soot radiation and molecular emissions of H2O and CO2 in combustion products [5]. Low-carbon chemical flames exhibit distinct spectral bands at 1.87 μm and 2.7 μm for H2O, and 2.7 μm and 4.3 μm for CO2. Leading institutions have developed comprehensive spectral databases, including NIST’s Chemistry WebBook and AIST’s SDBS (Spectral Database for Organic Compounds), which provide extensive reference data for combustion analysis [6,7]. The GEISA spectral line parameter database covers 58 molecules (145 isotopic variants) with over 6.7 million entries spanning 10−6 to 35,877 cm−1. The HITRAN database from AFGL contains high-resolution spectral data for 49 atmospheric components and pollutants, serving as essential references for quantitative gas analysis [8,9,10,11,12,13,14,15]. The HITEMP database specializes in high-temperature gas spectroscopy, showing improved accuracy for H2O, CO2, CO, NO, and OH through recent updates. However, uncertainties persist in some parameters [16,17,18,19]. For instance, CO2 spectral parameters in HITRAN and HITEMP databases show measurement uncertainties exceeding 20% for most line intensities, with only minor exceptions below 20% [20]. The applicability of these spectral databases for pollutant identification under fire conditions requires further validation. Current research predominantly focuses on common combustion products (H2O and CO2), while spectral analyses of characteristic pollutants like NOx and SOx remain limited. The interference mechanisms between combustion products also require deeper investigation.
Current research on the spectral characteristics of low-carbon chemical combustion focuses on two approaches: (1) Radiation heat transfer modeling using spectral data [21,22,23,24,25,26,27,28]. For instance, the RADCAL software V3.8.5 employs narrow-band methods to calculate combustion radiation, while flame spectral data are used to derive parameters like mass burning rates and thermal radiation values. However, these studies lack in-depth analysis of spectral features specific to combustion pollutants, limiting their applicability for pollution detection. (2) Differentiating pollutants via flame spectral signatures [29,30]. Techniques such as wavelet analysis distinguish fuel flames (e.g., oil) from other combustibles (e.g., alcohol, wood, paper), and spectral features combined with GOCI-Rrc satellite imagery identified carbonaceous aerosols as the primary pollutants in the Tianjin Port explosion. In smoke-free fire scenarios, molecular radiation from H2O, CO2, and CO dominates total fire radiation [31,32]. Recent studies highlight significant differences in infrared spectral properties of combustion-derived pollutants, enhancing remote sensing capabilities for fire pollution detection [33,34]. For example, Parent observed prominent absorption bands in atmospheric low-temperature gases within the 200–2200 cm−1 range, likely caused by high-temperature CO2 emissions during combustion [35]. Bordbar demonstrated that optimized filter responses amplify fire-induced spectral signals by 49–514% compared to blackbody emitters [36,37,38,39,40,41]. Challenges remain, however, due to spectral overlaps in certain chemical flames. Qiao visualized tail flame structures in rocket engines, identifying H2O and CO2 emission bands—consistent with findings from LNG flame studies [37]. Additionally, molecular similarities among combustion products create spectral interference effects [42,43,44,45].
Current research predominantly focuses on the spectral radiation properties of known chemical combustion processes, with limited attention to low-carbon chemical fire pollution from an aerospace remote sensing perspective. The existing satellite bands are concentrated in the range of 0.443–2.20 μm and 10.9–12.0 μm. The characteristic bands of SO2 (7.5 μm) and NO (5.5 μm) identified in this study can fill the monitoring blind area of satellites in the 2.2–10.9 μm range, realize the direct identification of characteristic pollutants from low-carbon chemical fires, and avoid the high error of traditional satellites relying on the indirect inversion of CO2. Meanwhile, the laboratory spectral resolution is as high as 0.12 cm−1 with an error of less than 15%, which can provide ground truth values for satellite low-resolution data (e.g., 0.35 cm−1 for Sentinel-2 MSI and 3 cm−1 for MODIS). It is worth mentioning that the “dual-band combination logic” proposed in this paper can help satellites distinguish between traditional hydrocarbon fires and low-carbon chemical fires.
There are significant differences in spectral characteristics between low-carbon chemical fires and traditional hydrocarbon fuel fires (such as those of petroleum and natural gas). Most studies emphasize common combustion byproducts such as H2O, CO2, and soot, while neglecting spectral analyses of characteristic pollutants like NOx and SOx. This gap limits the applicability of existing findings for detecting and distinguishing specific pollutants from diverse chemical fires. The combustion of low-carbon chemicals (e.g., CS2, acetonitrile) releases sulfur-containing/nitrogen-containing pollutants such as SO2 (7.5 μm) and NO (5.5 μm), which can serve as exclusive identification markers. In contrast, traditional hydrocarbon fires only rely on the common bands of CO2 and H2O. In low-carbon chemical fires, CO2 and SO2 have overlapping bands in the 4.3–4.6 μm and 7–10 μm ranges, and the high-temperature decomposition of NO2 dynamically changes the spectral characteristics. However, the interference in traditional hydrocarbon fires only comes from the slight overlap between H2O and CO2, making the analysis less difficult than that for low-carbon chemicals. It should be noted that existing studies only mention the spectral overlap between NOx/SOx and CO2, but do not quantify the degree of interference at high temperatures (e.g., 1023 K) and the effectiveness of deconvolution methods (such as NNLS), which cannot support the collaborative identification of multiple components.
In summary, to address this, our study employed a custom-built high-temperature gas spectral testing platform to extract spectral signatures of gas-phase products from low-carbon chemical pool fires. We established a stable high-temperature combustion environment, investigated spectral interference mechanisms among combustion products, and developed a high-precision spectral radiation model. These advancements provide a foundational framework for remote sensing technologies to identify and monitor low-carbon chemical fire pollutants.

2. Materials and Methods

2.1. High-Temperature Gas Spectral Testing Platform

Low-carbon chemical pool fires exhibit significant turbulent behavior, leading to substantial fluctuations in combustion product concentrations, temperatures, and spectral radiation. These variations hinder accurate measurements of flame field parameters such as temperature, concentration, and spectral data. To address this, a stable high-temperature gas environment was essential. This study developed a custom spectral testing platform (the platform is illustrated in Figure 1) to generate controlled high-temperature combustion conditions. Experiments were conducted in an enclosed chamber to measure spectral radiation from both individual and mixed gas-phase combustion products under stable conditions, providing critical spectral signature data for model development. The setup utilized a Fourier-transform infrared spectrometer, Bruker Optik GmbH, Ettlingen, Germany and an HFY-203D blackbody radiator, Shanghai Fuyuan Optoelectronics Technology Co., Ltd., Shanghai, China.
The emissivity of the HFY-203D blackbody is 0.995 (in compliance with JJG 856-2015 “Calibration Specification for Blackbody Radiation Sources” [46]). This blackbody is used for both lens transmittance measurement and gas spectral calibration: the blackbody radiation spectrum is collected at 800 K to establish a linear relationship between “radiance–spectral signal”, ensuring that the unit of gas spectral radiance is an absolute unit (W/(cm2·μm·sr)).
The Fourier Transform Infrared Spectrometer (FT-IR) produced by PerkinElmer (Shelton, CT, USA) (see Table 1 for instrument parameters) uses the Blackman–Harris apodization function. The spectral bands with high instrument noise in the range of 350–600 cm−1 are discarded. During the test, the spectral wavenumber range is 700–8000 cm−1, the spectral resolution is 4 cm−1, the cumulative number of scans is 2, the scan type is set to single-beam, and a high-precision liquid nitrogen-cooled MCT detector is used.
To minimize background radiation interference from daylight, all experiments were conducted at night under sunlight-free conditions. The platform consists of four independent modules: heating, spectral data acquisition, gas distribution, and water cooling. Connected via standard connectors, it can be disassembled for transportation and on-site assembly outdoors. The spectrometer can perform remote data collection via a 10 m optical fiber to avoid high-temperature interference. The water cooling system can be replaced with an air cooling module (the cooling efficiency is reduced to 80% of that of water cooling, but the weight is reduced by 60%), which is suitable for outdoor scenarios without water sources; after modification, the cost of a single outdoor experiment is approximately 30% higher than that in the laboratory, but it enables in situ spectral collection of fires involving CS2, acetonitrile, etc.
During platform assembly, we observed that the water-cooled flange provided limited cooling to the lens. When the gas in the chamber was heated to 800 °C, the monocrystalline silicon lens temperature reached 200 °C, causing significant changes in lens transmittance and inherent radiation spectra, which compromised spectral measurement accuracy. To address this, we quantified the interference effects of lens transmittance and self-emission on experimental results. To maintain a constant radiative background during high-temperature lens transmittance measurements, the HFY-203D blackbody furnace was stabilized at 800 °C. This temperature was selected based on the furnace’s stable radiative brightness and minimal thermal fluctuations at 800 °C. The lens transmittance is calculated using Equation (1).
τ v = E 2 + E 3 / E 1
τ v is the lens transmittance at varying wavenumbers, E1 denotes the blackbody radiative energy measured through the chamber, E2 represents the radiative energy measured through the chamber after inserting the lens, and E3 corresponds to the lens’s inherent radiative energy.
The measurement process for E1, E2, and E3 is as follows:
(1)
Measurement process of E1: Turn on the blackbody furnace and slowly heat it up at a temperature gradient of 100 °C. When the temperature reaches the target temperature of 800 °C, keep it constant. Then, open the valve at the air inlet and then open the valve at the air outlet of the gas distribution system to introduce inert gas and exhaust the air inside cavity 2. Nitrogen is selected as the background gas in this experiment. Move the rotating chassis to position the heating chamber in front of the blackbody and Fourier transform infrared (FT-IR) spectrometer. By adjusting the angle of the rotating chassis and its position on the slide rail, ensure that the radiation from the blackbody can pass through the chamber and fully enter the field of view of the FT-IR spectrometer. After completing the above operations, turn on the FT-IR spectrometer to scan and obtain E1.
(2)
Measurement process of E2: Maintain the position of each part on the platform during the measurement process of E1. Gently pull the handle to open the lens heating chamber, place the lens to be tested inside, adjust the sliding block, and press the lens firmly against the high-temperature heating plate to ensure uniform heating. Then, open the valve at the air inlet, and then open the valve at the air outlet of the gas distribution system to introduce inert gas or nitrogen to exhaust the air inside the cavity. Turn on the lens heating temperature control box to raise the lens to the target temperature. After completing the above operations, use an FT-IR spectrometer to scan and obtain E2.
(3)
Measurement process of E3: Move the blackbody furnace to the other side of the slide rail to ensure that it is not within the field of view of the FT-IR spectrometer, and there is no light source within the field of view. Rotate the rotating chassis by 180°, placing the lens end within the field of view of the FT-IR spectrometer. After completing the above operations, use the FT-IR spectrometer to scan and obtain E3.
There is a certain degree of uncertainty in E3 (lens radiant energy). This is because key parameters have an impact on accuracy: temperature (+0.42%/K, increasing to +0.65%/K at high temperatures) > incident angle (−1.8%/°) > transmittance fitting error (±1.5% per 1% error) > surface contamination (−0.9% per 1% decrease in transmittance). This error can be optimized (e.g., by integrating a precise temperature control module, mechanical positioning, segmented fitting, etc.).

2.2. Spectral Denoising and Smoothing Processing

During high-temperature gas spectral testing, the monocrystalline silicon lens within the field of view partially absorbs the gas’s spectral radiance C1(v), which is then detected by the FT-IR spectrometer. The total signal entering the spectrometer comprises C1(v) and the lens’s inherent radiation E3. To derive the true radiative brightness L(v) of the gas, the lens-induced interference must be eliminated by isolating Cn(v), the corrected spectral radiance. This correction is achieved using Equation (2):
C n v = C 1 v / τ v + E 3
Cn(v) is the initial spectral radiance measured by the FT-IR spectrometer, C1(v) is the spectral radiation value obtained by interference from high-temperature lenses, τ(v) denotes the transmittance of the high-temperature lens, E3 represents the inherent radiative energy of the lens.
In addition to interference from the high-temperature lens, reflected light from components such as the quartz tube, flange, and furnace heating filaments also enters the FT-IR spectrometer. This introduces baseline drift, excessive spectral noise, and spectral peaks which are difficult to distinguish, complicating data interpretation. To mitigate these effects, a background spectrum must first be acquired using a gas with negligible radiative features. Nitrogen, a diatomic molecule, exhibits extremely weak spectral signatures—its radiative brightness at equivalent temperatures is approximately five orders of magnitude lower than that of carbon dioxide [47]. Thus, nitrogen spectra at varying temperatures were selected as the background reference. The correction is calculated using Equation (3):
C 0 v = C 0 v C N v
C0(v) is the corrected spectral radiative energy of the target high-temperature gas after interference removal, CN(v) represents the radiative energy of the high-temperature nitrogen gas used for background correction, C0′(v) denotes the uncorrected radiative energy of the gas prior to background subtraction.
Despite background correction, residual baseline drift and spectral noise persist in the high-temperature gas spectra. To enhance the visibility of characteristic spectral peaks, this study employs the AirPLS algorithm—an adaptive iteratively reweighted penalized least squares method (Parameters of the AirPLS algorithm: regularization parameter λ = 106, number of iterations = 50, stopping tolerance = 10−6; Whittaker smoothing window size = 15; Gaussian convolution FWHM = 0.1 cm−1, and the convolution parameters are set based on the spectral line half-width of the HITRAN 2020 database (Ontar Corporation, North Andover, MA, USA))—to suppress noise and oscillations. This approach preserves the overall spectral trend while sharpening peak resolution. However, this algorithm has certain limitations. When the concentration of a certain component is extremely low (e.g., NO 0.5%) and the signal of its characteristic peak (5.5 μm) is close to the baseline noise (<0.005 W/(cm2·μm·sr)), the iterative weight of the algorithm will underestimate the residual of the peak, resulting in a peak area error of 15–20% and a detection rate reduced to 85%. The Whittaker smoothing window (15) causes peak shape broadening for narrow peaks (e.g., NO 5.5 μm) and suppresses edge signals for wide peaks (e.g., SO2 9.0 μm), resulting in an increase in peak height error from ±2.8% for a single gas to ±5.1% for mixed gases. Moreover, this algorithm can only remove baseline noise and cannot distinguish extremely weak overlapping peaks (e.g., CO2 4.35 μm and SO2 4.4 μm). It needs to rely on NNLS deconvolution (based on HITRAN basis functions) for further separation, with the residual controlled to be <8%.
The iterative application of the Whittaker smoothing algorithm progressively improves baseline correction accuracy with minimal distortion to spectral features. Additionally, Gaussian convolution was applied to the processed spectra using line-by-line parameters from the HITRAN database to further refine spectral resolution.
Whittaker Smoothing [48] is defined by Equation (4):
min _ L v W L v y 2 2 + λ D L v 2 2
L(v) is the input spectral data, W is a diagonal weight matrix (W = diag(w)), w is the weight vector, y is the initial estimated spectrum, λ is the smoothing parameter, and D is the difference matrix.
AirPLS Weight Adjustment: The algorithm dynamically updates weights based on residuals to suppress spectral peaks during baseline correction. The weighting function in Equation (5) is defined as:
w i = 0 d i 0 e x p i d i d s s n d i < 0
wi is the weight of the i-th data point, di is the residual for the i-th point, calculated as the difference between the observed value and the Whittaker-smoothed estimate, dssn is the sum of the absolute values of all negative residuals.
The experiment tested the transmittance of the lens within the temperature range of 353–483 K. The transmittance of the lens was calculated using Equations (1)–(3), as shown in Figure 2. The transmittance of the lens remained stable at around 0.535 in the 1.3 μm–6.6 μm wavelength range, and the temperature difference was not significant. There is a significant decrease in the wavelength range of 6.65 μm–6.95 μm, and the lens transmittance varies greatly at different temperatures between 8.6 μm–12 μm.
The lenses at both ends of the high-temperature gas testing platform gradually increased in temperature as the gas heated, though this rise was moderated by the water-cooled flanges. Infrared thermometry revealed that lens temperatures ranged from 353 K to 483 K during gas heating.
Lens transmittance within this temperature range was calculated using Equation (1). As shown in Figure 3, transmittance remained stable at approximately 0.535 across the 1.3–6.6 μm wavelength band, with minimal variation between temperatures. A sharp decline occurred between 6.65 and 6.95 μm, while significant temperature-dependent differences emerged in the 8.6–12 μm range.
During the experiment, the lens transmittance was measured 5 times per temperature point (353–483 K, with an interval of 30 K). Through calculation using the error propagation formula, there is a small error in the final radiance. The following is the table of lens transmittance characterization errors, see Table 2:

2.3. Single- and Mixed-Gas High-Temperature Spectral Testing

High-temperature spectral testing of individual and mixed gases was conducted using a dedicated platform. The study focused on four combustion-derived gases—CO2, SO2, NO, and NO2—generated from low-carbon chemicals (carbon disulfide and acetonitrile). The purity of gases used in the experiment was as follows: CO2 99.99%, SO2 99.9%, NO 99.5%, NO2 99.5%, N2 99.999%. The calibration accuracy of MFC (model: Alicat MCR-500, Alicat Scientific, Tucson, AZ, USA) is ±0.5% FS, with a calibration cycle of 3 months. The pipeline material is 316 stainless steel (corrosion-resistant, to avoid gas adsorption). The gas flow rate is controlled at 500 sccm, the mixing volume of the gas chamber is 500 mL, and the residence time is about 10 s (to ensure uniform gas mixing). The humidity in N2 is measured to be <0.1 ppm by a dew point meter (model: Testo 645, Testo SE & Co., KGaA, Titisee-Neustadt, Germany), and the content of CO/CO2 in N2 is measured to be <0.1 ppm by gas chromatography (model: Agilent 7890A, Agilent Technologies, Inc., Santa Clara, CA, USA).
Tests were carried out at standard atmospheric pressure (1 atm) and temperatures ranging from 773 K to 1073 K. Secondary spectral peaks outside characteristic bands were filtered to enhance the visibility of key features. The heating chamber’s high-temperature zone spanned 18 cm. Under active water-cooling and high flow rates, the gas was treated as an isothermal medium with an 18 cm optical path. Simulated spectra were generated using the HITRAN database via Python 3.10.2, accounting for temperature, concentration, pressure, and path length. Due to HITRAN’s line-by-line computational approach, there are some slight discrepancies between simulations and FT-IR measurements. Simulated data (4 cm−1 resolution, 18 cm path length, 1 atm pressure) served as preliminary references for validating the experimental results.
Low-carbon chemical combustion generates characteristic pollutants (SO2, NO, NO2) alongside CO2 as a common byproduct. These four gases were selected for single-component high-temperature spectral testing. For mixed-gas experiments, NO was excluded due to its reactivity with atmospheric oxygen (forming NO2) and the thermal instability of NO2 at elevated temperatures. Instead, SO2, CO2, and N2 were combined based on their stability and relevance to carbon disulfide (CS2) combustion. As shown in Equation (6), complete CS2 combustion in air produces CO2 and SO2 in a 1:2 molar ratio. Three gas mixtures were tested to reflect this stoichiometry: CO2:SO2 = 1:2, CO2:N2 = 1:2, and N2:SO2 = 1:2.

2.4. Simulated Combustion Spectral Testing

A stable combustion environment is critical for constructing reliable high-temperature gas spectral radiation models. Spectral testing of individual and mixed gases under controlled conditions revealed significant mutual interference in mixed-gas spectra. To address these challenges, this study established stable low-concentration gas mixtures at temperatures ranging from 773 to 1073 K, leveraging combustion field data (component concentrations and temperatures) from low-carbon chemicals. The experiments systematically characterized spectral interference patterns in mixed gases, providing foundational insights for refining our spectral radiation models.
To replicate the component concentrations and temperatures of low-carbon chemical combustion fields, gas inflow ratios were calculated based on the stoichiometry of CS2 combustion in air (Equation (6)) and experimental data from CS2 pool fire studies (Mishra et al. stated in their report that the ratio of combustion products of CS2 is CO2:SO2 ≈ 1:2.1 [49]). Two gas mixtures were defined to approximate CS2 combustion products: CO2:SO2:N2 = 0.8%:1%:98.2% and CO2:SO2:N2 = 0.5%:1%:98.5%, reflecting variations in combustion efficiency and dilution effects.
C S 2 + 3 O 2 = C O 2 + 2 S O 2
Gas mixture ratios for acetonitrile combustion were derived from its complete combustion stoichiometry Equation (7) and experimental studies on acetonitrile pool fires (Alzueta et al. stated in their report that the ratio of combustion products of acetonitrile is CO2:NO2 ≈ 2:1 [50]). Two simulated combustion conditions were established: CO2:NO:N2 = 1%:0.5%:98.5% and CO2:NO2:N2 = 0.8%:0.5%:98.7%, representing variations in oxidation efficiency and nitrogen dilution effects observed in controlled environments.
4 C 2 H 3 N + 15 O 2 = 8 C O 2 + 6 H 2 O + 4 N O 2
Since NO and O2 are prone to form NO2 at high temperatures, which interferes with the spectral measurement of NO, this experiment eliminates the possibility of O2 presence through strict repeated purging. Due to the thermal decomposition of NO2 at temperatures > 423 K (2NO2 ⇌ 2NO + O2, with a decomposition rate > 90% at 1073 K), which leads to unstable concentration, NO2 and NO are not tested simultaneously in the mixture. During the experiment, a rapid heating method (3–5 min) is adopted to avoid the thermal decomposition of NO2, and the spectral collection is carried out during the rapid heating process.

2.5. Developing a Spectral Radiation Model for High-Temperature Gases

Spectral data for high-temperature gases were collected at 50 K intervals within the 773–1073 K range. To extrapolate radiation spectra across this temperature domain, the relationship between gas radiation and temperature was analyzed through spectral fitting. Unlike gases, the radiation intensity of solids and liquids increases with thermal excitation. Blackbody radiation depends solely on its intrinsic temperature.
Gas molecules emit or absorb light at specific wavelengths only during transitions between discrete energy levels. Consequently, gases show negligible reflectivity (R(v) ≈ 0) and near-total transmittance (T(v) ≈ 100%) across most spectral bands. As derived from Equation (8), absorptivity remains near zero in these regions, aligning with Kirchhoff’s law of thermal radiation.
The interdependence of transmittance, absorptivity, and reflectivity is defined by Equation (8):
R v + A v + τ v = 100 %
R(v) is the reflectivity, A(v) denotes the absorptivity, and τ(v) represents the transmittance.
High-temperature gases serve as the primary source of radiative spectra. Kirchhoff’s law of thermal radiation establishes that under equilibrium conditions, a material’s absorptivity A(v,T) equals its emissivity E(v,T). Equation (9):
A v , T = E v , T
This principle enables the spectral radiance of high-temperature gases to be quantified through their emissivity and corresponding blackbody radiation characteristics under thermodynamic equilibrium. Equation (10):
L v = L B v , T E v = 2 h c 2 v 3 e x p h c v k T 1 E v
v: Wavenumber, T: Temperature, E(v): Spectral emissivity at wavenumber v, c: Speed of light, LB(v,T): Blackbody spectral radiance at wavenumber v and temperature T, h: Planck’s constant, k: Boltzmann’s constant.
Equation (10) demonstrates that spectral radiance is proportional to absorptivity. When gas absorption A(v) = 0 within a spectral band, the gas exhibits zero radiance L(v) regardless of blackbody radiation intensity in that band, resulting in no characteristic spectral peaks. Non-zero absorption A(v) ≠ 0 produces distinct emission peaks, creating a discontinuous line spectrum specific to the gas composition. Peak positions are determined by molecular structure, quantum energy levels, and environmental conditions. Using CubicSpline interpolation [51], we reconstructed spectra across temperature gradients with >95% accuracy in experimental validation, achieving complete positional agreement of spectral peaks.
Equation (11) establishes the proportionality between spectral radiance and absorptivity. If a gas exhibits zero absorptivity A(v) within a spectral band, its radiance L(v) remains zero regardless of blackbody radiation intensity in that band, eliminating characteristic spectral peaks. Non-zero absorptivity generates distinct emission peaks, which define the gas-specific spectral fingerprint. Since blackbody radiance depends solely on temperature, the gas radiation spectrum can be fully determined by quantifying its emissivity E(v). Under thermal equilibrium, Kirchhoff’s law equates gas absorptivity A(v,T) to emissivity E(v,T) [Equation (10)]. The Lambert–Beer law quantifies this relationship, linking absorptivity to gas concentration, temperature, and optical path length.
A v , T = 1 e x p [ R a v c ]
R: Optical path length, a(v): Spectral absorption coefficient, c: Gas concentration.
The core assumptions of the Lambert–Beer Law are medium uniformity (uniform spatial distribution of temperature and concentration) and optically thin conditions (weak light absorption, no multiple scattering/absorption saturation). Due to the influence of turbulent characteristics, the concentration of flame combustion products, temperature, and radiation spectrum fluctuate greatly. Particles such as soot and ash produced by incomplete combustion cause multiple scattering of light in the medium, leading to a calculated concentration higher than the actual value. In this study, a stable high-temperature gas environment of combustion products is created in a closed space, which can ensure the validity of the Lambert–Beer Law.
For a high-temperature gas layer in combustion environments with optical path length R, the absorbed radiance varies with differential path length across the 0-RR range.
d L v d R = a v L v
Equations (10)–(12) collectively define the relationships between absorptivity A(v,T), emissivity E(v,T), gas radiance L(v), and blackbody radiance LB(v,T). Substituting Equation (12) yields the differential variation of emitted radiance with optical path length.
d L v d R = a v L B v , T = 2 h c 2 v 3 exp h c v k T 1 a v
Integrating these results through Equations (12) and (13) provides the total radiance profile across the optical path.
d L v d R = a v 2 h c 2 v 3 exp h c v k T 1 L v
The spectral absorption coefficient, quantifying light attenuation per unit path length, relates directly to absorptivity through the fundamental identity.
A v , T = a v R
Equations (16)–(22) formalize the radiative transfer framework. By synthesizing Equations (9), (10), (14) and (15), we derive the explicit relationship between gas radiance L(v) and absorptivity A(v,T).
L v = L B v , T A v , T = 2 h c 2 v 3 exp h c v k T 1 A v , T
The absorptivity A(v,T) depends on three critical parameters: optical path length R, spectral absorption coefficient a(v), and gas concentration c.
It should be noted that when nitrogen is used for background subtraction, the radiance of nitrogen in the 4–10 μm band is only 1% of that of the experimental target spectrum. Under normal experimental conditions (high-purity nitrogen, matched environmental parameters), the risk of masking the weak signal of the target gas is extremely low, but there is a slight risk under extreme conditions (insufficient nitrogen purity, mismatch between background and sample environment, extremely low signal-to-noise ratio of weak signals).
Within a differential flame layer dR, the spectral transmittance τ(v) is defined.
d τ v = a v ,   R d R
Integrating across the full optical path (0-R) yields the total transmittance τv.
τ v = 0 d R a v ,   R d R
Multiply both sides of Equation (14) by exp (τ V) to couple emissivity and transmissivity. This culminates in the comprehensive radiative transfer equation for combustion systems [Equation (19)].
L v = e x p τ v [ L v + 0 τ 2 h c 2 v 3 e x p h c v k T 1 e x p ( τ v ) d τ v ]
The temperature- and thickness-dependent variability of τv across gas species renders traditional analytical integration methods impractical. By unifying Equations (16) and (19), we derive a generalized radiative equation under thermodynamic equilibrium [Equation (20)], circumventing the need for iterative summation.
L v = 2 h c 2 v 3 e x p h c v k T 1 { 1 e x p [ R a v c ] }
Equation (20) confirms that blackbody radiance LB(v,T) depends exclusively on temperature. The spectral absorption coefficient a(v), intrinsic to molecular structure, combined with readily measurable optical path length d, enables full spectral reconstruction when temperature T, a(v), and concentration c are known.
This model is based on the following assumptions: (1) Local Thermodynamic Equilibrium (LTE): within the temperature range of 773–1073 K, the energy level distribution of gas molecules conforms to the Boltzmann distribution; (2) Neglecting scattering: combustion gas is dominated by absorption/emission, with a scattering coefficient <10−5 cm−1, which can be neglected; (3) Bilateral symmetry: the gas chamber/flame is divided into 12 symmetric regions, and only the parameters of 6 regions on one side need to be calculated; (4) Isothermal sub-regions: the temperature fluctuation within each region is <5 K; (5) Constant pressure: the pressure is maintained at 1 atm during the experiment, with a pressure fluctuation <±0.01 atm; (6) Optically thin outside spectral bands: in non-characteristic spectral bands (e.g., 1–1.3 μm), the gas absorption rate is <0.01, which is regarded as optically thin. The above assumptions are valid within the temperature (773–1073 K) and concentration (0.5–1%) ranges of this study and have been verified by experiments.
Within our high-temperature gas spectral testing platform, T and c are predefined controlled parameters. We resolve a(v) using the HITRAN molecular database, leveraging its validated spectral line parameters for precise absorption coefficient determination.
Figure 3 reveals distinct thermal profiles: the testing platform exhibits radially decreasing temperatures from its heating core, whereas low-carbon chemical pool fires show minimum temperatures at the flame base with increasing values toward intermediate zones. Both systems maintain bilateral symmetry, enabling computational simplification through single-side spectral parameter analysis.
Leveraging flame symmetry, we developed a spectral radiation model for stable combustion systems. As shown in Figure 4, the flame and testing platform are divided into 12 symmetrical zones (6 per side). Absorption coefficients and transmittance for one lateral set of 6 zones are derived from the HITRAN database, exploiting bilateral equivalence. The custom radiative transfer model synthesizes platform-wide spectra. For instance, Zones 1 and 12 exhibit identical radiance. However, Zone 1 radiation enters the FT-IR spectrometer directly, while Zone 12 emission is attenuated by cumulative transmittance through Zones 1–11. This path-dependent attenuation is quantified by multiplying Zone 12 radiance with the product of transmittance values from intervening zones.
L 0 n = L n i = 1 n 1 τ v i
L(n): Spectral radiance of Zone n (1 ≤ n ≤ 12 (As shown in Figure 4)), τv(n): Transmittance of Zone n.
Exploiting the bilateral symmetry within the radiative zones, we adapt Equation (14) to derive radiance expressions for both lateral halves. Zones 1–6 (1 ≤ n ≤ 6) follow the computational framework of Equation (21), expressed as L0(n). For Zones 7–12 (7 ≤ n ≤ 12), the radiance calculation incorporates mirror-symmetric optical paths, formalized as:
L 1 n = L 13 n i = 1 6 τ v i i = 1 n 7 τ v 7 i
Equation (22) demonstrates that radiance values for Zones 7–12 can be derived using HITRAN-calculated transmittance and radiance data from Zones 1–6. This symmetry-driven approach reduces the spectral model’s computational load, achieving nearly 50% faster processing times.
Integrating Equations (21) and (22) yields the comprehensive spectral radiance formula for the total system:
L = n = 1 n = 6 ( L n i = 1 n 1 τ v i ) + n = 7 n = 13 ( L 13 n i = 1 6 τ v i * i = 1 n 7 τ v 7 i )
Per Equation (20), with predefined temperature T, concentration c, and optical path length d, determining the absorption coefficient a(v) enables full computation of L(n) values in Equation (23).
For example, when a fire occurs, there are 5 unknown quantities (temperature, optical path, and concentrations of 3 gases). Four characteristic bands are selected to ensure that the number of bands is greater than the number of unknown quantities (CO2 2.7/4.35 μm, SO2 7.5 μm, NO 5.5 μm). Referring to the matrix analysis conditions in Equations (20)–(23), it can be determined that an inversion condition number less than 10 indicates a stable inversion combination.

3. Results and Discussion

3.1. High-Temperature Single-Gas Spectral Analysis

Figure 5 compares experimental results with HITRAN-simulated spectra for four gases: SO2 and NOx (characteristic low-carbon combustion pollutants) and CO2 (universal combustion product). All exhibit distinct spectral fingerprints within their high-temperature emission profiles. Raw spectra underwent three-stage preprocessing: lens interference removal, radiometric calibration, and adaptive iteratively reweighted Penalized Least Squares (AirPLSs) baseline correction. Spectral measurements across 773–1073 K were conducted using a custom high-temperature gas testing platform. Post-acquisition processing eliminated instrumental artifacts, low-frequency baseline drift, and stochastic noise through manual curation and Whittaker smoothing.
Experimental CO2 emission spectra demonstrate strong agreement with HITRAN simulations, exhibiting characteristic peaks near 2.0 μm, 2.7 μm (highest intensity), 4.35 μm (secondary intensity maximum), and 9.3 μm. SO2 measurements align with simulated spectra in peak positions (2.8 μm, 4.05 μm, 4.35 μm, 7.35 μm, 9.0 μm), though radiance discrepancies occur at 4.05–4.35 μm. The 7–10 μm bands show precise alignment, with dominant emission at 4.05 μm, 7.35 μm, and 9.0 μm. NO experimental and simulated spectra exhibit matching spectral peaks at 2.4 μm and 5.5 μm, with near-identical radiance magnitudes. The 5.5 μm band dominates radiative output. NO2 measurements validate key spectral features at 2.3 μm, 3.6 μm, 6.3 μm, and 7.5 μm. The 3.6 μm and 6.3 μm bands demonstrate maximum radiance, providing critical wavelength targets for radiative model development. All experimental data confirm sufficient accuracy for spectral fingerprint identification. Figure 6 systematically catalogs the dominant spectral signatures for each investigated gas species, see Table 3.
It can be seen from the table that the Pearson correlation coefficients of all gas characteristic bands are higher than 0.97, and the peak height error is less than 7%, which indicates a strong consistency between the experimental spectra and HITRAN simulations.
Critically, thermal dissociation of NO2 into NO and O2 initiates above 150 °C (423 K), driven by positive entropy changes favoring gaseous product formation. Elevated temperatures accelerate this decomposition, significantly affecting spectral measurements of pure NO2 at high temperatures. c1 and c2 [Figure 4] reveal notable deviations between experimental and HITRAN-simulated spectra: 3.4–3.7 μm band: Measured radiance underperforms simulations, 5.9–6.8 μm band: Radiance growth rate diminishes with temperature increase, Post 973 K: Absolute radiance values decrease. Analytical results indicate that progressive NO2 depletion in the heating zone reduces both gas concentration and radiative output. At 973 K, near-complete decomposition occurs, with residual spectra originating from partially dissociated NO2 in thermally insulated regions.

3.2. High-Temperature Gas Mixture Spectral Analysis

To mitigate interference from NO2 decomposition (thermally unstable above 423 K) and NO oxidation dynamics, the study focused on SO2/CO2/N2 mixtures. Three stoichiometric ratios were investigated: CO2:SO2 = 1:2, CO2:N2 = 1:2, and N2:SO2 = 1:2. The experimental results are summarized below:
The CO2:N2 (1:2) mixture exhibits spectral signatures nearly identical to pure CO2. Comparative analysis (Figure 5(a1) and Figure 7(a’1)) reveals reduced overall radiance with decreased CO2 concentration. Radiance attenuation occurs 23% earlier in the 2.6–2.9 μm band compared to 4.2–4.7 μm. HITRAN simulations accurately replicate both spectral features and concentration-dependent attenuation patterns.
For the N2:SO2 (1:2) mixture, characteristic SO2 peaks at 2.8 μm, 4.05 μm, 4.35 μm, 7.35 μm, and 9.0 μm remain stable despite a reduced concentration. While 3.9–4.8 μm radiance decreases proportionally to SO2 volume fraction, the 7–10 μm band shows <5% variation, confirming wavelength-dependent concentration sensitivity. HITRAN effectively models this behavior.
The CO2:SO2 (1:2) mixture demonstrates spectral overlap in two critical regions: 4.3–4.6 μm: CO2/SO2 peak superposition, 7–10 μm. Broadband feature merging, Experimental data show <8% deviation in these regions versus HITRAN simulations, suggesting minor unmodeled gas interaction effects. Both systems maintain baseline agreement across 1.9–2.0 μm, 2.6–3 μm, and 3.9–4.7 μm bands.
To quantify the material contribution of the overlapping bands (4.3–4.6 μm, 7–10 μm) of CO2 and SO2, 1023 K (the core temperature of CS2 fires) was selected as the representative temperature. Based on the HITRAN 2020 database, the basis functions of pure CO2 (1% concentration) and pure SO2 (2% concentration) (1 atm, 18 cm optical path) were constructed, and deconvolution was performed using the constrained non-negative least squares (NNLS) method. In the 4.3–4.6 μm band, CO2 contributes 62% (95% confidence interval: 59–65%) and SO2 contributes 38% (35-41%), with the strong peak of CO2 at 4.35 μm dominating this band. In the 7–10 μm band, SO2 contributes 82% (79–85%) and CO2 contributes 18% (15–21%), with the strong peaks of SO2 at 7.5 μm/9.0 μm being absolutely dominant. The relative deviation of the deconvolution residual (measured value–fitted value) is <8%. As shown in Figure 8, the maximum absolute residual in the 4.3–4.6 μm band is 0.008 W/(cm2·μm·sr) (deviation: 7.2%), and the maximum in the 7–10 μm band is 0.005 W/(cm2·μm·sr) (deviation: 4.8%). This verifies the effectiveness of the HITRAN basis functions and the NNLS method in separating overlapping bands, and also proves the quantitative reliability of the “deviation < 8%” mentioned in the paper.
Concentration sensitivity is wavelength dependent. Through the analysis of the “peak height vs. mixing ratio” relationship (Figure 9) and detection limit, the peak height of the SO2 band represented by 7.5 μm has a strong linear correlation with concentration. When the concentration changes by 10 times, the peak height stability is >95%, and the detection limit is 0.1% (SNR ≥ 3), which is suitable for designing a fixed filter with a center of 7.5 μm and a bandwidth of 0.5 μm. For the peak height of the CO2 band represented by 7.5 μm, the linearity (R2) between peak height and concentration is 0.99 at low concentrations (<1%) with a detection limit of 0.09%. However, the linearity decreases to R2 = 0.95 at high concentrations (>5%), which is suitable for designing an adjustable bandwidth filter (bandwidth: 0.2–0.5 μm) to dynamically match the CO2 concentration range at different stages of the fire.

3.3. High-Temperature Gas Spectral Validation

The combustion environment was simulated using four gas mixtures: CS2 combustion analogs: CO2:SO2 = 0.8%:1% (a) and 0.5%:1% (a′), C2H3N combustion analogs: CO2:NO = 1%:0.5% (b) and 0.8%:0.5% (b′), Spectral profiles were obtained under controlled high-temperature conditions (Figure 10, panels a, a′, b, b′). Complementary FT-IR measurements of actual CS2 (a″) and C2H3N (b″) pool fires provided benchmark spectra for comparative analysis with simulated gas mixtures (Figure 10a″,b″).
Comparative analysis with Figure 5 reveals CO2’s dominant emission shifts from dual peaks (2.7 μm and 4.35 μm) in pure gas to single 4.35 μm dominance in low-concentration CS2 combustion products, showing 63% attenuation at 2.7 μm. SO2 spectral behavior transitions from tri-modal peaks (4.05/7.35/9.0 μm) in isolation to 7.5 μm dominance (82% intensity retention) under CS2 combustion conditions, establishing this wavelength as key for low-carbon SO2 detection.
0.3% CO2 reduction causes 41% peak attenuation in 4.2–4.7 μm band versus <5% variation in SO2’s 7.1–8 μm band, demonstrating decoupled concentration effects between co-existing species. The testing platform achieves 0.12 cm−1 resolution in 7.1–8 μm band versus 0.35 cm−1 for actual CS2 fires due to FT-IR’s temporal averaging of flame fluctuations. Spectral congruence exceeds 89% across other wavelengths, validating simulation fidelity.
C2H3N combustion shows CO2’s persistent 4.35 μm dominance (78% intensity retention) with 2.7 μm attenuation. While NO’s 2.4 μm peak overlaps with CO2, its distinct 5.5 μm signature (S/N ratio > 4.2) emerges as a reliable detection marker under low-concentration conditions.
C2H3N simulations show 22% lower 4.35 μm radiance versus actual fires due to 210 K temperature differential. The 5–6 μm band maintains <15% deviation, while peak intensity ratios across other wavelengths show >92% agreement, confirming model robustness.
To compare the laboratory spectra and the actual pool fire spectra, the SO2 7.5 μm (characteristic band) and CO2 4.35 μm (common band) of CS2 were selected, and Gaussian convolution (FWHM = 0.35 cm−1) was used to simulate the actual pool fire resolution. The changes in peak height/peak area are shown in Figure 10. The decrease in resolution leads to a slight attenuation of peak height (5–10%) and a smaller attenuation of peak area (3–6%). This is because the SO2 7.5 μm peak is narrower and more sensitive to resolution.
The NO 5.5 μm (characteristic band) of acetonitrile was selected as the analysis object, and the temperature was adjusted to the actual pool fire range (±200 K). The change in peak height after temperature adjustment is shown in Figure 11. Temperature has a significant impact on peak height (±15–20%/100 K), and CO2 4.35 μm is more sensitive than NO 5.5 μm, see Figure 12.

3.4. Analog Computation and Validation of High-Temperature Spectral Radiation Model

Our radiation model integrates three critical datasets: 1. Temperature-concentration profiles from high-temperature testing platform. 2. Low-concentration spectral measurements via FT-IR. 3. Low-carbon chemical flame spectra. Combined with HITRAN-derived combustion product parameters, this synthesis achieves >92% predictive accuracy across validation cases. The model enables combustion diagnostics through spectral inversion, successfully extrapolating to outdoor fire scenarios with 85–89% field detection accuracy. This establishes a critical foundation for remote sensing identification of low-carbon chemical fires in complex environments.

3.4.1. Gas Spectral Interference

While low gas concentrations in flame spectra minimize combustion product interference, optimizing model accuracy requires explicit analysis of HITRAN’s spectral interference handling. Using N2 as a diluent, we simulated three mixtures (CO2:SO2 = 1:2, CO2:N2 = 1:2, N2:SO2 = 1:2) at 1023 K, focusing on 3.9–4.8 μm and 7–10 μm bands.
HITRAN simulations exhibit spectral line crossover in 4.18–4.8 μm bands, resolved through peak overlap correction. In the 9–10 μm range, SO2-dominated profiles show 12–15% radiance enhancement in CO2:SO2 mixtures versus N2:SO2 controls, indicating that HITRAN employs additive peak summation. This alignment with FT-IR interference mitigation strategies (±8% deviation) validates HITRAN’s utility for high-temperature spectral modeling.

3.4.2. Parameter Acquisition

The absorption coefficient a(v) and transmittance τv form critical inputs for the radiative model. Using HITRAN, these parameters were calculated for: CS2 combustion products: 723–1023 K (6 cm gas column at 1023 K; 2 cm otherwise). C2H3N combustion products: 773–1073 K (6 cm column at 1073 K; 2 cm otherwise)
(I) Absorption Coefficients
For CO2:SO2:N2
CO2: Peak absorption at 4.35 μm (a(v)max = 0.35 cm−1), weaker 2.7 μm peak.
SO2: Dominant 7.5 μm band (a(v)max = 0.05 cm−1).
Temperature increase reduces absorption coefficients, most notably at 4.35 μm (15–22% decrease across 723–1023 K).
For CO2:NO:N2
CO2: Maintains 4.35 μm dominance (a(v)max = 0.35 cm−1).
NO: Primary peak at 5.5 μm (a(v)max = 0.05 cm−1).
Doubling CO2 concentration increases 4.35 μm absorption by 40% with 18% spectral broadening. Thermal equilibrium ensures absorptivity-emissivity equivalence, causing corresponding emission band widening.
(II) Transmittance
CO2:SO2:N2 mixture exhibits (Near-unity transmittance (>98%) outside absorption bands):
CO2: Minimum τv = 0.23 cm−1 at 4.35 μm.
SO2: Minimum τv = 0.86 cm−1 at 7.5 μm.
CO2:NO:N2 mixture shows:
Enhanced CO2 attenuation: τv = 0.19 cm−1 at 4.35 μm.
NO: τv = 0.89 cm−1 at 5.5 μm.
Transmittance decreases by 8–12% per 100 K temperature increase. Doubling gas column length reduces τv by 35–42% in characteristic bands.

3.5. Model Validation

In the previous section, a high-temperature gas spectral testing device was used to simulate the combustion environment of low-carbon chemicals, and temperature and concentration values close to the flame combustion field were set. By substituting the set temperature and concentration values into the self-constructed flame spectral radiation model (see Appendix A), the radiation spectra of the corresponding high-temperature mixture gas of combustion products were obtained through model calculation, and the model accuracy was compared and verified with the measured spectra of FT-IR spectrometer.
Previous work [52] established multi-scale correlations for CS2 combustion product concentrations through post-processing of thermal measurement data. The high-temperature zone concentration is derived via:
c 0 = c 1 V 1 V 2
c0: Gas concentration in flame zone, (V1 = 1 m3) c1: Measured concentration in sampling volume (V2 = 0.32 × 0.0752 × 3.14 m3 = 0.0175 m3).
Mishra’s scaling law extends this framework to SO2 emissions:
c S O 2 p p m = 909.5 d 0.5 t 0.1649
cSO2: SO2 concentration (ppm), d: Pool fire diameter (m), t: Combustion duration (s).
The experimental platform simulated combustion environments of low-carbon chemicals using temperature and concentration parameters matching real flame conditions. Model validation involved calculating radiation spectra through the developed framework and comparing results against FT-IR spectrometer measurements.
Validation cases included:
1. CS2 pool fire analogs (Peak temperature: 1023 K).
Mixture 1: CO2 0.8%/SO2 1%/N2 98.2%.
Mixture 2: CO2 0.5%/SO2 1%/N2 98.5%.
2. C2H3N pool fire analogs (Peak temperature: 1073 K)
Mixture 1: CO2 1%/NO 0.5%/N2 98.5%.
Mixture 2: CO2 0.8%/SO2 0.5%/N2 98.7%.
In this study, “model accuracy” is defined as the complement of the relative error of the peak area between the experimentally measured spectrum and the model-calculated spectrum: Accuracy = 1 − |(A(exp) − A(cal))/A(exp)| × 100%, where A(exp) is the experimental peak area and A(cal) is the model-calculated peak area. The baseline is the spectral baseline after nitrogen background correction at the same temperature.
Analysis of CS2 in Figure 13 reveals close alignment between experimental measurements and model calculations in spectral peak positions. At a gas composition of CO2:SO2:N2 = 0.5%:1%:98.5%, the measured radiance peak for CO2 at 4.35 μm was 0.087 W/(cm2·μm·sr), compared to the modeled value of 0.079 W/(cm2·μm·sr). For SO2 at 7.5 μm, the measured radiance peak was 0.03 W/(cm2·μm·sr) versus the modeled 0.029 W/(cm2·μm·sr). When the concentration was adjusted to CO2:SO2:N2 = 0.8%:1%:98.2%, the CO2 radiance peak at 4.35 μm measured 0.096 W/(cm2·μm·sr) (modeled: 0.093 W/(cm2·μm·sr)), while SO2 at 7.5 μm showed 0.032 W/(cm2·μm·sr) experimentally versus 0.029 W/(cm2·μm·sr) computationally. Model accuracy for CO2 characteristic bands reached 89.8–83.4% at 2.7 μm and 82.2–85.6% at 4.35 μm. For SO2 bands at 4 μm and 7.5 μm, accuracies were 82.2–85.6% and 96.7–90.6%, respectively.
For C2H3N in Figure 9, spectral peak positions showed strong agreement between measurements and models. At CO2:NO:N2 = 1%:0.5%:98.5%, CO2 radiance at 4.35 μm measured 0.112 W/(cm2·μm·sr) (modeled: 0.105 W/(cm2·μm·sr)), while NO at 5.5 μm exhibited 0.0048 W/(cm2·μm·sr) experimentally versus 0.0051 W/(cm2·μm·sr) computationally. At CO2:NO:N2 = 0.8%:0.5%:98.7%, CO2 radiance at 4.35 μm was 0.101 W/(cm2·μm·sr) (modeled: 0.106 W/(cm2·μm·sr)), with NO at 5.5 μm measuring 0.0046 W/(cm2·μm·sr) experimentally compared to 0.0051 W/(cm2·μm·sr) computationally. Model accuracies for CO2 reached 85.1–79.2% at 2.7 μm and 93.8–95.3% at 4.35 μm. For NO’s characteristic band at 5.5 μm, accuracies were 94.1–90.2%.
There are some band-specific uncertainties, which are reflected in Table 4.

4. Conclusions

This study established a dedicated experimental platform for analyzing high-temperature gas spectra in low-carbon chemical pool fires, with successful applications to carbon disulfide and C2H3N combustion. Complementary lens transmittance measurements (773–1073 K) effectively eliminated systemic errors in raw spectral data collection. Systematic experiments characterized individual gases (SO2, NO, NO2, CO2) and gas mixtures (CO2:SO2 = 1:2, CO2:N2 = 1:2, N2:SO2 = 1:2). Lens transmittance remained stable at ≈0.535 across 1.3–6.6 μm (773–1073 K), but significantly decreased at 6.65–12 μm wavelengths. Advanced signal processing combining airPLS baseline correction with Whittaker smoothing effectively preserved spectral trends while enhancing peak resolution, achieving optimal noise reduction without signal distortion. Distinct characteristic bands can be identified by self-made experimental platform and processing technology: CO2 (2.7, 4.35 μm), SO2 (4.05, 7.5, 9.0 μm), NO (5.5 μm), and NO2 (3.6, 6.3 μm). While minor spectral interference was observed in gas mixtures, comprehensive understanding of cross-gas interference mechanisms requires further investigation. Compared with real fire accidents, there are certain deviations in the assumptions of this model. When the local temperature of the fire is >1400 K, the NOx energy level distribution deviates from LTE (Local Thermodynamic Equilibrium), which will lead to a radiance deviation of ±5%. The temperature gradient of real flames can reach 50 K/cm, resulting in a peak height error of ±8% compared with the laboratory condition (<5 K/cm). The continuous absorption of soot in the 7–10 μm range in real fires will lead to a ±10% error in the SO2 peak height. But the developed radiation model demonstrated excellent agreement with experimental data, accurately replicating both spectral peak positions and profile shapes across validation scenarios. Model accuracy ranges: CS2 mixtures: 83.4–96.9%, C2H3N mixtures: 79.2–95.3%, CS2 flames: 81.2–95.6%, C2H3N flames: 83.4–96.9%.
When a real fire occurs, this study can provide some accurate recommended bands for spectral detection. For example, for CS2-type fires, 7.5 μm (SO2) is recommended as the main detection band; for nitrile-type fires, 5.5 μm (NO) is recommended as the main detection band. During the process, 4.35 μm (CO2) is detected as a fire identifier to confirm the fire, and the 2.7 μm band is detected to eliminate the interference of H2O.

5. Strengths and Limitations

This study developed an original high-temperature gas spectroscopy platform that successfully measured radiation spectra of combustion products under elevated temperatures. This experimental innovation enabled systematic analysis of spectral interference mechanisms between combustion byproducts. Numerical simulations achieved >95% accuracy in reproducing spectral curves across 773–1073 K, establishing a robust foundation for multi-temperature radiation modeling while effectively mitigating turbulence-related interference from flame dynamics. In terms of multi-component overlap of known target components, the model has strong processing capability and controllable accuracy. It can effectively handle the identified characteristic components (CO2, SO2, NO, NO2) in combustion flue gas. The absorption cross-sections of these components can be accurately obtained from the HITRAN database, and the model has completed basis function calibration for their typical overlapping bands (4.3–4.6 μm, 7–10 μm, 4.5–5.0 μm). When the component concentration is in the range of “detection limit − 10%” (SO2 0.1–10%, NO 0.05–5%), the separation error is <7% and the residual relative deviation is <8% (e.g., for the overlapping separation of CO2 and SO2 in the 4.3–4.6 μm band, the contribution ratio is 62%/38% and the residual is 0.008 W/(cm2·μm·sr)). Even if the component ratio changes dynamically (e.g., CO2:SO2 changes from 1:0.1 to 1:10), a reasonable solution can still be obtained through stoichiometric ratio constraints (e.g., CO2:SO2 ≈ 1:2 in CS2 combustion) and peak position uniqueness verification (e.g., calibration of the pure band of NO 5.5 μm), avoiding separation ambiguity caused by ratio fluctuations.
For chemical fire monitoring by UAVs (Unmanned Aerial Vehicles) or satellites, the accuracy level of 83–97% exactly covers the core needs of the entire process from “preliminary risk screening to precise emergency response”. It not only overcomes the industry pain point of “decision-making errors caused by low accuracy (<80%)” but also adapts to the scenario positioning of different platforms through hierarchical accuracy, providing key data support for the early detection, accurate identification, rapid response, and precise assessment of chemical fires.
While the developed model demonstrates high precision within controlled parameters, its current scope remains constrained to laboratory-scale conditions. Practical implementation faces some principal challenges: (1) Existing satellite spectral resolution and bandwidth capabilities fall short of experimental instrument precision, necessitating advancements in remote sensing technology. (2) Field applications must account for dynamic environmental variables including wind patterns, atmospheric conditions, and climatic factors that were intentionally excluded from current model parameters. (3) The model does not directly incorporate the processing of secondary byproducts of combustion, although it has the capability for extended processing. When encountering the molecular structure/absorption cross-section/characteristic band of unknown components, the basis functions cannot be calibrated, resulting in the lack of basis for NNLS deconvolution.

Author Contributions

G.J.: Formal analysis, Investigation, Validation, Writing—original draft; Z.C.: Conceptualization, Funding acquisition, Methodology, Resources, Supervision, Writing—review and editing; Y.L.: Formal analysis; P.L.: Formal analysis; Q.L.: Formal analysis; L.Z.: Formal analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Guangxi Key Technologies R&D Program (Grant No. AB25069093) and partially funded by the National Natural Science Foundation of China (General Program, Grant No. 21976043). This work was supported by the Guangxi Science and Technology Program (Guike AD25069074).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Acknowledgments

This research was made possible by the generous support of the Guangxi Key Laboratory of Environmental Pollution Control Theory and Technology, Guilin University of Technology, and the University Engineering Research Center of Watershed Protection and Green Development, Guangxi (Guilin 541006, China).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

import numpy as np
import matplotlib.pyplot as plt
from scipy.interpolate import CubicSpline
from scipy.interpolate import interp1d
plt.rcParams[’font.family’] = ’Microsoft YaHei’
a_1_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-773K.txt’
a_2_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-823K.txt’
a_3_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-873K.txt’
a_4_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-923K.txt’
a_5_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-973K.txt’
a_6_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-1023K.txt’
a_7_path = ’C:/Users/Administrator/Desktop/High temperature gas data processing/After spectral baseline correction and HITRAN smoothing filtering/C.S=0.5%.1%-New/C.S=0.5%.1%-1073K.txt’
a_1_data = np.loadtxt(a_1_path, dtype=float, usecols=1)[1:7292][::-1]
a_2_data = np.loadtxt(a_2_path, dtype=float, usecols=1)[1:7292][::-1]
a_3_data = np.loadtxt(a_4_path, dtype=float, usecols=1)[1:7292][::-1]
a_4_data = np.loadtxt(a_5_path, dtype=float, usecols=1)[1:7292][::-1]
a_5_data = np.loadtxt(a_6_path, dtype=float, usecols=1)[1:7292][::-1]
a_6_data = np.loadtxt(a_7_path, dtype=float, usecols=1)[1:7292][::-1]
temperatures = [500, 550, 650, 700, 750, 800]
spectra_data = [a_1_data, a_2_data,a_3_data, a_4_data, a_5_data, a_6_data]
input_temperature = int(input("Please enter the temperature: "))
index = np.argmax(np.array(temperatures) > input_temperature) - 1
temp1, temp2 = temperatures[index], temperatures[index + 1]
data1, data2 = spectra_data[index], spectra_data[index + 1]
cs = CubicSpline([temp1, temp2], [data1, data2], axis=0)
spectrum_fit = cs(input_temperature)
spectrum_reverse = np.flip(spectrum_fit)
y = np.loadtxt(a_3_path, dtype=float, usecols=1)[1:7292][::-1]
y1 = np.flip(y)
x = np.linspace(710, 8000, len(spectrum_reverse))
i = 10000 / x
plt.xlim(1, 10)
plt.plot(i, spectrum_reverse, color=’blue’, label=’CO2:SO2=0.5%:1%-873K Simulated spectrum’)
plt.plot(i, y1, ’--’, color=’red’, label=’CO2:SO2=0.5%:1%-873K Measured spectrum’)
plt.legend()
plt.xlabel("Wavelength/μm",)
plt.ylabel("Radiance/W/(cm$^2$μm sr)")
plt.show()

References

  1. Santana, J.A.D.; Arana, Y.C.; Gómez, O.G.; Furka, D.; Furka, S.; Orozco, J.L.; Di Benedetto, A.; Russo, D.; Portarapillo, M.; Febles, J.S. Fire and Explosion Economic Losses (FEEL) Index: A new approach for quantifying economic damages due to accidents in hydrocarbon storage sites. Process. Saf. Environ. Prot. 2022, 165, 77–92. [Google Scholar] [CrossRef]
  2. Shao, R.; Pan, H.; Huang, J. Safety risk assessment of chemical production process based on local and global objectives. J. Loss Prev. Process. Ind. 2022, 79, 104827. [Google Scholar] [CrossRef]
  3. Cui, W.; Wei, Y.; Ji, N. Global trends of waste-to-energy (WtE) technologies in carbon neutral perspective: Bibliometric analysis. Ecotoxicol. Environ. Saf. 2024, 270, 115913. [Google Scholar] [CrossRef]
  4. Gupta, P.; Shukla, A.K.; Shukla, D.P. Sentinel 2 based burn severity mapping and assessing post-fire impacts on forests and buildings in the Mizoram, a north-eastern Himalayan region. Remote. Sens. Appl. Soc. Environ. 2024, 36, 101279. [Google Scholar] [CrossRef]
  5. Raj, P.K. Large hydrocarbon fuel pool fires: Physical characteristics and thermal emission variations with height. J. Hazard. Mater. 2006, 140, 280–292, Erratum in J. Hazard. Mater. 2007, 149, 246. [Google Scholar] [CrossRef]
  6. Linstrom, P.J.; Mallard, W.G. The NIST Chemistry WebBook: A Chemical Data Resource on the Internet. J. Chem. Eng. Data 2001, 46, 1059–1063. [Google Scholar] [CrossRef]
  7. Saito, T.H.K.; Yanagisawa, M. Spectral Database for Organic Compounds (SDBS). 2006. Available online: https://sdbs.db.aist.go.jp/ (accessed on 8 October 2025).
  8. Stolarczyk, N.; Thibault, F.; Cybulski, H.; Jóźwiak, H.; Kowzan, G.; Vispoel, B.; Gordon, I.; Rothman, L.; Gamache, R.; Wcisło, P. Evaluation of different parameterizations of temperature dependences of the line-shape parameters based on ab initio calculations: Case study for the HITRAN database. J. Quant. Spectrosc. Radiat. Transf. 2019, 240, 106676. [Google Scholar] [CrossRef]
  9. Zhu, M.; Zhang, F.; Li, W.; Wu, Y.; Xu, N. The impact of various HITRAN molecular spectroscopic databases on infrared radiative transfer simulation. J. Quant. Spectrosc. Radiat. Transf. 2019, 234, 55–63. [Google Scholar] [CrossRef]
  10. Hashemi, R.; Gordon, I.E.; Tran, H.; Kochanov, R.V.; Karlovets, E.V.; Tan, Y.; Lamouroux, J.; Ngo, N.H.; Rothman, L.S. Revising the line-shape parameters for air- and self-broadened CO2 lines toward a sub-percent accuracy level. J. Quant. Spectrosc. Radiat. Transf. 2020, 256, 107283. [Google Scholar] [CrossRef]
  11. Karlovets, E.; Gordon, I.; Hashemi, R.; Kochanov, R.; Hargreaves, R.; Rothman, L. Addition of the line list for carbon disulfide to the HITRAN database: Line positions, intensities, and half-widths of the 12C32S2, 32S12C34S, 32S12C33S, and 13C32S2 isotopologues. J. Quant. Spectrosc. Radiat. Transf. 2020, 258, 107275. [Google Scholar] [CrossRef]
  12. Delahaye, T.; Armante, R.; Scott, N.; Jacquinet-Husson, N.; Chédin, A.; Crépeau, L.; Crevoisier, C.; Douet, V.; Perrin, A.; Barbe, A.; et al. The 2020 edition of the GEISA spectroscopic database. J. Mol. Spectrosc. 2021, 380, 111510. [Google Scholar] [CrossRef]
  13. Gordon, I.; Rothman, L.; Hargreaves, R.; Hashemi, R.; Karlovets, E.; Skinner, F.; Conway, E.; Hill, C.; Kochanov, R.; Tan, Y.; et al. The HITRAN2020 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 2021, 277, 107949. [Google Scholar] [CrossRef]
  14. Mellau, G.C.; Makhnev, V.Y.; Gordon, I.E.; Zobov, N.F.; Tennyson, J.; Polyansky, O.L. An experimentally-accurate and complete room-temperature infrared HCN line-list for the HITRAN database. J. Quant. Spectrosc. Radiat. Transf. 2021, 270, 107666. [Google Scholar] [CrossRef]
  15. Perrin, A.; Manceron, L.; Flaud, J.-M.; Kwabia-Tchana, F.; Armante, R.; Roy, P.; Doizi, D. The new nitrogen dioxide (NO2) linelist in the GEISA database and first identification of the ν1+2ν33 band of 14N16O2. J. Mol. Spectrosc. 2021, 376, 111394. [Google Scholar] [CrossRef]
  16. Alberti, M.; Weber, R.; Mancini, M.; Fateev, A.; Clausen, S. Validation of HITEMP-2010 for carbon dioxide and water vapour at high temperatures and atmospheric pressures in 450–7600 cm−1 spectral range. J. Quant. Spectrosc. Radiat. Transf. 2015, 157, 14–33. [Google Scholar] [CrossRef]
  17. Hargreaves, R.J.; Gordon, I.E.; Rothman, L.S.; Tashkun, S.A.; Perevalov, V.I.; Lukashevskaya, A.A.; Yurchenko, S.N.; Tennyson, J.; Müller, H.S. Spectroscopic line parameters of NO, NO2, and N2O for the HITEMP database. J. Quant. Spectrosc. Radiat. Transf. 2019, 232, 35–53. [Google Scholar] [CrossRef]
  18. Huang, X.; Freedman, R.S.; Tashkun, S.; Schwenke, D.W.; Lee, T.J. AI-3000K Infrared line list for hot CO2. J. Mol. Spectrosc. 2023, 392, 111748. [Google Scholar] [CrossRef]
  19. Yang, Y.; Peng, Z.; Zheng, D.; Zhang, M.; Ding, Y.; Du, Y. High-temperature line strength and line shape parameters measurements of Ar- and N2-perturbed CO2 lines near 4.18 µm in a shock tube. J. Mol. Spectrosc. 2024, 400, 111890. [Google Scholar] [CrossRef]
  20. Phillips, W.J.; Plemmons, D.H.; Galyen, N.A. HITRAN/HITEMP Spectral Databases and Uncertainty Propagation by Means of Monte Carlo Simulation with Ap-Plication to Tunable Diode Laser Absorption Diagnostics. 2011. Available online: https://itea.org/images/pdf/conferences/2011-Test-Instrumentation-presentations/ITEA%20Monte%20Carlo-TDLAS.pdf? (accessed on 8 October 2025).
  21. Wakatsuki, K.; Jackson, G.; Hamins, A.; Nyden, M. Effects of fuel absorption on radiative heat transfer in methanol pool fires. Proc. Combust. Inst. 2007, 31, 2573–2580. [Google Scholar] [CrossRef]
  22. Zhang, B.; Laboureur, D.M.; Liu, Y.; Gopalaswami, N.; Mannan, M.S. Experimental Study of a Liquefied Natural Gas Pool Fire on Land in the Field. Ind. Eng. Chem. Res. 2018, 57, 14297–14306. [Google Scholar] [CrossRef]
  23. Qian, Z.; Xiao-Jun, L.; Lei, H. Video Image Fire Recognition Based on Color Space and Moving Object Detection. In Proceedings of the 2020 International Conference on Artificial Intelligence and Computer Engineering (ICAICE), Beijing, China, 23–25 October 2020. [Google Scholar]
  24. Baek, J.; Alhindi, T.J.; Jeong, Y.-S.; Jeong, M.K.; Seo, S.; Kang, J.; Shim, W.; Heo, Y. Real-time fire detection system based on dynamic time warping of multichannel sensor networks. Fire Saf. J. 2021, 123, 103364. [Google Scholar] [CrossRef]
  25. Buffington, T.; Bilyaz, S.; Ezekoye, O.A. Brain-STORM: A deep learning model for computationally fast transient high-rise fire simulations. Fire Saf. J. 2021, 125, 103443. [Google Scholar] [CrossRef]
  26. Stoellinger, M.; Roekaerts, D. Using a flamelet generated manifold method in transported probability density function modeling of soot formation and thermal radiation. Comput. Fluids 2022, 245, 105567. [Google Scholar] [CrossRef]
  27. Tsai, K.-C.; Cheng, L.-H.; Yeh, C.-F.; Tseng, T.-P.; Huang, L.-J.; Yeh, S.-H.; Hsu, H.-T.; Lin, C.-H.; Lai, C.-H.; Brimblecombe, P.; et al. Fire dynamics, fire generated pollutants, and acute health implications: A scaling law analysis of heavy-fuel-oil pool fires. Fuel 2025, 381, 133210. [Google Scholar] [CrossRef]
  28. Wang, L.; Ding, P.; Yan, S.; Li, J.; Zhang, J.; Wang, C.; Cao, Z.; Zhou, Q.; Mu, J.; Meng, G.; et al. Soot volume fraction measurements in aero-engine model combustor outlet using two-color laser-induced incandescence. Chin. J. Aeronaut. 2025, 38, 103247. [Google Scholar] [CrossRef]
  29. Liu, H.-T.; Chen, Z.-L.; Yang, Y.; Yin, W.-Q.; Liu, Q. Research on Heat Transfer Characteristics in Oil Pool Fire Based on Spectral Analysis. Guang Pu Xue Yu Guang Pu Fen Xi 2016, 36, 3297–3302. [Google Scholar]
  30. Jiang, B.; Ji, Q.; Teng, G. Using GOCI to detect the diffusion of pollution after the TianJin harbour explosion. Int. J. Remote. Sens. 2017, 39, 1744–1753. [Google Scholar] [CrossRef]
  31. Hadjipanayis, M.; Beyrau, F.; Lindstedt, R.; Atkinson, G.; Cusco, L. Thermal radiation from vapour cloud explosions. Process. Saf. Environ. Prot. 2015, 94, 517–527. [Google Scholar] [CrossRef]
  32. Paul, C.; Roy, S.; Sailer, J.; Brännström, F.; Ahmed, M.M.; Trouvé, A.; Bordbar, H.; Hostikka, S.; McDermott, R. Detailed radiation modeling of two flames relevant to fire simulation using Photon Monte Carlo—Line by Line radiation model. J. Quant. Spectrosc. Radiat. Transf. 2024, 329, 109177. [Google Scholar] [CrossRef]
  33. Yang, C.S.-C.; Jin, F.; Trivedi, S.; Hommerich, U.; Nemes, L.; Samuels, A.C. Long-wave infrared laser-induced breakdown spectroscopy of complex gas molecules in the vicinity of a laser-induced plasma. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 294, 122536. [Google Scholar] [CrossRef]
  34. Zhang, K.; Ren, D.; Han, Y.; Zhu, C. Experimental and numerical study on the combustion and infrared radiation characteristics of Mg/PTFE/Viton pyrotechnic compositions. Infrared Phys. Technol. 2023, 133, 104841. [Google Scholar] [CrossRef]
  35. Parent, G.; Erez, G.; Collin, A.; Suzanne, M.; Thiry-Muller, A.; Weber, M.; Faure, E.; Boulet, P. Spectral radiation emitted by kerosene pool fires. Fire Saf. J. 2019, 108, 102847. [Google Scholar] [CrossRef]
  36. Bordbar, H.; Alinejad, F.; Conley, K.; Ala-Nissila, T.; Hostikka, S. Flame detection by heat from the infrared spectrum: Optimization and sensitivity analysis. Fire Saf. J. 2022, 133, 103673. [Google Scholar] [CrossRef]
  37. Raj, P.K. LNG fires: A review of experimental results, models and hazard prediction challenges. J. Hazard. Mater. 2006, 140, 444–464, Corrigendum in J. Hazard. Mater. 2007, 143, 603. [Google Scholar] [CrossRef] [PubMed]
  38. Vinnem, J.E. Risk analysis and risk acceptance criteria in the planning processes of hazardous facilities—A case of an LNG plant in an urban area. Reliab. Eng. Syst. Saf. 2010, 95, 662–670. [Google Scholar] [CrossRef]
  39. Qi, R.; Raj, P.K.; Mannan, M.S. Underwater LNG release test findings: Experimental data and model results. J. Loss Prev. Process. Ind. 2011, 24, 440–448. [Google Scholar] [CrossRef]
  40. Niu, Q.; He, Z.; Dong, S. IR radiation characteristics of rocket exhaust plumes under varying motor operating conditions. Chin. J. Aeronaut. 2017, 30, 1101–1114. [Google Scholar] [CrossRef]
  41. Qiao, Y.; Nie, W.S.; Wu, G.Y.; Feng, S.J. Study on Radiation Characteristics of Multi-Nozzle Liquid Rocket Engine Plumes. Missiles Space Veh. 2017, 4, 53–58. (In Chinese) [Google Scholar]
  42. Yang, B.; Li, W.; Xie, J.; Zhu, X.; Wang, F.; Yang, Y.; Li, Z. Hydrogen sulfide measurement of combustion gaseous product using ultraviolet absorption spectroscopy. Measurement 2023, 214, 112766. [Google Scholar] [CrossRef]
  43. Huang, W.; Gao, G.; Zhang, M.; Ruan, S.; Cai, T. Simultaneous measurement of temperature and C2H4 concentration in hydrocarbon flames using interference-immune differential absorption spectroscopy at 3.3 μm. Infrared Phys. Technol. 2024, 142, 105556. [Google Scholar] [CrossRef]
  44. Chang, K.; Liu, M.; He, Z.; Luo, Z.; Cheng, Q. Radiative characteristics detection and generation mechanism analysis of soot and PAHs during the combustion process of raw coal and tar coal. Fuel 2025, 382, 133767. [Google Scholar] [CrossRef]
  45. Ruan, S.; Gao, G.; Yang, Y.; Yang, H.; Cai, T. Measurement and validation of high-temperature CO spectral line parameters near 2.3 μm using TDLAS for combustion diagnostics. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2025, 329, 125577. [Google Scholar] [CrossRef]
  46. JJG 856-2015; Radiation Thermometers. General Administration of Quality Supervision, Inspection and Quarantine (AQSIQ): Beijing, China, 2015.
  47. Karman, T.; Gordon, I.E.; van der Avoird, A.; Baranov, Y.I.; Boulet, C.; Drouin, B.J.; Groenenboom, G.C.; Gustafsson, M.; Hartmann, J.-M.; Kurucz, R.L.; et al. Update of the HITRAN collision-induced absorption section. Icarus 2019, 328, 160–175. [Google Scholar] [CrossRef]
  48. Parra-Verde, E.R.; Gutiérrez-Vega, J.C. Steady-state solutions of the Whittaker–Hill equation of fractional order. Commun. Nonlinear Sci. Numer. Simul. 2024, 131, 107812. [Google Scholar] [CrossRef]
  49. Mishra, S.; Vishwakarma, P.K.; Sharma, A.; Mishra, K.B. Experimental investigation of small-scale CS2 (carbon disulphide) pool fires. Process. Saf. Environ. Prot. 2021, 145, 203–210. [Google Scholar] [CrossRef]
  50. Alzueta, M.U.; Guerrero, M.; Millera, Á.; Marshall, P.; Glarborg, P. Experimental and kinetic modeling study of oxidation of acetonitrile. Proc. Combust. Inst. 2021, 38, 575–583. [Google Scholar] [CrossRef]
  51. Salem, S.I.; Higa, H.; Ishizaka, J.; Pahlevan, N.; Oki, K. Spectral band-shifting of multispectral remote-sensing reflectance products: Insights for matchup and cross-mission consistency assessments. Remote. Sens. Environ. 2023, 299, 113846. [Google Scholar] [CrossRef]
  52. Peng, H.; Ning, J.; Chen, Z.; Tang, J.; Liu, L.; Chen, L. Construction of CS2 Combustion Flame Spectral Radiation Model and Inversion of Characteristic Pollution Product Concentration. Spectrosc. Spectr. Anal. 2022, 42, 672–677. [Google Scholar]
Figure 1. High-Temperature Gas Spectral Testing Platform.
Figure 1. High-Temperature Gas Spectral Testing Platform.
Toxics 13 00877 g001
Figure 2. Lens temperature variation during device heating.
Figure 2. Lens temperature variation during device heating.
Toxics 13 00877 g002
Figure 3. Changes in transmittance of high-temperature lenses at different temperatures.
Figure 3. Changes in transmittance of high-temperature lenses at different temperatures.
Toxics 13 00877 g003
Figure 4. Temperature Distribution Analysis in the Experimental Setup for Low-Carbon Chemical Pool Fires. The internal temperature of the high-temperature gas spectral testing device gradually decreases from the core heating zone outward, while the temperature of the low-carbon chemical pool fire at the center of the flame is lower than the outer flame temperature, and the overall temperature distribution increases first and then decreases from the flame core outward. The temperature structures of the two are generally similar and both have the property of left-right symmetry. Therefore, the same spectral radiation model can be used to calculate the single-sided gas spectral parameters according to the symmetrical structure during the calculation process.
Figure 4. Temperature Distribution Analysis in the Experimental Setup for Low-Carbon Chemical Pool Fires. The internal temperature of the high-temperature gas spectral testing device gradually decreases from the core heating zone outward, while the temperature of the low-carbon chemical pool fire at the center of the flame is lower than the outer flame temperature, and the overall temperature distribution increases first and then decreases from the flame core outward. The temperature structures of the two are generally similar and both have the property of left-right symmetry. Therefore, the same spectral radiation model can be used to calculate the single-sided gas spectral parameters according to the symmetrical structure during the calculation process.
Toxics 13 00877 g004
Figure 5. Four types of single high-temperature gas spectra ((a)—CO2, (b)—NO, (c)—NO2, (d)—SO2, 1—Measured spectra, 2—HITRAN database spectra) Experimental conditions: FT-IR resolution 0.12 cm−1 (7.1–8 μm band), number of scans 32, temperature range 773–1073 K, pressure 1 atm, optical path 18 cm; gas purity: CO2 99.99%, SO2 99.9%, NO 99.5%, NO2 99.5%.
Figure 5. Four types of single high-temperature gas spectra ((a)—CO2, (b)—NO, (c)—NO2, (d)—SO2, 1—Measured spectra, 2—HITRAN database spectra) Experimental conditions: FT-IR resolution 0.12 cm−1 (7.1–8 μm band), number of scans 32, temperature range 773–1073 K, pressure 1 atm, optical path 18 cm; gas purity: CO2 99.99%, SO2 99.9%, NO 99.5%, NO2 99.5%.
Toxics 13 00877 g005
Figure 6. Position of spectral characteristic peak. The band marked in red indicates that it is considered a key band that can be used for identifying characteristic pollutants. The black marked band indicates that it is considered a secondary band of characteristic pollutants, but can still be used to identify pollutants.
Figure 6. Position of spectral characteristic peak. The band marked in red indicates that it is considered a key band that can be used for identifying characteristic pollutants. The black marked band indicates that it is considered a secondary band of characteristic pollutants, but can still be used to identify pollutants.
Toxics 13 00877 g006
Figure 7. Spectral data of three specific ratios of mixed gases ((a)—CO2:N2 = 1:2, (b)—CO2:SO2 = 1:2, (c)—N2:SO2 = 1:2). Experimental temperature 1023 K, background gas is 99.999% nitrogen. ’1 represents the spectral data obtained from the experiment, and ’2 represents the spectral data in the HITRAN database.
Figure 7. Spectral data of three specific ratios of mixed gases ((a)—CO2:N2 = 1:2, (b)—CO2:SO2 = 1:2, (c)—N2:SO2 = 1:2). Experimental temperature 1023 K, background gas is 99.999% nitrogen. ’1 represents the spectral data obtained from the experiment, and ’2 represents the spectral data in the HITRAN database.
Toxics 13 00877 g007
Figure 8. Quantitative residual plots ((a)—4.3–4.6 μm (CO2), (b)—7–10 μm (SO2); temperature—1023 K).
Figure 8. Quantitative residual plots ((a)—4.3–4.6 μm (CO2), (b)—7–10 μm (SO2); temperature—1023 K).
Toxics 13 00877 g008
Figure 9. Relationship between Peak Height and Mixing Ratio (L-SO2, R-CO2).
Figure 9. Relationship between Peak Height and Mixing Ratio (L-SO2, R-CO2).
Toxics 13 00877 g009
Figure 10. Spectral profiles of combustion products, CS2 analogs: (a) (CO2:SO2 = 0.8%:1%), (a′) (CO2:SO2 = 0.8%:1%),(a″) (actual CS2 pool fire), C2H3N analogs: (b) (CO2:NO = 1%:0.5%), (b′) (CO2:NO = 0.8%:0.5%), (b″) (actual C2H3N pool fire).
Figure 10. Spectral profiles of combustion products, CS2 analogs: (a) (CO2:SO2 = 0.8%:1%), (a′) (CO2:SO2 = 0.8%:1%),(a″) (actual CS2 pool fire), C2H3N analogs: (b) (CO2:NO = 1%:0.5%), (b′) (CO2:NO = 0.8%:0.5%), (b″) (actual C2H3N pool fire).
Toxics 13 00877 g010
Figure 11. Effect of Resolution Convolution on Peak Height/Peak Area (1023 K CS2).
Figure 11. Effect of Resolution Convolution on Peak Height/Peak Area (1023 K CS2).
Toxics 13 00877 g011
Figure 12. Effect of Temperature on Peak Height (Acetonitrile NO 5.5 μm).
Figure 12. Effect of Temperature on Peak Height (Acetonitrile NO 5.5 μm).
Toxics 13 00877 g012
Figure 13. Comparison of Accuracy of High-Temperature Gas Spectral Radiation Models under Simulated Combustion Environment.
Figure 13. Comparison of Accuracy of High-Temperature Gas Spectral Radiation Models under Simulated Combustion Environment.
Toxics 13 00877 g013
Table 1. Technical Parameters of FT-IR Spectrometer.
Table 1. Technical Parameters of FT-IR Spectrometer.
Parameter NameParameter Value
Spectral Range350–8300 cm−1
Spectral Resolution0.5, 1, 2, 4, 8, 16, 32 cm−1
Number of Scans2, 4, 8, 16 times
Field of View10–15 mrads
DetectorDTGs detector, liquid nitrogen cooled MCT detector
Measurement Speed1–25 spectra per second
Scan TypeBackground, interferogram, sample, single-beam
Table 2. Lens Transmittance Characterization Error Table.
Table 2. Lens Transmittance Characterization Error Table.
Error Source1.3–6.6 μm8.6–12 μm
Lens Transmittance Measurement Error±0.012±0.042
E3 Measurement Error±5%±5%
Final Radiance Uncertainty±7.3%±9.8%
Table 3. Quantitative indicators of single-gas spectra and HITRAN simulations at 1023 K.
Table 3. Quantitative indicators of single-gas spectra and HITRAN simulations at 1023 K.
GasCharacteristic BandRMSEMAEPearson
r
Peak Center Error Δλ (μm) Peak Height Error (%)
CO22.70.0080.0060.987±0.02±3.2
CO24.350.0120.0090.982±0.03±4.5
SO27.50.0050.0040.991±0.01±2.8
NO5.50.0030.0020.995±0.01±1.9
NO23.60.0150.0110.973±0.04±6.7
Table 4. Uncertainty of model radiance.
Table 4. Uncertainty of model radiance.
Source of UncertaintyValue
Radiometric calibration±7.3%
Baseline calibration (AirPLS)±3.5%
Lens transmittance±4.2%
HITRAN parameters±5.1%
Total±10.2%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiang, G.; Chen, Z.; Liang, Y.; Li, P.; Liu, Q.; Zhou, L. Investigating the Spectral Characteristics of High-Temperature Gases in Low-Carbon Chemical Pool Fires and Developing a Spectral Model. Toxics 2025, 13, 877. https://doi.org/10.3390/toxics13100877

AMA Style

Jiang G, Chen Z, Liang Y, Li P, Liu Q, Zhou L. Investigating the Spectral Characteristics of High-Temperature Gases in Low-Carbon Chemical Pool Fires and Developing a Spectral Model. Toxics. 2025; 13(10):877. https://doi.org/10.3390/toxics13100877

Chicago/Turabian Style

Jiang, Gengfeng, Zhili Chen, Yaquan Liang, Peng Li, Qiang Liu, and Lv Zhou. 2025. "Investigating the Spectral Characteristics of High-Temperature Gases in Low-Carbon Chemical Pool Fires and Developing a Spectral Model" Toxics 13, no. 10: 877. https://doi.org/10.3390/toxics13100877

APA Style

Jiang, G., Chen, Z., Liang, Y., Li, P., Liu, Q., & Zhou, L. (2025). Investigating the Spectral Characteristics of High-Temperature Gases in Low-Carbon Chemical Pool Fires and Developing a Spectral Model. Toxics, 13(10), 877. https://doi.org/10.3390/toxics13100877

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