Elsevier

Combustion and Flame

Volume 226, April 2021, Pages 200-210
Combustion and Flame

An experimental investigation of furfural oxidation and the development of a comprehensive combustion model

https://doi.org/10.1016/j.combustflame.2020.12.015Get rights and content

Abstract

The oxidation of furfural has been studied experimentally in a jet-stirred reactor (JSR) under fuel-lean (Φ = 0.4) and fuel-rich conditions (Φ = 2.0) in the temperature range of 650–950 K; in addition, laminar burning velocity data have been measured at T = 473 K and p = 1 bar within a wide fuel-air range. From the JSR experiments, 13 species profiles have been identified and quantified by GC–MS and GC. A detailed kinetic reaction model involving 382 species and 2262 reactions was developed by exploiting the experimental data base provided within the present work as well as experimental data reported in literature. The rate coefficients of reactions of H abstraction, H addition as well as of decomposition of furfural were calculated by quantum chemical methods at CBS-QB3 level. A general agreement was achieved when simulating the experimental data. Rate of production analysis as well as sensitivity analysis were performed to get a deeper insight into the combustion of furfural, e.g. for the jet-stirred reactor data at around 50% fuel conversion, as well as sensitivity analysis of laminar flame speeds conducted for a fuel-air ratio Φ = 0.9, 1.2, and 1.6. According to the analysis, the main consumption pathways of furfural oxidation were identified as H abstraction reactions of the R-CHO (aldehyde) group by H, OH, O, and HO2 to produce a furfural radical (furfural-6). At pyrolysis condition, the dominant pathways within the furfural decay were found to occur via ring opening by splitting the Csingle bondO bond followed by isomerization to form α-pyrone (C5H4O2). Even more, the measured laminar flame speed data are well reproduced by the reaction model developed within the present work. The experimental data base as well as the developed reaction model will assist and contribute to a more detailed understanding of the combustion behavior of furfural and of furan derivatives as well.

Introduction

Nowadays, new energy sources are urgently needed because of the worldwide energy shortage and environmental pollution caused by the use of fossil fuels. Biofuels in general, such as ethanol or biodiesel [1,2], have attracted much attention due to their renewability (enabling a lower CO2 footprint) and large storage capacity, besides further benefits to the environment by a reduced emission pattern, e.g. lower amounts of soot particles (in the case of oxygenated fuels). In this context, furan derivatives (see Fig. 1), produced from non-edible feedstocks, especially 2-methylfuran (MF) and 2,5-dimethylfuran (DMF), were demonstrated to have additional benefits as next generation alternative fuels due to their significant advantages over traditional biofuels [3], [4], [5], [6], [7]. For instance, DMF-diesel blends produce less soot compared to blends of diesel and n-butanol [3,5]. Furthermore, the octane number of DMF is higher than that of ethanol [8]. Also, the heating values of furan, MF, and DMF are larger than that of ethanol [4] and thus, similar to that of gasoline. Sudholt et al. [9] investigated derived cetane numbers (DCN) of several furan derivatives, with furan, MF, DMF, among other, to characterize their ignition behavior. According to DCN and bond dissociation energies (BDE) at CBS-QB3 level, they found that the ignition behavior of the furan derivatives is governed by the ring structure, while the chain structure has a negligible effect.

Furfural is generally produced from C5 sugars (such as arabinose and xylose) being constituents of hemicellulose derivative [10]. Furfural is a feedstock of several furan derivatives [4,8] and thus, an important intermediate within furan derivatives oxidation [11]. Moreover, furfural, as a non-petroleum based chemical feedstock, is also a precursor for the production of valerate esters, pentanols, and 1-octanol as well as various C10–C15 coupling products [12], [13], [14]. Furfural also shows some interesting combustion properties. Within furfural combustion, a reduced production of soot was reported compared to hydrocarbon fuels [15]. Blending furfural in the tri-propylene glycol mono-methyl ether (C10H22O4), being a possible fuel additive, was shown to enhance the ignition reactivity compared to blending with furan or methyl substituted furans [16]. Therefore, studies focusing on the oxidation process at lean conditions are necessary for studying the combustion properties of furfural as an additive to diesel. Furthermore, furfural is also an important compound within the thermal conversion process of biomass [17], [18], [19], [20], [21]. Pyrolysis experiments as well as those at fuel-rich conditions are needed to investigate its conversion pathways. Hence, research on furfural combustion is meaningful for its efficient usage as well as to understand more comprehensively the combustion of furan derivatives and to assess better biomass pyrolysis.

Some earlier work has been performed to identify the major furfural consumption pathways, through theoretical calculation and experiments, as summarized in Table 1. Zhao et al. [22] reported that the furfural oxidation which they studied at ambient pressure was initiated via hydroxyl radicals reactions by using quantum chemistry and kinetic calculations. The furfural oxidation was also investigated by Thorton et al. [23] in a JSR combined with gas chromatography (GC), within temperatures between 1000 and 1300 K at atmospheric pressure. They identified as the main fuel consumption pathway the H abstraction of furfural forming a furfural radical which reacts further decomposing into a furan radical and CO. Due to the lack of high-precision experimental measurements and quantum chemistry calculations, Thorton et al. [23] proposed a preliminary mechanism for furfural oxidation; however, no detailed kinetic model was provided. Vasiliou et al. [24] investigated the thermal decomposition of furfural in a heated micro tubular flow reactor within the temperature range of 1200–1800 K, at 75–150 torr. They found that the ring-opening pathway has a lower energy barrier compared to the formation pathway of carbenic intermediates and proposed the initial fuel decomposition process to be an unimolecular decomposition reaction step: furfural (+ M) = furan + CO (+ M). Vermeire et al. [25] studied the furfural pyrolysis at atmospheric pressure in a jet-stirred reactor (JSR) combined with GC and GC-mass spectrometry (GC–MS). The important intermediates α-pyrone and furan were observed and quantified, besides others. Based on experimental results and the calculated potential energy surface of furfural initiation reactions, they concluded that furfural decomposes mainly through a ring-opening isomerization step forming formylvinylketene (cis-O=C=CH–CH=CH–CH=O). Li et al. [26] investigated the thermal decomposition of furfural in a flow tube reactor at 30 torr by synchrotron vacuum ultraviolet photoionization and mass spectrometry. Also, they calculated the H addition and H abstraction reactions via H radical at the CBS-QB3 level.

Although there are several studies available on furfural combustion as discussed above, none of them was dedicated to studying the detailed decomposition pathways of furfural. Thus, additional investigation on furfural is desirable to be able to predict more comprehensively its combustion behavior and to reveal its dominant reaction pathways by the construction of a comprehensive chemical-kinetic reaction model being validated by experimental data.

In the present work, furfural oxidation was investigated in a JSR at atmospheric pressure in the temperature range of 650–950 K, at two equivalence ratios Φ: 0.4 (fuel-lean) and 2.0 (fuel-rich). Furthermore, the laminar burning velocity of furfural in air was measured over a Φ-range from 0.6 to 1.8 at a preheat temperature of T = 473 K and at ambient pressure, p = 1 bar. Based on the experimental data, a comprehensive detailed chemical kinetic reaction model was developed to simulate the furfural oxidation (present work) and its pyrolysis (experimental data taken from Vermeire et al. [25]) as well as the laminar flame speed data (present work). Also, rate of production (ROP) and sensitivity analyses were used to identify the dominant consumption pathways of furfural. Furthermore, reaction rate expressions of major initiation reactions within furfural consumption have been determined by applying quantum chemical calculations, at CBS-QB3 level found to be appropriate for these types of fuels [29], [30], [31], [32].

Section snippets

Jet-stirred reactor experiment

The oxidation of furfural was carried out in a jet-stirred reactor developed and constructed at Chinese Academy of Sciences (CAS). The experimental setup and procedure have been described in detail in previous studies [33,34]; hence, only a brief description is given here. The volume of the JSR is 64.3 cm3, its inlet diameter 50 mm, respectively. For the experimental conditions applied in the present work, the average of the reaction time (τ) was around 1.50 s; the total flow rate was 1000 sccm

Modeling

The furfural oxidation studied within the JSR was simulated using the PSR code of Chemkin-Pro [40]. The laminar flame speed was calculated using Cantera [41] assuming a free flame and taking into account the multi-component diffusion model as well as thermo-diffusion. Mesh points were refined to achieve equal solution tolerance leading to about 130 mesh points. A detailed kinetic mechanism involving 382 species and 2262 reactions was newly developed (see also SM). This mechanism considers H

Results and discussion

In the oxidation process of furfural, the major species determined from the JSR experiments were CO and CO2. Several minor species including hydrocarbons (methane, acetylene, ethylene, propene, allene, and benzene) and oxygenated species (acetaldehyde, acrolein, and furan) were also measured. Therein, acetaldehyde (CH3CHO) and acrolein (C2H3CHO) were newly detected within the oxidation of furfural. The measured mole fractions of furfural and products as well as the predicted ones including

Conclusions

Furfural oxidation was experimentally investigated in a jet stirred reactor at two fuel equivalence ratios (Φ = 0.4 and Φ = 2.0) over a temperature range of 650–950 K at atmospheric pressure. It was found that the peak values of light hydrocarbon intermediates and oxygenated intermediates as quantitatively detected and determined by GC increased with decreasing equivalence ratio. Furthermore, the measurement of the laminar burning velocity was performed at ambient pressure (1 bar) and a high

Declaration of Competing Interest

None.

Acknowledgments

The authors thank for financial support from the National Natural Science Foundation of China (No. 51976216/51888103), the Ministry of Science and Technology of China (2017YFA0402800), Beijing Municipal Natural Science Foundation (20JQC0019) and the Recruitment Program of Global Youth Experts. The support from the Alexander-von-Humboldt Research Group Linkage Program is gratefully acknowledged.

References (53)

  • S. Wang et al.

    Lignocellulosic biomass pyrolysis mechanism: a state-of-the-art review

    Prog. Energy Combust. Sci.

    (2017)
  • Z. Li et al.

    A theoretical study on the mechanism of xylobiose during pyrolysis process

    Comput. Theor. Chem.

    (2017)
  • M.M. Thorton et al.

    Oxidation of furan and furfural in a well-stirred reactor

    Symp. (Int.) Combust.

    (1988)
  • K.P. Somers et al.

    A comprehensive experimental and detailed chemical kinetic modelling study of 2,5-dimethylfuran pyrolysis and oxidation

    Combust. Flame

    (2013)
  • L. Zhou et al.

    A detailed kinetic study on oxidation of benzyl alcohol

    Combust. Flame

    (2019)
  • Y.X. Liu et al.

    Combustion study of a surrogate jet fuel

    Combust. Flame

    (2019)
  • J.J. Weng et al.

    Experimental and kinetic investigation of 1,2,4-trimethylbenzene oxidation at low temperature

    Proc. Combust. Inst.

    (2017)
  • B.Y. Wang et al.

    New insights in the low-temperature oxidation of acetylene

    Proc. Combust. Inst.

    (2017)
  • S. Richter et al.

    Experimental and modeling study of farnesane

    Fuel

    (2018)
  • T. Kick et al.

    An experimental and modeling study of burning velocities of possible future synthetic jet fuels

    Energy

    (2012)
  • G. Andrews et al.

    Determination of burning velocities: a critical review

    Combust. Flame

    (1972)
  • Y.X. Liu et al.

    A wide-range experimental and modeling study of oxidation and combustion of n-propylbenzene

    Combust. Flame

    (2018)
  • Z.Y. Tian et al.

    An experimental and kinetic investigation of premixed furan/oxygen/argon flames

    Combust. Flame

    (2011)
  • D. Liu et al.

    Combustion chemistry and flame structure of furan group biofuels using molecular-beam mass spectrometry and gas chromatography – part I: Furan

    Combust. Flame

    (2014)
  • Q. Zhao et al.

    Pressure-dependent kinetics on benzoyl radical+ O2 and its implications for low temperature oxidation of benzaldehyde

    Combust. Flame

    (2020)
  • K.P. Somers et al.

    A high temperature and atmospheric pressure experimental and detailed chemical kinetic modelling study of 2-methyl furan oxidation

    Proc. Combust. Inst.

    (2013)
  • Cited by (15)

    • Experimental and kinetic modeling study of ethylene oxidation at elevated pressure

      2022, Combustion and Flame
      Citation Excerpt :

      The separated products and intermediates will be detected by quadrupole mass spectrometry. The results will be compared with the NIST database to achieve qualitative detection of oxygenated species and qualified using the effective carbon number method [21–23]. The detection thresholds of FID and TCD detectors are 0.1 and 10 ppm, respectively.

    View all citing articles on Scopus
    1

    Both contributed equally.

    View full text