Dynamics of laminar ethylene lifted flame with ozone addition

https://doi.org/10.1016/j.proci.2020.09.027Get rights and content

Abstract

The effect of ozone (O3) addition on ethylene (C2H4) non-premixed jet flames is investigated. Stable C2H4 lifted flames are established with oxygen/nitrogen co-flow at reduced oxygen content conditions. It is observed in the experiments that after O3 addition, the flame liftoff height could either increase or decrease, depending on the initial value of the flame liftoff height before O3 is added. Formaldehyde (CH2O) planar laser-induced fluorescence (PLIF) measurement shows that prompt ozonolysis reaction between C2H4 and O3 produces large amounts of CH2O upstream of the flame. In contrast to previous studies of O3 addition on lifted flames—with saturated hydrocarbon fuels in which O3 decomposition dominates—the ozonolysis reaction between C2H4 and O3 changes the chemical composition of flow even at room temperature. Such chemical reaction causes the simultaneous increase of both the triple flame propagation speed of lifted flame and the axial jet velocity along the stoichiometric contour, which also therefore changes the dynamic balance between these two values to stabilize the flame. While the increase of triple flame propagation speed tends to decrease the flame liftoff height, the increase of axial jet velocity along the stoichiometric contour tends to have the opposite effect. A competing kinetic-dynamic process forms, and the final location of the flame depends on the degree of ozonolysis reaction, which is determined by the initial flame liftoff height before O3 is added.

Introduction

Ozone (O3) addition has great potential in actively enhancing and controlling combustion [1], [2], [3]–4]. Considerable increases in flame propagation speeds have been observed with O3 addition. These experiments include those using both non-premixed laminar lifted flame [5] and premixed Bunsen burner [6], [7], [8], [9]–10] configurations with saturated hydrocarbon fuels, such as methane (CH4), propane (C3H8) and syngas (H2/CO). It is well accepted that the production of O atoms through O3 decomposition within pre-heat zone plays a vital role in the observed flame speed enhancement, while the reactions between O3 and saturated fuels are virtually negligible. However, with unsaturated hydrocarbon fuels, the addition of O3 could lead to an entirely different outcome, owing to the prompt exothermic reaction between O3 and fuel, i.e., the ozonolysis reaction. The difference between the reaction rate constants of alkene and alkane fuels with O3 could be as large as kalkene,O3/kalkane,O3106 [11]. Recently, Gao et al. [11] observed direct auto-ignition using a non-premixed ethylene (C2H4) jet with O3 addition in an oxidizer co-flow at room temperature and pressure conditions without an external ignition source. Significant enhancement on flame propagation by continuous formation of auto-ignition kernels upstream was observed. This unique phenomenon indicated that the ozonolysis reaction could be considered as an alternative means to modify combustion processes at low-temperature conditions. Different from conventional fuel oxidation which is confined in the flame zone, the high reactivity between unsaturated hydrocarbon and O3 renders immediate reaction and heat release far upstream upon mixing, and therefore affecting the flame both kinetically and dynamically. Such a feature (combustion region) has not been studied before.

The influence of O3 addition on laminar flame speed (SL) of C2H4 was investigated in premixed configurations [10,12,13]. It was reported that at atmospheric conditions, adding O3 would decrease SL, while at sub-atmospheric conditions, enhancement in SL was measured. Based on kinetic simulations, Gao et al. [10] concludes that the reaction of C2H4/O3 has a positive effect on SL by producing reactive species such as formaldehyde (CH2O) and hydrogen (H2). Meanwhile, the premixed configuration induces continuous heat loss to the burner wall which decreases the total enthalpy of the system, thus lowering SL. At low-pressure, the ozonolysis reaction is suppressed given a limited residence time upstream of the flame zone. Therefore, SL would be increased via O3 decomposition, similar to the case with alkanes. Nevertheless, the effect of the ozonolysis reaction on flame dynamics has not been investigated in detail.

In this work, the effect of O3 addition on C2H4 flame dynamics is studied using a non-premixed laminar co-flow jet flame burner. The lifted laminar jet flame has been studied extensively [14], [15], [16], [17], [18], [19]–20]. The stabilization of lifted flames is achieved by the dynamic balance between the propagation speed of the triple flame (Stri), which is located at the base of the lifted flame, and the axial flow velocity (ust) just upstream of the flame along the stoichiometric contour. Therefore, the liftoff height (HL) of the flame could be a sensitive indicator of the change in either Stri or upstream condition. In our experiments, by tuning the co-flow composition, a stable C2H4 lifted flame is obtained in the laminar regime. Various flame dynamic behaviors are observed as a function of fuel jet velocity (uf). Specifically, in the low uf region, HL decreases with O3 addition, while instead it increases in high uf region. Planar laser-induced fluorescence (PLIF) of CH2O and numerical simulations are performed to understand the highly coupled kinetic and dynamic process.

Section snippets

Experiment setup

Figure 1 schematically illustrates the experimental system, which consists of a laminar jet flame burner and the PLIF set-up. All experiments are undertaken at room temperature and pressure (T = 298 K, P = 101 kPa). The burner used here is the same one used in the study by Ombrello et al. [5,21] comprising a slender, aerodynamically shaped quartz fuel nozzle with 0.69 mm inner diameter at the exit on the tip, and a cylindrical quartz shield with inner diameter of 90 mm and length of 460 mm to

Computational framework

Numerical simulations are conducted using ANSYS Fluent [22] with the SIMPLEC algorithm. An axisymmetric 2-D computational domain is adopted with a size (axial × radial) of 450 mm × 45 mm. A gravitational acceleration of 9.79 m/s2 is given in the negative axial direction so that buoyancy effect could be accounted for. The built-in CHEMKIN-CFD solver of laminar finite-rate chemistry is enabled. The detailed USC Mech II model [23] is firstly reduced (for shorter computational time) using Global

Stable lifted flames without O3 addition

Stable C2H4 laminar lifted flames are obtained with pure fuel jet and the O2/N2 mixture as co-flow. Figure 2 shows a series of photographs of stable C2H4 lifted flames at different uf but constant co-flow velocity, UCO = 0.013 m/s, and constant oxidizer composition of 12.7% O2 + 87.3% N2. The uf changes from 2.98 m/s as an attached diffusion flame to finally 5.23 m/s.

As the flame detaches from the fuel nozzle, a clear triple flame structure appears with a relatively weak lean premixed wing and

Conclusions

In this study, the effect of O3 addition on non-premixed C2H4 lifted jet flames is investigated, using experiments, including PLIF of CH2O, and numerical simulations. Stable, laminar lifted C2H4 flames are established in a co-flow geometry and a wide range of liftoff height is obtained. Numerical simulations and experimental results are in qualitative agreement. With O3 addition in the co-flow, the change of flame liftoff height shows sensitivity on its initial value. If the initial flame

Declaration of Competing Interest

None.

Acknowledgments

WS acknowledges Prof. Yiguang Ju from Princeton University for the loan of the burner. This work is supported by the Air Force Office of Scientific Research (FA9550-16-1-0441)

References (31)

  • T. Ombrello et al.

    Combust. Flame

    (2010)
  • Z.H. Wang et al.

    Combust. Flame

    (2012)
  • X. Liang et al.

    Int. J. Hydrogen Energy

    (2013)
  • T.M. Vu et al.

    Combust. Flame

    (2014)
  • X. Gao et al.

    Combust. Flame

    (2015)
  • S. Chung et al.

    Combust. Flame

    (1991)
  • B. Lee et al.

    Combust. Flame

    (1997)
  • Y.-C. Chen et al.

    Combust. Flame

    (2000)
  • S. Ghosal et al.

    Combust. Flame

    (2001)
  • S. Chung

    Proc. Combust. Inst.

    (2007)
  • T. Ombrello et al.

    Combust. Flame

    (2010)
  • X. Gao et al.

    Combust. Flame

    (2016)
  • A. Mohammadi, H. Kawanabe, T. Ishiyama, M. Shioji A. Komada, Study on Combustion Control in Natural-Gas PCCI Engines...
  • H. Nishida et al.

    J. Propuls. Power

    (2006)
  • I. Ekoto et al.

    SAE Int. J. Fuels Lubr.

    (2018)
  • Cited by (9)

    • Combustion machine learning: Principles, progress and prospects: Combustion machine learning

      2022, Progress in Energy and Combustion Science
      Citation Excerpt :

      Convolutional and recurrent layers in these networks enable the direct use of complex datasets such as visual images and time-series data as feature sets. Kuzhagaliyeva et al. [638] reported that deep LSTM networks make accurate (75%) predictions of engine preignition when fed with time-series data extracted from low-cost pressure sensors. In the same work, 1D convolutional networks produced more accurate (79%) predictions when fed with datasets preprocessed via PCA.

    • Introduction

      2022, Artificial Intelligence and Data Driven Optimization of Internal Combustion Engines
    View all citing articles on Scopus
    View full text