Numerical simulations of a second-order chemical reaction in a plane turbulent channel flow

https://doi.org/10.1016/j.ijheatmasstransfer.2010.05.051Get rights and content

Abstract

We analyze numerical simulations of a second-order chemical reaction (Da = 1) in a fully developed turbulent plane channel flow at a low Reynolds number (Reτ = 180). The reactive plume is formed when a reactant A is released through a line source into the channel flow doped with reactant B. Two different inlet pre-mixing conditions and line source heights are considered. Direct Numerical Simulations (DNS) and Stochastic Fields (SF) methods have been used and compared for these different conditions. The results obtained using SF are sensitive to the particular value of the turbulent Schmidt number (ScT) selected to model the turbulent dispersion. It has been found that if a representative value of ScT extracted from DNS is used in the SF method, both DNS and SF, give similar results.

Introduction

Scalar transport in turbulent flows plays a fundamental role in a large number of industrial/engineering/environmental applications including chemical reactors, heat exchangers, mixing and separation processes and dispersion of pollutants in the atmosphere. Some of these applications may include chemical reactions between different reactants. The first studies of the scalar dispersion in homogeneous and isotropic turbulence were carried out by Taylor [1], [2], Uberoi and Corrsin [3] and Townsend [4]. The same problem was studied by Sawford and Hunt [5] using a Lagrangian approach in order to determine the effect of molecular diffusion on the scalar fluctuations. Thermal plume statistics dependences on molecular diffusivity and viscosity were confirmed experimentally by Stapountzis et al. [6]. Probability Density Function (PDF) methods [7] and Direct Numerical Simulations (DNS) [8] were used to study the dispersion from line sources in isotropic turbulence.

The dispersion of scalars from line sources in homogeneous turbulent shear flows to determine the effect of anisotropy on scalar dispersion was experimentally studied by Stapountzis and Britter [9], Karnik and Tavoularis [10] and Chung and Kyong [11]. These authors provided information about mean scalar distributions, fluctuation intensities and velocity–temperature correlations. Numerical results were reported by Wilson et al. [12] and Cho and Chung [13].

Scalar dispersion in turbulent channel flows is an example of mixing in a non-isotropic turbulent flow. Shlien and Corrsin [14] studied the dispersion from line sources in boundary layers finding that the mean temperature profile approached an asymptotic form when normalized adequately. Paranthoen et al. [15] found a rescaling scheme that collapsed the mean temperature profile into a single curve but it was not valid for the fluctuation intensities. Additional experimental measurements of dispersion in turbulent boundary layers can be found in Legg et al. [16], Veervalli and Warhaft [17], Bara et al. [18], Tong and Warhaft [19] and Vincont et al. [20].

Numerical studies of scalar transport at low Reynolds numbers were carried out by Lyons et al. [21], Papavassiliou and Hanratty [22], Na and Hanratty [23] and Kontomaris and Hanratty [24]. The last one studied the effects of molecular diffusivity on the dispersion of a passive scalar released from a point source located at the centerline of a turbulent channel flow. Using the Lagrangian Scalar Tracking (LST) method, Mito and Hanratty [25] studied the dispersion of point sources located at different distances away from a wall. Le and Papavassiliou [26] used LST to track heat markers released form line sources at different locations away the wall for a wide range of Pr numbers.

Chemical reaction and DNS/LST have been combined by Mitrovic and Papavassiliou [27] and Nguyen and Papavassiliou [28], [29] to study the chemical reaction (first- and second-order) of reactive markers released from walls in turbulent channel flow configurations.

DNS of dispersion in a pipe from point sources was studied by Brethouwer et al. [30]. Vrieling and Nieuwstadt [31] simulated single and double line sources of passive scalars in fully-developed channel flows. Experimental work was reported by Lavertu and Mydlarski [32].

Other authors studied experimentally the dispersion from line sources of conserved and reactive scalars in homogeneous turbulence [33], [34] and numerically the influence of the mixing process on the chemical reaction rates [35], [31].

Fabregat et al. [36] simulated the dispersion of a plume in a developed channel flow at Reτ = 180 where the scalar was released through a line source located at the inlet. These authors simulated the development of a hot plume resulting from the coupling of heat and momentum transport equations via the buoyancy term obtaining a plume drifting towards the top channel wall.

The application of the Reynolds decomposition to the mass transport equations of the chemical species involved in a second-order chemical reaction produces a source term, w˙, than can written as:w˙=kCA¯CB¯+CACB¯

The covariance term, CACB¯ is difficult to model. This term cannot be ignored when the reaction term is fast enough and segregation is important [37]. Moment methods used a direct closure for the covariance term [38], [39], [40] or Large-Eddy Simulation (LES) results to parametrize the reaction rate [41]. The main problem of these methods is related with the extra unclosed terms associated to complex reactions that are not easy to model. The Conditional Moment Closure (CMC) [42], [43], [44] does not require the modelization of CACB¯ and it has been applied successfully to combustion and atmospheric dispersion problems.

PDF methods [45], based on a transport equation for a joint probability density function of the reacting scalars, represent an alternative to simulate turbulent reacting flows. In these methods, micromixing has to be modeled but the chemistry term is closed. Monte Carlo methods are usually used in the numerical solutions and the computational cost is usually high due to the large number of Monte Carlo samples required [46].

An alternative and novel PDF method is the Stochastic Fields (SF) method. This method developed by Valiño [47] has been used to solve the transport of reactive species in different flow configurations as turbulent mixing layers [48], street canyons [49], exhaust plumes from aircrafts [50], plumes in grid-generated turbulence [51] and combustion [52]. The Stochastic Fields method allows to bypass the modelization of the turbulent covariance term. There is no limitation on the Damkhöler number and it can deal with complex mechanisms of reaction.

In this study, DNS and the SF method are compared and used to simulate the dispersion of pollutants, which react following a second-order chemical reaction, in a pressure driven channel flow at a low Reynolds number and a Damkhöler number equal to unity. The physical and mathematical models are presented in Section 2 and the numerical methods are described in Section 3. The results are presented and discussed in Section 4.

Section snippets

Physical and mathematical models

The channel configuration consists in two parallel, infinite and smooth walls separated a distance 2δ. The fluid flows between these walls driven by a constant mean pressure gradient. Fig. 1 shows a schematic representation of the computational domain. The coordinate system origin is located at point O and the coordinates xi = x, y, z correspond to streamwise, spanwise and wall-normal directions, respectively. Both streamwise and spanwise directions are considered homogeneous in terms of momentum

Direct Numerical Simulations

The size of the computational domain must be large enough in the streamwise and spanwise directions to capture the representative structures of the flow. According to this, the domain size has been set to Lx × Ly × Lz = 8πδ × 2πδ × 2δ (see Fig. 1).

The domain has been discretized using a staggered grid with a resolution of 258 × 130 × 130 control volumes. The grid has been stretched near the walls to increase the resolution in the wall boundary layer where velocity gradients are larger. The control volume

Instantaneous concentration fields

Fig. 3(a) shows instantaneous concentration fields for species A (top row), B (middle row) and C (bottom row), respectively, for the case DNS1 at different selected positions: x/δ = 6.0 (left column), x/δ = 12.5 (middle column) and x/δ = 24.0 (right column). In this case, the premixed condition is used at the inlet of the channel and consequently reaction starts immediately and product C is found everywhere in the plume of reactant A.

Fig. 3(b) shows the corresponding contours for the case DNS2. In

Conclusions

A second-order chemical reaction in a fully-developed channel flow has been studied using Direct Numerical Simulations (DNS) and the Stochastic Fields Method (SFM). The reaction occurs in the plume formed when a reactant is released through a line source located at the channel inlet into a background fluid doped with the other reactant. Different source heights and reactants pre-mixing conditions have been studied and the comparison between the two numerical methods allowed to quantize the

Acknowledgements

This study was financially supported by the Spanish Ministry of Science and Technology under project DPI2009-11204.

References (59)

  • G. Brethouwer et al.

    Direct numerical simulation of turbulent mixing of a passive scalar in a pipe flow

    Eur. J. Mech. B/Fluids

    (1999)
  • A.J. Vrieling et al.

    Turbulent dispersion from nearby point sources – interference of the concentration statistics

    Atmos. Environ.

    (2003)
  • A. Fabregat et al.

    Dispersion of a buoyant plume in a turbulent pressure-driven channel flow

    Int. J. Heat Mass Transfer

    (2009)
  • J.P. Meeder et al.

    Large-eddy simulation of the turbulent dispersion of a reactive plume from a point source into a neutral atmospheric boundary layer

    Atmos. Environ.

    (2000)
  • A.Y. Klimenko et al.

    Conditional moment closure for turbulent combustion

    Prog. Energy Combust. Sci.

    (1999)
  • R.J. Brown et al.

    Experiments on a reacting plume. Conditional concentration statistics

    Atmos. Environ.

    (1998)
  • M. Cassiani et al.

    A PDF micromixing model of dispersion for atmospheric flow. Part 1: development of the model, application to homogeneous turbulence and to neutral boundary layer

    Atmos. Environ.

    (2005)
  • G. Hauke et al.

    Computing reactive flows with a field monte carlo formulation and multi-scale methods

    Comput. Methods Appl. Mech. Eng.

    (2004)
  • A. Garmory et al.

    Simulations of the dispersion of reactive pollutants in a street canyon considering different chemical mechanisms and micromixing

    Atmos. Environ.

    (2009)
  • A. Garmory et al.

    Micromixing effects in a reacting plume by the stochastic fields method

    Atmos. Environ.

    (2006)
  • L. Valencia et al.

    Turbulent Rayleigh–Bénard convection of water in cubical cavities: a numerical and experimental study

    Int. J. Heat Mass Transfer

    (2007)
  • J. Pallarès et al.

    Eduction of near-wall flow structures responsible for large deviations of the momentum-heat transfer analogy and fluctuations of wall transfer rates in turbulent channel flow

    Comput. Fluids

    (2007)
  • G.I. Taylor

    Diffusion by continuous movements

    Proc. Lond. Math. Soc. A

    (1921)
  • G.I. Taylor

    Statistical theory of turbulence. IV – diffusion in a turbulent air stream

    Proc. R. Soc. Lond. A

    (1935)
  • M.S. Uberoi et al.

    Diffusion form a line source in isotropic turbulence

    Proc. R. Soc. Lond. A

    (1935)
  • A.A. Townsend

    The diffusion behind a line source in homogeneous turbulence

    Proc. R. Soc. Lond. A

    (1954)
  • B.L. Sawford et al.

    Effects of turbulence structure, molecular diffusion and source size on scalar fluctuations in homogeneous turbulence

    J. Fluid Mech.

    (1986)
  • H. Stapountzis et al.

    Structure of the temperature field downwind of a line source in grid turbulence

    J. Fluid Mech.

    (1986)
  • M.S. Anand, S.B. Pope, Diffusion behind a line source in grid turbulence, in: Proceedings of the Fourth Symposium on...
  • Cited by (12)

    • The assessment of water pollution by chemical reaction products from the activities of industrial facilities: Numerical study

      2021, Journal of Cleaner Production
      Citation Excerpt :

      In (J. A. Denev et al., 2006), a reference criterion is carefully selected to provide reference data concerning the following phenomena: flow field, mixing of passive scalars, and three chemical reactions. Fabregat et al. (2010) analyze the numerical simulation of a second order chemical reaction in a channel flow at a low Reynolds number (Re = 180), where two different inlet premixing conditions are considered. Direct numerical simulation (DNS) and stochastic field (SF) methods have been used and compared for these different conditions.

    • Characterization of the reacting laminar flow in a cylindrical cavity with a rotating endwall using numerical simulations and a combined PIV/PLIF technique

      2016, International Journal of Heat and Mass Transfer
      Citation Excerpt :

      Due to its ability to mix and transport species faster than molecular diffusion and laminar flows, turbulent flows have important implications in many industrial processes. For this reason, the great majority of authors perform their experimental and computational studies in turbulent regime [2,4–6]. Fukushima et al. [7] investigated the mixing of a passive scalar contaminant with a surrounding fluid in a turbulent free jet.

    • Conditional statistics in a planar liquid jet with a second-order chemical reaction

      2015, International Journal of Heat and Mass Transfer
      Citation Excerpt :

      However, because of the high computational cost of DNS, its application is limited to flows at low Reynolds numbers and in simple geometries. DNS has been employed for investigating chemical reactions in turbulent flows and developing models for turbulent reactive flows (e.g., [3–8]). From a practical point of view, the Reynolds-averaged approach, in which the Reynolds-averaged governing equations are solved by using turbulent models, is useful for predicting turbulent flows because of its low computational cost.

    • Effect of the instantaneous turbulent flow structures on the particle distribution near the wall of a channel

      2014, European Journal of Mechanics, B/Fluids
      Citation Excerpt :

      The Poisson equation resulting from the coupling between the velocity and pressure fields is solved with an efficient parallel multigrid solver. More details about the code can be found in Fabregat et al. [19]. The conditional pattern recognition technique used is based on the detection of extreme values of the fluctuations of the local wall shear stress in the instantaneous distributions on the walls of the channel.

    • Macro- and micromixing in a plane turbulent channel flow with a second-order chemical reaction

      2013, Computers and Fluids
      Citation Excerpt :

      The numerical results were validated with the experimental data [1]. Fabregat et al. [11] performed and analysed direct numerical simulations of a second-order chemical reaction with a Damkhöler number, Da = 1 and Sc = 0.7, in an incompressible fully developed turbulent plane channel flow at a low Reynolds number (Reτ = 180). An additional non-reactive scalar (T) was added in the numerical simulation in order to use it as marker of the non-reactive mixing.

    View all citing articles on Scopus
    View full text