You are on page 1of 23

Flow, Turbulence and Combustion 67: 5779, 2001.

2001 Kluwer Academic Publishers. Printed in the Netherlands.

57

Mixing and Entrainment of Transitional


Non-Circular Buoyant Reactive Plumes
X. JIANG and K.H. LUO
Department of Engineering, Queen Mary and Westfield College, University of London,
London E1 4NS, U.K.; E-mail: x.jiang@qmw.ac.uk
Received 17 April 2001; accepted in revised form 8 October 2001
Abstract. Three-dimensional spatial direct numerical simulation is used to investigate the evolution
of reactive plumes established on non-circular sources. Simulations are performed for three cases:
a rectangular plume with an aspect ratio of 2:1, a square plume, and the square plume in a corner
configuration. Buoyancy-induced large scale vortical structures evolve spatially in the flow field. A
stronger tendency of transition to turbulence is observed for the free rectangular plume than the free
square case due to the aspect ratio effect. Dynamics of the corner square plume differs significantly
from the corresponding free case due to the enhanced mixing by the side-wall effects. A turbulent
inertial subrange has been observed for the free rectangular and corner square plumes. Mean flow
properties are also calculated. The study shows significant effects of source geometry and side-wall
boundary on the flow transition and entrainment of reactive plumes.
Key words: buoyancy, DNS, entrainment, mixing, non-premixed flame, reaction, transition.

1. Introduction
Reactive jets and plumes are encountered in many industrial and environmental
applications. Understanding their mixing and entrainment properties is of both
practical and fundamental importance. The spatial development of a jet depends
on its initial momentum and the surrounding environment. For a plume, buoyancy
effects due to density inhomogeneity are also of importance. Major sources of
density inhomogeneity include temperature inhomogeneity due to heat release,
differences in density of chemical species and phase changes. All of these may
occur in one reactive flow.
Despite the numerous experimental and computational studies [5, 7, 12], dynamics of jets and plumes is far from being well understood. For instance, effects
of the geometry of jet nozzle or plume source have not been fully investigated. On
the other hand, optimizing the nozzle geometry is potentially an efficient technique
of passive flow control that can improve combustion efficiency, enhance heat and
mass transfer, and reduce undesired emissions at a relatively low cost. Non-circular
jet nozzles, for example, have been used to improve the mixing and entrainment of
reactive and non-reactive systems [12]. For reactive plumes, another outstanding

58

X. JIANG AND K.H. LUO

issue is the side-wall effects, which is important in applications such as fire control
but has not been fully investigated.
Dynamics of reactive plumes is a difficult problem to tackle. The complexities
include the effects of plume source geometry and side-walls, together with the coupling between fluid dynamics and combustion through buoyancy. Buoyancy effects
are important to low speed reactive flows. In experiments, it is difficult to separate
the individual effects. Accurate measurements of the instantaneous quantities and
flow field data are not easy either. In analytical approaches, it is very difficult to take
into account the non-parallel effects due to strong entrainment and the effects of
source geometry and surroundings. Numerical approaches based on the ReynoldsAveraged NavierStokes equations (RANS) have been used in the modelling of
reactive plumes for many years [5]. However, the RANS approach is inappropriate to investigate unsteady reactive plumes involving transition, intermittency and
turbulence.
Direct numerical simulation (DNS) which directly resolves all the relevant time
and length scales described in the NavierStokes equations provides a possibility
to study the complex phenomena occurring in reactive plumes. Recently, a direct
and large-eddy simulation of the transition of plane plumes in a confined enclosure
was reported for a non-reacting flow with Boussinesq approximation for the effect
of buoyancy by Bastiaans et al. [1], in which no clear turbulent inertial subrange
was present. The present authors [17] performed a spatial DNS of the near field
dynamics of a rectangular reactive plume. The spatially developing reactive plume
showed a tendency of transition to turbulence under the effects of combustioninduced buoyancy. However, the effects of plume source geometry and side-walls
were not fully investigated.
The main objective of this study is to examine the plume source geometry
and side-wall effects on the near field dynamics of a buoyant reactive plume,
which belongs to diffusion-controlled non-premixed flames. Chemical reactions
produce heat sources heterogeneously distributed. Under the buoyancy effects due
to temperature inhomogeneity, the reactive plume evolves spatially with transition
occurring downstream. A comparative study of three cases has been conducted.
The first case performed is a rectangular reactive plume with an aspect ratio of 2:1
in an open boundary domain, while the second case performed is a square reactive
plume with the same cross-sectional area as the rectangular case. The third case
considered is the square reactive plume in a corner configuration. The simulations
describe details of the reactive plume evolution through transition to turbulence.
Results are discussed in terms of instantaneous quantities, history of the streamwise
velocities and energy spectra, vorticity transport, and time-averaged statistics.
The rest of the paper is organized as follows. Mathematical description of the
physical problem and numerical method used for the spatial DNS are presented in
the next section, followed by a discussion on the simulation results from a comparative study of the effects of source geometry and side-walls on the dynamics of
non-circular buoyant reactive plumes. Finally, conclusions are drawn.

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

59

2. Numerical Approach
2.1. M ATHEMATICAL DESCRIPTION
Flow transition to turbulence in reactive jets and plumes occurs in the near field,
which is the region close to the jet nozzle or plume source. In this study, the near
field of transitional buoyant reactive plumes is considered, where a fuel jet issues
from the non-circular nozzle vertically into an oxidant ambient. Combustion occurs
when fuel/oxidizer mixing takes place and a non-premixed flame is established
above the inlet plane.
The governing equations used to describe the reactive flow field above the inlet
plane are the compressible time-dependent NavierStokes equations with temperature-dependent viscosity. The conservation laws for mass, momentum, energy and
chemical species formulated in non-dimensional forms are adopted. Major refer
= 9.81 m/s2 , magnitude of the
ence quantities used in the normalization are: gref
gravitational acceleration; Lref = L0 , width of the fuel jet along the major axis of

= Ta , the ambient temperature; wref


= w0 , the maxthe rectangular source; Tref

imum velocity of the fuel jet at the inlet; ref = a , fuel viscosity at the ambient

= a , fuel density at the ambient temperature. Accordingly time


temperature; ref

= L0 /w0 . Here, the subscripts 0 and a refer to the


is non-dimensionalized with tref
fuel jet (source) and ambient respectively and the superscript stands for dimensional quantities. The dynamic viscosity is chosen to be temperature-dependent
according to = a (T /Ta )0.76 .
The non-dimensional governing equations can be written in a vector form as
U E F G
+
+
+
+ S = 0.
t
x
y
z

(1)

In the Cartesian coordinate system, the z-direction is assumed to be the streamwise


(vertical) direction, the x-direction is aligned with the minor axis of the rectangular
fuel jet at the inlet, while the y-coordinate is along the major axis. The terms
major and minor used here refer to the dimensions of the rectangular inlet.
For square geometry, this difference vanishes. The velocity components in the x,
y, and z directions are represented by u, v, and w respectively. In Equation (1),
vectors U, E, F, G and S are defined as U = (, u, v, w, ET , Yf , Yo )T
with the superscript T representing transposition,

u2 + p xx

uv xy

uw

xz
E =
,
(ET + p)u + qx uxx vxy wxz




Yf
1

uYf Re Sc x

 Yo 
1
uYo Re Sc x

60

X. JIANG AND K.H. LUO

uv xy

2
v
+ p yy

vw yz
F =
(ET + p)v + qy uxy vyy wyz




Yf
1

vY

f
Re Sc
y



o
vYo Re1Sc Y
y

uw
xz

vw yz

w 2 + p zz
G =
(ET + p)w + qz uxz vyz wzz




wYf Re1Sc zf



o
wYo Re1Sc Y
z

(a )gz

S =
.
Fr

(a )wgz

h
Fr

f
o

and

As shown, the gravitational effect is expressed as buoyancy terms (a )gz /Fr


and (a )wgz /Fr in the streamwise momentum equation and energy equation,

L0 ) and gz = 1
respectively, where Froude number is defined as F r = w0 2 /(gref
is the gravity imposed in the downward vertical direction. The governing equations
also include the perfect gas law for the mixture.
In Equation (1), ET = [e + (u2 + v 2 + w 2 )/2] is the total energy with e representing the internal energy per unit mass, q represents the heat flux components,
Re and Sc represent the Reynolds and Schmidt numbers, respectively, Y stands
for mass fraction of the chemical species. Currently, a complete three-dimensional
(3D) DNS of reactive flows with multi-species transport and complex chemistry is
still out of reach [2], due to the excessive computer resources required. In this study,
a one-step global reaction f Mf +o Mo p Mp with finite-rate Arrhenius kinetics is presumed for the chemistry, where Mi and i represent the chemical symbol
and stoichiometric coefficient for species i, respectively. The subscripts f , o, p
stand for fuel, oxidizer and product, respectively. The reaction rate takes the form

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

61

of T = Da(Yf /Wf t)f (Yo /Wo )o exp[Ze(1/T 1/Tfl )] after normalization


[21], where Da, Ze, and Tfl are the Damkhler number, Zeldovich number, and
flame temperature, respectively, Wi stands for the molecular weight of species i.
The reaction rates for individual species are f = f Wf T , o = o Wo T
and p = p Wp T . The heat release rate in the energy equation is given by
h = Qh T with Qh representing the heat of combustion.
2.2. N UMERICAL METHODS
The equations are solved using a sixth-order compact finite difference scheme [19]
with spectral-like resolution for evaluation of the spatial derivatives. This scheme
allows more flexibility in the specification of boundary conditions with minimal
loss of accuracy compared with spectral methods. A third-order accurate fullyexplicit compact-storage RungeKutta scheme is used to advance the equations in
time. The time step is limited by the CourantFriedrichsLewy (CFL) condition for
stability. It is further limited by the reaction rate, an increase in local mass fraction
of product of more than 0.01% is prevented for one time step. For DNS of reactive
flows, the time step is usually limited by the chemical restraint.
The specification of the boundary conditions is performed by using the general
formulation of characteristic boundary conditions for DNS of NavierStokes equations by Poinsot and Lele [24]. The 3D computational domain is bounded by the
inflow and outflow boundaries in the streamwise direction. At the inflow boundary,
temperature is treated as a soft variable while other variables are imposed with
their initial values. A soft variable means that the temperature is allowed to
fluctuate around the prescribed value, which is associated with the characteristic
wave variations at this boundary. This treatment guarantees the numerical stability
of the high-order numerical scheme near the boundary, while the fluctuation of
temperature with time is less than 0.5% of its prescribed value in the simulations
performed.
The outflow boundary condition needs careful attention since vortices are being
convected through this boundary and the flow field outside the domain is unknown.
At the outflow boundary, non-reflecting characteristic boundary condition [25] is
used. In order to completely eliminate the spurious wave reflections, a sponge layer
[15] is applied next to the outflow boundary. This is because that the non-reflecting
characteristic boundary condition is based on a one-dimensional formulation, but
the flow near the outlet is of multi-dimensional nature due to the convection of
vortices. The computational results in the sponge layer are not truly physical and
therefore not used in the data analysis.
In the cross-streamwise directions, the flow field is either open or bounded by
the side-walls. For the open boundary, the entrainment boundary condition [15] is
used, which allows entrainment of the ambient fluid into the computational domain.
For the side-wall boundary condition, the wall is assumed to be impermeable and
non-slip. The wall temperature is determined from the characteristic form of the

62

X. JIANG AND K.H. LUO

energy equation by applying the local one-dimensional inviscid relations for a wall
boundary [24], which allows variations of the wall temperature.
Atop-hat profile is assumed for the streamwise velocity of the fuel jet on the
inlet plane, which is given by
{1 + tanh[20 (sx |x Lx /2|)]} {1 + tanh[20 (sy |y|)]}
,
w=
[1 + tanh(20 sx ) + tanh(20 sy ) + tanh(20 sx ) tanh(20 sy )]
while the cross-streamwise velocity components are taken as zero on the inlet
plane. The inlet plane is of the dimension of (0 Lx , Ly /2 Ly /2), where Lx
and Ly stand for the inlet domain lengths in the minor and major axis directions,
respectively, and s stands for the half-width of the fuel source on the inlet plane.
The fuel and oxidizer are unmixed on the inlet plane and the fuel jet is centered in
the inlet domain. In the simulations, the fuel temperature at the source is assumed
to be 3, which was chosen to ensure auto-ignition of the mixture [10, 11, 26].
Ignition occurs automatically when the fuel and oxidizer mix with each other. The
temperature and streamwise velocity profiles on the inlet plane are linked with the
CroccoBusemann relation.
The flow field inside the domain is initialized with the inlet conditions. Buoyant
jets and plumes display an intrinsic, absolutely unstable flow instability [14, 16,
20, 22] in which vortices evolve naturally in the flow field due to the gravitational
effect [16]. For such as an absolute instability, there is no need to apply continuous
external perturbations at the inflow boundary for the development of flow vortices
[15, 16, 18]. In order to isolate the absolute buoyancy instability from the convective shear instability, external perturbations were not used at the inflow boundary
in the simulations performed. The external stimulus to initiate flow vortices in
the numerical simulations can arise from a mismatch between the imposed initial
conditions and solutions to the flow field. For the present spatial simulations, initial
conditions are of minor importance after the initial transient period.
3. Results and Discussion
In the plume near field, buoyancy effects such as flow acceleration on the resolution
set a limit on the minimum value of the Froude number that can be prescribed in
a DNS under a certain number of grid points [6]. In this study, the Froude number
Fr = 1.5 used has been chosen so that the buoyant reactive flow field can be fully
resolved. The nominal flow Reynolds number used in the simulations is Re =
1000, which is based on the inlet reference quantities. The species are assumed to
be of equal diffusivity. The ratio of specific heats, Prandtl and Schmidt numbers
used in the simulations are chosen to be constants: = 1.4, Pr = 1 and Sc = 1.
The parameters used for the chemical reaction are: Da = 6, Ze = 12, Tfl = 6, and
Qh = 1650. These values are chosen to mimic the behavior of a relatively low heat
release combustion within the computational resources available.
For the free rectangular case, a computational domain of the size of 3 6
8 is used, which has been chosen to minimize the effects of boundaries on the

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

63

simulation results. The area of the fuel jet is 0.51.0 on the inlet plane. For the free
square case, the domain size used is 4.254.258.50 which has approximately the
same cross-sectional area as the rectangular case, while the fuel jet source area is
approximately 0.71 0.71. The results presented next are obtained from a uniform
grid system with 108 216 288 nodes for the free rectangular case and a uniform
grid system with 152 152 304 nodes for the free square case. The corner
square case performed is of the same size as the free square case with side-wall
boundaries located at x = 0 and y = Ly /2. For this case, more grid points are
needed to fully resolve the flow field between the wall and plume. A grid system
with 188 188 376 nodes is used for this corner square case. For a sufficiently
small reference length scale (e.g. Lref 1 cm), the grid points used in these three
cases are sufficient to resolve the energy spectra of the relevant physical problems.
In the simulations, the sponge layer used to prevent the spurious wave reflections
from the outflow boundary is located between z = 7.0 and the end of the domain
in the streamwise direction. For clarity, some results shown in the figures are for a
box which is smaller than the computational domain used.
Simulations have been performed on the massively parallel computer Cray T3E1200E in Manchester by using 72 processors for the free rectangular case, 76
processors for the free square case and 94 processors for the corner square case. In
this study, grid independence and time-step independence tests were performed. A
spatial resolution study was performed by adding 20% more grid points in each direction for the corner square case, while a temporal resolution study was performed
by reducing the time step by half. In both cases, history of the vorticity extrema in
the computational domain and typical instantaneous velocity profiles showed no
appreciable changes. Therefore the results presented next are considered as grid
and time-step independent. In the following, instantaneous quantities are shown
first to discuss the dynamics and structure of the transitional flow fields, followed
by the time-averaged flow statistics of the reactive plumes.

3.1. I NSTANTANEOUS FLOW STRUCTURE AND MIXING CHARACTERISTICS


Buoyancy-driven large scale vortical structures are the characteristics of both reactive and non-reactive plumes in the transitional regime, and the breakdown of
these vortical structures leads to turbulence in the flow field [3, 4]. Figure 1 shows
the instantaneous 3D visualizations of the vorticity magnitude at t = 23.0 and
t = 24.0 of the corner square reactive plume. A diagram showing the inlet plane
and the Cartesian coordinates is also included. The three components of the vorticity are x = w/y v/z, y = u/z w/x, and z = v/x u/y,
respectively. The quantity shown in Figure 1 is (x2 +y2 +z2)1/2 . It can be observed
that large vortical structures evolve spatially in the flow field.
Figures 24 show the contour plots on the center plane x = Lx /2 of the
corner square plume at the two different times, corresponding to the 3D plots in
Figure 1. In the near field, buoyant jets and plumes display a puffing or flick-

64

X. JIANG AND K.H. LUO

Figure 1. Instantaneous 3D visualizations of the vorticity magnitude at t = 23.0 and t = 24.0


of the corner square reactive plume.

Figure 2. Fuel mass fraction contours on the x = Lx /2 plane at t = 23.0 and t = 24.0 of the
corner square reactive plume (15 contours between the minimum and maximum as indicated).

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

65

Figure 3. Reaction rate contours on the x = Lx /2 plane at t = 23.0 and t = 24.0 of the
corner square reactive plume (15 contours between the minimum and maximum as indicated).

Figure 4. Temperature contours on the x = Lx /2 plane at t = 23.0 and t = 24.0 of the corner
square reactive plume (15 contours between the minimum and maximum as indicated).

66

X. JIANG AND K.H. LUO

Figure 5. Temperature contours on the z = 1.0 and z = 5.0 planes at t = 24.0 of the corner
square reactive plume (15 contours between the minimum and maximum as indicated).

ering phenomenon [3, 4, 1518, 20, 22], which is associated with the formation
and convection of buoyancy-induced large vortical structures. The flow exhibits
approximately a periodic behavior near the plume source due to the puffing or
flickering. The times t = 23.0 and t = 24.0 shown in these figures are within one
pulsation period of the plume. The necking phenomenon close to the inlet plane
can be seen clearly. One prominent feature in Figures 24 is that the flow field is
less coherent downstream due to the breakdown of the large vortical structures,
which indicates the emergence of small scale turbulence in the flow field.
For a non-premixed flame, the flow must satisfy two criteria for a significant
reaction to occur: both the fuel and oxidizer must be well mixed at a given point
in the field and the temperature must be high enough at that point. The chemical
reactivity of the corner square plume is evident from Figures 2 and 3, which show
the fuel mass fraction and reaction rate T contours, respectively. The fuel is being
consumed by the chemical reaction when the fuel/oxidizer mixing takes place. The
unburnt fuel being convected out of the computational box for this corner square
case is less than 20%. The chemical reaction is rather weak at the downstream
location z = 6.0 near the outlet as shown in the reaction rate contours. Chemical
heat release also increases the local temperature in the reaction zone, as shown in
Figure 4. From the temperature contours, a disorganized flow regime downstream
of the reactive plume characterized by small scales due to the breakdown of large
scale structures is also evident.
Figure 5 shows temperature contours on the z = 1.0 and z = 5.0 planes at
t = 24.0 of the corner square plume. The complex 3D structures in the noncircular reactive plume are evident. It is noticed that the corner square reactive
plume remains symmetric about the bisector of the domain. Unlike the simulations of non-buoyant reactive jets performed by Grinstein and Kailasanath [10,

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

67

Figure 6. Temperature contours on the z = 1.0 and z = 5.0 planes at t = 24.0 of the free
square reactive plume (15 contours between the minimum and maximum as indicated).

Figure 7. Temperature contours on the z = 1.0 and z = 5.0 planes at t = 24.0 of the free
rectangular reactive plume (15 contours between the minimum and maximum as indicated).

11], external perturbations were not used at the inflow boundary in the simulations
performed, in order to isolate the absolutely unstable buoyancy instability [14, 16,
20, 22] from the shear instability. Therefore the flow remains symmetric due to
the geometric symmetry of the physical problem and absence of external disturbances. Nevertheless, transition to turbulence still occurs in the flow field due to
the buoyancy and 3D vortex stretching effects.
For comparison, Figures 6 and 7 show the temperature contours of the free
square and free rectangular cases at the two different vertical locations at t = 24.0,
corresponding to those shown in Figure 5 for the corner square case. At the location
z = 1.0 which is close to the fuel source, both the free square and free rectangular
reactive plumes resemble their original shapes specified on the inlet plane. At the

68

X. JIANG AND K.H. LUO

downstream location z = 5.0, complex structures are developed with the spatial
development of the flow field, which are characteristics of the non-circular plumes.
For the comparison between the two square cases shown in Figures 5 and 6, it is
noticed that the reactive plume in a corner configuration spreads more in the lateral
directions than the free case at the downstream location. This is mainly because of
the recirculation zones formed between the wall and plume [1]. With the presence
of impermeable and non-slip side-wall boundaries, mixing between the reactive
plume and its ambient is enhanced by the strong recirculation, therefore the corner
square plume has larger spreading.
For the comparison between the two free cases shown in Figures 6 and 7,
it is also noticed that the free rectangular reactive plume spreads more than the
free square case at the downstream location. This can be explained by the selfinduced BiotSavart vortex-ring deformation associated with non-circular nozzle
geometries [9, 12], which is responsible for the complex 3D structures in the flow
field. The differences between the free rectangular and free square cases originate
from the aspect ratio effect of the rectangular source. Gutmark and Grinstein [12]
summarized the application of stability theory to non-circular jets, which identified the aspect ratio effects on the vortex flow evolution. For the BiotSavart
deformation of vortex-ring dynamics, the self-induced velocity responsible for the
vortex-ring deformation is proportional to the local curvature Vd C log(1/ )b
in a thin, incompressible, inviscid vortex tube, where C is the local curvature of
the tube, is the local cross-section of the vortex tube, and b is the binormal to
the plane containing the tube. The self-induced Biot-Savart deformation is stronger
in the free rectangular case than that in the free square case with the same crosssectional area, due to the differences in curvature. A stronger vortex deformation
in the rectangular case therefore leads to a more vortical flow field and in turn a
stronger mixing and spreading at the plume downstream.
In the near field, a plume entrains its ambient fluids mainly through the large
scale vortical structures. Figure 8 shows a comparison of the instantaneous crossstreamwise velocities on the x = Lx /2 plane at z = 5.0. The three different times
shown in the figure correspond approximately to one pulsation period of the free
rectangular and free square cases, which is about +t = 2.0. Therefore the curves
at t = 22.0 and t = 24.0 are almost overlapping for these two cases. For the corner
square case, the pulsation period at this downstream location differs from the two
free cases. From this figure, it is observed that the cross-streamwise velocities of
the free rectangular and corner square cases are much higher than that of the free
square case, which indicate stronger entrainment in these two cases.

3.2. M IXING TRANSITION AND VORTEX DYNAMICS


In the present non-circular buoyant reactive plumes, the 3D vortex deformation
associated with the non-circular geometry and side-wall effects leads to the breakdown of large vortical structures into small scales. The emergence of turbulence in

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

69

Figure 8. Comparison of the instantaneous cross-streamwise velocities on the x = Lx /2


plane at z = 5.0 of the free rectangular, free square, and corner square reactive plumes (
t = 22.0; t = 23.0; t = 24.0).

70

X. JIANG AND K.H. LUO

the non-circular buoyant reactive plumes can be testified by spectral analysis of the
flow field.
Figure 9 shows the comparison of the history of the centerline streamwise velocities at z = 5.0 of the three cases. The velocity fluctuations downstream of the
plume shown in Figure 9 are associated with the spatial development of the flow
field since there is no velocity fluctuations at the inlet. At the later stage (approximately after t = 12.0) when the flow is more developed, it can be observed that the
downstream centerline velocities of the free rectangular and corner square plumes
fluctuate much more than the centerline velocity of the free square plume. The
downstream centerline velocities of the free rectangular and corner square cases are
also lower than that of the free square case. This is mainly because of the stronger
mixing and larger spreading associated with the stronger vortex deformation in
the free rectangular case and the side-wall effects in the corner square case. With
stronger mixing, the density inhomogeneity and buoyancy in the free rectangular
and corner square cases decay faster, therefore the buoyancy acceleration decays
faster in these two cases. The important feature in Figure 9 is that the variations of
the downstream centerline velocity of the free rectangular and corner square cases
are less coherent than that of the free square case, which indicates the emergence
of small scale turbulence in the flow field.
The energy spectra of the three cases determined from the history of the centerline streamwise velocities at z = 5.0 using Fourier analysis are shown in Figure 10,
which is expressed in logarithmic scales (to the base 10) of the non-dimensional
frequency (Strouhal number St = f L0 /w0 ) and kinetic energy. The spectral
analysis used velocity data of the developed reactive plumes after the initial stage
of the simulation. It is observed that the most energetic mode for each case occurs
at a low frequency, which is the puffing or flickering frequency of the buoyant
reactive plume associated with the convection of large vortical structures. For the
free rectangular and corner square cases, the flows are more energetic than the
free square case and high frequency harmonics are developed in the flow field. A
prominent feature in Figure 10 is the development of these high frequency harmonics, which is associated with the emergence of high frequency small scales in the
flow field at the downstream location. The flow turbulence can be measured by the
Kolmogorov cascade theory (cf. [9]), which states a power law correlation between
the energy and frequency: E(St) St5/3 . In Figure 10, the Kolmogorov power
law is plotted together with the energy spectra. It can be seen that the behavior of
the free rectangular and corner square reactive plumes at the downstream location
approximately follows the 5/3 power law. The spectra shown reflect the energy
fluctuations of the high frequency harmonics in the free rectangular and corner
square cases.
Flow transition to turbulence is the direct consequence of the breakdown of
large scale vortical structures due to strong vortex interactions, especially the interactions between the streamwise vorticity and cross-streamwise vorticity [9]. The
occurrence of flow transition to turbulence downstream of the free rectangular

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

71

Figure 9. Comparison of the history of the centerline streamwise velocities at z = 5.0 of the
free rectangular, free square, and corner square reactive plumes.

72

X. JIANG AND K.H. LUO

Figure 10. Comparison of the energy spectra of the centerline streamwise velocity variations
at z = 5.0 of the free rectangular, free square, and corner square reactive plumes (
Kolmogorov cascade theory E(St) St 5/3 ).

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

73

and corner square cases is due to the much stronger vortex interactions in these
two cases in comparison with the free square case. For the flow configuration
investigated, there are only vortex rings very close to the plume source since the
streamwise vorticity z is zero on the inlet plane. The spatial development of
streamwise vorticity above the inlet plane and the strong vortex interactions further
downstream involving all the three vorticity components are essential to the flow
transition to turbulence [9, 17].
In buoyant flows with gravitational effect, the governing equation for vorticity
transport can be written in a vector form as
D
1
= ()V (V) + 2 (p)
Dt



1
1 a
.
+ 2 (g) +
Fr

(2)

The terms on the right-hand side of Equation (2) are the vortex stretching term,
dilatation term, baroclinic torque, gravitational term, and viscous term, respectively. Among these five terms, it is known that the dilatation term and viscous
term mainly attenuate flow vorticity [8], therefore chemical heat release reduces the
mixing and entrainment in non-buoyant reactive mixing layers [13, 23]. However,
their importance in the vorticity transport is relatively less than the other terms for
buoyant plumes with moderately high Reynolds numbers and low Froude numbers
[15, 16]. Therefore, these two terms are not discussed further in the following. In
this vorticity transport equation, a transport term on the right-hand side promotes
the flow vortical level if this term and the vorticity are of the same sign, while the
term attenuates flow vorticity if they are of opposite signs.
The stretching term, baroclinic torque and gravitational term are the major
terms in the vorticity transport budget of a buoyancy-driven flow. Axisymmetric
and planar simulations [15, 16] identified that the effect of gravitational term is to
promote the flow vortical level. This term is responsible for the absolute instability
of buoyant jets and plumes that can initiate flow vorticity. For the flow configuration studied here, the three components of this term are a /y /( 2 Fr),
a /x /( 2 Fr), and 0, in x, y, and z directions, respectively. This implies that
the gravitational term does not contribute to the streamwise vorticity directly. In
this study, it was found that the gravitational term mainly promotes the flow vorticity in the cross-streamwise directions. The stretching term and baroclinic torque
can either promote or destroy the cross-streamwise vorticity depending on the
local flow structure. These trends are consistent with the previous observations
on axisymmetric buoyant reactive plumes [15, 16], but the vortex stretching term
for 3D cases studied here is very significant. The vortex stretching is completely
absent in 2D planar simulations, while the stretching in axisymmetric simulations
is less significant.
Figure 11 shows the streamwise vorticity contours on the z = 1.0 and z = 5.0
planes at t = 24.0 of the corner square reactive plume. The streamwise vorticity

74

X. JIANG AND K.H. LUO

Figure 11. Streamwise vorticity contours on the z = 1.0 and z = 5.0 planes at t = 24.0 of the
corner square reactive plume (15 contours between the minimum and maximum as indicated;
solid line: positive; dotted line: negative).

Figure 12. Contours of the major transport terms of the streamwise vorticity on the z = 1.0
plane at t = 24.0 of the corner square reactive plume (15 contours between the minimum and
maximum as indicated; solid line: positive; dotted line: negative).

shows a distribution which is symmetric but with a sign change across the symmetry plane of the corner configuration. For the streamwise vorticity, the major
transport terms are the 3D vortex stretching and the baroclinic torque. Figure 12
shows contours of the major transport terms of the streamwise vorticity shown in
Figure 11a.
From Figure 12, it is observed that the vortex stretching x w/x+y w/y+
z w/z in z-direction is very significant in the transport of the streamwise vor-

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

75

Figure 13. Comparison of the time-averaged centerline temperature and streamwise velocities
of the free rectangular, free square, and corner square reactive plumes ( free rectangular;
free square; corner square).

ticity. A careful examination on the vortex stretching term distribution in Figure 12a shows that this term predominantly takes the same sign of the corresponding streamwise vorticity shown in Figure 11a. This indicates that the vortex
stretching term is a major source for the streamwise vorticity. For the baroclinic
torque shown in Figure 12b, it can either promote or destroy streamwise vorticity,
which is of the same nature as its contribution to the cross-streamwise vorticity. It is
also noticed that this term is less significant than the stretching term in the transport
budget. Analysis on streamwise vorticity transport at different times and different
locations and for other cases follow the same trend. Therefore it can be concluded
that the streamwise vorticity in the non-circular buoyant reactive plumes is mainly
caused by the 3D vortex stretching. The development of streamwise vorticity due to
vortex stretching leads to the coexistence of vortex rings and spirals in the reactive
plume, and consequently, strong vortex interactions occur downstream which are
responsible for the flow transition.
3.3. F LOW STATISTICS
Flow statistics were obtained by averaging the flow quantities over six pulsation
periods after the initial transients had been convected out of the computational
domain. Figure 13 shows the comparison of the time-averaged centerline temperature and streamwise velocities of the free rectangular, free square, and corner
square reactive plumes, while Figure 14 shows the comparison of the entrainment
properties of these three cases. It is worth noting that the bumps in the profiles
in Figures 13 and 14 are mainly because of the existence of buoyancy-induced

76

X. JIANG AND K.H. LUO

Figure 14. Comparison of the entrainment properties of the free rectangular, free square, and
corner square reactive plumes ( free rectangular; free square; corner
square).

large scale vortical structures in the near field. Although the large scale vortices are
being convected by the mean flow, the time-averaged quantities still indicate their
existence due to the spatial nature of the simulation.
In the near field of a buoyant reactive plume, the temperature depends on the
reaction and mixing. Combustion heat release increases the local temperature. In
the meantime, large spreading and strong mixing with the ambient reduces the
local temperature through the convective heat transfer. For the time-averaged centerline temperature shown in Figure 13a, it maintains the initial temperature of the
fuel near the inlet plane (z < 1.0 in the streamwise direction) since no reaction
takes place due to the unmixedness of the mixture. The temperature then increases
due to the combustion heat release. Downstream, the time-averaged centerline
temperature behaves differently for the three cases due to the different mixing
characteristics. For the free rectangular and corner square cases, the temperature
decreases due to the large spreading and the fuel consumption. For the free square
case, however, the downstream temperature still increases due to the small spreading and the combustion heat release. It is worth noting that the plume centerline
does not correspond to the intense reaction zone, therefore the temperature shown
is lower than the highest temperature within the non-premixed flame.
For the reactive plume, the mean streamwise velocity strongly depends on the
buoyancy effects. The time-averaged centerline streamwise velocity increases at
locations close to the plume source due to the buoyancy acceleration. The buoyancy
acceleration decays downstream because of the plume spreading and mixing with
the ambient. The free rectangular and corner square cases have larger spreading

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

77

than the free square case as shown before, therefore the downstream streamwise
velocities in these two cases decay faster.
Figure 14a shows the comparison of the streamwise mass fluxes
L
y /2 Lx

w dx dy
Ly /2 0

of the free rectangular, free square, and corner square reactive plumes, while Figure 14b shows the comparison of the entrainment rates which are the gradients
of the streamwise mass fluxes in the vertical direction of the three cases. Compared with the entrainment rates around 0.3 of square non-buoyant jets reported by
Grinstein and Kailasanath [10], the entrainment rate of the square buoyant plume
studied here is about three times of this value due to the strong buoyancy effects.
This trend is consistent with the comparison between the near field entrainment
properties of circular buoyant plumes and non-buoyant jets [7]. It is worth noting
that there is no comparable experimental data available for the non-circular buoyant
reactive plumes in the near field, and thus, a direct comparison with experiments
is not feasible. However, the mean properties qualitatively follow the trend of
buoyancy enhanced entrainment in the near field [7].
In Figure 14, the higher entrainment rate of the free rectangular case compared
with the free square case is due to the aspect ratio effects [12]. Although the
corner square reactive plume is half-confined in the cross-streamwise directions,
the entrainment of the corner square case is significantly higher than that of the
free square case. This can be explained by the side-wall enhanced mixing in the
corner case. From a macro-scale mixing point of view, the corner square case has
more length scales compared to the one obvious length scale (the fuel jet width) in
the free square case. Also, there is only one symmetry plane in the corner square
case (the bisector of the domain) compared to the many in the free square case.
For the corner case, the plume spreading is greatly enhanced by the recirculation
zones formed between the wall and plume as discussed in Section 3.1. This leads to
enhanced mixing at large scales in the corner square case. For the mixing at small
scales, the corner square case is much more turbulent than its free counterpart as
discussed in Section 3.2, which leads to enhanced small scale mixing in the corner
case. The side-wall enhanced mixing is also evident in the sample instantaneous
mixing characteristics shown in Figure 8.
4. Conclusions
In this paper, spatial direct numerical simulations of non-circular reactive plumes
with buoyancy effects have been performed for three cases. A one-step chemical
reaction with Arrhenius kinetics is used. The simulations are devoted to a better
understanding of the effects of plume source geometry and side-wall confinement
on the flow structure and dynamics of reactive plumes. Particular attention has been

78

X. JIANG AND K.H. LUO

given to the effects on the flow transition to turbulence and mixing and entrainment
in non-circular reactive plumes.
Complex 3D structures are observed for the buoyant reactive plumes established
on non-circular sources. Large scale vortical structures develop naturally in the
flow field due to the gravitational effect. Analysis on the vorticity transport shows
that 3D vortex stretching leads to the development of streamwise vorticity in the
flow field. Vortex interactions and the consequent breakdown lead to the emergence
of turbulence in the buoyant reactive plumes. A turbulent inertial subrange has been
observed downstream of the free rectangular and corner square reactive plumes.
The geometry and side-wall effects on the mixing and entrainment of noncircular reactive plumes are significant. Compared with the square geometry, the
rectangular geometry with an aspect ratio of 2:1 promotes the plume entrainment
of the ambient fluids due to the aspect ratio effects. By creating recirculation zones
between the wall and plume and promoting transition to turbulence, the side-walls
in the corner square case enhance the mixing and entrainment of the reactive plume
at both large and small scales.
Acknowledgment
This work was funded by the UK Engineering and Physical Sciences Research
Council under Grant No. GR/L67271.
References
1.

2.
3.
4.
5.
6.
7.
8.
9.
10.
11.

Bastiaans, R.J.M., Rindt, C.C.M., Nieuwstadt, F.T.M. and van Steenhoven, A.A., Direct and
large-eddy simulation of the transition of two- and three-dimensional plane plumes in a
confined enclosure. Internat. J. Heat Mass Transfer 43 (2000) 23752393.
Candel, S., Thevenin, D., Darabiha, N. and Veynante, D., Progress in numerical combustion.
Combust. Sci. Technol. 149 (1999) 297337.
Cetegen, B.M. and Ahmed, T.A., Experiments on the periodic instability of buoyant plumes
and pool fires. Combust. Flame 93 (1993) 157184.
Cetegen, B.M., Dong, Y. and Soteriou, M.C., Experiments on stability and oscillatory behavior
of planar buoyant plumes. Phys. Fluids 10 (1998) 16581665.
Cox, G., Basic considerations. In: Cox, G. (ed.), Combustion Fundamentals of Fire. Academic
Press, London (1995) pp. 130.
Elghobashi, S., Zhong, R. and Boratav, O., Effects of gravity on turbulent nonpremixed flames.
Phys. Fluids 11 (1999) 31233135.
Gebhart, B., Jaluria, Y., Mahajan, R.L. and Sammakia, B., Buoyancy-Induced Flows and
Transport. Hemisphere, New York (1988) pp. 657697.
Givi, P., Model-free simulations of turbulent reactive flows. Prog. Energy Combust. Sci. 15
(1989) 1107.
Grinstein, F.F. and DeVore, C.R., Dynamics of coherent structures and transition to turbulence
in free square jets. Phys. Fluids 8 (1996) 12371251.
Grinstein, F.F. and Kailasanath, K., Three-dimensional numerical simulations of unsteady
reactive square jets. Combust. Flame 100 (1995) 210.
Grinstein, F.F. and Kailasanath, K., Exothermicity and relaminarization effects in unsteady
reactive square jets. Combust. Sci. Technol. 114 (1996) 291312.

TRANSITIONAL NON-CIRCULAR BUOYANT REACTIVE PLUMES

12.
13.
14.
15.
16.
17.

18.
19.
20.
21.
22.
23.
24.
25.
26.

79

Gutmark, E.J. and Grinstein, F.F., Flow control with noncircular jets. Annu. Rev. Fluid Mech.
31 (1999) 239272.
Hermanson, J.C. and Dimotakis, P.E., Effects of heat release in a turbulent, reacting shear layer.
J. Fluid Mech. 199 (1989) 333375.
Huerre, P. and Monkewitz, P.A., Local and global instabilities in spatially developing flows.
Annu. Rev. Fluid Mech. 22 (1990) 473537.
Jiang, X. and Luo, K.H., Spatial direct numerical simulation of the large vortical structures in
forced plumes. Flow Turbul. Combust. 64 (2000) 4369.
Jiang, X. and Luo, K.H., Combustion-induced buoyancy effects of an axisymmetric reactive
plume. Proc. Combust. Inst. 28 (2000) 19891995.
Jiang, X. and Luo, K.H., Spatial DNS of the flow transition of a rectangular buoyant reacting
free-jet. In: Lindborg, E., Johansson, A., Eaton, J., Humphrey, J., Kasagi, N., Leschziner, M.
and Sommerfeld, M. (eds), Proceedings of the 2nd International Symposium on Turbulence
and Shear Flow Phenomena Vol. III. Royal Institute of Technology (KTH), Stockholm (2001)
pp. 297302.
Katta, V.R., Goss, L.P. and Roquemore, W.M., Numerical investigations of transitional H2 /N2
jet diffusion flames. AIAA J. 32 (1994) 8494.
Lele, S.K., Compact finite difference schemes with spectral-like resolution. J. Comput. Phys.
103 (1992) 1642.
Lingens, A., Reeker, M. and Schreiber, M., Instability of buoyant diffusion flames. Exp. Fluids
20 (1996) 241248.
Luo, K.H., Combustion effects on turbulence in a partially premixed supersonic diffusion flame.
Combust. Flame 119 (1999) 417435.
Maxworthy, T., The flickering candle: transition to a global oscillation in a thermal plume. J.
Fluid Mech. 390 (1999) 297323.
McMurtry, P.A., Riley, J.J. and Metcalfe, R.W., Effects of heat release on the large-scale
structure in turbulent mixing layers. J. Fluid Mech. 199 (1989) 297332.
Poinsot, T.J. and Lele, S.K., Boundary conditions for direct simulations of compressible viscous
flows. J. Comput. Phys. 101 (1992) 104129.
Thompson, K.W., Time dependent boundary conditions for hyperbolic systems. J. Comput.
Phys. 68 (1987) 124.
Vervisch, L. and Poinsot, T., Direct numerical simulation of non-premixed turbulent combustion. Annu. Rev. Fluid Mech. 30 (1998) 655691.

You might also like