You are on page 1of 12

Pergamon

Mathl.

Comput.

PII: SO8957177(96)00148-3

Modelling Vol. 24, No. 8, pp. 165-176, 1996 Copyright@1996 Elsevier Science Ltd Printed in Great Britain. All rights reserved 08957177/96 $15.00 + 0.00

The Baroclinic Effect in Combustion


G. A. BATLEY Department of Fuel and Energy Department of Applied Mathematics Leeds University, Leeds LS2 QJT, United Kingdom A. C. MCINTOSH Department of Fuel and Energy Leeds University, Leeds LS2 QJT, United Kingdom J. BRINDLEY Department of Applied Mathematics Leeds University, Leeds LS2 QJT, United Kingdom
baroclinic effect is due to the nonalignment of pressure and density gradients, and its result is to induce vorticity production. Because of the steep density gradients and the almost universal presence of ambient pressure signals in combustion systems, the baroclinic effect is a crucial mechanism for the production of turbulence. In this paper, we provide a review of the results of numerical simulations of baroclinic interactions between planar pressure signals and cylindrical laminar flame fronts. In the first section, the evolution of a flame front alfected by a single baroclinic impulse is considered. In the second section, we consider the effect of a double baroclinic impulse on a growing flame ball within a shock tube and examine the effect of viscosity on the evolution of the overall burning rate.

Abstract-The

Keywords-Baroclinic,

Vorticity, Turbulence,

Reaction progress variable, Favre averaging.

1. INTRODUCTION
Many of the early investigations into the baroclinic effect have been experimental. Studies by Markstein [l] in particular, demonstrated the possibility that a baroclinically induced vorticity field can eventually lead to the turbulent break up of a laminar flame. In the most famous of his experiments, that author introduced a planar shock through the open end of a shock The combined effect of the incident tube just after a point ignition event within the tube. shock, and the signal reflected by the opposite closed end of the tube, was to induce a strong vorticity field coincident with the growing spherical flame bubble. This vorticity field rapidly distorted the laminar flame front, and the eventual turbulent break-up of this front was clearly demonstrated. Further demonstrations of the baroclinic effect have been given by Scarinci [2], and Scarinci and Thomas [3]. These authors examined single shock interactions with spherical and cylindrical laminar flame fronts. Although this single passage is insufficient to cause such rapid transformation to turbulence and the same increase in overall burning rates, the experiments are useful in that they provide more convenient cases, for theoretical and numerical modelling. In this paper, we describe the results of numerical simulations of two simple baroclinic interactions. In both cases, the curved laminar flame front is cylindrical, and the pressure disturbance is planar. In the first type of interaction, the pressure disturbance encounters the flame once in an open region, thereby, inducing a roughly annular vorticity field within the flame front. The subsequent
Typeset 165

by &&-T@

166

G. A. BATLEY etd.

behaviour of the flame follows closely the behaviour of cylindrical vortex sheets as examined by Rottman and Stansby [4]. The second, more complex scenario, involves the double interaction of the pressure signal with the flame front within a shock tube which is closed at one end. In this case, the induced vorticity field is more complicated, in some regions. with the amplitude approximately doubled by the second interaction the overall behaviour is The importance of viscous diffusion in determining

highlighted by considering the results of a similar simulation with Prandtl number increased from unity to sixteen. It is observed that the increased viscosity slows the evolution to small scale features, thereby reducing the burning surface area, and therefore, the overall burning rate.

2. NUMERICAL

SCHEME
at the

The code used for this work is second order accurate in both space and time. The Godunov scheme employed divides the z-y plane into grid-cells, with the variable values calculated cell centroids. A nonlinear Riemann solver is then used to calculate the hyperbolic fluxes at the

cell boundaries. The first order time step takes the distributions within each cell to be uniform. The second order time step then gives second order space accuracy by using adjacent midcell values to calculate the flow quantity gradients. In order to maintain monotonicity, an averaging function is employed in the regions with large second derivatives to reduce the accuracy of the scheme to first order. Diffusive fluxes are then also calculated at the cell boundaries, and the reaction terms evaluated using cell-centroid values. The details of the hyperbolic part of the numerical scheme are described by Falle [5] for cylindrical geometry.

3. METHODOLOGY
The methodology is described in considerable detail elsewhere [6,7]. Briefly, the cylindrical flames are set up from ignition using a simplified version of the code designed for cylindrical symmetry. When the flame front has expanded to the required radius, the distribution is transformed to a 2-d grid, and a short length scale pressure step is introduced propagating towards the flame. The set of parameters used in this work is as follows: T, = 3OOK, Tb = 1500K, ToI = 0.2, e = 10, = 5 x 107s- 7 LewisNo. (= 9; y) =l , = 1, &I = 0.1 Jm-ls-lK-l, PU = 1.17kgnY3, and therefore, 4b

(3.1) (3.2) (3.3) (3.4) (3.5) (3.6) (3.7) (3.8) (3.0) (3.10)

Prandtl No. (=

= 9 x 10S5m2s-l.

PUCP

(Note that a single unimolecular decomposition reaction is assumed, and the rate of reaction is assumed to obey an Arrhenius temperature depence, i.e., -g .) ( > and the reaction zone thickness is For this parameter set, the planar flame speed is 7.75cms-, about 4 mm. A grid size of 0.4 mm is, therefore, used for these simulations. Reaction Rate = k,Cexp

Baroclinic

Effect

167

4. RESULTS
4.1. Single Passage of the integration disturbance domain. In the left hand immediately plot ( at t = 0.085s above the cylindrical

Figure

1 shows the full extent the planar pressure

after ignition)

is clearly visible

flame front. After 0.25ms (middle flame region inducing an annular vorticity field induces

plot of Figure I) the pressure disturbance has traversed the vorticity field shown in the middle plot of Figure 2. This above the flame ball to be drawn through the 1, and the continuing evolution of the density divides the original flame ball into two distinct drawn into spiral structures is highly within each of of the as in the [6], this behaviour by [4], and indeed reminiscent

cold gas from the region

flame region (see the right hand plot of Figure field shown in Figure 3). This cold gas eventually burning behaviour regions, with the flame front vortex itself, these regions. As pointed paper, out in an earlier the initial paper

being

of cylindrical

sheets as described vorticity

it is easy to see that variation around

the case of the latter

field does obey a sinusoidal

flame front. As the evolution continues, the final set of three plots in Figure 3 shows that the cold gas drawn into the two components is heated by the hot burnt gas, and begins to react. The final situation is that of two separate rotating reacting regions.

0.050

0.025

Figure 1. The initial evolution of the density distribution during the interaction between a positive pressure step with fractional amplitude 0.3 and a cylindrical flame ball of radius 2cm. l=O.W5s

I - 0.0@5258

t =0.06558

Figure 2. The initial evolution of the vorticity distribution during the interaction between a positive pressure step with fractional amplitude 0.3 and a cylindrical flame ball of radius 2cm.

168

G. A. BATLEY I = 0.06575s 0.150 -

et al.

t =0.086s

t =0.06625s

I = 0.0665s

t =0.06675s

t =0.067s

0.125 S! 5 E 0.100

(b)
I = 0.06725s 0.125 -

t =0.0675s

I =0.06775s

Figure 3. The continuing evolution of the density distribution ball of radius 2cm.

after the interaction

between a positive pressure step with fractional amplitude 0.3 and a cylindrical flame

4.2. Double Passage We now examine the effect of a double passage of a planar pressure step across a cylindrical flame front by modelling a growing flame ball within a shock tube. As shown in the left hand plot of Figure 4, the incident pressure signal is introduced through the open end and propagates down the tube interacting with the flame ball. The pressure signal is then reflected by the closed bottom end of the tube and retraverses the flame region further affecting the vorticity field. Much has been written about the crucial role of viscosity in direct numerical simulations involving significant vorticity fields. In particular, finite difference techniques rely on the presence of viscosity to retard the evolution to small scale structures, which would otherwise, continue past the point at which the grid can resolve them. Here, a second run has been carried out, with the only change being the increase in viscosity by a factor of sixteen. Figure 5 shows the density distributions at t = 1.5 ms, 2 ms, and 2.5 ms, during both the low and high viscosity integrations, respectively. At 1.5ms, after the introduction of the pressure signal (see the left hand plots of Figure 5), small closed loop filaments are seen to have been formed during the simulation of the low viscosity flame, whereas these features have been surpressed in the high viscosity case. The

Baroclinic
t=o

Effect
t =0.5-m
t=lms

169

0.175

0.150

0.125

0100

Figure 4. The initial evolution of the density distribution during the interaction between a positive pressure step with fractional amplitude 0.3 and a cylindrical flame ball of radius 3 cm. Grid size &r = dy = 0.4 mm.

t=1.5ms

t=2ms

t =2.5ms

Figure 5. The evolution of the density distribution between t = 1.5 ms and t = 2.5 ms, during the simulations of the low viscosity PT = 1 (top), and high viscosity PT = 16 (bottom) flames. Grid size da: = dy = 0.4 mm.

plots at t = 2 ms (middle plots of Figure 5) and t = 2.5 ms (right hand plots of Figure 5) further demonstrate this tendency. In each case, the two symmetric components are seen to be split into lower, large burning regions, and upper, small components. Within the upper components, very small scale features have developed in the low viscosity case. Cold fuel is drawn into the hot burnt gas, thereby, increasing the overall burning surface. However, when the viscosity is increased, evolution to these smallest scale features is clearly surpressed. An obvious fact which

170

G. A. BATLEY et al.

time(ms) Figure 6. The evolution of the total reaction rate during the two runs with low viscosity (Pr = l), and high viscosity (Pr = 16), respectively.

has been made use of in modelling mean turbulent reaction rates (see, for example, [8,9]) is that the overall burning rate within a given volume (G) is proportional to the total surface area of the flame front. The effect of viscosity in controlling the evolution toward small scale structures is, therefore, crucial in determining the overall reaction rate. In order to quantify this effect, the evolution of the total reaction rate, defined as R=

G k,pCe-EA~T,

is plotted in Figure 6, for both the low and high viscosity flames. This figure clearly shows the higher values attained in the case of lower viscosity. After 4.5ms, the two converge as the available fuel is reduced faster for lower viscosity.

5. MEAN

REACTION

RATE

The transition from laminar to turbulent combustion is poorly understood, and we are, therefore, interested in gaining some insight using numerical results of this kind of simulation. To this end an analagous simulation has been performed to that described above, with the Arrhenius reaction term replaced by a function of the reaction progress variable, c(= 1 - C). Numerical integrations can then be carried out in order to calculate mean, and fluctuating quantities. In this case, the previous reaction rate expression R = k,CeAEAIT, is replaced by

R=5

x 104c4 (1 - c) .

By using this approach, comparisons can be made between a numerically calculated average reaction rate, and the laminar flamelet expression of Bray, Libby and Moss [g], in which

Mean quantities are calculated by evaluating numerically integrals of the form 4 = l;htW,

Baroclinic

Effect

171

using Simpsons rule. Favre averaged quantities are calculated using the expression

Figures 7 and 8 show, respectively,

the values of the numerically

calculated

Favre averaged

reaction rate (R), and the expression ~(1 -c),

calculated using the numerically integrated values

of F. (Note that At is here equal to 2 x 10m5.) It is important to note that no account has been taken of the details of the B-M-L expression, so an exact numerical comparison between the two distributions is inappropriate, and the best we can do is to give a qualitative comparison.

Figure 7. This figure shows the values of the numerically integrated Favre averaged reaction rate R.

0.0000

0.0496

Figure 8. This figure shows the values of the expression E (I - CZ)

In fact, what these two plots show, is that while boundaries of the burning regions match quite closely for the two expressions, the overall agreement is very poor. In particular, the directly integrated expression is near zero within each component, while the value of the expression E (1-F) is quite significant in the same regions. Several different values for At have been tried, but in each

172

G. A. BATLEY et al.

case the two expressions show the same lack of correspondence. In addition, unless the value of At is less than a few time steps, we have found that fluctuating components turn out to be much larger than Favre averaged values. We must, therefore, conclude that this numerical integration procedure Instead Although modelled the ground flames. is incapable no simulations the propagation of capturing the transition from laminar involving to turbulent already combustion. flames. This lays we are interested in examining of cylindrical similar interactions fully turbulent /c-c approach.

of these interactions turbulent

have been carried pressure

out as yet, we have successfully with fully turbulent

flames using a standard

work for considering

baroclinic

wave interactions

6. TURBULENT
In this modelling. section, we describe numerical solutions The full set of equations is as follows:

FLAMES
to turbulent flame equations used for /C-E

apl3i~ W) + at
dXj

~pc4

axj

a(p)
at

+ ap&&
dXj -=-

a (pij + p/u:) axj

In the above set,

k is the turbulent kinetic energy (= f&/2), and E is the turbulence dissipation rate. The usual notation for conventional, and Favre averaging, as well, as fluctuation terms, has been adopted. In this work, the terms -j%, and -C,@2/k have been omitted from the k, and c equations, respectively, so that k, and E remain fixed in uniform regions. The Reynolds stress tensor is modeled using the usual Boussinesq approach, i.e., 4 86 pk--pt--+-pt 3 ax -/Lt(g+n) diffusion, i.e., 2 3 E ay

pz1;2c[i= ( and a simple gradient

model is used for turbulent

where crt is the turbulent viscosity) is given by

Prandtl/Schmidt

number

for 9.

The turbulent

viscosity

(or eddy

Baroclinic Effect

173

The following

set of constants

is adopted:

c, = 0.09,
C,r = 1.44, C& = 1.92, UC = 0.7, Ue = 1.0, 0k = 1.0, 0, = 1.30. Another important aspect of the modelling, is how to deal with the mean reaction model, is rate term G. for the

We have used a modification flame surface

of the Bray-Libby-Moss The final expression

which uses an expression

area per unit volume.

where the subscript R refers to quantities measured within the reactants, 5: is the laminar burning speed, g is a constant derived from the pdf. of crossing lengths (= l.5), c is a reaction progress variable (which in our case is equal to l- fuel mass fraction), er, is the mean cosine of the angle between the mean contour, and the instantaneous flame at the crossing point (= 0.5), and .&, is the principal large length scale of the flame (= CE.751c1.5/e). The flame is initiated by adopting the following initial distribution of the reaction progress variable (c) : .
c=

(l- ($4)exp(-10(~)10).
p=Pu 1+7c

The initial

density

distribution

is then given by

The remaining flow quantities have an where T is the heat release parameter (Ta - T,)/T,. initially uniform distribution. Although some success has been achieved in modeling the propagation of cylindrical laminar flames, we have been restricted so far, to extremely large values for the initial turbulent kinetic energy, which is usually expected to be less than unity. Figures 9 and 10 show the evolution of the flow quantities during the propagation of two flames. Both integrations used a grid size of 0.2 mm, and the initial turbulence length scale was taken to be 0.5 mm. The flame shown in Figure 9 had an initial k value equal to 64, while in Figure 10 this value was equal to 4. Note that as with a laminar flame the propagation speed, ST, and length scale, ZF, of a turbulent flame are proportional to the square root of the turbulent mass diffusivity. As described above, the diffusivity (DT) obeys DTo(C. We also have the defining formula for the turbulence i Y %n this work, the initial value of I$, remains e length scale, &,, which is given by

= C;.75k1.5 c .

fixed at 0.5 mm. Hence, we have &cc&,

174 Density

G. A. BATLEYet

al.

Radial Velocity t-am

0.2 00 005 0.1 r (m) 015 02

-10 00 005 01 r (m) 015 02

710 1

Pressure
l-am

Reaction Progress Variable

f
5

610

m SlO-

t-am

% ClOt

0 3108 g 210-

t-1llU

1'10'
0.0

005

0.1 r(m)

0.15

1 02

00

0.05

01 r(m)

015

I 0.2

(4
Turbulent Kinetic Energy t=*
3416 2c.w 2.OW

Tuttwlence Dissipation Rate

5 5 w

1 5w 1 Ow 5010* 0.01~

010

005

0:1
r Cm)

015

02

00

005

01 r(m)

015

0.2

0.54 0.52

Turbulence Length Scale

WC-J-

t-am

-c
t

moBQ)-

0.50

E 5043
0.46 0.44 I
0.0
1 I , !

d [I
I

400

0.05

01

015

02

04 00

005

01 f (m)

0;s

--I 02

r (ml

(b)
Figure 9. A propagating cylindrical turbulent flame with T = 4, k = 64 m2s-2, I!,, = 0.5 mm initially, dr = 0.2 mm and SOL = 0.3 me--l. and

and therefore, that ST,lFO:k


0.25

For the two flames shown here, the ratios of flame speeds, and length scales are, therefore, expected to be equal to 2. This corresponds quite well with the results shown in the two latter plots. Of course because the k, and E distributions do not remain uniform, these ratios are not expected to be exact because the diffusion coefficients clearly vary with k, and E.

Baroclinic

Effect

175

Radial Velocity

-IO--

0.1
r(m)

0,;s

0.2

00

005

0.1 0)

0.15

0.2

Pressure

1.0

Reaction Progress Variable

: .E 0.6 s 2 06

i? t 04 s 'ij 0 0.2 $

01
r(m)

01 r(m)

0.15

0.2

(4
77
6-

Turbulent Kinetic Energy


t=4m t-m

6.10' 5'10' 4'10'

Turbulence Dissipation Rate

f5 E AC 4-

r x g w

3'10' 2'10'

2 0.0

I 0.05

1 0.t r(m)

I 0.15

I 0.2

Ovd 00 005 01 r(m) 015 02

Turbulence Length Scale


I
1.0

1000

Reaction Rate

-g

0.6

E _r
0.6

.
0.0 r(m)

02

Figure 10. A propagating cylindrical turbulent flame with T = 4, k = 4m29-2, i y = 0.5 mm initially, dr = 0.2 mm and SOL = 0.3 ms-

and

176

G. A. BATLEY et al.

7. CONCLUSIONS
Baroclinic interactions between planar pressure signals, and curved laminar flame fronts, have been successfully modelled in this paper. In the case of the single interaction we have seen that the flame is split into two components, and that fairly well defined spiral structures form within each. This early behaviour is dominated by purely play an important role in controlling the tightening effects of the chemical reaction, and thermal/mass is shown by the combustion of the material drawn In the case of the double baroclinic interactions Eulerian effects, although the viscosity does of the spiral arms. It is also clear that the diffusion are significant within into the burnt gas regions. within 2-3ms. This

the shock tube, we have again shown the

dramatic effect of the induced vorticity field in distorting the laminar flame front and splitting the flame into separate components. We have also shown that increased viscosity retards the formation of very small scale features, thereby, reducing the overall burning surface area. We have succeeded bulent kinetic energy in setting required up propagating to maintain turbulent cylindrical flames, although the turto numerical stability is very high. It is intended

investigate further these turbulent flames with lower kinetic energies by altering the initial distribution of the reaction progress variable, and to consider the effect of further imposed pressure interactions.

REFERENCES
1. G.H. Markstein, Non-Steady Flame Propagation AGARDogmph 75, Pergamon Press, Oxford, (1964). 2. T. Scarinci, The generation of vorticity by shock waves in a reactive flow, B. Eng. Thesis, McGill University, Montreal, Canada, (1990). 3. T. Scarinci and G.O. Thomas, Some experiments on shock-flame inters&ion, Report DET905, University College of Wales, Aberystwyth, (1992). 4. J.W. Rottman and P.K. Stansby, On the 6-equations for vortex sheet evolution, J. Fluid Mech. 247, 527-549, (1993). 5. S.A.E.G. Falle, Self-similar jets, Mon. Not. R. Ask. Sot. 250, 581-596, (1991). 6. G.A. Batley, A.C. McIntosh, J. Brindley and S.A.E.G. Falle, A numerical study of the vorticity field generated by the baroclinic effect due to the propagation of a planar pressure wave through a cylindrical premixed laminar flame, J. Fluid Mech. 274, 217-237, (1994). 7. G.A. Batley, A.C. McIntosh and J. Brindley, Baroclinic distortion of laminar flames, Proc. Roy. Sot. A 452, 199-221, (1996). 8. K. Bray, A. Libby and J.B. Moss, Flamelet crossing frequencies and mean reaction rates in premixed turbulent combustion, Comb. Sci and Tech 41, 143-172, (1984). 9. K. Bray, A. Libby and J.B. Moss, Flamelet Crossing frequencies and mean reaction rates in premixed turbulent combustion, Comb. Sci. and Tech. 41, 143-172, (1984). 10. Bray, Libby and Moss, Unified modelling approach for premixed turbulent combustion-Part I: General formulation, Comb. and Flame 61, 87-102, (1985).

You might also like