You are on page 1of 10

Proceedings of Eurotherm78 – Computational Thermal Radiation in Participating Media II

5-7 April 2006, Poitiers, France

AN EXPLORATORY INVESTIGATION OF
RADIATION STATISTICS IN HOMOGONEOUS
ISOTROPIC TURBULENCE

by C.B. da SILVA(*), I. MALICO(**) , P.J. COELHO(*) and J.C.F. PEREIRA(*)

(*)Instituto Superior Técnico, Technical University of Lisbon, Mechanical Engineering Department


Avenida Rovisco Pais, 1049-001 Lisboa, Portugal.
(**)Universidade de Évora, Physics Department, R. Romão Ramalho, 59, 7000-671 Évora, Portugal.
coelho@navier.ist.utl.pt

Abstract
A fundamental study of radiation statistics in homogeneous isotropic turbulence is presented. A
pseudo-spectral code is used to simulate isotropic turbulence by means of DNS of the full Navier-
Stokes equations. The instantaneous scalar data is used to calculate the radiation intensity along a line
of sight using the statistical narrow band model. The mean, variance, skewness and flatness of
radiation intensity were obtained for conditions observed downstream of the flame tip of a piloted
turbulent jet flame, where the statistics of the flow field are close to the ones found in isotropic
turbulence. The joint probability density function between the temperature and the radiation intensity
is presented, as well as the spectra for the radiation intensity. The present one way coupling
philosophy used to connect isotropic turbulence data with radiation computations shows the correct
trends and allows one to study the detailed effects of the turbulent characteristics upon the structure of
the radiation intensity field.

Nomenclature Reλ Reynolds number σ Stefan-Boltzman


D Diameter s Coordinate along an optical constant
E Energy spectrum path τ Transmissivity
F Flateness S Skewness
I Radiation intensity Sc Schmidt number Subscripts
k Parameter of SNB model tref Reference time in the DNS b blackbody
kp Wave number where data bank rad radiation
forcing is applied T Temperature ref reference
kmax Maximum resolved wave Vc Characteristic velocity s species
number x Molar fraction; λ Taylor microscale
r
K Wave number vector Axial coordinate ν wavenumber;
L Size of the computational molecular viscosity
box for isotropic Greek symbols
turbulence γ/δ Parameter of SNB model
L11 Turbulence integral scale ∆ν Narrow band width
N Number of grid nodes ε Dissipation rate of turbulent
along an optical path kinetic energy
p Total pressure η Kolmogorov micro-scale
P Forcing intensity ηB Batchelor micro-scale
r Radial coordinate κ Absorption coefficient
1. Introduction
Most flows of practical relevance are turbulent. In some of them, radiative heat transfer
plays an important role, e.g., in most combustion equipment (furnaces, boilers, engines) and
in fires. Presently, it is widely accepted that the interaction between turbulence and radiation
(TRI) yields a significant increase of the radiative heat fluxes in comparison with laminar
flows. It has been reported that TRI may enhance mean radiation levels by more than 100% in
comparison with estimates based on mean scalar properties (see, e.g., [1]). However, our
knowledge about such interaction is rather limited. As a consequence, the interaction between
turbulence and combustion is often ignored or accounted for by assuming that it is only due to
the influence of temperature fluctuations on the emissive power of the medium.
A more accurate approach to account for the TRI requires the solution of the radiative heat
transfer equation with the mean radiation intensity as the dependent variable. Still, there are
no models to approximate the correlation between fluctuations of the radiation intensity and
fluctuations of the absorption coefficient. This correlation is generally ignored, assuming that
the individual eddies are homogeneous, optically thin and statistically independent. This is the
so-called optically thin eddy approximation. Nevertheless, the error of this approximation is
unknown. It has been claimed that the error introduced by this approximation is small, even if
it does not hold for wavelengths in the vicinity of the centre of the most important absorption
bands of radiatively participating gaseous species.
Another powerful method to account for the TRI consists of the solution of the radiative
transfer equation for instantaneous values of scalars (temperature and species concentrations).
This requires a sufficiently large number of solutions based on instantaneous scalar
distributions to ensure statistically meaningful results. Moreover, the instantaneous scalar
distributions are not available, and need to be generated in such a way that the mean values,
variance and relevant spatial and temporal auto-correlations and cross-correlations are
satisfied. Stochastic models for the generation of time series of scalar distributions have been
developed for this purpose [1-4]. The large number of realizations implied by the stochastic
methods prevents their application to practical engineering calculations.
Direct numerical simulation (DNS) is a powerful tool to provide fundamental and reliable
insight on turbulent flows, although it cannot be applied to engineering calculations owing to
the computational requirements. Homogeneous isotropic turbulence consists in the simplest
possible flow configuration in which the Navier-Stokes equations are solved in a box with
periodic boundary conditions in the three spatial directions. Instantaneous fields from DNS of
isotropic turbulence show the existence of intense regions of vorticity in the form of tubes
(“worms”) with radii and length of the order of the Kolmogorov and integral scales,
respectively [5-6]. DNS of isotropic turbulence has been used to study some detailed aspects
of turbulence physics, e.g., to test the hypothesis underlying the Kolmogorov-Obukhov
theory, and to analyze the local energy cascade in the inertial range. Other works used DNS
of isotropic turbulence to develop and assess new turbulence models. More recently, some
works used direct numerical simulations of isotropic turbulence to study fundamental issues
related to combustion, e.g. the analysis of flamelets in premixed turbulent combustion [7].
Recently, DNS has been used to investigate TRI in a premixed combustion system [8]. In
the present study, a classical pseudo-spectral code for the simulation of isotropic turbulence in
non-reactive flows [9] is used, which allows the computation of forced and freely decaying
isotropic turbulence by means of DNS of the full Navier-Stokes equations. The DNS fields
are completely characterised by their variance and by the shape of their three-dimensional
spectra for the kinetic energy and scalars, such as temperature and chemical species.
Combustion is not considered in the present work, so that the temperature and species
concentration fields are simply transported by the turbulent flow. The instantaneous fields of
temperature and chemical species are used as input data for the calculation of the radiation
intensity using a ray tracing method and a statistical narrow band model (SNB). This allows
the calculation of statistical data of the radiation intensity field, namely the mean, variance,
skewness and flatness of radiation intensity, joint probability density function between the
temperature and the radiation intensity and spectra for the radiation intensity. The present
paper presents and discusses this statistical information. There is no influence of the radiative
transfer calculations on the DNS simulation, i.e., there is only a one-way coupling between
turbulence and radiation. The goal here is to prepare the ground for future works where the
effects of the turbulent statistics on the radiative properties of the medium will be thoroughly
investigated.

2. Direct numerical simulation of isotropic turbulence


The numerical code used in the present DNS simulations is a standard pseudo-spectral
code in which the temporal advancement is made with an explicit 3rd order Runge-Kutta
scheme. Details about the DNS solver may be found in [10]. The physical domain consists in
a periodic cubic box of side 2π and the simulations were fully dealised using the 3/2 rule.
The instantaneous field of a passive scalar is taken from a DNS simulation of statistically
steady (forced) homogeneous isotropic turbulence using N = 192 collocation points in each
direction. Table 1 lists the details of the simulation. Both the velocity and scalar large scales
were forced in order to sustain the turbulence using the method described by Alvelius [11].
The forcing was imposed on 3 wave numbers concentrated on kp = 3. The same data bank was
recently used by da Silva and Pereira [9]. After an initial transient that lasts about 10 tref,
where tref = (Vc kp)-1, Vc = (P/kp)1/3, and P is the forcing intensity [11], the flow reaches a state
where all the turbulence quantities are statistically stationary. The analysis was made using 10
instantaneous fields taken from this region, separated by about 0.5 tref.
Notice that L > 4 L11 in all simulations, where L is the box size and L11 is the integral scale,
so that the size of the computational domain does not affect the larger flow structures [12].
Also, to ensure a good resolution of the dissipative scales we have kmax η > 1.5 and kmax
ηΒ > 1.5 in all simulations, as recommended by Pope [12], where η = (ν3/ε)1/4 and ηB =
η/Sc1/2 are the Kolmogorov and Batchelor micro-scales, respectively.
Figure 1 shows the kinetic energy spectra for the present simulation. That the dissipative
scales are indeed being well resolved is attested by the small upturn at the end of the wave
number range. The velocity spectrum has a –5/3 range which shows the existence of an
inertial range region. Moreover, the values of the skewness and flatness of the velocity
derivative oscillate around about -0.5 and 4.0 respectively, and are quite close to the ones of
Jimenez et al. [6] for similar Reynolds numbers.

Reλ ν Sc kmax η kmax ηΒ L11 η ηB S F Smix


95.6 0.006 0.7 1.8 2.1 1.24 2.8×10-2 3.3×10-2 -0.49 +4.63 -0.46
Table 1 - Details of the DNS (Reλ - Reynolds number based on the Taylor micro-scale and
rms of the velocity fluctuations; S - Skewness of the velocity derivative; F - Flatness of the
velocity derivative; S mix = S (∂ u ∂ x ) (∂ θ ∂ x )2 - Mixed derivative skewness).
108

106 Reλ=95.6
(-5/3)
4
10

E(k)/(εν5)1/4
102

0
10

10-2

10-4

10-6
0.5 1 1.5 2

Fig. 1 - Energy spectrum for the present simulation.

3. Radiative transfer calculations


The radiative transfer equation for a non-scattering medium may be written as
d Iν
= −κν Iν + κν I bν (1)
ds
After some algebra, the integration of this equation along a line of sight yields [13]
 s  s  s 
Iν (s ) = Iν (0 ) exp −
 ∫0 κν (s ′) ds ′  +
 0 ∫I bν (s ′) κν (s ′) exp −
 ∫ s′ κν (s′′) ds′′  ds′ =
(2)
s ∂ τ (s ′ → s )
= Iν (0) τν (0 → s ) + ∫0 I bν (s ′) ν
∂ s′
ds ′

In the present work, the radiative properties of the medium are evaluated using the SNB
model [14]. Moreover, a cold boundary is assumed, i.e. Iν(0) = 0. Therefore, Eq. (2) is
integrated over a narrow band, so that the mean radiation intensity over a narrow band of
width ∆ν is given by
1 s ∂ τ ∆ν (s ′ → s )
I ∆ν (s ) =
∆ν ∫∆ν Iν (s ) dν = ∫0 I b,∆ν (s ′)
∂ s′
ds ′ (3)

In the case of a homogeneous gas layer at total pressure p, the mean transmissivity is given
by [14]
 γ  δ 
τ ∆ν (0 → s ) = exp − 2  1 + x s p s k − 1
  (4)
 δ  γ 
where k and 1/δ are parameters of the model, which were taken from the data of Soufiani and
Taine [15]. For a given absorbing species, these parameters are tabulated as a function of
temperature and spectral location. The parameter γ is a function of temperature, pressure and
species concentration, also given in [15]. In the case of a non-homogeneous medium, the
Curtis-Godson approximation [16] is employed.
The integrals in Eqs. (3) and (4) are numerically evaluated using Simpson’s rule, and the
parameters k and 1/δ are interpolated from the tabulated data using cubic splines. This keeps
the order of accuracy of the numerically evaluated radiation intensity consistent with the
order of accuracy of the DNS solver. Therefore, Eq. (3) is discretized as
N ( )
I b,∆ν (s n ) + 4 I b,∆ν s n+1 2 + I b,∆ν (s n +1 )
I ∆ν (s N +1 ) = ∑ 6
×
(5)
n =1
[τ ∆ν (s n+1 → s N +1 ) − τ ∆ν (sn → s N +1 )]
where N is the total number of grid points along the optical path in the DNS calculations. In
the DNS calculations, the boundaries of the computational domain are periodic. Therefore,
the instantaneous scalar data at the first point n =1, where the radiation intensity is
determined, are identical to those at n = N+1. The temperature and the species concentration
at points sn+1/2 are not available, and so they are interpolated from the DNS data using again
cubic splines.

4. Computational details
The computational domain in the DNS calculations is a cubic box of side 2π. The selection
of the size of the radiation domain, also taken as a cubic box, is not so straightforward. Here,
the size of the radiation domain was chosen in such a way that the ratio of the size of the box
in the DNS calculations to the integral length scale, L11, is equal to the ratio of the size of the
radiation domain to the integral length scale in the radiation calculations The implications of
this choice deserve further examination, which will be investigated in the future.
In order to rescale the data from isotropic turbulence simulations obtained in a cubic box
of side 2π into a radiation domain with a different size, kinematic similarity between the two
flows was assumed. Therefore, the instantaneous temperature field used in the radiation
r
computations, Trad( x ), was determined using the instantaneous temperature field from the
r
DNS data, TDNS( x ), through

r r r ′2 >
< Trad
Trad ( x ) =< Trad ( x ) > +TDNS ( x ) (6)
′2 >
< TDNS
r r
where <Trad( x )> is the mean temperature at point x , and < Trad ′ 2 > and < TDNS
′ 2 > are the
variance of the temperature fields from the radiation and from the isotropic turbulence
simulation, respectively. Similarly, for the field of the molar fraction of the absorbing species,
we use

r r r < x ′s2,rad >


x s ,rad ( x ) =< x s ,rad ( x ) > + x s , DNS ( x ) (7)
< x ′s2,DNS >

It was assumed that the temperature and the absorbing species fields are fully correlated.
This is consistent with combustion models that relate the instantaneous thermochemical state
of the gaseous mixture to a single scalar, typically mixture fraction, e.g., the laminar flamelet
model. The experimental data reveals a very strong correlation between the temperature and
the molar fractions of H2O and CO2, which supports this assumption. Although radiation from
H2O could easily be included in the calculations, only radiation from CO2 is considered here.
This is mainly to reduce the computational requirements. Accordingly, the same scalar field
from the DNS calculations was used to prescribe both temperature and molar fraction of CO2,
r r
by setting TDNS( x ) = xs,DNS( x ).
The present radiative transfer calculations consider a purely idealized scenario, and no
extrapolation to practical applications is envisaged. Nevertheless, to ensure that the input data
considered in the radiative calculations are typical of those found in real flames, the mean and
rms of temperature and CO2 molar fractions were prescribed from the experimental data
available for a piloted CH4/air turbulent jet diffusion flame, the so-called flame D [17-19].
The mean and variance of these scalars, which are taken from a point in the flame, are
prescribed over the entire radiation domain. The integral length scale in the radiation
calculations is also taken from the experimental data [19].
The experimental data at x/D = 75, and for several radial locations, was used. This station
is located downstream of the flame tip, and it is the station at a largest distance from the
burner exit where experimental data is available. Hence, this is where the topology and
statistics of the flow are closer to the ones found in isotropic turbulence [20]. Radiative
calculations are performed for every radial location using the instantaneous scalar data
obtained from a unique DNS simulation by means of Eqs. (6) and (7). The mean and variance
of temperature and CO2 molar fraction used in these equations, which correspond to a single
point in flame D, are the same over all the radiation domain.
It is important to stress that the values predicted here cannot be compared with the
measurements and predictions reported in [19], not only because the radiation from H2O was
not taken into account in the present work, but also because the values in that reference were
obtained for the actual flame, in which the mean temperature and the species concentration
change along the optical path in the radial direction. On the contrary, the present values are
obtained assuming that the mean temperature and CO2 concentration taken from the
measurements at a fixed r/D remain unchanged in the radiation domain, and only the
instantaneous values change. This is consistent with the direct numerical simulation of
homogeneous isotropic turbulence.
Under the conditions of homogeneous and isotropic turbulence, the statistical data
computed from a temporal series of scalar data along a single optical path parallel to a
coordinate axis is identical to the statistical data calculated from all optical paths parallel to
the coordinate axes at a given time. The statistical data reported below was obtained from the
DNS data at a given time, using all the available optical paths parallel to the coordinate axes,
which are statistically indistinguishable. This means that 3×N2 ≈ 1.1×104 samples are used to
obtain the results described below. It is assumed that there is no radiation entering the
radiation domain, as stated above. The influence of this assumption on the computed
statistical data will be the subject of further research.

5. Results and discussion


Radiation statistics are computed for the mean spectral radiation intensity over a
narrowband of width 25 cm-1, centred at 2.7 µm, for an optical path parallel to a coordinate
axis, evaluated using Eq. (5). Hereafter, this mean spectral radiation intensity over a
narrowband is denoted by I, for the sake of simplicity. Additional calculations were
performed for the mean spectral radiation intensity over a narrow band centred at 4.3 µm, and
for the total radiation intensity. However, the results of these calculations are not shown here,
since they do not provide any additional useful information. Moreover, although experimental
radiation statistics, e.g., PDFs and power spectral densities, are available in the literature [1,
2, 19, 21], they are not comparable with the present results, which are only applicable to
homogeneous isotropic turbulence.
2 -1
103 < I > [W/m sr cm ]
'2 2 4 2 -2
2
< I > [W /m sr cm ]
10
S(I)
F(I)
101

100

10-1

-2
10

10-3

10-4

10-5

-6
10
0 2 4 6 8 10

r/D
Fig. 2 - Mean, variance, skewness and flatness of radiation intensity as predicted by the
present model using the jet flame D data at x/D=75.

Figure 2 shows the mean, variance, skewness and flatness of the mean spectral radiation
intensity over the narrowband centred at 2.7 µm. The variance of the radiation intensity is
defined by <I’2>, where the brackets represent an average over all the samples from the
radiation domain computations. The skewness, S(I) = <I´3>/<I´2>3/2, represents the degree of
symmetry of the probability density function (PDF) of the radiation intensity (S = 0 for a
symmetric PDF), and the flatness factor, F(I) = <I´4>/<I´2>2, is a measure of the degree of
intermittence of a given variable (F = 3 for a Gaussian function).
Both the mean and variance of the mean spectral radiation intensity over a narrowband
attain their maxima at the centre of the jet, where the temperature and CO2 concentration are
highest, as shown in Fig. 2. The temperature and the CO2 molar fraction decrease along the
radial direction in the flame, and so do <I> and <I’2>. However, the turbulence intensity,
taken as ′ 2 > < Trad > , increases along the radial direction. This implies an increase of
< Trad
the influence of turbulence on radiation as r/D increases, as shown in Table 2. This trend is
also observed in [19], although the values cannot be compared, as explained above.
If there were no fluctuations of the absorption coefficient of the medium, then the
temperature self-correlation would be fully responsible for the influence of turbulence on
radiation. The temperature self-correlation is given by [22]
<T4 > < T ′2 > < T ′3 > < T ′4 >
= 1+ 6 +4 + (8)
< T >4 < T >2 < T >3 < T >4
The two first terms on the right side are generally dominant. If the PDF of temperature is
Gaussian, then the third term on the right side is zero and the last one is equal to 3 ×
(< T ′ 2 >/<T>2)2. Table 2 confirms the importance of the temperature self-correlation, and
demonstrates that the fluctuations of the absorption coefficient of the medium contribute to
enhance the influence of turbulence on radiation. The mean spectral radiation intensity over a
narrowband increases by 34% due to turbulence at r/D = 0, while it increases by 92% at r/D =
8.33. This supports the results from [19, 21], who concluded that the investigation of flame
radiation along chord-like paths is important to understand TRI, providing a more challenging
test to predictive models than diametric paths.
r/ D ′ 2 > < Trad >
< Trad < T 4 > < T >4 (a)
< T 4 > < T >4 (b)
< I ′2 > < I >

0 0.183 1.200 1.204 0.340


2.78 0.226 1.306 1.314 0.430
5.56 0.307 1.565 1.592 0.631
8.33 0.343 1.706 1.748 0.916
(a)
Computed using first two terms of Eq. (8)
(b)
Computed using Eq. (8) and assuming a Gaussian PDF for the temperature
Table 2 – Temperature self-correlation and ratio of the rms to the mean spectral radiation
intensity over a narrowband.

The skewness and flatness, on the other hand, seem to increase slightly with r/D. At the
centreline, S(I)=0.44 and F(I)=2.8, whereas at r/D=8.33 we have S(I)=1.50 and F(I)=5.2. It is
not clear why the asymmetry and intermittency of the radiation intensity increase with the
distance from the flame axis.
Figure 3 shows the contour plots of temperature and blackbody radiation intensity obtained
from the DNS computations in the first (y,z) plane of the domain. The contours of the mean
spectral radiation intensity over the narrowband in the same plane are also shown using the
flame data at x/D = 75 and r = 0. The correlation coefficient between T and I is equal to about
50%. The correlation coefficients at other r/D locations are similar. It should be noticed that
the radiation intensity at a grid node in the first (y,z) plane depends on the temperature and
CO2 molar fraction along the optical path in x direction, and not only on the local properties.
Therefore, it is not surprising that this correlation is not very high.
To have a global picture of this result, Figure 4 shows the joint PDF of T and I from all the
available samples. In agreement with the previous results, the joint PDF of these quantities
shows that the correlation between them is associated with extreme events of both variables.
The spectrum for the mean spectral radiation intensity over a narrowband at x/D=75 and
for several radial locations of flame D is shown in figure 5. The spectrum E(K) is defined by
r r E(K2D ) r ' r
< Iˆ( K ' ) Iˆ( K ) >= δ (K + K ) (9)
π K2D
r r
where Iˆ( K ) is the two dimensional Fourier transform of I ( x ) , given by
+∞ +∞ rr r r
r 1 r
Iˆ( K ) = ∫∫ I ( x ) exp(−ik .x )dKdx (10)
(2π ) 2 −∞ −∞

r r
and K = ( K1 , K 2 ) is the two-dimensional wave number vector of norm K = | K | . Thus, the
energy spectrum represents the spectral distribution of the energy associated to each wave
number. Figure 5 shows that the overall amount of radiative energy decreases with the radial
position, as expected, and in agreement with Fig. 2. It seems that the slope of the radiation
intensity spectrum, at the inertial range region, changes slightly with the distance from the jet
centreline. Moreover, the characteristic bump caused by the forcing at low wave numbers
seems to be attenuated near the jet centreline. These are only some of the issues we intend to
explore in future works.
2 -1
T (K) Ib (W/m sr cm ) I (W/m sr cm )
2 -1

1800 1.9E+05 0.18


1700 1.8E+05 0.17
1600 1.7E+05 0.16
1500 1.5E+05 0.15
1400 1.4E+05 0.14
1300 1.3E+05 0.13
1200 1.2E+05 0.12
1100 1.0E+05 0.11
1000 9.3E+04 0.1
900 8.0E+04 0.09
800 6.8E+04 0.08
5.6E+04 0.07
4.4E+04 0.06
3.1E+04 0.05
1.9E+04 0.04

Fig. 3 - Contours of temperature (left), blackbody radiation intensity (centre) and mean
spectral radiation intensity over a narrowband (right) in the first (y,z) plane of the domain.

10
0 r/D=0.0
r/D=2.7
0.2 10
-2
r/D=5.6
r/D=8.3
-4
10
I(W/m sr cm )
-1

-6
0.15 10

-8
E(K) 10
2

-10
10
0.1
-12
10

-14
10
0.05
-16
10

500 1000 1500 -18


10
T(K)
0 1 2
10 10 10

K
Fig. 4 - Joint PDF between the temperature Fig. 5 - Spectra for the mean spectral radiation
and the mean spectral radiation intensity over intensity over a narrowband for several radial
a narrowband obtained with data taken from locations obtained with the present model
three boundary planes of the turbulent box. using the flame D data at x/D=75.

6. Conclusion
Classical one point statistics, correlations and power spectra of radiation intensity were
computed from an idealized model combining direct numerical simulations of isotropic
turbulence, comprising temperature and concentration fields, coupled with radiative transfer
calculations. The statistics demonstrate that the turbulence has a strong influence on the
radiation for the studied cases, yielding an increase of radiation intensity due to turbulence
that ranges from 34% to 92%. The increase is largest away from the flame axis, where the
turbulence intensity is also highest. It is hoped that the numerical tools employed here, when
further exploited, will shed more light into the complex nature of the turbulence/radiation
interactions. In particular, the influence of the chemical composition of the medium, mean gas
temperature, Reynolds number, shape of the kinetic energy and temperature spectra, and
turbulence intensity will be investigated.

7. Acknowledgement
This work was developed within the framework of project POCI/EME/59879/2004, which
is financially supported by FCT-Fundação para a Ciência e a Tecnologia, programme POCI
2010 (29.82% of the funds from FEDER and 70.18% from OE).
REFERENCES
[1] KOUNALAKIS, M.E., GORE, J.P., FAETH, G.M. “Turbulence/radiation interactions in nonpremixed
hydrogen/air flames”, 22nd Symposium (Int.) on Combustion, pp. 1281-1290, 1988.
[2] KOUNALAKIS, M.E., SIVATHANU Y.R., FAETH, G.M. “Infrared radiation statistics of nonluminous
turbulent diffusion flames”, J. Heat Transfer, vol. 113, pp. 437-445, 1991.
[3] KRITZSTEIN, F., SOUFIANI, A. “Infrared gas radiation from a homogeneously turbulent medium”, Int.
J. Heat and Mass Transfer, Vol. 36, pp. 1749-1762, 1993.
[4] CHAN, S.H., PAN, X.C. “A general semicausal stochastic model for turbulence/ radiation interactions in
flames”, J. Heat Transfer, Vol. 119, pp. 509-516, 1997.
[5] VINCENT, A., MENEGUZZI, M. “The spatial structure and statistical properties of homogeneous
turbulence”, J. Fluid Mech., Vol. 225, pp. 1-20, 1991.
[6] JIMENEZ, J., WRAY, A.A., SAFFMAN, P.G., ROGALLO, R.S. “The structure of intense vorticity in
isotropic turbulence”, J. Fluid Mech., Vol. 255, pp. 65-90, 1993.
[7] GIRIMAJI, S.S., POPE, S.B., “Propagating surfaces in isotropic turbulence”, J. Fluid Mech., Vol. 234,
pp. 247-277, 1992.
[8] WU, Y., HAWORTH, D.C., MODEST, M.F., CUENOT, B. “Direct numerical simulation of turbulence/
radiation interaction in premixed combustion systems”, Proc. Combustion Institute, Vol. 30, pp. 639-646, 2005.
[9] DA SILVA, C.B., PEREIRA, J.C.F. “On the local equilibrium of the subgrid-scales: the velocity and
scalar fields”, Phys. Fluids, Vol. 17, 108103, 2005.
[10] CANUTO, C., HUSSAINI, M.Y., QUARTERONI, A., ZANG, T.A., “Spectral Methods in Fluid
Mechanics”, Springer-Verlag, New York, pp. 201-212, 1988.
[11] ALVELIUS, K. “Random forcing of three-dimensional homogeneous turbulence”, Phys. Fluids, Vol.11,
pp. 1880-1889, 1999.
[12] POPE, S. “Turbulent flows”, Cambridge University Press, 2000.
[13] MODEST, M.F. “Radiative Heat Transfer”, McGraw-Hill, New York, 2nd Edition, 2003.
[14] MALKMUS, W. “Random Lorentz band model with exponential tailed S-1 line-intensity distribution
function”, J. Optical Society of America, Vol. 57, pp. 323-329, 1967.
[15] SOUFIANI, A., TAINE, J. “High temperature gas radiative property parameters of statistical narrow-
band model for H2O, CO2 and CO, and correlated-k model for H2O and CO2”, Int. J. Heat and Mass Transfer,
Vol. 40, pp. 987-991, 1997.
[16] YOUNG,S.J. “Nonisothermal band model theory”, J. Quant. Spectroscopy and Radiative Transfer, Vol.
18, pp. 1-28, 1977.
[17] BARLOW, R.S. FRANK, J.H., “Effects of turbulence on species mass fractions in methane/air jet
flames”, Proc. Combustion Institute, Vol. 27, pp. 1087-1095, 1998.
[18] International Workshop on Measurement and Computation of Turbulent Nonpremixed Flames,
http://www.ca.sandia.gov/TNF/DataArch/FlameD.html.
[19] ZHENG, Y., BARLOW, R.S., GORE, J.P., “Spectral radiation properties of partially premixed
turbulent flames”, J. Heat Transfer, Vol. 125, pp. 1065-1073, 2003.
[20] DA SILVA, C.B., MÉTAIS, O. “On the influence of the coherent structures upon interscale interactions
in turbulent plane jets”, J. Fluid Mechanics, Vol. 473, pp. 103-145, 2002.
[21] ZHENG, Y., SIVATHANU, Y.R., GORE, J.P., “Measurements and stochastic time and space series
simulations of spectral radiation in a turbulent non-premixed flame”, Proc. Combustion Institute, Vol. 29, pp.
1957-1963, 2002.
[22] COX, G., “On radiant heat transfer from turbulent flames”, Combustion Science and Technology, Vol.
17, pp. 75-78, 1977.

You might also like