You are on page 1of 7

5th International Seminar on Flame Structure July 11-14, 2005

Institute of Chemical Kinetics & Combustion Novosibirsk, Russia

Thermodynamics of Combustion of TNT Products in a Chamber


Allen L. Kuhl University of California Lawrence Livermore National Laboratory P. O. Box 808 Livermore, California 94551 USA Introit We consider the explosion of a 1.5-g spherical PETN/TNT charge in air in a 6.6-l calorimeter [1]. The hot detonation products are rich in fuel ( C(s) , CO and H 2 ), and when they mix with air, they release 2,500 cal/g (in addition to the 1,100 cal/g released by the detonation) by a non-premixed turbulent combustion process [2]. Here we trace the evolution of combustion in Thermodynamic State Spaceto elucidate the mechanisms (constraints) that control this process. The analysis is formulated in the Le Chatelier diagram of Oppenheim [3], i.e., specific-internal-energy versus temperature plane, shown in Fig. 1.

Figure 1. Locus of states in the Le Chatelier diagram for the combustion of the detonation products gases generated by the explosion of a composite (PETN/TNT) charge in a 6.6-liter calorimeter. Chamber stoichiometry is c = 5.367 . The thermodynamic properties of the components k (Fuel-F, Air-A, Reactants-R, and Products-P) were computed by the Cheetah code [4], and fit with quadratics to establish the Equations of State (EOS) of the components:

uk (Tk ) = ak Tk2 + bk Tk + c k

( k = F, A,R,P )

(1)

These quadratic EOS functions (curves in Fig. 1) well approximate the thermodynamic values (circles in Fig. 1) and fully characterize the thermodynamic behavior of the system [4]. Combining this with a combustion model, we find a combustion locus in thermodynamic space. We show that this locus is a strong attractor to numerical simulations of the combustion field.
Thermodynamics

We divide the process into two steps: (i) mixing: of F and A to form R; (ii) combustion: material transformation from R to P in a thermodynamically-isolated chamber. Mixing is governed by:

Thermodynamics of Combustion of TNT Products in a Chamber Mass: Volume: Energy:


YF + YA = 1 YF v F + YA v A = v c YF uF + YA uA = uc ( c )

5th ISFS (2) (3) (4)

where the specific volume of the chamber is given by v c = Vc /(mF + mA ) , and c = YA /YF = 5.37 represents the chamber air/fuel ratio. Given the initial conditions ( TA = 300K & TF = 3,070K ), the constraint (4) specifies the chamber energy: uc ( c ) = 45 cal /g and reactants temperature TR = 1,032 K . Fuel and air combine in c proportions to form reactant R( c ) . Combustion at constant energy and volume (UV) is expressed by: Mass: Volume: Energy:
YR + YP = 1 YR v R + YP v P = v c uP = uR = uc ( c )

(5) (6) (7)

Constraint (7) provides the combustion locus: uP = uR = uc ( c ) depicted in Fig. 1 as the R-P line.
Gas Dynamics

We model evolution of the combustion field in the limit of large Reynolds Peclet and Damkhler numbers. Then the flow field for the mixture (m) is governed by the multicomponent gas-dynamic conservation laws: Mass: Momentum: Energy: Components:

t m + ( m u) = 0 t m u + ( m uu) = pm t m (um + u u /2) + m (um + u u /2)u = ( pm u)


& tYk + u Yk = k Ys

(8) (9) (10) (11)

with

Y
k

=1

where k = F , A, P and k = {1, s,(1+ s )} denotes the stoichiometric source/sink & coefficients based on the stoichiometric air/fuel ratio: s = 2.77 . The term Ys embodies kinetic effects. Here it is modeled by the fast-chemistry, large-Damkhler-number limit which is consistent with the gas dynamic model (8-11). This system is closed by Equations of State (EOS) for components-k (1), while mixture EOS is based on the ideal mixing laws for perfect gases:
Mixture EOS: Mixture Properties:
2 Tm = [bm + bm 4am (c m um )]/2am

&
k

pm = m Rm Tm
k

(12a,b) (13a,b,c,d)

am = Yk ak , bm = Yk bk , c m = Yk c k , Rm = Yk Rk
k k

The above model equations (8-11) were integrated using a high-order Godunov scheme, and Adaptive Mesh Refinement (AMR) was used to follow the thin reaction zones on the grid. We simulated the explosion and combustion of 1.5-g PETN/TNT charge in 6.6-l calorimeter [5]. Computed pressure histories are compared with experimental data in Fig. 2.

Thermodynamics of Combustion of TNT Products in a Chamber

5th ISFS

Figure 2. Comparison of pressure histories from numerical simulations with experimental data for explosion of 1.5-g PETN/TNT charges in 6.6-liter chamber.
Projection

The numerical solution for the 6.6-l calorimeter was projected from physical space to thermodynamic space via the following volume-averaging operators:
Mass:
Amagats Law: Daltons Law: Specific Energy: Specific Entropy:

M m ( t ) = M k (t )
k

where M k (t) =
where Vk (t) = where pk (t) =

Vc

k (t,x)dV

(14a,b) (15a,b) (16a,b)

Vm (t) = Vk (t)
k

Vc

N k (t,x)dV /N m pk (t,x)dV /Vc

pm (t) = pk (t)
k k

Vc

um (t) = uk (t) where uk (t) = sm (t) = sk (t) where sk (t) =


k k

Vc

k (t,x) uk (t,x) dV / M k (t) (17a,b)

Vc

k (t,x) sk (t,x) dV / M k (t) (18a,b)

Temperature: RT m (t) = Rk Tk (t) where Rk Tk (t) = where Vc denotes the volume of the chamber. Solution

Vc

k (t,x)Rk Tk (t,x) dV / M k (t) (19a,b)

Temporal Space The evolution of the mean thermodynamic variables from the numerical solution (averaged over Vc ) is presented in temporal space in Figs. 3. While component variables in general change over time due to combustion, conserved variables such as the mixture mass, volume and energy remain constantthereby illustrating conservation principles (2)-(7) for the system. In addition, Figs. 3b and 3d illustrate Amagats Law of partial volumes and Daltons Law of partial pressures, respectively, for the mixture.

Thermodynamics of Combustion of TNT Products in a Chamber

5th ISFS

(a) Mass
10 Mixture

(d) Pressure
10x10
6

9 Mixture 8 7

6 Mk (g)

Air Products

6 pk (bars) Products 5 4 Air 3

2 Fuel

2 1 Fuel

2 t (ms)

2 t (ms)

(b) Volume
7000 Mixture 6000

(e) Moles
0.4 Mixture 0.3

5000 Products
Nk 0.2 Products Air

Vk(cm )

4000

3000

Air

2000

0.1

1000

Fuel
0.0

Fuel

2 t (ms)

2 t (ms)

(c) Temperature
3000 2500 Products

(f) Specific Internal Energy


500 Air

Mixture

2000

Fuel
uk (cal/gk)

Mixture Tk (K) 1500 Air

Products

-500

Fuel

1000

500
-1000

2 t (ms)

2 t (ms)

Figure 3. Evolution of the mean thermodynamic variables in the 6.6-l chamber. Thermodynamic Space The mean solution is presented in Thermodynamic Space in Fig. 4. In this depiction, time appears only implicitly as the evolution parameter that links corresponding thermodynamic values, depicted as data points (circles) on those figures. Figure 4a presents the evolution of numerical solution in the u T plane. Computed points for fuel, air and products lie on the F, A and P lines corresponding to the equation of state functions (1). Computed points for the mixture rapidly (by 0.1 ms) approach the combustion locus: uc = 45 Cal / g denoted as

Thermodynamics of Combustion of TNT Products in a Chamber

5th ISFS

the R-P curve. Figure 4b presents the numerical solution in the p v plane. Computed points for the mixture lie along the combustion isochor: v c = 690 cc / g (red curve R-P). Figure 4c presents the numerical solution in the p T plane. Computed points for the mixture lie between the reactants isochor and products isochor. Figure 4d presents the numerical solution in the s T plane. While the reactants and products entropies obey the usual logarithmic dependence on temperature, the mixture entropy exhibits a linear dependence on temperaturein accordance with the linear mixing laws (18)-(19) of perfect gases. (a) u-T plane
400 200 P 0 -200 u (cal/g) -400 -600 -800 -1000 -1200 R
0.1 1ms 2 10

(c) p-T plane


10x10
A
6

P
10ms

2ms

P
1ms

p (bars)

u0,d

R 4
0.1ms

Fuel(t) Air(t) Products(t) Mixture(t) Combustion Locus

Reactants isochor (v=691cc/g) Products isochor (v=691cc/g) Mixture(t) Combustion Locus

500

1000

1500 T (K)

2000

2500

3000

3500

500

1000

1500 T (K)

2000

2500

3000

(b) p-v plane


10x10
6

(d) s-T plane


1.0
P
10ms 2ms Reactants isotherm (T=1032K) Products isotherm (T=2249K) Products(t) Mixture(t) Combustion Locus

9 8 7 6

0.8

P P
10ms 2ms

1ms

0.6 s (cal/g-K)
1ms

p (bars)

5 R 4 3 2 1 0
0.1ms

R 0.4
0.1ms

0.2

Reactants isochor (v=691 cc/g) Products isochor (v=691 cc/g) Mixture(t) Combustion Locus

500

1000

1500 v (cm /g)


3

2000

2500

3000

0.0

500

1000

1500 T (K)

2000

2500

3000

Figure 4. Evolution of the mean solution in thermodynamic state space. Numerical solution for the mixture rapidly approaches the combustion locus R-P

Chamber History
Utilizing the perfect gas mixing law for specific thermodynamic variables, one can construct the following model for the evolution of the thermodynamic solution of combustion in a chamber (denoted by subscript c):

Temperature: Pressure: Entropy:

Rc Tc ( t ) RR TR + YP ( t )[ RP TP RR TR ] pc ( t ) c Rc Tc ( t ) = pR + YP ( t )[ pP pR ] sc ( t ) sR + YP (t )[ sP sR ]

(20) (21) (22)

Thermodynamics of Combustion of TNT Products in a Chamber

5th ISFS

Here the model parameters ( TR = 1,032 K & TP = 2,249 K ; pR = 4.40 bars & pP = 8.98 bars ; sR = 0.25 cal / g K & sP = 0.78 cal / g K ) come from the thermodynamic solution, while the Products function YP ( t ) comes from the gas-dynamic solution. In the present case, the products function is well approximated by the exponential:

Products:

YP ( t ) = 1 e1.2t

where [t] = ms

(23)

This model is depicted in Fig. 5. One can see that the numerical solution (circles) lies close to the model functions (20)-(22), thereby confirming the model of this process. (a) Products function
1.0

(b) Temperature
3000
2500

0.8

2000

0.6
YP(t)
T(K)

1500

0.4
1000

0.2

2D AMR
fit=1-e
-1.2t

Tc(t)

Mixture

500

0.0

2
t (ms)

2
t (ms)

(c) Pressure
10x10
6

(d) Specific Entropy


1.0

9 8 7 6
p (bars)
0.6

0.8

5 4 3 2 1 0 pc(t) Mixture(t)

sk (cal/g-K)

0.4

0.2

sc(t) Mixture(t)

0.0

2 t (ms)

2
t (ms)

Figure 5. Comparison of computed mixture environment histories with the chamber model: (a) mass fraction of products produced, (b) temperature, (c) pressure, (d) entropy. Assessment To assess the validity of the model, we have compared the theoretical model predictions with experimental data; results are shown in Table 1. For the 0.5-g PETN booster charge, the model predicts 2.75 bars versus the measured value of 2.66 bars; their ratio suggests that 97% of the charge detonated. For the composite charge (0.5-g PETN + 1.0-g TNT) in nitrogen, the model predicts 4.4 bars versus the measured value of 3.6 bars; their ratio suggests that 82% of the charge detonated. For the composite charge in air, the model predicts 8.98 bars versus the measured value of 8.35 bars; their ratio suggests that 93% of the charge was consumed by combustion during the course of the experiment.

Thermodynamics of Combustion of TNT Products in a Chamber Table 1. Comparison of theoretical predictions with experimental data. CASE PETN booster charge: experiment (in N 2 ) Thermodynamic Model (in N 2 ) PETN/TNT Composite charge: experiment (in N 2 ) Thermodynamic Model (in N 2 ) experiment (in air) Thermodynamic Model (in air) Numerical simulation (in air @ 10ms) Rsum
pc (bars)

5th ISFS

p (%)
97% 82% 93% 95%

T (K)

2.660.07 2.75 3.600.2 4.40 8.350.28 8.98 8.55

1,032 K 2,249 K 2,156 K

Mixing and combustion in a thermodynamically-isolated chamber is controlled by uc : the initial energy in the system. It establishes not only the initial state (temperature) of the reactants and the final state (temperature) of the products, but also the locus of combustion states in thermodynamic space. Implementing the quadratic EOS model for stoichiometric combustion in the multi-component gas dynamic conservation laws allows one to simulate the evolution of combustion in physical space. Note that combustion in physical space always occurs under locally stoichiometric conditions ( s = 2.77 ), while the combustion locus represents global combustion effects for the chamber (i.e., at the chamber stoichiometry: c = 5.37 ). The most important contribution of this paper is that we show that the gas dynamic solution, when projected from physical space to thermodynamic space, rapidly approaches the combustion locus predicted by the Thermodynamic modelthereby demonstrating that the combustion locus is a strong attractor for this reactive system. This model is especially useful in cases where the kinetics of the system is not known. Acknowledgements This work was performed under the auspices of the U. S. Department of Energy by the University of California, Lawrence Livermore National Laboratory under Contract No. W7405-Eng-48. It was sponsored by the Defense Threat Reduction Agency under IACRO # 05-4071. References [1] Neuwald, P., Reichenbach, H. & Kuhl, A.L. (2003) Shock-Dispersed Fuel Charges Combustion in Chambers and Tunnels, Energetic Materials, 34th ICT Conf., 13.1-14. [2] Kuhl, A. L., Ferguson, R.E & Oppenheim, A.K. (1999) Gasdynamics of Combustion of TNT Products in Air, Archivum Combustionius 19:1-4, 67-89. [3] Oppenheim, A. K. and Maxon, J. A. (1991) Thermodynamics of Combustion in an Enclosure, Progress in Aeronautics and Astronautics, AIAA, New York, pp. 365-382. [4] Kuhl, A.L, Howard, M and Fried, L. (1999) Thermodynamic Model of Afterburning in Explosions Energetic Materials, 34th ICT Conf., 74.1-74.14. [5] Bell, J. B., Day, M., Kuhl, A.L., Beckner, V., & Neuwald, P. (2005) Numerical Simulation of Combustion of TNT/PETN in a Calorimeter, 20th ICDERS.

You might also like