Sei sulla pagina 1di 12

Competition Eects in Combustion Chemistry and Their Role in Detonation Initiation

Shannon Browne California Institute of Technology Pasadena, CA 91125 May 2004

Abstract This study considers the issues of approximate modeling of chemical kinetics in detonation waves and the prediction of dynamic detonation parameters such as initiation energy. Several models have been proposed for the initiation of detonation waves. Due to the complexity of the coupling between uid mechanics and chemistry, these models have relied on simple single-step reaction mechanisms. Recently Short, Quirk, and Kapilla have proposed a three-step chemistry model that includes initiation, branching, and termination components. They propose that initiation and failure can be related to competition between branching and termination and the existence of a crossover temperature. This chemistry model is a mathematical model and proposes to reduce a complicated chemical mechanism to three generic reactions that do not depend on specic chemical properties. A program of research is proposed to examine the validity of their model by comparing detonations based on detailed chemistry with detonations based on their 3-step model using realistic values of the input parameters.

Introduction

Direct initiation refers to any method of detonation initiation that uses a strong shock wave. Commonly, high explosives or electrical discharges provide the necessary energy to create a strong spherical blast wave in the test medium. If the exothermic chemical reactions and the wave remain coupled as the blast wave expands and decays, then a detonation has successfully been initiated. The coupling of the shock wave and exothermic chemical reactions involves both gas dynamics and chemistry. This study initially will focus on the role of the chemical reaction mechanism in determining the critical conditions for direct initiation success. The combustion process in a detonation wave can be conceived of as a convected chain-branching/thermal explosion. The crucial feature of high temperature combustion chemistry that may play a role in this problem is the competition between branching chain reactions and chain termination reactions. The historical viewpoint [10] is that if termination reactions proceed more quickly than branching reactions, an explosion can be quenched. This eect may play a role in causing the detonation initiation process to fail. The critical parameter controlling this behavior can be expressed as a crossover temperature where the chain-branching and chain-termination reaction rates are equal. A simplied view of the situation is that above this temperature, branching dominates, while below, termination may quench the reaction. The situation with realistic combustion chemistry is more subtle and although chain termination may alter the explosion time, peroxide chemistry may ultimately still enable an explosion to occur [16].
Submitted Candidacy

in Partial Fulllment of the Requirements for the Degree of Doctor of Philosophy Committee: Dr. Chris Brennen (chair), Dr. Joseph Shepherd (advisor), and Dr. David Goodwin

The proposed study plans to address three things. Determining whether this crossover temperature is a real chemical and physical phenomenon, establishing how well this three-step model reects nature, and investigating if this model can be used to nd critical conditions for direct initiation in real systems. In this report, rst we present a discussion of previous work on detonation failure in Section 2.1. Then we give two formulations of the ZND model of steady one-dimensional waves in Section 2.2. As the focus of the study is the chemistry aspect of detonations, combustion chemistry is illustrated for the hydrogen-oxygen system in Section 2.3. Once we have established a foundation, we describe more thoroughly the crossover eect in the context of specic fuel-oxidizer systems in Section 3. Section 4 explicitly gives the model that Short, Quirk, and Kapilla [17, 18] propose as well as graphical results. Finally, the concluding Section 5 explains the work proposed to achieve the goals of this study.

2
2.1

Background
Previous Work

During the 1950s, Zeldovich et al. [21] and Manson and Ferrie [12] were the rst to directly initiate a spherical detonation experimentally. In their experiments, they found that there is a minimum energy that must be provided in order to successfully detonate their test mixture. Because the source energy is an independent variable, knowing the value of this critical initiation energy would dictate under what circumstances there would be direct initiation success. Critical initiation energy is termed by Lee as a dynamic detonation parameter [8] and is dependent on the structure of the detonation. Detonation structure is unsteady and multidimensional and for this reason, the most successful models for critical initiation energy are empirically based on extensive experimental data. These models include the Zeldovich et al. criterion [21], the Lee et al. model [9], and the Sichel Model [19]. Litcheld et al. [11] rst showed that the trend of the critical initiation energy as a function of equivalence ratio is a U-shaped curve. The minimum critical initiation occurs at approximately stoichiometric conditions while very rich or very lean mixtures require large amounts of energy to successfully initiate. The main diculty in determining the critical initiation energy is that there is not a robust theoretical model for detonation structure. Currently the interaction between the shock and the reaction zone is not completely understood. Extensive eort has been invested in developing theoretical and numerical models that can accurately simulate detonation structure. The complexity of the coupled chemical reaction and uid mechanics requires that simplications be made in order to analyze the initiation process. This is accomplished by neglecting certain aspects of the uid mechanics and the chemistry. For example, manipulating the energy equation using a single-step reaction model, one can derive the temperature-reaction zone structure equation [3] for one-dimensional ows: DT Dt

cp

= (1 M 2 ) + c2 ek k W k Wk

(1) (2) (3) (4)

j w2 (D w) Rc x dD w 1 P +w w + dt t t

This relation shows that there are three major contributions to the temperature change (1); these are the energy released by chemical reaction (2), the wave curvature (3), and the unsteadiness (4). It has been shown by Eckett, Quirk and Shepherd [3] that the unsteadiness of the leading shock wave plays

Reaction 1. H + O2 OH + O 2. H + O2 + M HO2 + M 3. H2 + M 2H + M 4. OH + H2 H2 O + H 5. O + H2 OH + H 6. O + H2 O 2OH

A 5.13 1016 2.10 1018 2.23 1012 1.17 1009 1.80 1010 6.00 1008

n 0.82 1.00 0.5 1.30 1.00 1.30

Ea 16507 0 92600 3626 8826 0

Table 1: Partial Hydrogen Oxidation Mechanism and Rate Constants a role in causing failure in direct initiation. They proposed a model, the critical decay rate theory, based on the balance between terms 2 and 4. Another theory of direct initiation is based on considering the structure of slightly curved quasisteady detonation waves [6]. In this model, the dominant balance is between chemical heat release and front curvature, terms 2 and 4. They propose that excessive curvature prevents the appearance of a sonic point at the rear of the reaction zone. The absence of a rear sonic plane allows acoustic disturbances to weaken the leading shock wave and cause the initiation to fail. Yet a third explanation, proposed by Short and Quirk [18], is that the chemical energy release, term 2, vanishes causing the shock wave and reaction zone to decouple. This model is discussed in more detail in Section 4.

2.2

ZND Model of Steady One Dimensional Waves

The primary governing equations for any uid mechanics problem are the conservation equations for mass, momentum, energy, and species. Zeldovich [22], von Neumann [20], and Doering [2] proposed that detonation structure can be modeled by considering a leading nonreactive shock followed by a steady ow with chemical reaction. This has come to be called the ZND model of detonation. The ZND model can be expressed in terms of a system of ordinary dierential equations. A key part of this formulation is the thermicity, , which is the sum of two terms: the dierence in the number of moles of products and reactants and the energy absorbed or released from chemical bonds [4, pg. 78]. The dierential equations are less dicult to solve than the mixed set of algebraic expressions and ordinary dierential equations that Zeldovich, von Neumann, and Doering rst proposed. Around the turn of the century, Chapman [1] and Jouguet [7] hypothesized that the steady detonation velocity is the minimum velocity consistent with the jump conditions. The speed is now termed the Chapman-Jouguet (CJ) detonation velocity, DCJ . The standard model of a propagating detonation used by most analysts is a ZND structure moving at the CJ speed in a uniform mixture.

2.3

Detailed Reaction Mechanism

In a combustion system, collisions between dierent species cause reactions and the breaking and reforming of bonds, ultimately converting reactants to products. This is described by a reaction mechanism containing a set of reactions and reaction rates. Table 11 shows a selection of reactions and rates for a hydrogen-oxygen combustion mechanism [16] An explosive chemical reaction is due to both thermal and chain reactions. There are four primary reaction types in a branching chain: chain initiation, chain branching, chain propagation, and chain termination. A chain initiation reaction is a collision resulting in a free radical. In the hydrogen mechanism (table 1), reaction 3 is the initiation reaction. Here the hydrogen molecule reacts with another molecule to produce two hydrogen radicals. Once sucient free radicals exist, chain branching and chain propagating reactions continue the chain. Chain propagating reactions produce one radical per input radical while chain branching reactions
1 Units

are moles, cubic centimeters, seconds, Kelvin, and calories/mole

produce more than one radical per input radical. Reaction 4 of the hydrogen mechanism is chain propagating while reactions 1, 5, and 6 are chain branching. In each of these branching reactions, two radicals are produced for each radical that reacts. Chain termination reactions occur when one or more radicals react to form a stable species. Reaction 2 of the hydrogen mechanism is a termination reaction. Unlike reaction 1, the extra reactant molecule removes enough energy to H and O2 to allow them to bond and form HO2 . The rate of reaction is proportional to the product of reactant concentrations. The proportionality constant, the reaction rate coecient, has a modied Arrhenius form. The molecularity of the reaction is also a crucial factor in determining the reaction rate. Molecularity is the number of molecules that interact in the reaction step [5, pg. 37]. In the hydrogen mechanism, all of the reactions except reaction 2 are bimolecular. Reaction 2, the termination reaction, is trimolecular. The rate of termination reactions, which are trimolecular, increase more rapidly with pressure than bimolecular reactions.

3
3.1

Crossover Eect
Crossover Region

There exists a region of the temperature-pressure plane where the rates of reactions 1 and 2 (Table 1) are equal. Above this region, reaction 1, the branching reaction, proceeds more quickly than reaction 2, the termination reaction. To determine this region in the specic case of hydrogen combustion, consider the two reaction rates. R1 = A1 T n1 exp Ea1 RT [H][O2 ] R2 = A2 T n2 exp P RT Ea2 Ea1 RT (5) Ea2 RT [H][O2 ][M]

[M] = A1 A2

R2 = R1 P =

R T 1+n1 n2 exp

In gure 1, the broken lines depict R1 = R2 for several values of . The corresponding expression for the pressure as a function of temperature for these lines is P = 0.048 T 1.18 exp 16507 2T .

Figure 1 also shows a collection of gaseous hydrogen-air detonation post-shock states. Each point on the plot corresponds to a dierent overdrive value. The CJ detonation is noted. This plot indicates that the crossover eect may play a role in detonation initiation. For each case, the initial pressure was one atmosphere and the initial temperature was 300 K. The post-shock states plotted in gure 1 were calculated with the Stanjan program [14]. This program works with a thermodynamic data le to calculate equilibrium states for species mixtures. It allows the user to specify the initial state including initial pressure, temperature, species mass fractions, and species phases. It then provides twenty-two dierent options for calculating an equilibrium state. The two important options for this study calculate the post-shock state and the Chapman-Jouguet detonation velocity. After choosing an option, Stanjan calculates the pressure, specic volume, specic energy, specic enthalpy, specic entropy, and species mass fractions of the equilibrium state.

3.2

Induction Zone

Detonations have two major components: the shock wave and the reaction zone. The reaction zone usually consists of an induction zone that is almost thermally neutral followed by an exothermic recombination zone [15]. 4

2400

Temperature (K)

2200 2000 1800 1600 1400 1200 1000 10

Branching Regime H + O2 OH + O
CJ Point =0.5 =1 =2
Termination Regime H + O2 + M HO2 + M

20

Pressure (atm)

30

40

50

60

Figure 1: Stoichiometric hydrogen-air detonation post-shock states with respect to the chemical crossover region Figure 2 illustrates these elements of the detonation structure. In this gure, the induction length coincides with the location of maximum thermicity, .

Induction 4.0E+06 Zone 2800 Thermicity (1/s) 3.0E+06 2.0E+06 1.0E+06 -6.0E+01 0 2600 2400 2200 2000 1800 T2 0.025 0.05 0.075 Distance (cm) 1600 1400 0.1 Temperature (K)

Figure 2: Induction length based on the maximum thermicity. It also shows the spatial temperature prole. Initial State was 300 K, 1 atm, and stoichiometric hydrogen-air. The proles in gure 2 were calculated with the ZND program [16]. This program works with a chemical mechanism, the Chemkin library, an ordinary dierential equation solver, and a root nder to spatially march from the post-shock state to the sonic plane where M = 1. It allows the user to specify 5

the initial state including initial pressure, temperature, species mass fractions, and species phases as well as the shock speed and tolerances for the ordinary dierential equations solver. It returns spatial proles for the mach number, thermodynamic properties, velocity, thermicity, species mass fractions, and average molecular weight. Figure 2 was created with the GRI Mechanism version 3.0 [13].

3.3

Crossover Temperature

The overall activation energy of a combustion system quanties the relative ease with which the overall reaction will proceed. If the activation energy of the overall process decreases, the reaction proceeds more easily, but if the activation energy increases, the reaction will have more resistance and proceed more slowly. An Arrhenius plot graphically shows the activation energy. Induction length, , can be expressed as a function of temperature through the following simple model. = tind w2 k = AT n exp Ea RT

d = k(1 ) (6) dt In this model, equals 0 in the unburned reactants and 1 in burned products. It is apparent in gure 2 that the induction zone occurs at approximately constant temperature equal to the post-shock temperature. Integrating equation 6 for a nite time with this in mind gives ln() = Ea 1 + ln (A w2 ) R T2

Now a linear relationship between the logarithm of the induction length and the inverse of the post-shock temperature is clear, and the slope of this line is the eective activation energy, Ea /R. Dierent shock velocities will give dierent post-shock states. If there is a change in the slope of the Arrhenius plot, ln() vs. 1/T2 , then the systems eective activation energy has changed. This is an indication of a crossover eect from one chemical regime to another. Figures 3 and 4 are two Arrhenius plots. Figure 3 shows a stoichiometric mixture of hydrogen and air while Figure 4 shows a stoichiometric mixture of methane and air. It is apparent that hydrogen exhibits a change of slope while methane does not. This change of slope for hydrogen-air is associated with a post-shock temperature of approximately 1530 K. These gures were created using calculated by the ZND program [16]. DCJ and T2 were given by the Stanjan program [14]. Again the calculations relied on the GRI Mechanism version 3.0 [13] and a thermodynamics data le that includes all of the pertinent species.

Short, Quirk, & Kapilla Method

Short, Quirk, and Kapilla [17, 18] propose that traditional one-step chemistry models are not adequate for detonation simulation. They propose an alternative reaction model that consists sequentially of a chain-initiation step, a chain-branching step, and chain-termination step. The Short, Quirk, and Kapilla model2 is: 1
I

F Y F + Y 2Y YP
2 nondimensional

kI = exp kB = exp kC = 1

1
B

1 1 TI T 1 1 TB T

(7) (8) (9)

quantities are indicated with an *

10 10 10 10 10

ln() ln(1/s)

-1

CJ Detonation

-2

0.4

0.5

0.6 0.7 0.8 1000/T2 (1/K)

0.9

Figure 3: Arrhenius plot for stoichiometric hydrogen-air with varying overdrive values.

10 ln() ln(1/s) 10 10 10 10

CJ Detonation

-1

-2

0.4

0.5

0.6 0.7 0.8 1000/T2 (1/K)

0.9

Figure 4: Arrhenius plot for stoichiometric methane-air with varying overdrive values. The Short, Quirk, and Kapilla [17, 18] method of analyzing their three-step chemistry model uses dimensionless versions of the ZND equations with respect to the post-shock state. The three reaction rates they use are:
R = f kI I R = f ykB B

R = y C

With only three species, fuel, F, Y, and P, the product mass fraction can be expressed as a function of the other mass fractions, i.e. p = 1 f y. Consequently, there are only two species equations in this 7

system. (R + R ) df I B = dx w dy R + R R I B C = dx w

After applying the algebraic constraints, Short, Quirk and Kapilla also solve the following non-dimensional algebraic equations. P = where a = ( + 1) b = 2( + 1)
2 2 c = 2M2 ( 1)q + ( 1 2M2 )

b2 4ac 2a

(P 1) 1 =1

T =

w =

M2

q = Q(1 f ) Qy Besides the spatially varying thermodynamic properties, this system depends on certain constant values. , M2 , TI , TB , I , B , Q, DCJ , and d, are also indirectly important to the system. Short, Quirk, and Kapilla specify Q rather than specifying M2 , but M2 can be determined from Q using the following system of equations. M2 = ( 1)D2 + 2 2D2 ( 1) D= dDCJ DCJ = H +1+ H

H=

1 ( 2 1) ho 1 2 P1

o 1 QP2 1 = P1 P1 2 1 D2 +1

P2 2(D 1) =1+ P1 +1

1 =1 2

2 1

The resulting detonation proles for species mass fraction, temperature, pressure, and density are shown in gures 58. The mixed system was solved by a Matlab script employing a forward Euler numerical method with the following values for the constants [17, 18]: Q=3
I

= 1/20

= 1/8 TI = 3 TB = 0.82 = 1.2 d = 1.2

Once the published gures could be reproduced, the thermicity equations were non-dimensionalized. The proles were constructed again with the dimensionless thermicity equations. dP w = dx d = dx w dw = dx

We have looked at both the initial state3 and the post-shock state as the reference state.
3Q

= 8.33

= 1/37.5

= 1/10

TI = 3

P2 1 P1 2

TB = 0.82

P2 1 P1 2

= 1.2

d=1

1 mass fraction 0.8 0.6 0.4 0.2 0 0 2 4 position 6 8 Fuel Radicals

Figure 5: Mass fraction proles based on the three step model.

1.5 temperature 1.4 1.3 1.2 1.1 1 0.9 0 2 4 position 6 8

Figure 6: Temperature prole based on the three step model. The reference state is the post-shock state. Results of solving the mixed algebraic-dierential formulation of the ZND model.

Proposed Work

The goal is a comprehensive model of direct initiation of detonation that can quantitatively predict the critical initiation energy. To do this we must construct simple but robust models for the chemical kinetics and include these in models of detonation structure. This study proposes to examine how the chemistry, both realistic and modeled, aects the shock wave/reaction zone interaction. The chemical regime of the post-shock state is one factor in determining if the reaction will proceed or decouple from the shock wave. The post-shock temperature and pressure will allow us to quantify which regime the combustion system is in. We propose to study the existence and usefulness of this crossover temperature in real chemical and physical situations. We plan to carry out simulations using the three-step model with parameters appropriate to realistic chemical mixtures. We then plan to compare these results with experimental data to determine how well the model emulates nature. This will allow us to comment on the possible applications of this reduced 9

1.1 pressure 1 0.9 0.8 0.7 0 2 4 position 6 8

Figure 7: Pressure prole based on the three step model. The reference state is the post-shock state. Results of solving the mixed algebraic-dierential formulation of the ZND model.

1.1 1 density 0.9 0.8 0.7 0.6 0.5 0 2 4 position 6 8

Figure 8: Density prole based on the three step model. The reference state is the post-shock state. Results of solving the mixed algebraic-dierential formulation of the ZND model. chemistry mechanism. Initially, the three-step method needs to be related back to dimensional physical parameters so dimensional proles can be created. In addition, Arrhenius plots using this chemistry model need to be created to see if the change in slope is similar to those seen in Figures 3 and 4. One of the most challenging aspects of this model is the generic nature of the species. The numerical results from the ZND and Stanjan programs depend on specic properties of the involved species properties not included in the model. These properties include molecular weights, a variable gas constant, and a variable ratio of specic heats. In addition, to accommodate this generic three-step chemistry model, we must analyze simulation results to see if there exist common threads between dierent combustion systems. An extensive study must then include dierent fuels, oxidizers, stoichiometric conditions, initial shock speeds, and initial 10

conditions. Thus far hydrogen and methane mixed with both air and oxygen have been studied for equivalence ratios ranging from 0.5 to 3.0 and overdrive values from 0.8 to 1.4. The hydrogen systems have also been studied at an equivalence ratio of 0.26 which corresponds to 10% hydrogen, the lowest detonable fuel percentage. The initial conditions have yet to be varied. All of the simulations discussed start from a pressure of 1 bar and a temperature of 300 K. The numerical simulation programs ZND and Stanjan will generate reaction zone properties which can then be compared against the three-step chemistry model results. In the long term, we propose to compare unsteady simulations of initiation using the three step model with experimental results and simulations using other approximate reaction mechanisms. Nomenclature a, b, c A A [A] c cp d D DCJ e Ea f F h ho H j k kB kC kI M M n p P P q Q R Rc constants pre-exponential factor modied A concentration of species A frozen sound speed specic heat capacity overdrive factor detonation wave velocity Chapman-Jouguet detonation wave velocity specic internal energy activation energy fuel mass fraction fuel specic enthalpy heat of reaction dimensionless heat of reaction based on initial state geometry constant reaction rate coecient branching reaction rate coecient termination reaction rate coecient initiation reaction rate coecient mach number generic reactant molecule empirically determined power product mass fraction pressure product dimensionless chemical energy dimensionless heat of reaction based on post-shock state gas constant position of the shock in the xed reference frame R RB RC RI t tind T TB TI v w Wk W x y yk Y
I B

k k ( )1 ( )2 ( )k

reaction rate branching reaction rate termination reaction rate initiation reaction rate time induction time temperature branching crossover temperature initiation crossover temperature specic volume uid velocity in shock-still frame molecular weight of species k average molecular weight of mixture distance radical mass fraction mass fraction of species k Radical proportionality constant induction length induction length based on maximum thermicity initiation reaction activation energy branching reaction activation energy ratio of specic heats rate of progress variable sonic parameter molar rate of production of species k mass rate of production of species k density thermicity initial state post-shock state species k

References
[1] D. L. Chapman. On the rate of explosion in gases. Philos. Mag., 14:10911094, 1899. 3 [2] W. Doering. On detonation processes in gases. Annals of Physics, 43:421436, 1943. 3 11

[3] C. A. Eckett, J. J. Quirk, and J. E. Shepherd. The role of unsteadiness in direct initiation of gaseous detonations. Journal of Fluid Mechanics, 421:147183, 2000. 2 [4] W. Fickett and W. Davis. Detonation Theory and Experiment. Dover Publications, INC., 1979. 3 [5] I. Glassman. Combustion. Academic Press, 3 edition, 1996. 4 [6] L. He and P. Clavin. On the direct intiation of gaseous detonations by and energy source. Journal of Fluid Mechanics, 277:227248, 1994. 3 [7] E. Jouguet. On the propogation of chemical reactions in gases. J. de Mathematiques Pures et Appliquees, 1:347425, 1905. continued in 2:5-85 (1906). 3 [8] J. H. S. Lee. Dynamic parameters of gaseous detonations. Annual Review of Fluid Mechanics, 16:311336, 1984. 2 [9] J. H. S. Lee, R. Knystautas, and C. Guirao. Proceedings of the First Sepcialist Meeting on Fuel-Air Explosions, chapter SM Study No. 16, pages 157187. University of Waterloo Press, Waterloo, Canada, 1982. 2 [10] B. Lewis and G. von Elbe. Combustion, Flames and Explosions of Gases. Academic Press, second edition, 1961. 1 [11] E. L. Litcheld, M. H. Hay, and D. R. Forshey. Direct electrical initiation of freely expanding gaseous detonation waves. Proceedings of the Ninth Symposium (International) on Combustion, pages 282286, 1963. 2 [12] N. Manson and F. Ferrie. Contribution to the study of spherical detonation waves. Proceedings of the Fourth Symposium (International) on Combustion, pages 486494, 1953. 2 [13] University of California Berkeley. http://www.me.berkeley.edu/gri mech. GRI-Mechanism 3.0. 6 [14] W. Reynolds. The element potential method for chemical equilibrium analysis: Implementation in the interactive program stanjan. Technical report, Mechanical Engineering Department, Stanford University, 1986. 4, 6 [15] J. E. Shepherd. http://www.galcit.cal-tech.edu/EDL/projects/JetA/Glossary.html. Explosion Dynamics Laboratory Glossary. 4 [16] J. E. Shepherd. Chemical kinetics of hydrogen-air-diluent detonations. Progress in Astronautics and Aeronautics series, 106:263293, 1986. 1, 3, 5, 6 [17] M. Short, A. K. Kapila, and J. J. Quirk. The chemical-gas dynamic mechanism of pulsating detonation wave instability. Philosophical Transactions of the Royal Society of London, 357:3621 3637, 1999. 2, 6, 7, 8 [18] M. Short and J. J. Quirk. On the nonlinear stability and detonability limit of a detonation wave for a model three-step chain-branching reaction. Journal of Fluid Mechanics, 339:89119, 1997. 2, 3, 6, 7, 8 [19] M. Sichel. A simple analysis of the blast initiation of detonation. Acta Astronaut, 4:409424, 1977. 2 [20] J. von Neumann. John von Neumann, collected works, volume 6, chapter Theory of detonation waves. Macmillian, New York, 1942. 3 [21] Y. B. Zeldovich, S. M. Kogarko, and M. N. Simonov. An experimental investigation of spherical detonation gases. Sov. Phys. Tech. Phys., 1(8):16891713, 1956. 2 [22] Ya. B. Zeldovich. On the theory of the propagation of detonations in gaseous systems. Zh. Eksp. Teor. Fiz., 10:542568, 1940. (English Translation: NACA TM 1261, 1960). 3 12

Potrebbero piacerti anche