21.11.2014 Views

Untitled - Aerobib - Universidad Politécnica de Madrid

Untitled - Aerobib - Universidad Politécnica de Madrid

Untitled - Aerobib - Universidad Politécnica de Madrid

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Aerothermochemistry<br />

Gregorio Millán Barbany<br />

Catedrático <strong>de</strong> Mecánica <strong>de</strong> Fluidos y Aerodinámica<br />

<strong>de</strong> la Escuela <strong>de</strong> Ingenieros Aeronáuticos<br />

Miembro <strong>de</strong> la Real Aca<strong>de</strong>mia <strong>de</strong> Ciencias<br />

Reedición conmemorativa <strong>de</strong>l 50 ◦ aniversario <strong>de</strong> la edición original<br />

Escuela Técnica Superior <strong>de</strong> Ingenieros Aeronáuticos<br />

<strong>Universidad</strong> <strong>Politécnica</strong> <strong>de</strong> <strong>Madrid</strong><br />

Asociación <strong>de</strong> Ingenieros Aeronáuticos <strong>de</strong> España<br />

<strong>Madrid</strong>, 2009


Aerothermochemistry, Gregorio Millán Barbany<br />

Edición conmemorativa <strong>de</strong>l 50 o aniversario <strong>de</strong> la edición original <strong>de</strong> 1958 publicada<br />

por el INTA, abril 2009<br />

Editores científicos: Manuel Rodríguez Fernán<strong>de</strong>z, Carlos Vázquez Espí<br />

Diseño <strong>de</strong> la cubierta: Javier Leonés Ranz<br />

c○ 2009 Escuela Técnica Superior <strong>de</strong> Ingenieros Aeronáuticos<br />

Plaza <strong>de</strong>l Car<strong>de</strong>nal Cisneros 3, 28040 <strong>Madrid</strong>, España<br />

ISBN 13:978-84-86402-08-2<br />

D.L. GU-114-2009


PRESENTACIÓN<br />

Amable Liñán<br />

Es un gran placer para mí presentar esta re-edición literal <strong>de</strong> la monografía<br />

Aerothermochemistry <strong>de</strong> Gregorio Millán (1919-2004). La re-edición por la Escuela<br />

<strong>de</strong> Ingenieros Aeronáuticos se produce cuando acaban <strong>de</strong> cumplirse 50 años <strong>de</strong> la<br />

publicación por el INTA, en ciclostilo, <strong>de</strong> los 800 ejemplares <strong>de</strong> la edición original.<br />

Cuando Gregorio Millán cursaba sus estudios <strong>de</strong> Ingeniería Aeronáutica, que<br />

inició en 1941, se estaba produciendo un cambio revolucionario en esta ingeniería. Se<br />

acaba <strong>de</strong> iniciar el <strong>de</strong>sarrrollo <strong>de</strong> los aerorreactores, sin los cuales era impensable que<br />

los aviones pudiesen alcanzar velocida<strong>de</strong>s transónicas o supersónicas; simultáneamente,<br />

para la viabilidad <strong>de</strong> los cohetes <strong>de</strong> son<strong>de</strong>o o <strong>de</strong> los misiles balísticos, se impulsó el<br />

<strong>de</strong>sarrollo <strong>de</strong> los motores cohete. La Mecánica <strong>de</strong> Fluidos, disciplina central <strong>de</strong> las<br />

Ciencias Aeronáuticas y <strong>de</strong>terminante <strong>de</strong>l diseño <strong>de</strong> estos motores, fue elegida por<br />

Gregorio Millán como objeto <strong>de</strong> su actividad docente e investigadora.<br />

En la formación <strong>de</strong> Gregorio Millán había tenido un papel crucial el inusual<br />

ambiente docente <strong>de</strong> la Aca<strong>de</strong>mia Militar <strong>de</strong> Ingenieros Aeronáuticos, que se reflejaba<br />

en la preocupación <strong>de</strong>l profesorado por el papel <strong>de</strong> las ciencias básicas y aplicadas en<br />

el <strong>de</strong>sarrollo <strong>de</strong> la Ingeniería Aeronáutica. Este ambiente docente era here<strong>de</strong>ro <strong>de</strong>l que<br />

ya se daba en la antecesora <strong>de</strong> la Aca<strong>de</strong>mia, la Escuela Superior Aerotécnica, <strong>de</strong>s<strong>de</strong><br />

su creación en 1928, bajo la dirección <strong>de</strong> Emilio Herrera. Para las enseñanzas básicas,<br />

Herrera había conseguido la colaboración, que se mantuvo en la Aca<strong>de</strong>mia Militar, <strong>de</strong><br />

los profesores universitarios españoles más eminentes (muchos <strong>de</strong> ellos, igual que el<br />

propio Emilio Herrera, miembros <strong>de</strong> la Aca<strong>de</strong>mia <strong>de</strong> Ciencias).<br />

Uno <strong>de</strong> estos profesores era Esteban Terradas, que fue Presi<strong>de</strong>nte <strong>de</strong>l Patronato<br />

<strong>de</strong>l INTA <strong>de</strong>s<strong>de</strong> su creación. Terradas se propuso impulsar el <strong>de</strong>sarrollo en España<br />

<strong>de</strong> las Ciencias Aeronáuticas, invitando a los científicos extranjeros mas prestigiosos<br />

en estas ciencias a impartir ciclos <strong>de</strong> conferencias. Entre ellos Teodoro von Kármán<br />

que en 1948 vino a España, por primera vez, para hablar <strong>de</strong> Aerodinámica Transónica<br />

y Supersónica y sobre Turbulencia. De entonces nació la colaboración fructífera <strong>de</strong><br />

Gregorio Millán con von Kármán, quien orientó la actividad docente e investigadora<br />

posterior <strong>de</strong> Millán y también las investigaciones <strong>de</strong>l Grupo Español <strong>de</strong> Combustión.<br />

Von Kármán, que se consi<strong>de</strong>ra con justicia el padre <strong>de</strong> las Ciencias Aeronáuticas<br />

americanas, había sido el Director <strong>de</strong> los Guggenheim Aeronautical Laboratories<br />

<strong>de</strong>l Instituto Tecnológico <strong>de</strong> California (Caltech). Poco antes <strong>de</strong> la última Guerra<br />

Mundial inició su preocupación por el <strong>de</strong>sarrollo <strong>de</strong> los cohetes, y <strong>de</strong>spués por los


aerorreactores, siendo el creador <strong>de</strong>l Jet Propulsion Laboratory. Comprendiendo que<br />

el análisis <strong>de</strong> los procesos <strong>de</strong> combustión era esencial para el diseño <strong>de</strong> estos motores y<br />

que <strong>de</strong>bía hacerse con el apoyo <strong>de</strong> la Dinámica <strong>de</strong> Fluidos, se embarcó en el proyecto<br />

<strong>de</strong> establecer el marco multidiciplinar apropiado. Para ello, consiguió la colaboración<br />

<strong>de</strong>l Profesor Saul Pennner <strong>de</strong>l Caltech y <strong>de</strong> Gregorio Millán, Ingeniero <strong>de</strong>l INTA y<br />

Profesor <strong>de</strong> la Aca<strong>de</strong>mia (luego Escuela) <strong>de</strong> Ingenieros Aeronáuticos.<br />

La Aerothermochemistry se basa en el ciclo <strong>de</strong> conferencias que Teodoro von<br />

Kármán impartió en la Sorbona durante el curso 1951-1952, en cuya preparación y<br />

<strong>de</strong>sarrollo contó con la ayuda <strong>de</strong> Gregorio Millán. Por el interés suscitado por estas<br />

conferencias <strong>de</strong> von Kármán, el Air Research and Development Command (ARDC)<br />

<strong>de</strong> las Fuerzas Aéreas <strong>de</strong> Estados Unidos ofreció a Gregorio Millán, en 1954, un contrato<br />

para apoyar la redacción y actualización, mediante un programa <strong>de</strong> investigación,<br />

<strong>de</strong> las conferencias <strong>de</strong> la Sorbona. Para este proyecto, Gregorio Millán contó con la<br />

colaboración <strong>de</strong> un grupo <strong>de</strong> ingenieros y profesores <strong>de</strong> la Escuela <strong>de</strong> Ingenieros Aeronáuticos,<br />

que formaron el Grupo Español <strong>de</strong> Combustión. Este grupo incluía a Segismundo<br />

Sanz Aránguez, Jesús Salas Larrazábal, Carlos Sánchez Tarifa, José Manuel<br />

Sendagorta e Ignacio Da Riva. Al grupo se sumaron pronto los profesores Francisco<br />

García Moreno y Pedro Pérez <strong>de</strong>l Notario, y yo mismo que empecé como becario en<br />

1958. Más tar<strong>de</strong>, creció notablemente el número <strong>de</strong> participantes (entre ellos, Enrique<br />

Fraga, Antonio Crespo, José Luis Urrutia y Juan Ramón Sanmartín) en los proyectos<br />

<strong>de</strong> investigación <strong>de</strong>l Grupo. También se ampliaron las fuentes <strong>de</strong> subvención, que incluyeron<br />

el US Forest Service <strong>de</strong>l Departamento <strong>de</strong> Agricultura americano, así como<br />

la European Space Research Organization (ESRO).<br />

El ARDC facilitó la difusión internacional, a través <strong>de</strong> laboratorios universitarios<br />

y centros <strong>de</strong> investigación, <strong>de</strong> los 800 ejemplares <strong>de</strong> la edición original <strong>de</strong> la<br />

Aerothermochemistry. A pesar <strong>de</strong> ello, hace ya bastante tiempo que no se encuentran<br />

disponibles ni accesibles copias <strong>de</strong>l original. El valor histórico y la actualidad <strong>de</strong> la Aerotermoquímica<br />

<strong>de</strong> Millán nos ha animado a hacer el esfuerzo <strong>de</strong> transcribir el original<br />

en TEX y rehacer las figuras para una re-edición. Esta tarea no hubiese sido posible sin<br />

la <strong>de</strong>cisión espontánea <strong>de</strong> Manuel Rodríguez Fernán<strong>de</strong>z <strong>de</strong> iniciar esa transcripción,<br />

llegando a completar casi la tercera parte. Este impulso inicial animó a las autorida<strong>de</strong>s<br />

académicas <strong>de</strong> la Escuela a finalizar la tarea, encargando a Carlos Vázquez Espí la<br />

coordinación y supervisión <strong>de</strong> la transcripción y <strong>de</strong> las correcciones, labor en la que<br />

colaboraron Eva Villacieros y los estudiantes <strong>de</strong> la ETSI Aeronáuticos Alfredo Giralda,<br />

Ramón Lacruz y David Marchante. En la re-edición se han eliminado erratas <strong>de</strong>l<br />

original, se han rehecho la mayoría <strong>de</strong> las figuras con los métodos actuales <strong>de</strong> cálculo,


que no cambiaron los resultados, y se han incluido algunas anotaciones significativas.<br />

Por todo el apoyo recibido, tenemos que agra<strong>de</strong>cer a la <strong>Universidad</strong> <strong>Politécnica</strong> <strong>de</strong><br />

<strong>Madrid</strong> y a la Asociación <strong>de</strong> Ingenieros Aeronáuticos la publicación en forma <strong>de</strong> libro<br />

<strong>de</strong> la monografía.<br />

En la Aerothermochemistry aparece por primera vez el marco multidisciplinar<br />

coherente, que es necesario para el análisis <strong>de</strong> los procesos <strong>de</strong> combustión. Este<br />

análisis, como había anticipado von Kármán, no pue<strong>de</strong> hacerse sin ampliar las leyes<br />

<strong>de</strong> la Mecánica <strong>de</strong> Fluidos con las <strong>de</strong> la Teoría <strong>de</strong> los Fenómenos <strong>de</strong> Transporte, la<br />

Termodinámica <strong>de</strong> Mezclas y la Cinética <strong>de</strong> las Reacciones Químicas. 1 Aunque la<br />

monografía <strong>de</strong> Millán contribuyó <strong>de</strong>cisivamente a establecer la necesidad <strong>de</strong>l tratamiento<br />

multidisciplinar en la investigación <strong>de</strong> los Procesos <strong>de</strong> Combustión y <strong>de</strong> la Aerodinámica<br />

Hipersónica, el nombre Aerothermochemistry fue <strong>de</strong>splazado finalmente<br />

por el <strong>de</strong> Ciencias <strong>de</strong> la Combustión.<br />

La importancia <strong>de</strong> la Aerothermochemistry para la Escuela Española <strong>de</strong> Mecánica<br />

<strong>de</strong> Fluidos ha sido trascen<strong>de</strong>ntal, porque en ella se abordan problemas <strong>de</strong> frontera <strong>de</strong><br />

la Mecánica <strong>de</strong> Fluidos con un lenguaje mo<strong>de</strong>rno y unas técnicas avanzadas. Ha sido<br />

ejemplar, tanto para la enseñaza <strong>de</strong> la Mecánica <strong>de</strong> Fluidos como para la investigación<br />

y el uso <strong>de</strong> la dinámica <strong>de</strong> los fluidos en ámbitos multidisciplinares.<br />

Por el cariño con el que von Kármán acogió y valoró las aportaciones <strong>de</strong>l Grupo<br />

Español <strong>de</strong> Combustión, von Kármán eligió <strong>Madrid</strong> como se<strong>de</strong> <strong>de</strong>l Primer Congreso<br />

Internacional <strong>de</strong> Ciencias Aeronáuticas, que se celebró también en 1958. A este<br />

Congreso acudieron las personalida<strong>de</strong>s más relevantes <strong>de</strong> las Ciencias Aeronáuticas;<br />

entre ellos, Geoffrey Taylor, Leónidas Sedov y James Lighthill. Lighthill, por ejemplo,<br />

habló <strong>de</strong> los efectos en flujos hipersónicos <strong>de</strong> las reacciones químicas <strong>de</strong> disociación<br />

<strong>de</strong> las moléculas <strong>de</strong>l aire. Von Kármán fue también esencial en la organización en<br />

1960 <strong>de</strong> un Curso sobre Ciencia y Tecnología <strong>de</strong>l Espacio, en el INTA, que precedió a<br />

la creación <strong>de</strong> la CONIE y a nuestra participación en la ESRO. Por las aportaciones<br />

<strong>de</strong> von Kármán a las Ciencias Aeronáuticas, poco antes <strong>de</strong> su muerte en Aachen en<br />

1963, recibió <strong>de</strong> las manos <strong>de</strong>l Presi<strong>de</strong>nte Kennedy la primera Medalla Nacional <strong>de</strong><br />

Ciencias americana y su imagen fue recogida en uno <strong>de</strong> sus sellos.<br />

Gregorio Millán agra<strong>de</strong>ce en el prólogo la valiosa colaboración <strong>de</strong> Carlos Sánchez<br />

Tarifa, José Manuel Sendagorta e Ignacio Da Riva a la Aerothermochemistry.<br />

1 Conviene advertir al lector <strong>de</strong> la Aerothermochemistry que, por ejemplo, fue en las conferencias <strong>de</strong> la<br />

Sorbona don<strong>de</strong> apareció por primera vez la forma general <strong>de</strong> la ecuación <strong>de</strong> la energía para la dinámica<br />

<strong>de</strong> fluidos reactantes. En ella se incorporan explícitamente los intercambios <strong>de</strong> energía química, térmica y<br />

cinética, el flujo <strong>de</strong> calor y el trabajo <strong>de</strong> las fuerzas <strong>de</strong> presión y <strong>de</strong> viscosidad; todos ellos involucrados en<br />

el funcionamiento <strong>de</strong> los motores térmicos.


Los tres fueron <strong>de</strong>terminantes en el <strong>de</strong>sarrollo <strong>de</strong> la Escuela Española <strong>de</strong> Mecánica <strong>de</strong><br />

Fluidos. José Manuel Sendagorta (responsable <strong>de</strong> <strong>de</strong>spertar en mí la vocación por la<br />

investigación) sustituyó a Gregorio Millán en la enseñanza <strong>de</strong> la Mecánica <strong>de</strong> Fluidos,<br />

que tuvo que <strong>de</strong>jar cuando la fundación y el <strong>de</strong>sarrollo <strong>de</strong> Sener le absorbieron todo<br />

su tiempo; pero consiguió convertir pronto a Sener en una <strong>de</strong> las gran<strong>de</strong>s empresas<br />

españolas <strong>de</strong> Ingeniería. Primero Sánchez Tarifa y más tar<strong>de</strong> Millán, cuando <strong>de</strong>jó la<br />

dirección <strong>de</strong> Babcock & Wilcox, fueron llamados por Sendagorta para contribuir al<br />

<strong>de</strong>sarrollo <strong>de</strong> Sener. Sánchez Tarifa fue el responsable <strong>de</strong> la enseñanza <strong>de</strong> Propulsión<br />

en nuestra Escuela, y director, en el INI, <strong>de</strong>l proyecto y ensayo <strong>de</strong>l primer motor <strong>de</strong><br />

reacción español; <strong>de</strong>s<strong>de</strong> Sener dirigió nuestra participación en el diseño <strong>de</strong>l motor <strong>de</strong>l<br />

Eurofighter. La labor investigadora experimental en Combustión <strong>de</strong> Sánchez Tarifa fue<br />

seminal, inspirando la labor <strong>de</strong> algunos <strong>de</strong> los investigadores más distinguidos proce<strong>de</strong>ntes<br />

<strong>de</strong> nuestra Escuela. Ignacio Da Riva, pronto catedrático <strong>de</strong> Aerodinámica, sólo<br />

<strong>de</strong>jó su fructífera <strong>de</strong>dicación, incansable y ejemplar, a la enseñanza e investigación<br />

cuando la muerte le sobrevino en clase. Fue mi maestro y compañero en la investigación,<br />

y puntal esencial en la Escuela Española <strong>de</strong> Mecánica <strong>de</strong> Fluidos.<br />

Las contribuciones científicas <strong>de</strong> Millán, que están recogidas en parte en este<br />

libro, fueron muy importantes y contribuyeron a su elección en 1961 como Miembro<br />

<strong>de</strong> la Real Aca<strong>de</strong>mia <strong>de</strong> Ciencias. Su influencia en la enseñanza es difícil <strong>de</strong> imaginar<br />

para las nuevas generaciones, pero es inolvidable para mí; que aprendí Mecánica <strong>de</strong><br />

Fluidos <strong>de</strong> él y <strong>de</strong> José Manuel Sendagorta, y adopté las notas para clase que nos<br />

<strong>de</strong>jaron. Yo aprendí en la Aerothermochemistry la base teórica <strong>de</strong> la Combustión; y fui<br />

iniciado en la investigación sobre las llamas <strong>de</strong> difusión por Gregorio Millán en las<br />

reuniones semanales que los componentes <strong>de</strong>l Grupo <strong>de</strong> Combustión teníamos con él<br />

en su etapa (1957-1961) <strong>de</strong> Director General <strong>de</strong> Enseñanzas Técnicas. El lector pue<strong>de</strong><br />

encontrar en la reseña necrológica 2 que escribí para la Revista <strong>de</strong> la Real Aca<strong>de</strong>mia<br />

<strong>de</strong> Ciencias, un breve resumen <strong>de</strong> las aportaciones <strong>de</strong> Millán a la política educativa y<br />

también <strong>de</strong> su labor como gestor a nuestro <strong>de</strong>sarrollo tecnológico.<br />

Espero que esta re-edición <strong>de</strong> la Aerothermochemistry sirva <strong>de</strong> mo<strong>de</strong>sto homenaje<br />

a los méritos <strong>de</strong> las aportaciones <strong>de</strong> Gregorio Millán y sus colaboradores a las<br />

Ciencias Aeronáuticas Españolas.<br />

2 In Memoriam: D. Gregorio Millán Barbany, Rev. R. Acad. Cienc. Exact. Fis. Nat., Vol 99, No. 1, pp.<br />

183-185, 2005.


INSTITUTO NACIONAL DE TÉCNICA AERONÁUTICA<br />

ESTEBAN TERRADAS<br />

AEROTHERMOCHEMISTRY<br />

(A report based on the course conducted by Prof. Theodore von Kármán<br />

at the University of Paris)<br />

BY<br />

GREGORIO MILLÁN<br />

PROFESSOR OF AERODYNAMICS<br />

AT<br />

ESCUELA TÉCNICA SUPERIOR DE INGENIEROS AERONÁUTICOS<br />

MADRID (SPAIN)<br />

January, 1958<br />

The research reported in this document has been sponsored by the AIR<br />

RESEARCH AND DEVELOPMENT COMMAND, UNITED STATES AIR<br />

FORCE, un<strong>de</strong>r contract No. 61 (514)-441, through the European Office, ARDC.


PREFACE<br />

by<br />

Theodore von Kármán<br />

I have great pleasure in introducing the report of Prof. Gregorio Millán on<br />

“Aerothermochemistry”. This word refers to problems whose solution necessitates<br />

the application of fundamental principles of Thermodynamics and Chemistry especially<br />

chemical kinetics. In other words they are flow problems with exchange of heat<br />

and production of heat by means of chemical reaction. The phenomena of high speed<br />

flight ma<strong>de</strong> it necessary for the aeronautical engineer to be fully familiar in addition to<br />

Fluid Mechanics, i.e. Aerodynamics proper, with the fundamentals and applications<br />

of Thermodynamics. This need created the science of Aerothermodynamics. The <strong>de</strong>velopment<br />

of jet propulsion introduced problems which have direct bearing in mo<strong>de</strong>rn<br />

aircraft and missile <strong>de</strong>sign and necessitate the un<strong>de</strong>rstanding of changes in the composition<br />

of the flowing medium, as dissociation and chemical reactions. Thus Aerothermochemistry<br />

<strong>de</strong>als with the interaction between chemical changes and the merely<br />

aerodynamic phenomena as pressure and velocity distribution, momentum transfer<br />

and alike and between the phenomena of Aerothermodynamics, as shockwaves, heat<br />

transfer, diffusion etc. This novel extension in the scope of the Aeronautical Sciences<br />

necessitates a thorough study of branches of Science, which the aeronautical engineer<br />

in general gave little attention during his professional education. This report is written<br />

for aeronautical engineers, in that the author does not assume that the rea<strong>de</strong>r is<br />

acquainted with the kinetic theory of gases and with chemistry beyond quite general,<br />

i<strong>de</strong>as.<br />

A few chapters of the report are based on lectures which I gave as Fullbright<br />

guest professor at the Sorbonne in Paris in the scholastic year 1951/52. This course<br />

had the same general aim to acquaint the aeronautical engineer including myself with<br />

the problems introduced by the phenomena of combustion occurring in flowing media.<br />

The author however completed, and I believe very successfully, the presentation of the<br />

subject by including a systematic treatment of the Thermodynamics of gas mixtures,<br />

of the theory of chemical equilibrium, the elements of chemical kinetics, theory of<br />

transfer phenomena in gases and gas mixtures. In addition the problems of flame stabilization,<br />

combustion of liquid droplets and diffusion flames, similarity which were<br />

only touched on in my lectures are treated in this report systematically.<br />

I am sure that the report will be helpful to many aeronautics engineers in their


aim to become acquainted with Aerothermochemistry and it may also contribute some<br />

suggestions to those engaged in pursing the subject in their further researches.<br />

For this reason the report inclu<strong>de</strong>s rather exten<strong>de</strong>d bibliographic material after<br />

each chapter.<br />

I would like to mention without commitment that if I can secure the valuable<br />

collaboration of Prof. Millán, I am playing with the i<strong>de</strong>a to publish a book on the<br />

fundamentals of Aerothermochemistry.


FOREWORD<br />

During the aca<strong>de</strong>mic term 1951–52, Professor Theodore von Kármán conducted<br />

a course on Aerodynamics of Combustion at the University of Paris. This<br />

course was mainly inten<strong>de</strong>d for aeronautical engineers with the objective of encouraging<br />

the study of the problems risen by the new propulsion systems <strong>de</strong>veloped by<br />

Aeronautics, which necessitates the assistance of the theories and methods of Aerodynamics,<br />

Thermodynamics and Chemistry of Combustion. The new Science born<br />

to <strong>de</strong>al with these problems nee<strong>de</strong>d a name and Prof. von Kármán suggested it be<br />

called AEROTHERMOCHEMISTRY for its analogy with other classical science of<br />

Aeronautics, such as Aeroelasticity or Aerothermodynamics. In a few years this <strong>de</strong>signation<br />

has been wi<strong>de</strong>ly accepted.<br />

The author was invited by Prof. von Kármán to participate as his assistant in<br />

<strong>de</strong>veloping this course. Due to the amount of bibliography that has to be reviewed in<br />

addition to the numerous studies and exercises required in its preparation it was not<br />

possible at the time to write the lectures.<br />

Later, when the European Office of the ARDC was activated it Brussels and<br />

in view of the general <strong>de</strong>mand for the Lectures by Prof. von Kármán, the said Office<br />

offered the author the opportunity of preparing a Report on Aerothermochemistry<br />

based upon the course conducted by Prof. von Kármán, through a research contract<br />

sponsored by ARDC.<br />

Work was started in 1953 but for different reasons it had to be interrumped<br />

in several occasions. At the same time investigation was progressing very rapidly in<br />

the different fields of Aerothermochemistry and it was necessary to incorporate them<br />

into this report, in or<strong>de</strong>r to keep it current. For this reason, initial planning had to be<br />

continuously exten<strong>de</strong>d and finally carried far beyond its original intention.<br />

The author wishes to express his <strong>de</strong>ep appreciation to Prof. von Kármán who<br />

was a constant source of inspiration and encouragement and whose advise in preparing<br />

this report proved invaluable. At the same time he trusts that Prof. von Kármán will<br />

excuse this report if it fails to reflect the excellencies of the course on which it is based.<br />

Special acknowledgment is given to the contribution of the European Office<br />

ARDC, and the Instituto Nacional <strong>de</strong> Técnica Aeronáutica Esteban Terradas for the<br />

help received in preparing this report for publication.<br />

Finally, I wish to acknowledge the valuable collaboration of Carlos Sánchez


Tarifa, Manuel <strong>de</strong> Sendagorta and Ignacio Da Riva, aeronautical engineers of the<br />

INTA, in the research work that ma<strong>de</strong> possible the publication of this report, and to<br />

Mrs. <strong>de</strong> los Casares for her technical translation.<br />

<strong>Madrid</strong>, January 1958


Contents<br />

1 Thermochemistry 1<br />

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Thermodynamic functions of an i<strong>de</strong>al gas. . . . . . . . . . . . . . . 6<br />

1.3 Mixture of gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.4 Calculation of the thermodynamic functions . . . . . . . . . . . . . 16<br />

1.5 Chemical reactions in a mixture of gases . . . . . . . . . . . . . . . 21<br />

1.6 Chemical equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

1.7 Case of a mixture of i<strong>de</strong>al gases . . . . . . . . . . . . . . . . . . . . 24<br />

1.8 Chemical kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2 Transport phenomena in gas mixtures 37<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.2 Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.3 Viscosities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

2.4 Thermal conductivity . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3 General equations 59<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.2 Equation of continuity . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.3 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

v


3.4 Energy equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

3.5 General equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

3.6 Entropy variation . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

3.7 One-dimensional motions . . . . . . . . . . . . . . . . . . . . . . . 72<br />

3.8 Stationary, one-dimensional motions . . . . . . . . . . . . . . . . . 75<br />

3.9 The case of only two chemical species . . . . . . . . . . . . . . . . 77<br />

3.10 Stationary, one-dimensional motion of i<strong>de</strong>al gases with heat addition 79<br />

3.11 Appendix: Notation and repertoire of vectorial and tensorial formulae 84<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

4 Combustion Waves 89<br />

4.1 Detonation and <strong>de</strong>flagration . . . . . . . . . . . . . . . . . . . . . . 89<br />

4.2 Kinds of <strong>de</strong>tonations and <strong>de</strong>flagrations . . . . . . . . . . . . . . . . 91<br />

4.3 Velocity of the burnt gases . . . . . . . . . . . . . . . . . . . . . . . 94<br />

4.4 Propagation velocity . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

4.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

4.6 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

4.7 In<strong>de</strong>terminacy of the solution . . . . . . . . . . . . . . . . . . . . . 103<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

5 Structure of the combustion waves 105<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

5.2 Wave equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

5.3 Characteristic times . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

5.4 Limiting form of the wave equations . . . . . . . . . . . . . . . . . 111<br />

5.5 Detonations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

5.6 Deflagrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

5.7 Transition from <strong>de</strong>flagration to <strong>de</strong>tonation . . . . . . . . . . . . . . 126<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128


6 Laminar flames 131<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

6.2 Equation for the combustion wave (two chemical species) . . . . . . 134<br />

6.3 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 135<br />

6.4 Modification of the conditions at the “cold boundary” . . . . . . . . 137<br />

6.5 Propagation velocity of the flame . . . . . . . . . . . . . . . . . . . 140<br />

6.6 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140<br />

6.7 Reaction velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . 145<br />

6.8 Flame equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147<br />

6.9 Solution of the flame equations . . . . . . . . . . . . . . . . . . . . 148<br />

6.10 Structure of the combustion wave . . . . . . . . . . . . . . . . . . . 159<br />

6.11 Ignition temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

6.12 General equations for the combustion wave . . . . . . . . . . . . . . 162<br />

6.13 Ozone <strong>de</strong>composition flame . . . . . . . . . . . . . . . . . . . . . . 169<br />

6.14 Hydrazine <strong>de</strong>composition flame . . . . . . . . . . . . . . . . . . . . 176<br />

6.15 Flame propagation in Hydrogen-Bromine mixtures . . . . . . . . . . 187<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203<br />

7 Turbulent flames 207<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207<br />

7.2 Turbulent combustion theories . . . . . . . . . . . . . . . . . . . . . 210<br />

7.3 Turbulence generated by the flame . . . . . . . . . . . . . . . . . . 218<br />

7.4 Comparison with experimental results . . . . . . . . . . . . . . . . 219<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219<br />

8 Ignition, flammability and quenching 221<br />

8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221<br />

8.2 Ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222<br />

8.3 Flammability limits . . . . . . . . . . . . . . . . . . . . . . . . . . 225


8.4 Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227<br />

9 Flows with combustion waves 231<br />

9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231<br />

9.2 Conditions that must be satisfied by the jump across a flame front. . . 231<br />

9.3 Normal flame front . . . . . . . . . . . . . . . . . . . . . . . . . . 235<br />

9.4 Inclined flame front . . . . . . . . . . . . . . . . . . . . . . . . . . 236<br />

9.5 Entropy jump across the flame front . . . . . . . . . . . . . . . . . . 239<br />

9.6 Vorticity across the flame . . . . . . . . . . . . . . . . . . . . . . . 242<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244<br />

10 Aerothermodynamic field of a stabilized flame 245<br />

10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245<br />

10.2 Tsien method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250<br />

10.3 Method of Fabri-Siestrunck-Fouré . . . . . . . . . . . . . . . . . . 255<br />

10.4 Cylindrical chambers . . . . . . . . . . . . . . . . . . . . . . . . . 258<br />

10.5 Chamber with slowly varying cross-section . . . . . . . . . . . . . . 259<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260<br />

11 Similarity in combustion. Applications 261<br />

11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261<br />

11.2 Dimensionless parameters of Aerothermochemistry . . . . . . . . . 262<br />

11.3 Scaling of rockets . . . . . . . . . . . . . . . . . . . . . . . . . . . 266<br />

11.4 Scaling of rockets for non-steady conditions . . . . . . . . . . . . . 270<br />

11.5 Flame stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . 274<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279<br />

12 Diffusion flame 281<br />

12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281<br />

12.2 General equations for laminar diffusion flames . . . . . . . . . . . . 289


12.3 Boundary conditions on the flame . . . . . . . . . . . . . . . . . . . 291<br />

12.4 Simplified equations . . . . . . . . . . . . . . . . . . . . . . . . . . 292<br />

12.5 Solutions of the simplified system . . . . . . . . . . . . . . . . . . . 296<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298<br />

13 Combustion of liquid fuels 301<br />

13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301<br />

13.2 Atomization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302<br />

13.3 Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304<br />

13.4 Combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305<br />

13.5 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307<br />

13.6 Continuity equations . . . . . . . . . . . . . . . . . . . . . . . . . . 308<br />

13.7 Energy equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310<br />

13.8 Diffusion equations . . . . . . . . . . . . . . . . . . . . . . . . . . 313<br />

13.9 Combustion velocity of the droplet . . . . . . . . . . . . . . . . . . 314<br />

13.10 Integration of the equations . . . . . . . . . . . . . . . . . . . . . . 316<br />

13.11 Numerical application . . . . . . . . . . . . . . . . . . . . . . . . . 319<br />

13.12 Comparison with experimental results and limitations of the theory . 322<br />

13.13 Influence of convection . . . . . . . . . . . . . . . . . . . . . . . . 324<br />

13.14 Combustion of fuel sprays . . . . . . . . . . . . . . . . . . . . . . . 325<br />

13.15 Droplet evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . 325<br />

13.16 Appendix: Application of Probert’s method . . . . . . . . . . . . . . 331<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333


Chapter 1<br />

Thermochemistry<br />

1.1 Introduction<br />

Aerothermochemistry <strong>de</strong>als with mixtures of reactant gases of variable composition.<br />

Such variation in composition is due to the reactions between the various chemical<br />

species forming the mixture and to their mutual diffusion. In general, the pressure<br />

of these mixtures is not larger than several atmospheres 1 and their temperatures<br />

range from ambient temperature to 3 000 or 4 000 K. Un<strong>de</strong>r such conditions the mean<br />

distance between molecules is large respect to their size, and their potential energy<br />

of interaction is negligible compared to the kinetic energy of the molecular motion.<br />

Thereby, the mixture can generally be treated as an i<strong>de</strong>al gas. In those cases where it<br />

is necessary to take into consi<strong>de</strong>ration the <strong>de</strong>viations in the behavior of the gas with<br />

respect to that of an i<strong>de</strong>al gas, this can be attained by including correcting terms in the<br />

thermodynamic equations of mixture. For example, its state equation can be approximated<br />

by the virial equation proposed by Kamerlingh Onnes<br />

p<br />

ρR g T = 1 + ρB (T ) + ρ2 C (T ) + . . . (1.1)<br />

where p, ρ and T are, respectively, the pressure, <strong>de</strong>nsity and absolute temperature of<br />

the gas. R g is its particular constant and B (T ), C (T ) , . . . are the second, third, etc.<br />

coefficients of the virial, respectively. These can be calculated from the potential ϕ<br />

of interaction of the molecules. For example, if both molecules belong to the same<br />

species and if the potential <strong>de</strong>pends only on the distance r between their centers of<br />

gravity but not on their relative orientation, the second coefficient, which generally is<br />

1 Some ballistic problems are exclu<strong>de</strong>d (See Ref. [1]) as well as <strong>de</strong>tonations of con<strong>de</strong>nsed explosives<br />

(See Ref. [2]) where pressures can be very large.<br />

1


2 CHAPTER 1. THERMOCHEMISTRY<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

ϕ/ε<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0<br />

r/σ<br />

Figure 1.1: Dimensionless Interaction Potential of Lennard-Jones, ϕ/ε, versus dimensionless<br />

distance between molecules, r/σ.<br />

the only nee<strong>de</strong>d except for very high pressures, is expressed in the form 2<br />

B (T ) = 2π N M<br />

∫ ∞<br />

0<br />

(<br />

1 − exp(− ϕ(r)<br />

kT ) )<br />

r 2 dr, (1.2)<br />

where N = 6.0288 × 10 23 mol −1 is the Avogadro number, M is the molar mass<br />

of the gas and k = 1.38047 × 10 −16<br />

erg/grad is the Boltzmann constant. It has<br />

experimentally been checked that the interaction potential of Lennard-Jones<br />

( (σ ) 12 ( σ<br />

) ) 6<br />

ϕ (r) = 4ε − , (1.3)<br />

r r<br />

represents, suitably, the actual behavior of many gases. In Eq. (1.3), σ is the radius<br />

of the molecules and ε is a constant which has energy dimensions. Fig. 1.1 represents<br />

Eq. (1.3). There, it is seen that if the distance r between the molecules is larger than<br />

several times the radius of the molecule, the molecules attract each other whilst if<br />

r < 1.12σ they reject with a force that increases very rapidly as the molecules get<br />

closer. By substituting (1.3) into (1.2) one verifies that B (T ) can be expressed in the<br />

form<br />

where<br />

2 See Ref. [3].<br />

B (T ) = b 0<br />

M B∗ (T ∗ ) , (1.4)<br />

b 0 = 2 3 πNσ3 (1.5)


1.1. INTRODUCTION 3<br />

GAS ε/k (K) σ (Ȧ) b 0 (cm 3 /mol)<br />

Air 99.2 3.522 55.11<br />

N 2 95.05 3.698 63.78<br />

O 2 117.5 3.580 57.75<br />

CO 100.2 3.763 67.22<br />

CO 2 189 4.468 113.90<br />

NO 131 3.170 40.00<br />

N 2 O 189 4.590 122.00<br />

CH 4 148.2 3.817 70.16<br />

CH-CH 185 4.221 94.86<br />

CH 2=CH 2 199.2 4.523 116.70<br />

C 2H 6 243 3.954 78.00<br />

C 3H 8 242 5.367 226.00<br />

n-C 4 H 10 297 4.971 155.00<br />

i-C 4 H 10 313 5.341 192.18<br />

n-C 5 H 12 345 5.769 242.19<br />

n-C 6 H 14 413 5.909 260.25<br />

n-C 7 H 16 282 8.880 884.00<br />

n-C 8 H 18 320 7.451 521.79<br />

n-C 9H 20 240 8.448 760.53<br />

Cyclohexane 324 6.093 285.33<br />

C 6H 6 440 5.270 84.62<br />

CH 3 OH 507 3.585 58.12<br />

C 2 H 5 OH 391 4.455 111.53<br />

CH 3 Cl 855 4.455 48.49<br />

CH 2 Cl 2 406 4.759 135.96<br />

CHCl 3 327 5.430 201.95<br />

C 2 N 2 339 4.380 105.99<br />

COS 335 4.130 88.86<br />

CS 2 488 4.438 110.26<br />

Table 1.1: Force constants for the Lennard-Jones potential.<br />

and<br />

T ∗ = kT ε . (1.6)<br />

Table 1.1, taken from Ref. [3], gives the values of b 0 and ε/k for several gases.<br />

These values have been <strong>de</strong>termined from experimentation. Fig. 1.2 gives the values<br />

for B ∗ (T ∗ ) taken from the same reference. 3 To get an i<strong>de</strong>a on the or<strong>de</strong>r of magnitu<strong>de</strong><br />

of coefficients B (T ) and C (T ), and, consequently, on the <strong>de</strong>viations that are to be<br />

expected in the actual behavior of gases respect to their i<strong>de</strong>al mo<strong>de</strong>l, Figs. 1.3 and<br />

1.4 give the values of these coefficients for some gases. As can be observed B (T ) is<br />

first negative and then becomes positive, tending slowly to zero for T → ∞. Such<br />

behavior of B (T ) is common for all gases. It is seen, for example, that in nitrogen<br />

for p = 25 kg/cm 2 and T = 800 K, B (T ) = 1 cm 3 /gr. Therefore, the error ma<strong>de</strong><br />

3 Ref. [3] also inclu<strong>de</strong>s a table for the calculation of C (T ).


4 CHAPTER 1. THERMOCHEMISTRY<br />

with the assumption that in such a state nitrogen behaves as an i<strong>de</strong>al gas is one per<br />

cent. Therefore un<strong>de</strong>r the conditions of the present study it is justified to consi<strong>de</strong>r the<br />

gases as i<strong>de</strong>al. Hereinafter, unless the contrary is ma<strong>de</strong> explicit, such an assumption<br />

is adopted.<br />

0.5<br />

0.0<br />

T * = 100ξ<br />

T * = 10ξ<br />

−0.5<br />

T * = ξ<br />

B * (T * )<br />

−1.0<br />

−1.5<br />

−2.0<br />

−2.5<br />

0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0<br />

ξ<br />

Figure 1.2: B ∗ as a function of T ∗ . Note the two changes of scale.<br />

8<br />

H 2<br />

C 7<br />

H 16<br />

4<br />

O 2<br />

N 2<br />

B(T) (cm 3 /gr)<br />

0<br />

−4<br />

CH 4<br />

CO 2<br />

−8<br />

−12<br />

0 400 800 1200 1600 2000 2400 2800 3200 3600<br />

T(K)<br />

Figure 1.3: Virial coefficient B (cm 3 /gr) , as function of temperature for several gases.


1.1. INTRODUCTION 5<br />

14<br />

12<br />

10<br />

N 2<br />

CH 4<br />

O 2<br />

CO 2<br />

CO<br />

C(T) (cm 6 /gr 2 )<br />

8<br />

6<br />

4<br />

2<br />

H 2<br />

(0.1 C(T))<br />

C 7<br />

H 16<br />

(0.1 C(T))<br />

0<br />

0 400 800 1200 1600<br />

T(K)<br />

Figure 1.4: Virial coefficient C (cm 3 /gr) 2 as a function of temperature for several gases.<br />

In the following paragraphs we will first briefly summarize the thermodynamic<br />

functions of i<strong>de</strong>al gases and of their mixtures. A more extensive study of the subject<br />

can be found, for example in the works mentioned in Refs. [4] and [5], where, the thermodynamic<br />

functions for the case of air gas are also inclu<strong>de</strong>d. A complete repertoire<br />

of thermodynamic functions of the gases and their mixtures including the influence of<br />

some virial coefficients, can be found in Ref. [6]. Further on, the form in which these<br />

thermodynamic functions can be obtained is briefly consi<strong>de</strong>red. Finally, reference is<br />

ma<strong>de</strong> to the thermodynamic equilibrium and to the chemical kinetics of gas reactions.<br />

Throughout the following paragraphs the mixture will be assumed to be in thermodynamic<br />

equilibrium, 4 at rest and that its state and composition are homogeneous. The<br />

effects of motion and of lack of homogeneity will be studied in chapter 2.<br />

It is generally advantageous in Aerothermochemistry to refer the thermodynamic<br />

functions to the unit mass, which is done in the following instead of referring<br />

then to a mole, as usually done in Thermodynamics. The thermodynamic functions<br />

(internal energy, enthalpy, entropy and free energy) when referred to the unit mass<br />

will be <strong>de</strong>noted with small letters. If referred to a mole the same letters in capital will<br />

be used.<br />

4 Obviously except in the paragraph <strong>de</strong>voted to Chemical Kinetics.


6 CHAPTER 1. THERMOCHEMISTRY<br />

1.2 Thermodynamic functions of an i<strong>de</strong>al gas.<br />

The equation of state<br />

The state equation of an i<strong>de</strong>al gas is given by the expression<br />

p<br />

ρ = R gT, (1.7)<br />

where<br />

R g = R M g<br />

(1.8)<br />

is the particular constant of the gas, where R = 8.3144 joule/mol/grad = 1.9872 cal/mol/grad<br />

is the universal constant of the gases.<br />

Internal Energy<br />

The internal energy of the gas is<br />

∫ T<br />

u = u 0 + c v (T ) dT. (1.9)<br />

T 0<br />

Here u 0 is the internal energy of the gas at temperature T 0 , c v is the specific heat at<br />

constant volume, which <strong>de</strong>pends on temperature as will be analyzed in §4, and T 0 is a<br />

reference temperature. In particular, if c v is constant from T 0 to T , we have<br />

u = u 0 + c v (T − T 0 ) , (1.10)<br />

otherwise, in many applications c v is substituted by a mean value ¯c v . In such case<br />

(1.9) takes the approximate form<br />

u = u 0 + ¯c v (T − T 0 ) . (1.11)<br />

Enthalpy<br />

The gas enthalpy h is <strong>de</strong>fined by<br />

h = u + p ρ . (1.12)<br />

Making use of the state equation (1.8) and of equation (1.9), one obtains for (1.12)<br />

h = h 0 +<br />

∫ T<br />

T 0<br />

c p (T ) dT, (1.13)


1.2. THERMODYNAMIC FUNCTIONS OF AN IDEAL GAS. 7<br />

where h 0 is the enthalpy of the gas at the temperature T 0 , and c p = c v +R g its specific<br />

heat at constant pressure. As before, if c p is constant or is substituted by a mean value<br />

¯c p , we have<br />

h = h 0 + ¯c p (T − T 0 ) . (1.14)<br />

Since only differences in energy can be measured a convention becomes necessary<br />

to <strong>de</strong>termine u 0 and h 0 . This convention is as follows:<br />

1) A standard state is adopted for each chemical substance, <strong>de</strong>fined by a value p 0<br />

of the pressure and a value T 0 of the temperature.<br />

2) The formation enthalpies of the chemical elements in their stable phase in the<br />

standard state are zero.<br />

3) If any of these substances is a gas its formation enthalpy for T = T 0 and p → 0<br />

were equal to that of the actual gas.<br />

For mo<strong>de</strong>rate pressures, this enthalpy differs slightly from the enthalpy of the actual<br />

gas at temperature T 0 and at pressure p 0 . At present the values generally adopted for<br />

p 0 and T 0 are those proposed by G. M. Lewis [7], namely T 0 = 298.16 K, that is<br />

25 ◦ C, and p 0 = 1 atm. Formerly values of temperature slightly smaller were used.<br />

The reduction from one state to another can be easily performed. The previous convention<br />

fixes the values of the internal energy and of the enthalpy of any substance as<br />

per the first principle of Thermodynamics. The value of h 0 can be obtained, for example,<br />

by measuring the heat of reaction at pressure p 0 and at temperature T 0 , when the<br />

substance forms from its elements or from other substances for which the formation<br />

enthalpies are known. Therefore, in the preceding formulae u 0 and h 0 are the internal<br />

energy and the formation enthalpy of gas respectively. Between both the following<br />

relation exists<br />

h 0 = u 0 + p 0<br />

ρ 0<br />

. (1.15)<br />

Tables 1.2, 1.3 and 1.4 give the formation enthalpies of some substances and<br />

their stable phase in the state of reference. Such values have been taken from Ref. [4],<br />

pp. 98 and following, were additional information can be found. Tables NBS-NACA<br />

of thermal properties of the gases [8] are particularly interesting.<br />

Entropy<br />

The entropy s of the gas is<br />

( ) ∫ p T<br />

c p (T )<br />

s = s 0 − R g ln + dT, (1.16)<br />

p 0 T 0<br />

T


8 CHAPTER 1. THERMOCHEMISTRY<br />

Inorganic Substances<br />

Substance State h 0 s 0 ∆G 0<br />

H 2 gas 0.00 5.482 0<br />

H gas 51 675.60 27.176 48 575<br />

Br 2 liquid 0.00 0.228 0<br />

Br gas 3 342.25 0.543 19 690<br />

HBr gas -107.01 0.586 -12 720<br />

O 2 gas 0.00 1.531 0<br />

O gas 3 697.44 2.404 54 994<br />

O 3 gas 708.33 1.183 39 060<br />

H 2O gas -3 208.15 2.504 -54 635<br />

H 2 O liquid -3 792.02 0.928 -56 690<br />

N 2 gas 0.00 1.634 0<br />

N gas 6108.37 2.614 81 476<br />

NH 3 gas -648.19 2.701 -3 976<br />

NO gas 719.81 1.678 20 719<br />

NO 2 gas 175.86 1.249 12 390<br />

N 2 O 4 gas 25.09 0.790 23 491<br />

N 2O gas 442.79 1.195 24 760<br />

HNO 3 liquid -657.04 0.591 -19 100<br />

C(graphite) solid 0.113 0<br />

C(diamond) solid 37.72 0.049 685<br />

CO gas -943.09 1.689 -32 808<br />

CO 2 gas -2 137.06 1.16 -94 259<br />

Table 1.2: Formation enthalpies (cal/gr),<br />

(cal/mol). Inorganic substances.<br />

entropies (cal gr −1 K −1 ) and free energies<br />

where s 0 is the entropy of the gas at pressure p 0 and at temperature T 0 . The third law of<br />

Thermodynamics assigns zero values to the entropies of all substances at the absolute<br />

zero of temperature. This <strong>de</strong>termines the values of s 0 without the need of a convention<br />

as for the enthalpy case. 5<br />

various substances.<br />

Tables 1.2-1.4 also inclu<strong>de</strong> the formation entropies of the<br />

Let s 0 be the entropy of the gas at temperature T and standard pressure p 0 that<br />

is<br />

∫ T<br />

s 0 c p (T )<br />

= s 0 + dT. (1.17)<br />

T 0<br />

T<br />

From this equation and Eq. (1.16) one obtains for s<br />

( ) p<br />

s = s 0 − R g ln , (1.18)<br />

p 0<br />

where s 0 <strong>de</strong>pends only on T , once p 0 is selected.<br />

5 See, i.e., Ref. [9] for complementary information and for discussion of the anomalies that appear in the<br />

application of the Third Law.


1.2. THERMODYNAMIC FUNCTIONS OF AN IDEAL GAS. 9<br />

Alcohols<br />

Substance Formula State h 0 s 0 ∆G 0<br />

Methanol CH 3OH g -1501.15 1.773 38 700<br />

Methanol CH 3 OH l -1780.04 0.946 -39 750<br />

Ethanol C 2H 5OH g -1220.80 1.463 -40 300<br />

Ethanol C 2 H 5 OH l -1440.39 0.834 -41 770<br />

1-Propanol C 3H 7OH l -1212.43 0.767 -40 900<br />

2-Propanol C 3 H 7 OH l -1279.00 0.716 -44 000<br />

1-Butanol C 4H 9OH l -1074.07 0.735 -40 400<br />

2-methyl-2-Propanol (CH 3 ) 3 COH l -1206.29 0.611 -47 500<br />

1-Pentanol C 5H 11OH l -976.33 0.691 -39 100<br />

2-methyl-2-Butanol C 2 H 5 COH(CH 3 ) 2 l -1094.32 0.622 -47 700<br />

Diphenyl Carbinol (C 6H 5) 2CHOH s -110.62 0.311 30 700<br />

Triphenyl Carbinol (C 6 H 5 ) 3 COH s 15.21 0.302 69 700<br />

Cydohexanol C 6H 11OH l -85.47 0.476 -34 300<br />

Ethylene Glycol CH 2 OHCH 2 OH l -1749.37 0.643 -77 120<br />

Glycerol CH 2OHCHOHCH 2OH l -1728.23 0.540 -113 600<br />

Phenol C 6 H 5 OH s -407.72 0.362 -11 000<br />

Thiophenol C 6 H 5 SH s 0.477<br />

Pyrocatechol C 6 H 4 (OH) 2 s -795.31 0.326 -51 400<br />

Resorcinol C 6 H 4 (OH) 2 s -795.31 0.321 -53 200<br />

Hydroquinone C 6 H 4 (OH) 2 s -795.31 0.304 -52 700<br />

Aniline C 6 H 5 NH 2 l 78.82 0.492 35 400<br />

Table 1.3: Formation enthalpies (cal/gr),<br />

(cal/mol). Alcohols.<br />

entropies (cal gr −1 K −1 ) and free energies<br />

Free Energy<br />

Gibbs free energy<br />

which can be written in the form<br />

g = h − T s, (1.19)<br />

g = g 0 + R g T ln<br />

( p<br />

p 0<br />

)<br />

, (1.20)<br />

where<br />

g 0 = h − T s 0 (1.21)<br />

is the free energy at standard pressure.<br />

gas<br />

The free energy G = Mg per mole is also called chemical potential µ of the<br />

( ) p<br />

µ = G = G 0 + RT ln , (1.22)<br />

p 0


10 CHAPTER 1. THERMOCHEMISTRY<br />

where<br />

G 0 = H − T S 0 = µ 0 , (1.23)<br />

and H, S 0 and µ 0 are, respectively, the molar enthalpy and entropy of the gas and its<br />

chemical potential at standard pressure.<br />

Hydrocarbons<br />

Substance Formula State h 0 s 0 ∆G 0<br />

Methane CH 4 g -1 115.14 2.774 -12 140<br />

Ethane C 2 H 6 g -673.01 1.824 -7 860<br />

Propane C 3 H 8 g -562.89 1.463 -56 14<br />

n-Butane C 4 H 10 g -512.94 1.275 -3 754<br />

n-Pentane C 5H 12 g -485.13 1.154 -1 960<br />

Ethylene C 2 H 4 g 445.49 1.87 1 6282<br />

Propilene C 3 H 6 g 115.95 1.516 14 990<br />

1-Butene C 4H 8 g 4.99 1.31 17 217<br />

cis-2-Butene - g -24.28 1.282 16 046<br />

trans-2-Butene - g -42.87 1.263 15 315<br />

Isobutene - g -59.59 1.251 14 582<br />

1-Pentene C 5H 10 g -71.30 1.185 18 787<br />

Acetylene C 2 H 2 g 2081.5 1.843 50 000<br />

Methylacetylene C 3 H 4 g 1106.26 1.48 46 313<br />

Cyclopentane C 5H 10 l -360.76 0.696 8 700<br />

Methylcyclopentane C 6 H 12 l -392.96 0.704 7 530<br />

Cyclohexane C 6 H 12 l -443.7 0.58 6 370<br />

Methylcyclohexane C 7H 14 l -462.92 0.604 4 860<br />

Benzene C 6 H 6 g 253.75 0.824 30 989<br />

Benzene C 6 H 6 l 150.02 0.529 29 756<br />

Toluene C 7H 8 g 129.7 0.829 29 228<br />

Toluene C 7H 8 l 31.12 0.57 27 282<br />

o-Xylene C 8 H 10 g 42.77 0.794 29 177<br />

o-Xylene C 8 H 10 l -55.02 0.555 26 370<br />

m-Xylene - g 38.81 0.805 28 405<br />

m-Xylene - g -57.22 0.568 25 730<br />

p-Xylene - g 40.41 0.793 28 952<br />

p-Xylene - l -54.99 0.557 26 310<br />

Durene C 6 H 2 (CH 3 ) 4 s -242.68 0.437 19 000<br />

Cumene C 6 H 5 CH(CH 3 ) 2 l -81.94 0.556 20 700<br />

Mesitylene C 6 H 3 (CH 3 ) 3 l -126.34 0.544 24 830<br />

Diphenyl C 6H 5-C 6H 5 s 135.34 0.319 57 400<br />

Diphenylmethane C 6 H 5 -CH 2 -C 6 H 5 s 117.22 0.340 63 600<br />

Triphenylmethane (C 6 H 5 ) 3 s 170.92 0.305 101 400<br />

Naphthalene C 10H 8 s 124.53 0.311 45 200<br />

Anthracene C 14 H 10 s 154.86 0.278 64 800<br />

Phenanthrene C 14 H 10 s 129.62 0.284 60 000<br />

Table 1.4: Formation enthalpies (cal/gr),<br />

(cal/mol). Hydrocarbons.<br />

entropies (cal gr −1 K −1 ) and free energies


1.3. MIXTURE OF GASES 11<br />

Helmholtz free energy<br />

Like in (1.20), here f can be expressed in the form<br />

f = u − T s. (1.24)<br />

( ) p<br />

f = f 0 + R g T ln , (1.25)<br />

p 0<br />

where<br />

f 0 = u − R g T s 0 (1.26)<br />

is the free energy at standard pressure, which <strong>de</strong>pends only on the gas temperature.<br />

Remark<br />

When <strong>de</strong>sired to refer the thermodynamic functions to a mol, the previous formulae<br />

should be changed as follows:<br />

1) c v and c p are the molar heats.<br />

2) R g must be substituted by the universal constant R.<br />

3) In the state equation (1.7), ρ must be substituted by the number c of moles of the<br />

gas per unit volume.<br />

1.3 Mixture of gases<br />

Now let us see the form taken by the thermodynamic functions in a mixture of gases.<br />

For the purpose we shall consi<strong>de</strong>r a mixture of gases formed by l different chemical<br />

species A i , (i = 1, 2, . . . , l). Let us assume that its state is such that the mixture<br />

behaves as an i<strong>de</strong>al gas. The problem <strong>de</strong>als with the expression of the thermodynamic<br />

functions by means of the corresponding functions of the chemical species of the<br />

mixture.<br />

Concentrations, mole fractions, <strong>de</strong>nsities and mass fractions<br />

Let c and c i , (i = 1, 2, . . . , l), be the number of moles of the mixture and of each<br />

species per unit volume respectively, and let X i be the mole fraction c i /c of species A i .


12 CHAPTER 1. THERMOCHEMISTRY<br />

Between these magnitu<strong>de</strong>s the following relations exist<br />

∑<br />

c j =c , (1.27)<br />

j<br />

∑<br />

X j =1 . (1.28)<br />

j<br />

Let ρ be the <strong>de</strong>nsity of the mixture, ρ i the partial <strong>de</strong>nsity of species A i , Y i its<br />

mass fraction and M i its molar mass. The following relations exist<br />

ρ i = ρY i , (1.29)<br />

∑ ∑<br />

ρ j = ρ, Y j = 1 , (1.30)<br />

j<br />

j<br />

c i = ρ M i<br />

Y i , (1.31)<br />

c = ρ ∑ j<br />

Y j<br />

M j<br />

, (1.32)<br />

The mean molar mass M m of the mixture is <strong>de</strong>fined by<br />

X i = ∑<br />

Y i/M i<br />

, (1.33)<br />

Y j /M j<br />

j<br />

Y i = ∑<br />

M iX i<br />

. (1.34)<br />

M j X j<br />

j<br />

M m = ∑ j<br />

M j X j . (1.35)<br />

From Eqs.<br />

following expression<br />

(1.30) and (1.35) one <strong>de</strong>duces for M m as a function of Y i the<br />

1<br />

M m<br />

= ∑ j<br />

Y j<br />

M j<br />

. (1.36)<br />

Equation of state of the mixture<br />

Since the number of moles of the mixture per unit volume is c, one has<br />

p = cRT = RT ∑ j<br />

c j = ∑ j<br />

p j , (1.37)<br />

where<br />

p j = c j RT = R gj T = X j p (1.38)


1.3. MIXTURE OF GASES 13<br />

is the partial pressure exerted by species A j when only this species occupies the volume<br />

of the mixture at temperature T and R gj is the gas constant of the said species.<br />

From (1.32), Eq. (1.37) can be written in the form<br />

p<br />

ρ = RT ∑ j<br />

Y j<br />

M j<br />

, (1.39)<br />

or else, in the form<br />

where<br />

R m = R ∑ j<br />

p<br />

ρ = R mT, (1.40)<br />

Y j<br />

M j<br />

=<br />

R M m<br />

= ∑ j<br />

R gj Y j (1.41)<br />

is the gas constant of the mixture. Opposite to R g , R m changes with the composition<br />

of the mixture. If the molar masses of the species are very different then the variations<br />

of R m with the composition of the mixture can be very ”large”. Otherwise, these<br />

variations are ”small”. For simplicity, in many aerothermochemical problems R m is<br />

assumed to be constant, taken a mean value.<br />

When nee<strong>de</strong>d to take into account the <strong>de</strong>viations of the state equation respect<br />

to Eq. (1.41) this can be attained by using an equation similar to (1.1) for the mixture.<br />

Thus, the second virial coefficient B m (T ) for the mixture is given by the expression 6<br />

∑∑<br />

B rs Y r Y s<br />

B m = M m √ . (1.42)<br />

Mr M s<br />

r<br />

s<br />

Here, if r = s, B rr is the second virial coefficient of species A r previously<br />

<strong>de</strong>fined. If r ≠ s, B rs is the second virial coefficient for the interaction potential<br />

between the molecules of species A r and A s . Since, in general, the constants relative<br />

to those mixed potentials are not available, several formulae have been proposed for<br />

B rs as a function of B rr and B ss . For example 7<br />

B rs = 1 2<br />

(√ √ )<br />

Mr Ms<br />

B rr + B ss . (1.43)<br />

M s M r<br />

The potential energy of an i<strong>de</strong>al gas due to interaction between molecules is<br />

zero. Therefore, each species adds to the value of any thermodynamic function ψ per<br />

unit mass of the mixture (internal energy, enthalpy, entropy, free energy, etc.) with<br />

a value which is in<strong>de</strong>pen<strong>de</strong>nt from the presence of other species. For A i , the said<br />

contribution to the value ψ i of the corresponding thermodynamic function per unit<br />

6 See Ref. [1], p. 153.<br />

7 See Ref. [6], p. 211.


14 CHAPTER 1. THERMOCHEMISTRY<br />

mass of the species at temperature T of the mixture and at partial pressure p i of A i<br />

multiplied by its mass fraction Y i . That is<br />

ψ = ∑ i<br />

ψ i Y i . (1.44)<br />

In the following, Eq. (1.44) is applied to the calculation of the thermodynamic<br />

functions of the mixture.<br />

Internal Energy<br />

The internal energy u per unit mass of the mixture is<br />

u = ∑ i<br />

u i Y i . (1.45)<br />

When taking into this expression the value<br />

u i = u 0i +<br />

∫ T<br />

obtained when (1.9) is particularized for species A i , it results<br />

u = u 0 +<br />

where<br />

∫ T<br />

T 0<br />

c vi (T ) dT (1.46)<br />

T 0<br />

c v (T ) dT, (1.47)<br />

u 0 = ∑ u 0i Y i (1.48)<br />

i<br />

is the energy of the mixture at temperature T 0 and<br />

c v = ∑ i<br />

c vi Y i (1.49)<br />

is the heat capacity at constant volume per unit mass of the mixture.<br />

Enthalpy<br />

Similarly<br />

with<br />

and<br />

∫ T<br />

h = h 0 + c p (T ) dT, (1.50)<br />

T 0<br />

h 0 = ∑ i<br />

c p = ∑ i<br />

h 0i Y i (1.51)<br />

c pi Y i . (1.52)


1.3. MIXTURE OF GASES 15<br />

Entropy<br />

Similarly<br />

∫ T<br />

c p (T )<br />

s = s 0 + dT − ∑<br />

T 0<br />

T<br />

j<br />

( )<br />

pj<br />

R gj Y j ln<br />

p 0<br />

(1.53)<br />

with<br />

s 0 = ∑ j<br />

s 0j Y j . (1.54)<br />

Equation (1.53) can also be written in the following form<br />

( ) p<br />

s = s 0 − R m ln − ∑ R gj Y j ln X j , (1.55)<br />

p 0<br />

j<br />

where<br />

s 0 = ∑ j<br />

s 0 jY j . (1.56)<br />

Once p 0 is fixed, this value <strong>de</strong>pends only on T and on the composition of the mixture.<br />

In (1.55), the two first terms of the right hand si<strong>de</strong> give the entropy that the gas would<br />

have at pressure p and at temperature T if instead of being a mixture it were a single<br />

chemical species. The remaining term is the entropy of mixing whose contribution is<br />

always positive.<br />

Free Energy<br />

Gibbs free energy<br />

g = h − T s, (1.57)<br />

where h and s are given by (1.50) and (1.55) respectively. g can also be expressed in<br />

the form<br />

( ) p<br />

g = g 0 + R m T ln + T ∑ p 0<br />

j<br />

R gj Y j ln X j , (1.58)<br />

where<br />

g 0 = h − T s 0 . (1.59)<br />

Helmholtz free energy<br />

where u and s are given by (1.47) and (1.55) respectively.<br />

f = u − T s, (1.60)


16 CHAPTER 1. THERMOCHEMISTRY<br />

As in the previous cases, f can be expressed in the form<br />

( ) p<br />

f = f 0 + R m T ln + T ∑ p 0<br />

j<br />

R gj Y j ln X j , (1.61)<br />

where<br />

f 0 = u − T s 0 . (1.62)<br />

Chemical Potentials<br />

The chemical potential µ i of species A i in the mixture is G i<br />

( )<br />

( )<br />

µ i = G i = G 0 pi<br />

p<br />

i + RT ln = G 0 i + RT ln + RT ln X i . (1.63)<br />

p 0 p 0<br />

1.4 Calculation of the thermodynamic functions<br />

The preceding formulas show that all thermodynamic functions of a gas can be <strong>de</strong>termined<br />

provi<strong>de</strong>d that one of them and the standard values for the others are known. For<br />

example, the heat capacity, either at constant pressure or at constant volume, and the<br />

formation enthalpy and entropy of the gas <strong>de</strong>termine all other thermodynamic functions.<br />

The law of variation of heat capacity as function of T is generally complicated.<br />

Fig. 1.5, for example, obtained from the tables in Ref. [10], shows the curves of c p<br />

versus T for some gases. These curves are usually approximated with empirical formulae.<br />

For example, the following Table 1.3, taken from Ref. [4] p. 28, gives some<br />

of these parabolic formulae as well as their maximum <strong>de</strong>viations within the range<br />

300 < T < 2 000K. The use of linear formulas is advisable for smaller ranges of T .<br />

Gas c p (cal gr −1 K −1 ) Max. error (%)<br />

H 2 3.440 + 0.033 × 10 −3 T + 0.1395 × 10 −6 T 2 1<br />

N 2 0.225 + 0.065 × 10 −3 T - 0.0123 × 10 −6 T 2 2<br />

O 2 0.196 + 0.086 × 10 −3 T - 0.0241 × 10 −6 T 2 1<br />

CO 0.223 + 0.075 × 10 −3 T - 0.0164 × 10 −6 T 2 2<br />

NO 0.207 + 0.081 × 10 −3 T - 0.0204 × 10 −6 T 2 2<br />

H 2 O 0.383 + 0.182 × 10 −3 T - 0.0191 × 10 −6 T 2 2<br />

CO 2 0.156 + 0.194 × 10 −3 T - 0.0562 × 10 −6 T 2 3<br />

NH 3 0.348 + 0.527 × 10 −3 T - 0.1040 × 10 −6 T 2 1<br />

C 2H 2 0.318 + 0.404 × 10 −3 T - 0.1020 × 10 −6 T 2 2 (for T > 400 K)<br />

CH 4 0.211 + 1.119 × 10 −3 T - 0.2620 × 10 −6 T 2 2<br />

Table 1.5: c p(T ) correlations for several gases.


1.4. CALCULATION OF THE THERMODYNAMIC FUNCTIONS 17<br />

1.4<br />

1.2<br />

1.0<br />

c p<br />

(cal/gr K)<br />

0.8<br />

0.6<br />

0.4<br />

C 2<br />

H 2<br />

CH 4<br />

CO 2<br />

H 2<br />

(0.1× c p<br />

)<br />

H 2<br />

O<br />

0.2<br />

CO, N 2<br />

O 2<br />

0.0<br />

0 400 800 1200 1600 2000 2400 2800<br />

T(K)<br />

Figure 1.5: Specific heat at constant pressure of several gases as function of temperature.<br />

When working at normal or not too high temperatures the <strong>de</strong>termination of<br />

the thermodynamic functions can be done through calorimetric measurements. On<br />

the other hand, when working at very high temperatures, as are those encountered<br />

in combustion processes, such measurements are very difficult when not impossible.<br />

In such case one has to resort to Statistical Mechanics and to the use of the results<br />

provi<strong>de</strong>d by the spectroscopic analysis of the gas. This method gives excellent results<br />

specially when <strong>de</strong>aling with simple molecules.<br />

The method is based on the formation of the so-called partition function of the<br />

gas. Once this function is known all thermodynamic functions are really computed. 8<br />

Quantum Mechanics shows that the energy of a particle confined within a domain<br />

can only take discrete set of values. These values are called “energy levels”. Let<br />

ε j (j = 1, 2, . . .) be the energy levels of the molecules, when the gas occupies volume<br />

V at temperature T . The following expression<br />

Q = ∑ j<br />

exp (−ε j /kT ) , (1.64)<br />

is called “partition function” of the gas molecules. The summation is exten<strong>de</strong>d to all<br />

energy levels. If one of them is <strong>de</strong>generate, it must be counted as many times as corresponds<br />

to its or<strong>de</strong>r of <strong>de</strong>generacy. That is, the number of discernible states in which<br />

this level can be attained. Fixing volume V and temperature T of the gas its partition<br />

8 For a full study on this matter see Refs. [11] and [12].


18 CHAPTER 1. THERMOCHEMISTRY<br />

function is <strong>de</strong>termined. Then, the problem lies in expressing the thermodynamic functions<br />

of the gas by means of its partition function. According to Statistical Mechanics<br />

the solution is as follows.<br />

a) Internal energy.<br />

b) Enthalpy.<br />

c) Entropy.<br />

h = R g T<br />

( ) ∂ ln Q<br />

u = R g T 2 . (1.65)<br />

∂T<br />

V<br />

[( ) ∂ ln Q<br />

+<br />

∂ ln T<br />

V<br />

s = R g<br />

[ln Q + T<br />

( ) ] ∂ ln Q<br />

. (1.66)<br />

∂ ln V<br />

T<br />

( ) ] ∂ ln Q<br />

. (1.67)<br />

∂ ln T<br />

V<br />

d) Free energy.<br />

d.1) Gibbs:<br />

g = R g T<br />

[<br />

ln Q −<br />

( ) ] ∂ ln Q<br />

. (1.68)<br />

∂ ln V<br />

T<br />

d.2) Helmholtz: f = −R g T ln Q. (1.69)<br />

e) Chemical Potential.<br />

G = M g g = RT<br />

[<br />

ln Q −<br />

( ) ] ∂ ln Q<br />

. (1.70)<br />

∂ ln V<br />

T<br />

Heat capacities at constant volume, c v , and at constant pressure, c p , are expressed<br />

by u and h respectively in the form<br />

c v = ∂u<br />

∂T ,<br />

c p = ∂h<br />

∂T . (1.71)<br />

This enables the calculation of their values as a function of Q by means of (1.65) and<br />

(1.66).<br />

Thus, the problem reduces to the <strong>de</strong>termination of the energy levels for the formation<br />

of Q. For the purpose, one must analyze the ways in which a gas molecule can<br />

store energy. Except when working at very high temperatures, where it is necessary to<br />

consi<strong>de</strong>r the states of electronic excitation in the molecule, this can be consi<strong>de</strong>red as<br />

a system of material points which are its atoms. Then, the energy of each molecule is<br />

the summation of the kinetic energies of its atoms plus the potential energy of the field<br />

of forces that hold them together. Be n the number of atoms in each molecule. Then,<br />

its number of <strong>de</strong>grees of freedom is 3n. Of which three are external <strong>de</strong>gree of freedom<br />

corresponding to the translational motion of the center of gravity of the molecule. The


1.4. CALCULATION OF THE THERMODYNAMIC FUNCTIONS 19<br />

remaining 3n − 3 are internal <strong>de</strong>grees of freedom and they correspond to its rotational<br />

and vibrational motions. Energy ε j of the molecule can always be expressed as<br />

ε j = ε t j + ε i j. (1.72)<br />

Here ε t j is the kinetic energy of the molecule due to the motion of its center of gravity,<br />

and ε i j is the internal energy. Moreover, the translational levels εt j are in<strong>de</strong>pen<strong>de</strong>nt<br />

from the internal levels ε i j . Therefore, the combination of either one of the εt j with<br />

either one of the ε i j gives the energy level ε j. Then, it is easily verified that the separation<br />

(1.72) for ε j corresponds to a factorization<br />

Q = Q t · Q i (1.73)<br />

for Q, where<br />

Q t = ∑ j<br />

exp ( −ε t j/kT ) (1.74)<br />

and<br />

Q i = ∑ j<br />

exp ( −ε i j/kT ) (1.75)<br />

are the translational and internal partition functions of the molecule respectively. All<br />

thermodynamic functions are linear and homogeneous in ln Q and in its <strong>de</strong>rivatives.<br />

Hence, when factorization (1.73) is substituted into them, these functions are separated<br />

into contributions from Q t and Q i respectively. Let Φ be one of these functions. Then,<br />

Φ = Φ t + Φ i , (1.76)<br />

where Φ t is the contribution to Φ from Q t , and Φ i is the contribution from Q i . This<br />

property simplifies the calculation of the thermodynamic functions.<br />

Translational partition function Q t is the same for all gases<br />

Q t = V<br />

( ) 3<br />

2πmkT<br />

2<br />

. (1.77)<br />

h 2<br />

Here, m is the mass of the molecule and h = 6.6238 × 10 −27 erg×s is the Planck’s<br />

constant. On the other hand Q i <strong>de</strong>pends on the structure of the molecule. At least in<br />

first approximation, ε i j can also be separated into contributions from internal energy<br />

of rotation ε r j and internal energy of vibration εv j<br />

ε i j = ε r j + ε v j . (1.78)<br />

Here, as before, ε r j and εv j are approximately in<strong>de</strong>pen<strong>de</strong>nt from each other. Therefore,<br />

separation (1.78) gives for Q i<br />

Q i = Q r · Q v . (1.79)


20 CHAPTER 1. THERMOCHEMISTRY<br />

Two cases can arise in the rotational motion of a molecule:<br />

1) If the molecule is linear, that is, if all its atoms lie on a straight line, the moment<br />

of inertia with respect to the axis of the molecule is zero and it cannot store<br />

energy in such a <strong>de</strong>gree of freedom. Hence, the rotational <strong>de</strong>grees of freedom of<br />

the molecule reduce to two.<br />

2) Whereas if the molecule is non-linear the number of its rotational <strong>de</strong>grees of<br />

freedom is 3.<br />

If the rotational coordinates are selected accordingly, 9 the energies corresponding<br />

to the rotational <strong>de</strong>grees of freedom are also separable. Then, except at temperatures<br />

un<strong>de</strong>r ambient, the distribution Q rj from the rotational <strong>de</strong>gree of freedom j to<br />

the rotational partition function Q r is<br />

√<br />

2kT Ij<br />

Q rj = 2π , (1.80)<br />

h<br />

where I j is the moment of inertia of the molecule for such <strong>de</strong>gree. Hence, we have:<br />

a) For linear molecules<br />

Q r = 8π2 kT I<br />

h 2 , (1.81)<br />

δ<br />

where I is the moment of inertia of the molecule respect to an axis passing<br />

through its center of gravity and normal to the axis of the molecule.<br />

b) For non-linear molecules<br />

Q r = 1 δ<br />

( 8π 2 ) 3<br />

2<br />

kT √Ix<br />

h 2 I y I z . (1.82)<br />

Here I x , I y and I z are the three moments of inertia of the molecule with respect<br />

to three orthogonal axis at its center of gravity.<br />

In (1.81) and (1.82) δ is a symmetry factor, that accounts for undisguised stable<br />

states due to symmetries of the molecule. For example, for linear molecules δ = 2.<br />

Thus, Q r is <strong>de</strong>termined when the moments of inertia of the molecule are known. Such<br />

moments are obtained from spectroscopic analysis.<br />

As for vibration energy a direct <strong>de</strong>termination of its levels through spectroscopic<br />

analysis is best. In<strong>de</strong>pen<strong>de</strong>nt contributions of the vibrational <strong>de</strong>grees of freedom<br />

can, otherwise, be obtained by assuming that their corresponding energies are<br />

separable. Which happens, for example, when vibration amplitu<strong>de</strong>s are small enough<br />

so that the potential energy of the molecule can be approximated by a quadratic function<br />

of the <strong>de</strong>viations of the atoms respect to their equilibrium positions. For this<br />

9 For which the principal axis of inertia of the molecule must be taken.


1.5. CHEMICAL REACTIONS IN A MIXTURE OF GASES 21<br />

case, a normal system of coordinates can be adopted where the vibrational energy is<br />

separated into contributions from the natural mo<strong>de</strong>s of the molecule. Be ν s the frequency<br />

of one of these mo<strong>de</strong>s. Un<strong>de</strong>r the assumption that the molecule behaves as an<br />

harmonic oscillator the contribution Q vs of this mo<strong>de</strong> to Q v is<br />

(<br />

Q vs = exp − hν ) ( (<br />

s<br />

1 − exp − hν )) −1<br />

s<br />

. (1.83)<br />

2kT<br />

kT<br />

Frequency ν s is obtained from the vibration spectra of the molecule.<br />

Finally, for very high temperatures the contribution of electronic excitation<br />

must be inclu<strong>de</strong>d in the value of Q. The corresponding energy levels are obtained<br />

from the spectra.<br />

The preceding analysis assumes that the rotational and vibrational energies are<br />

separable. Yet, this does not always happen, mainly for high temperatures. In fact, rotation<br />

stretches the molecule and this introduces coupling terms between the rotational<br />

and vibrational <strong>de</strong>grees of freedom, preventing a separate study of both contributions.<br />

Besi<strong>de</strong>s, the assumption that vibration is harmonic not always is justified, which forces<br />

the introduction of correcting terms to account for anharmonicity. 10<br />

1.5 Chemical reactions in a mixture of gases<br />

Now, let us assume that the l species A i of the mixture can react according to a system<br />

of r reaction equations of the form<br />

∑<br />

ν ijA ′ i ⇄ ∑<br />

i<br />

i<br />

ν ′′<br />

ijA i , (j = 1, . . . , r), (1.84)<br />

where ν ij ′ and ν′′ ij are the stoichiometric coefficients of species A i in the reaction j for<br />

the forward and backward reactions respectively. Let<br />

These coefficients must satisfy the following conditions<br />

ν ij = ν ′′<br />

ij − ν ′ ij. (1.85)<br />

∑<br />

ν ij M i = 0, (j = 1, . . . , r), (1.86)<br />

i<br />

which express that the mass of the mixture is not changed by chemical reactions.<br />

Let us consi<strong>de</strong>r the unit mass of the mixture and be ∆C ij the number of moles<br />

of species A i produced by the reaction j. Due to (1.84) the following relations exist<br />

10 For further information on this problem see Refs. [11] and [12].


22 CHAPTER 1. THERMOCHEMISTRY<br />

between ∆C ij<br />

∆C 1j<br />

ν 1j<br />

= ∆C 2j<br />

ν 2j<br />

= . . . = ∆C lj<br />

ν lj<br />

. (1.87)<br />

Let ξ j be the common value of these relations. ξ j is called the <strong>de</strong>gree of<br />

advancement of reaction j (De Don<strong>de</strong>r [13]). Thus <strong>de</strong>fined, ξ j has the dimensions<br />

mole/gr. The number of moles of species A i produced by reaction j is, therefore,<br />

∆C ij = ν ij ξ j , (1.88)<br />

and the number ∆C i of moles of A i produced by the r reactions is<br />

∆C i = ∑ j<br />

∆C ij = ∑ j<br />

ν ij ξ j . (1.89)<br />

If C i0 is the number of moles of species A i existing when the reactions start,<br />

the number C i that will exists when the <strong>de</strong>grees of advancement of the reactions are<br />

ξ j , (j = 1, 2, . . . , r), is<br />

C i = C i0 + ∑ ν ij ξ j . (1.90)<br />

j<br />

Similarly the mass fraction Y i of species A i produced by the r reaction is<br />

∑<br />

Y i = M i ν ij ξ j , (1.91)<br />

and if Y i0 is the initial mass fraction of this species, one has<br />

∑<br />

Y i = Y i0 + M i ν ij ξ j . (1.92)<br />

j<br />

j<br />

1.6 Chemical equilibrium<br />

Conditions of equilibrium<br />

Once the reactions are initiated they continue up to the point where the mixture reaches<br />

its state of equilibrium. Such an equilibrium is <strong>de</strong>termined by additional conditions<br />

which fix the state of the system. Let us assume, for example, that the process is<br />

adiabatic and takes place at constant pressure. Such a case has a great practical interest<br />

in the study of combustion processes. Then the First Law of Thermodynamics shows<br />

that the enthalpy of the mixture must be constant. Therefore, the conditions that must<br />

be satisfied by the mixture are as follows<br />

h = h 0 = const., p = p 0 = const. (1.93)


1.6. CHEMICAL EQUILIBRIUM 23<br />

Since the process is adiabatic the state of equilibrium is <strong>de</strong>termined by the<br />

condition that entropy s be maximum. The condition for this is that its first variation<br />

be zero. Moreover, from (1.93) the variations of h and p must also be zero. Therefore,<br />

the state of equilibrium is <strong>de</strong>termined by conditions<br />

δh = δp = δs = 0. (1.94)<br />

Thermodynamics shows 11 that the entropy variations of the mixture is given by<br />

the expression<br />

T δs = δh − V δp − ∑ i<br />

µ i<br />

M i<br />

δY i . (1.95)<br />

Here V is the volume of the unit mass of the mixture and µ i is the chemical potential<br />

of species A i in the mixture. This relation, by virtue of (1.94), reduces to<br />

∑ µ i<br />

δY i = 0. (1.96)<br />

M i<br />

i<br />

Variations of δY i cannot be arbitrary since Y i must satisfy conditions (1.92). From<br />

them one obtains<br />

δY i = ∑ j<br />

M i ν ij δξ j . (1.97)<br />

This expression when taken into (1.96) gives<br />

∑∑<br />

µ i ν ij δξ j = ∑<br />

i j<br />

j<br />

( ∑<br />

i<br />

µ i ν ij<br />

)<br />

δξ j = 0, (1.98)<br />

and since the variations δξ j are arbitrary this condition breaks into the following system<br />

of equilibrium equations<br />

∑<br />

µ i ν ij = 0, (j = 1, 2, . . . , r). (1.99)<br />

i<br />

The expression ∑ µ i ν ij is named affinity α j of reaction j. Therefore, the equilibrium<br />

is <strong>de</strong>termined by the condition that the affinities of the reactions become zero<br />

i<br />

α j = 0, (j = 1, 2, . . . , r). (1.100)<br />

System (1.100), together with conditions (1.92) and (1.93), <strong>de</strong>termine the values<br />

for the <strong>de</strong>grees of advancement of the various reactions and for the corresponding<br />

mass fractions of the species as well as for the temperature T of the mixture.<br />

11 See Ref. [4], p. 68.


24 CHAPTER 1. THERMOCHEMISTRY<br />

Remarks<br />

If the reactions take place at pressure and temperature (instead of enthalpy) both constant<br />

the conditions of equilibrium are<br />

δp = δT = δg = 0. (1.101)<br />

But<br />

δg = δh − sδT − T δs. (1.102)<br />

From Eqs. (1.95), (1.97), (1.101) and (1.102) conditions (1.99) are obtained again.<br />

Thus conditions of equilibrium are the same for both cases, with the only difference<br />

that here, temperature is given whilst in the previous case it must be calculated. The<br />

computation is done by <strong>de</strong>termining the conditions of equilibrium as a function of temperature,<br />

and then looking for the value of T which satisfies the additional condition<br />

h = h 0 .<br />

are<br />

Likewise, for constant volume adiabatic reactions the conditions of equilibrium<br />

δu = δV = δs = 0, (1.103)<br />

from which result again conditions (1.99).<br />

1.7 Case of a mixture of i<strong>de</strong>al gases<br />

For the case of a mixture of i<strong>de</strong>al gases the form taken by the conditions of equilibrium<br />

is specially simple. In fact, as for (1.63) system (1.99) reduces to<br />

∑<br />

G i ν ij = 0, (j = 1, 2, . . . , r), (1.104)<br />

i<br />

or else<br />

∑<br />

G 0 i ν ij + RT ∑<br />

i<br />

i<br />

( )<br />

pi<br />

ν ij ln = 0, (j = 1, 2, . . . , r), (1.105)<br />

p 0<br />

which together with (1.38) gives<br />

∏<br />

i<br />

X ν ij<br />

i =<br />

( ) (<br />

νj<br />

p0<br />

exp − ∆G0 j<br />

p<br />

RT<br />

)<br />

, (j = 1, 2, . . . , r), (1.106)<br />

where<br />

ν j = ∑ i<br />

ν ij (1.107)


∆G 0 j = ∑ i<br />

1.7. CASE OF A MIXTURE OF IDEAL GASES 25<br />

and<br />

(<br />

Hi − T Si<br />

0 )<br />

νij (1.108)<br />

is the increase in free energy due to reaction j at temperature T of the mixture and<br />

at standard pressure p 0 . Thus ∆G 0 j <strong>de</strong>pends only on T . Table 1.2 gives the values of<br />

for several species corresponding to their formation from the chemical elements<br />

∆G 0 j<br />

at the standard state. Therefore Eq. (1.106) can be written<br />

∏<br />

( ) νj<br />

X νij p0<br />

i = Kj 0 (T ) , (j = 1, 2, . . . , r), (1.109)<br />

p<br />

i<br />

where Kj 0 is called the equilibrium constant of reaction j <strong>de</strong>pending only on temperature<br />

T . The values of Kj 0 versus T are given in Fig. 1.6, for several typical reactions<br />

of interest in combustion problems.<br />

30<br />

25<br />

1<br />

2H2 +OH ↔ H2O<br />

20<br />

log 10<br />

(K p<br />

)<br />

15<br />

10<br />

5<br />

0<br />

CO<br />

H 2 + 1 2O2 ↔ H2O<br />

CO 2 +H 2 ↔ CO +H 2O<br />

+ 1 2O2 ↔ CO2<br />

1<br />

2N2 +H2O ↔ NO +H2O<br />

H 2O ↔ H 2 +O<br />

1<br />

2 N2 + 1 2O2 ↔ NO<br />

1<br />

2 H2 ↔ H<br />

−5<br />

1<br />

2 N2 ↔ N<br />

−10<br />

0 400 800 1200 1600 2000 2400 2800 3200 3600<br />

T(K)<br />

Figure 1.6: Equilibrium constant of reaction vs temperature, for several typical reactions of<br />

interest in combustion problems.<br />

The chemical equations (1.84) used for this calculation can be any ones within<br />

the following limitations: a) they must be linearly in<strong>de</strong>pen<strong>de</strong>nt; b) their number must<br />

be the maximum that can exist between species. This number can be <strong>de</strong>termined when<br />

the phase rule is applied. Be l ′ < l the number of components of the mixture, then<br />

the number of in<strong>de</strong>pen<strong>de</strong>nt reactions will be l − l ′ . In turn l ′ can be <strong>de</strong>termined as<br />

follows: let E j , (j = 1, 2, . . . , l), be the chemical elements forming the species and<br />

a ij be the number of atoms of element E j in species A i . The number of components


26 CHAPTER 1. THERMOCHEMISTRY<br />

of the mixture is the rank of the matrix of coefficients a ij .<br />

The coefficients of a<br />

principal minor of or<strong>de</strong>r l ′ of this matrix <strong>de</strong>termine the species that can be adopted<br />

as components of the mixture. A complete system of in<strong>de</strong>pen<strong>de</strong>nt chemical reactions<br />

can now be obtained by expressing each one of the remaining species as a linear<br />

combination of the said components in the form<br />

∑<br />

A j = b ji A i , (j = l ′ + 1, l ′ + 2, . . . , l). (1.110)<br />

l ′<br />

i=1<br />

Here, the first l ′ species are assumed to be the components of the mixture.<br />

For example in the combustion of mixtures of hydrocarbons and some of their<br />

compounds with oxygen and nitrogen, the combustion products are formed by the<br />

species: CO 2 , CO, H 2 O, HO, H 2 , H, O 2 , O, NO, N 2 and N. Therefore, one has:<br />

l = 11; l ′ = 4; and the number of in<strong>de</strong>pen<strong>de</strong>nt chemical reactions is 7. The following<br />

systems of reactions can be adopted (see Ref. [13]) since they are linearly<br />

in<strong>de</strong>pen<strong>de</strong>nt:<br />

C O 2 + H 2 ⇄ C O + H 2 O<br />

1<br />

2 N 2 + H 2 O ⇄ N O + H 2<br />

2 H 2 O ⇄ 2 H 2 + O 2<br />

H 2 O ⇄ O + H 2<br />

N 2 ⇄ 2 N<br />

H 2 ⇄ 2 H<br />

H 2 O ⇄ 1 2 H 2 + O H<br />

(1.111)<br />

The r equations (1.106), together with the l ′ equations which express the conservation<br />

of components in the reactions, <strong>de</strong>termine the equilibrium composition of<br />

the mixture, once its temperature is known. Thus in the preceding example system<br />

(1.111) must be completed with the four equations for the conservation of the number<br />

of atoms of carbon, oxygen, hydrogen and nitrogen. It is not always necessary to consi<strong>de</strong>r<br />

all possible reactions since <strong>de</strong>pending on the state of the mixture the fractions of<br />

some of the species may be neglected and the computation simplified.<br />

For the case of adiabatic equilibrium the temperature of the products is unknown.<br />

Hence, trial and error methods become necessary. In applying then, first a<br />

temperature is assumed and the equilibrium composition is <strong>de</strong>termined for it. Once<br />

this is known the corresponding temperature is obtainable through equation (1.10),<br />

h = h 0 , and if this temperature differs from the one previously assumed, the computation<br />

should be reviewed. Gay<strong>de</strong>n and Wolfhard [14] recommend the application of<br />

Damköhler and Edse’s method mentioned in Ref. [15]. The fundamentals and some<br />

of the applications of this method are <strong>de</strong>scribed in the book by Gay<strong>de</strong>n and Wolfhard.


1.7. CASE OF A MIXTURE OF IDEAL GASES 27<br />

0.30<br />

3200<br />

T b<br />

X i<br />

0.25<br />

0.20<br />

0.15<br />

H O 2<br />

CO (X /2)<br />

H CO CO 2 2<br />

O (X /2) 2 O2<br />

3000<br />

2800<br />

2600<br />

T(K)<br />

0.10<br />

0.05<br />

H<br />

O<br />

OH<br />

2400<br />

2200<br />

0.00<br />

2000<br />

0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20<br />

Y<br />

O2<br />

/ (Y<br />

O2<br />

) s<br />

Figure 1.7: Compositions and temperature of the combustion products obtained when burning<br />

a mixture of ethylene and oxygen as a function of the ratio of the mass<br />

fraction of oxygen to the corresponding stoichiometric mixture.<br />

Other methods such as those in Refs. [16] and [17] inclu<strong>de</strong> diagrams in or<strong>de</strong>r to reduce<br />

calculations as they are very cumbersome when the number of species is large.<br />

Then, specially when many cases must be solved, it becomes necessary to systematize<br />

the procedure. This is accomplished by giving computational schemes which enable<br />

the use of electronic computers. Full <strong>de</strong>scription of these methods can be found, for<br />

instance, in Refs. [18] and [19] where additional bibliography is inclu<strong>de</strong>d.<br />

Fig. 1.7, taken from Ref. [14], shows the compositions and temperature of the<br />

combustion products obtained when burning a mixture of ethylene and oxygen as a<br />

function of the ratio of the mass fraction of oxygen to the corresponding stoichiometric<br />

mixture.<br />

Fig. 1.8, taken from Ref. [17], gives the theoretical combustion temperatures<br />

for some hydrocarbons.<br />

Table 1.6, taken from Ref. [14], gives the adiabatic combustion temperatures<br />

for some typical mixtures, as well as these that would be obtained by neglecting dissociations.<br />

The book by Lewis and von Elbe mentioned in Ref. [20] gives the experimental<br />

adiabatic combustion temperatures for great number of mixtures. See Ref. [21]<br />

for the experimental technique used for the <strong>de</strong>termination of combustion temperatures.


28 CHAPTER 1. THERMOCHEMISTRY<br />

2500<br />

2400<br />

acetylene<br />

methylacetylene<br />

2300<br />

T b<br />

(K)<br />

2200<br />

ethene<br />

benzene<br />

propylene<br />

a:n−hexane<br />

1−butene<br />

b: n−pentane<br />

2100<br />

1−pentene<br />

c: n−butane<br />

d: propane<br />

cyclopentene<br />

metane a,b,c,d<br />

2000<br />

etane<br />

0.8 0.9 1.0 1.1 1.2 1.3 1.4<br />

Y / (Y ) F F s<br />

Figure 1.8: Theoretical combustion temperatures of several hydrocarbons in the air at atmospheric<br />

pressure.<br />

Fuel CH 4 CH 4 C 3H 8 C 3H 8 C 8H 18 C 8H 18 C 2H 2 H 2<br />

Mixture 8<br />

>< Fuel 0.2<br />

Y i O 2 0.8<br />

> :<br />

N 2 0<br />

Products<br />

H 2O<br />

O 2<br />

O<br />

H 2<br />

8> < OH<br />

X i<br />

CO<br />

> :<br />

CO 2<br />

NO<br />

N 2<br />

H<br />

0.3719<br />

0.0691<br />

0.0326<br />

0.0690<br />

0.1438<br />

0.1512<br />

0.1166<br />

0<br />

0<br />

0.0458<br />

0.1062<br />

0.3865<br />

0.5073<br />

0.2735<br />

0.0128<br />

0.0041<br />

0.0343<br />

0.0402<br />

0.0747<br />

0.0818<br />

0.0054<br />

0.4527<br />

0.0105<br />

0.2245<br />

0.7755<br />

0<br />

0.2988<br />

0.0665<br />

0.0375<br />

0.0648<br />

0.1378<br />

0.2064<br />

0.1372<br />

0<br />

0<br />

0.0510<br />

0.1738<br />

0.6005<br />

0.2257<br />

0.2627<br />

0.0483<br />

0.0230<br />

0.0495<br />

0.1010<br />

0.1612<br />

0.1232<br />

0.0100<br />

0.1880<br />

0.0331<br />

0.2218<br />

0.7782<br />

0<br />

0.2562<br />

0.0880<br />

0.0462<br />

0.0538<br />

0.1447<br />

0.2167<br />

0.1450<br />

0<br />

0<br />

0.0493<br />

0.1076<br />

0.3777<br />

0.5147<br />

0.1853<br />

0.0307<br />

0.0073<br />

0.0175<br />

0.0453<br />

0.0782<br />

0.1259<br />

0.0091<br />

0.2923<br />

0.0084<br />

0.0918<br />

0.2113<br />

0.6969<br />

0.0745<br />

0.0015<br />

0.0008<br />

0.0170<br />

0.0092<br />

0.1159<br />

0.0810<br />

0.0019<br />

0.6934<br />

0.0048<br />

0.0305<br />

0.1360<br />

0.8335<br />

0.1886<br />

0.00<br />

0.00<br />

0.1505<br />

0.00<br />

0<br />

0<br />

0.00<br />

0.5609<br />

Temperatures ( ◦ C)<br />

Dissociation 2 737 2 410 2 776 2 689 2 809 2 447 2 300 1 358<br />

No Dissociation 5 047 2 980 5 202 4 405 5 740 3 417 2 439 1 362<br />

0.00<br />

Table 1.6: Adiabatic combustion temperature for some typical mixtures.


1.8. CHEMICAL KINETICS 29<br />

1.8 Chemical kinetics<br />

In some of the Aerothermochemistry problems it is essential to know the reaction<br />

rates of the species forming a mixture of gases. 12<br />

<strong>de</strong>termined:<br />

1) The reactions that can take place between the species.<br />

2) The governing laws of their reaction rates.<br />

Such questions are the subject of Chemical Kinetics.<br />

For which the following must be<br />

In Ref. [22] an introduction to Chemical Kinetics can be found mainly from<br />

the viewpoint of jet propulsion. For a more complete study on this subject see Refs.<br />

[23], [24] and [25]. The present paragraph gives a brief introduction to this matter.<br />

First let us consi<strong>de</strong>r the case where the species react according to a single<br />

irreversible reaction<br />

∑<br />

ν iA ′ i → ∑<br />

i<br />

i<br />

ν ′′<br />

i A i , (1.112)<br />

and let ξ be the <strong>de</strong>gree of advancement of this reaction as was <strong>de</strong>fined in paragraph 5.<br />

The reaction rate of Eq. (1.112) is, by <strong>de</strong>finition,<br />

r = dξ<br />

dt = ˙ξ. (1.113)<br />

From here, keeping in mind (1.88), the number dC i / dt of moles of species A i<br />

produced per unit mass of the mixture and per unit time is<br />

dC i<br />

dt<br />

= (ν ′′<br />

i − ν ′ i) dξ<br />

dt . (1.114)<br />

Likewise, the mass dm i / dt of these species produced per unit mass and time is<br />

dm i<br />

dt<br />

= M i<br />

dC i<br />

dt<br />

= M i (ν ′′<br />

i − ν ′ i) dξ<br />

dt . (1.115)<br />

Aerothermochemistry <strong>de</strong>als, normally, with mass w i of A i produced per unit<br />

volume and per unit time, which is<br />

w i = ρ dm i<br />

dt<br />

= ρM i (ν ′′<br />

i − ν ′ i) dξ<br />

dt . (1.116)<br />

The problem lies in <strong>de</strong>termining the way in which w i <strong>de</strong>pends on the state and composition<br />

of the mixture. Such <strong>de</strong>termination is done through the so-called law of mass<br />

action of Guldberg and Waage [26].<br />

12 See, i.e., Chap. 6.


30 CHAPTER 1. THERMOCHEMISTRY<br />

This law states that the number dc i / dt of moles of A i produced within unit<br />

volume during unit time is<br />

dc i<br />

dt = (ν′′ i − ν ′ i) k ∏ j<br />

c ν′ j<br />

j . (1.117)<br />

Here, k is in<strong>de</strong>pen<strong>de</strong>nt from the composition and pressure of the mixture and <strong>de</strong>pends<br />

only on its temperature. It is called specific rate or rate coefficient of the reaction and<br />

its dimensions are ( cm 3 /mol ) ν−1<br />

s −1 , where<br />

ν = ∑ j<br />

ν ′ j (1.118)<br />

is called or<strong>de</strong>r or molecularity of the reaction.<br />

Since the molar mass of A i is M i , we have for w i , from (1.117)<br />

w i = M i<br />

dc i<br />

dt = M i (ν ′′<br />

i − ν ′ i) k ∏ j<br />

c ν′ j<br />

j . (1.119)<br />

This formula can also be expressed as a function of the mass fractions of the reacting<br />

species, by virtue of (1.31), in the form<br />

w i = ∏<br />

ρ ν<br />

M ν′ j<br />

j<br />

j<br />

kM i (ν ′′<br />

i − ν ′ i) ∏ j<br />

Y ν′ j<br />

j . (1.120)<br />

For example, in the case of a bimolecular reaction between species A 1 and A 2 which<br />

produces A 3 ,<br />

one has, due to (1.120),<br />

A 1 + A 2 → A 3 , (1.121)<br />

w 1<br />

= w 2<br />

= − w 3<br />

= dc 1<br />

M 1 M 2 M 3 dt = dc 2<br />

dt = − dc 3<br />

dt = −<br />

ρ2<br />

kY 1 Y 2 . (1.122)<br />

M 1 M 2<br />

The way in which the specific rate k <strong>de</strong>pends on temperature is given by Arrhenius’s<br />

law<br />

k = Ae −E/RT (1.123)<br />

where E is called activation energy of the reaction and A is a constant or the product<br />

of a constant by a power of the absolute temperature T<br />

A = BT α . (1.124)<br />

Here B is a constant called frequency factor and α is an exponent within zero and one.<br />

Arrhenius [27] <strong>de</strong>duced his law empirically as a generalization of van’t Hoff’s<br />

law for the variation of equilibrium constants as functions of temperature. Statistical


1.8. CHEMICAL KINETICS 31<br />

Mechanics 13 enables a theoretical <strong>de</strong>rivation of this law throughout the so-called Collisions<br />

Theory. For this <strong>de</strong>rivation it is enough to assume that for a reaction to occur<br />

the following is nee<strong>de</strong>d:<br />

1) Coinci<strong>de</strong>nce of reacting molecules in a collision.<br />

2) The collision energy in certain <strong>de</strong>grees of freedom must be larger than a given<br />

value which <strong>de</strong>pends on the activation energy of reaction.<br />

When calculating the number of collisions per unit volume and per unit time that satisfy<br />

the preceding conditions, one obtains expressions similar to the Arrhenius law for<br />

the influence of temperature and to the law of mass action for the influence of concentrations.<br />

For example, in the simple case of a bimolecular reaction of the type (1.121)<br />

the number n of collisions per unit volume and per unit time between molecules of<br />

the species A 1 and A 2 is<br />

n = N 2 σ 2 12c 1 c 2<br />

√8πRT<br />

( )<br />

M1 + M 2<br />

. (1.125)<br />

M 1 M 2<br />

Here σ 12 is the mean molecular diameter of both species. Let us assume that from the<br />

n collisions (1.125) only are active, that is, only produce reaction, those that have a<br />

relative kinetic energy of approximation larger than ε = E/N. 14 The number n a of<br />

these active collisions is<br />

n a = n e −E/RT . (1.126)<br />

Hence, the velocity dc 1 / dt of the reaction, in moles per unit volume and per unit<br />

time, is<br />

That is, due to (1.125),<br />

dc 1<br />

dt = dc 2<br />

dt = − dc 3<br />

dt = −n a<br />

N . (1.127)<br />

√<br />

dc 1<br />

dt = dc 2<br />

dt = − dc ( )<br />

3<br />

dt = M1 + M 2<br />

−Nσ2 12c 1 c 2 8πRT<br />

e −E/RT . (1.128)<br />

M 1 M 2<br />

Due to (1.31), this formula can be written in the form<br />

dc 1<br />

dt = dc 2<br />

dt = − dc 3<br />

dt =<br />

−<br />

√<br />

ρ2<br />

Nσ 2<br />

M 1 M<br />

12Y 1 Y 2 8πRT<br />

2<br />

( )<br />

M1 + M 2<br />

e −E/RT ,<br />

M 1 M 2<br />

(1.129)<br />

13 See Ref. [23].<br />

14 It is not essential that the kinetic energy of approximation of the collision be precisely the one larger<br />

than ε. The formulae holds as long as the collision energy in two separated <strong>de</strong>grees of freedom is larger<br />

than ε.


32 CHAPTER 1. THERMOCHEMISTRY<br />

which can be i<strong>de</strong>ntified with Eq. (1.122) by making<br />

√ ( )<br />

k = Nσ12<br />

2 M1 + M 2<br />

8πRT<br />

e −E/RT , (1.130)<br />

M 1 M 2<br />

that is, from Eq. (1.124),<br />

α = 1 2<br />

(1.131)<br />

and<br />

B = Nσ 2 12<br />

√<br />

8πR<br />

( )<br />

M1 + M 2<br />

. (1.132)<br />

M 1 M 2<br />

When comparing Collisions Theory with experimental results, an exact numerical<br />

agreement between the reaction rates predicted and observed is not to be expected,<br />

but an agreement of the or<strong>de</strong>rs of magnitu<strong>de</strong>. In fact, Collisions Theory does not inclu<strong>de</strong>,<br />

among others, the influence of the relative orientation of the molecules as they<br />

colli<strong>de</strong> which can <strong>de</strong>ci<strong>de</strong> the success of the collision. Such an effect can be of importance<br />

specially in molecules of complicated structure. It is taken into consi<strong>de</strong>ration<br />

by including in Eq. (1.129) a numerical factor P of orientation equal or smaller than<br />

unity<br />

dc 1<br />

dt = dc 2<br />

dt = − dc 3<br />

dt =<br />

− P<br />

√<br />

ρ2<br />

Nσ 2<br />

M 1 M<br />

12Y 1 Y 2 8πRT<br />

2<br />

( )<br />

M1 + M 2<br />

e −E/RT .<br />

M 1 M 2<br />

(1.133)<br />

Experimental reaction rates agree, fairly, with those predicted by Collision Theory<br />

only for a short number of reactions, yet in other cases experimental rates are as much<br />

as 10 8 times larger or smaller than those predicted by theory. Such a discrepancy can<br />

not be justified within the theory. These results that in the best cases Collision Theory<br />

represents approximately the actual facts only for a short number of reactions. A more<br />

correct formulation is provi<strong>de</strong>d by the so-called Theory of Absolute Reaction Rates.<br />

In this theory it is assumed that the passing of the reacting species to the products<br />

takes place through the formation of an activated complex resulting from the<br />

assembly of the reacting species. This complex is consi<strong>de</strong>red as located at the top of<br />

an energy barrier which separates the species from the products. The reaction rate<br />

is <strong>de</strong>termined by the velocity at which the activated complex crosses the barrier. It<br />

is, furthermore, assumed that such complex stands in equilibrium with the reacting<br />

species and that its dissociation in products is due to the action of one of the vibrational<br />

<strong>de</strong>gree of freedom. Then, by applying the laws of Statistical Mechanics the<br />

reaction rate can be computed when the structure of the activated complex is known.


1.8. CHEMICAL KINETICS 33<br />

This structure can be obtained from Quantum Mechanics or be postulated from the<br />

structure of the molecules. The Absolute Reaction Rate Theory has been successfully<br />

applied to the study of a great number of reactions well as to many other physical<br />

and physicochemical phenomena. A full <strong>de</strong>velopment of the theory can be found, for<br />

instance, in the works mentioned in Refs. [24] and [25].<br />

In the previous study it has been assumed that a single one directional reaction<br />

takes place within the mixture. Generally, all reactions proceed simultaneously in both<br />

senses at different rates. Thus, reaction (1.112) must be substituted by the following<br />

∑<br />

ν iA ′ i ⇄ ∑<br />

i<br />

i<br />

ν ′′<br />

i A i , (1.134)<br />

and the rate w i of production of species A i is given by the expressions<br />

(<br />

)<br />

∏ ∏<br />

w i = M i (ν i ′′ − ν i)<br />

′ k f c ν′ i<br />

− k b c ν′′ i<br />

. (1.135)<br />

Here k f and k b are the specific rates corresponding to the forward and backward reactions<br />

respectively.<br />

i<br />

On the other hand to be able to form the products, the reacting molecules must<br />

contact one another. This reduces the molecularity of a single elementary reaction to<br />

1, 2 or 3, since the probability of a collision of more than three molecules is practically<br />

nil. Consequently, when the reaction rate <strong>de</strong>pends in a complicated way on the<br />

concentrations, it can be assured that it proceeds from the combination of several elementary<br />

reactions. The basic problem of Chemical Kinetics is then the establishment<br />

of the set of elementary reactions that produce within the mixture. For example, the<br />

stoichiometric equation for the <strong>de</strong>composition of ozone is<br />

While the following is the actual system of reactions [28]:<br />

i<br />

i<br />

i<br />

2O 3 ⇄ 3O 2 . (1.136)<br />

O 3 +G k fi<br />

⇄<br />

k b1<br />

O+O 2 +G, (1.137)<br />

O+O 3<br />

k f2<br />

⇄<br />

k b2<br />

2O 2 , (1.138)<br />

O 2 +G k f3<br />

⇄<br />

k b3<br />

2O+G. (1.139)<br />

Here G is one of the atoms or molecules of the mixture. Each one of these reactants<br />

acts at its own reaction rate and the rate of consumption of the ozone results from


34 CHAPTER 1. THERMOCHEMISTRY<br />

the combination of the rates corresponding to elementary reactions. Denoting species<br />

O, O 2 and O 3 with subscripts 1, 2 and 3 respectively, one obtains for the rate of<br />

consumption of O 3<br />

− w 3<br />

M 3<br />

= k f1 cc 3 − k b1 cc 1 c 3 + k f2 c 1 c 3 − k b2 c 2 2. (1.140)<br />

This expression differs essentially from that obtainable from Eq. (1.136).<br />

The complexity of the system of the elementary reactions that produce within<br />

a mixture of reacting species originates an essential difficulty for the study of these<br />

problems which <strong>de</strong>pend on the reaction rates of the species. One of such problems<br />

is, for example, the calculation of the propagation velocity of a flame throughout a<br />

combustible mixture. 15 A great effort is being ma<strong>de</strong> by Chemical Kinetics towards the<br />

enlightenment of the elementary reactions that produce in each case and of their corresponding<br />

rates. Specially in the field of Combustion, the Fifth International Congress<br />

held in 1954 <strong>de</strong>voted the major part of its work to the study of Combustion Kinetics. 16<br />

In spite of the efforts the actual state of knowledge is still scant making impossible a<br />

full study of the process except for a short number of particular cases. 17<br />

Even when assuming that the system of reactions is known, it is usually too<br />

complicated to be fully inclu<strong>de</strong>d in the study of an Aerothermochemical problem. For<br />

this case the system must be simplified, for example with the steady state assumption,<br />

18 or else by adopting global reaction rates for the mixture such as<br />

w = Aρ n (1 − Y ) n e −E/RT , (1.141)<br />

where Y is the mass fraction of products and n is the molecularity of the overall<br />

reaction. In the following chapters frequent use will be ma<strong>de</strong> of similar expressions<br />

for the approximate study of several problems.<br />

References<br />

[1] Corner, J.: Theory of Interior Ballistics of Guns. John Wiley and Sons, Inc.<br />

New York, 1950.<br />

[2] Taylor, J.: Detonation in Con<strong>de</strong>nsed Explosives. Oxford, at the Clarendon<br />

Press, 1952.<br />

15 See chapter 6.<br />

16 See the Proceedings of the Congress, soon to be published.<br />

17 See chapter 6.<br />

18 See Ref. [28] where von Kármán and Penner have shown that the introduction of such an assumption<br />

for the oxygen atoms in the ozone flame is justified. Yet, for other cases it appears less justified. For<br />

example, in the Hydrogen–Bromine flame.


1.8. CHEMICAL KINETICS 35<br />

[3] Hirschfel<strong>de</strong>r, J. O., Curtiss, C. F. and Bird, R. B.: Molecular Theory of Gases<br />

and Liquids. John Wiley and Sons, Inc. New York, 1954.<br />

[4] Prigogine, I. and Defray, R.: Chemical Thermodynamics. Longmans, Green<br />

and Co., New York, 1954.<br />

[5] Guggenheim, E. A.: Thermodynamics. North Holland Publishing Comp. Amsterdam,<br />

1950.<br />

[6] Taylor, H. S. and Glastone: A Treatise on Physical Chemistry, Vol. II, States of<br />

Matter. D. Van Nostrand Comp. Inc., New York, 1951.<br />

[7] Lewis, G. N. and Randall, M.: Thermodynamics and the Free Energy of Chemical<br />

Substances. McGraw-Hill Book Co., New York.<br />

[8] National Bureau of Standards: Selected Values of Chemical Thermodynamic<br />

Properties. 1950.<br />

[9] Rossini, F. D.: Chemical Thermodynamics. John Wiley and Sons, Inc., New<br />

York, 1950.<br />

[10] The NBS-NACA Tables of Thermal Properties of Gases.<br />

[11] Mayer, J. E. and Mayer, M. G.: Statistical Mechanics. John Wiley and Sons,<br />

Inc., 1940.<br />

[12] Pitzer, K. S.: Quantum Chemistry. Prentice-Hall, Inc., New York, 1953.<br />

[13] Hirschfel<strong>de</strong>r, J. O., McClure, F. T., Curtiss, C. F. and Osborne, D. W.: Thermodynamic<br />

Properties of Propellant Gases. NDRC Rep. No. A-116, Nov. 23,<br />

1942.<br />

[14] Gay<strong>de</strong>n, A. G. and Welfhard H. G.: Flames. Their Structure Radiation and<br />

Temperature. Chapman and Hall Ltd., 1953.<br />

[15] Damköhler, G. and Edse, R.: Zeitschrift für Elektrochemi, Vol. 49, 1943, p. 178.<br />

[16] Huff, V. W. and Caluart, C. S.: Charts for the Computation of Equilibrium<br />

Composition of Chemical Reactions in the Carbon-Hydrogen-Oxygen-Nitrogen<br />

System at Temperatures from 2 000 to 5 000 K. NACA Technical Note No. 1653,<br />

July 1948.<br />

[17] Vichnievsky, R., Sale, B. and Marca<strong>de</strong>t, J.: Combustion Temperatures and Gas<br />

Composition. Jet Propulsion, Vol. 25, No. 3, March 1955, p. 105.<br />

[18] Huff, V. N., Gordon, S. and Morrell, V.: General Method and Thermodynamic<br />

Tables for Computation of Equilibrium Composition and Temperature of Chemical<br />

Reactions. NACA Technical Report No. 1037, 1951.<br />

[19] Brinialey, S. R.: Computational Methods in Combustion Calculations. Combustion<br />

Processes, Sec. C, Vol. II of High Speed Aerodynamics and Jet Propulsion,<br />

Princeton University Press, 1956, p. 64.


36 CHAPTER 1. THERMOCHEMISTRY<br />

[20] Lewis, B. and von Elbe, G.: Combustion, Flames and Explosions of Gases.<br />

Aca<strong>de</strong>mic Press., Inc. Publishers, New York, 1951.<br />

[21] Bundy, F. P. and Streng, H. M.: Measurement of Flame Temperature, Pressure<br />

and Velocity. Physical Measurement in Gas Dynamics and Combustion, Sec. I,<br />

Vol. IX of High Speed Aerodynamics and Jet Propulsion, Princeton University<br />

Press, 1954.<br />

[22] Penner, S. S.: Introduction to Chemical Kinetics. AGARD AG 5/p2, December<br />

1952.<br />

[23] Hinshelwood, C. N.: The Kinetics of Chemical Change. Oxford University<br />

Press, 1940.<br />

[24] Glastone, S., Laidler, K. J. and Eyring, H.: The Theory of Rate Processes.<br />

McGraw-Hill Book Company, Inc., New York, 1941.<br />

[25] Laidler, K. J.: Chemical Kinetics. McGraw-Hill Book Company, Inc., New<br />

York, 1950.<br />

[26] Taylor, H. S.: Fundamentals of Chemical Kinetics. Combustion Processes, Sec.<br />

D, Vol. II of High Speed Aerodynamics and Jet Propulsion, Princeton University<br />

Press, 1956, p. 101<br />

[27] Arrhenius, S.: Zeitschrift für Physikalische Chemic, 1889.<br />

[28] von Kármán, Th. and Penner, S. S.: Fundamental Approach to Laminar Flame<br />

Propagation. Selected Combustion Problems, Vol. I, AGARD, Paris, 1954,<br />

pp. 5-41.


Chapter 2<br />

Transport phenomena in gas<br />

mixtures<br />

2.1 Introduction<br />

When the state or composition of a mixture of gases are not uniform, a transport<br />

of mass, momentum and energy takes place between different points of the mixture<br />

tending to level the initial differences.<br />

At the neighborhood of each point transport <strong>de</strong>pends on the state and composition<br />

of the mixture and on the lack of uniformity from which it arises, according<br />

to laws <strong>de</strong>alt with in the present chapter. Even though these phenomena arise from<br />

molecular motion their <strong>de</strong>scription can be ma<strong>de</strong> through phenomenological variables<br />

except when the gas is very rarefied or when the variations in state and composition<br />

within distances comparable with the molecular mean free path are large. 1 If such<br />

cases are exclu<strong>de</strong>d, gas can be assimilated to a continuous medium where at each of<br />

its points <strong>de</strong>nsity, pressure, temperature, mass fractions and velocities of the species<br />

and mixture are <strong>de</strong>finable. Un<strong>de</strong>r such conditions, let P , Fig. 2.1, be a point, fixed or<br />

in motion at a given velocity ¯v, and dσ a surface element linked to it. Transport dF<br />

of mass of one of the species, or of momentum or energy of the mixture, through dσ<br />

during time dt is of the form<br />

dF = f dσ dt, (2.1)<br />

where f is the transport per unit surface and per unit time. f is a function of the orientation<br />

of dσ, <strong>de</strong>fined by unit vector ¯n normal to dσ, of the variables that <strong>de</strong>termine<br />

the state and composition of the mixture at the point and of their <strong>de</strong>rivatives. If P<br />

1 See chapter 3 §1.<br />

37


38 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

dσ<br />

v<br />

P<br />

n<br />

Figure 2.1: Schematic diagram to <strong>de</strong>fine the transport phenomena.<br />

moves at velocity ¯v of the mixture at the point 2 and mass transport of the species is<br />

consi<strong>de</strong>red, the phenomenon is called diffusion. By consi<strong>de</strong>ring the transport of momentum<br />

one obtains pressure and viscous stresses. Finally when energy transport is<br />

consi<strong>de</strong>red heat transfer is obtained.<br />

In the following paragraphs these three phenomena will be studied in sequence<br />

adopting as a starting point the expressions given for them by the Kinetic Theory of<br />

Gases. This theory states the problem as follows: let f i (¯v i , ¯x, t) be the velocity distribution<br />

function of species A i of the mixture. By <strong>de</strong>finition, the probable number<br />

of molecules of species A i whose velocities lie between ¯v i and ¯v i + d¯v i , and which<br />

are contained in volume element dV around point P of spatial coordinate ¯x, at instant<br />

t, is f i (¯v i , ¯x, t) dV dv i1 dv i2 dv i3 , where v i1 , v i2 and v i3 are the three components<br />

of ¯v i in a rectangular cartesian system of coordinates. Due to molecular collisions<br />

a statistical equilibrium establishes which <strong>de</strong>termines the values for f i . Such values<br />

are <strong>de</strong>termined by a system of Boltzmann integral equations. When the state of the<br />

mixture is uniform, the solution to this system is Maxwell’s velocity distribution for<br />

each species. Otherwise, the solution <strong>de</strong>pends not only on the state at the point but<br />

also on its rate of variation when passing from one point to another. This variation is<br />

<strong>de</strong>fined by the successive <strong>de</strong>rivatives of the variables of state and composition at the<br />

point. If such <strong>de</strong>rivatives are small, that is, when relative variations of <strong>de</strong>nsity, pressure,<br />

temperature, velocities and mass fractions at the point are small within distances<br />

of the or<strong>de</strong>r of the molecular mean free path, 3 an approximate solution to Boltzmann’s<br />

system can be obtained by means of the method of the perturbations <strong>de</strong>veloped by Enskog<br />

and Chapman. This method looks for solutions to Boltzmann’s system which<br />

differ slightly from Maxwell’s fundamental solution. Solutions show that transport of<br />

mass, momentum and energy <strong>de</strong>pend linearly on the first or<strong>de</strong>r <strong>de</strong>rivatives of the variables<br />

of state and composition at the point. Coefficients of such linear functions are<br />

those of diffusion, viscosity and thermal conductivity of the mixture <strong>de</strong>pending only<br />

on the state and composition at the point and on dynamics of molecular collisions. For<br />

a <strong>de</strong>tailed study on the subject see Refs. [1] and [2].<br />

2 See §2 of this chapter for <strong>de</strong>finition of ¯v.<br />

3 Such is normally the case with exceptions as, for example, insi<strong>de</strong> shock waves.


2.2. DIFFUSION 39<br />

Hereinafter, notation in chapter 1 will be kept, adopting the vectorial and tensorial<br />

notation <strong>de</strong>fined in the Appendix to chapter 3.<br />

2.2 Diffusion<br />

Let ¯v i be the velocity of species A i at P . 4 ¯v i is generally different from velocity ¯v of<br />

the mixture. The difference ¯v di is called diffusion velocity of A i<br />

¯v di = ¯v i − ¯v. (2.2)<br />

The following relations exist between ¯v, ¯v i and ¯v di<br />

¯v = ∑ i<br />

Y i¯v i , (2.3)<br />

∑<br />

Y i¯v di = ¯0. (2.4)<br />

i<br />

Flux dm i of A i through dσ during time dt, when P moves at velocity ¯v of the<br />

mixture, is obviously<br />

dm i = ρ i¯v di · ¯n dσ dt. (2.5)<br />

The problem lies in calculating flux vector ¯f i ,<br />

¯f i = ρ i¯v di . (2.6)<br />

Let us first consi<strong>de</strong>r the case of a binary mixture formed by species A 1 and A 2 .<br />

The problem has been solved by Enskog [3] and Chapman [4]. The flux of A 1 is<br />

¯f 1 = ρY 1¯v d1 = −ρ M (<br />

1M 2<br />

Mm<br />

2 D 12 ∇X 1 − (Y 1 − X 1 ) ∇p )<br />

∇T<br />

− D T 1<br />

p T . (2.7)<br />

Here D 12 and D T 1 are, respectively, the coefficients of diffusion and thermal diffusion<br />

5 of the mixture. Flux ¯f 2 of A 2 is, from (2.4) and (2.6),<br />

¯f 2 = − ¯f 1 . (2.8)<br />

4¯v i is <strong>de</strong>fined as follows: taking a volume element dV around P and forming the mean value of velocities<br />

¯v i , of the molecules A i on it. Such value is ¯v i<br />

¯v i =<br />

1 X<br />

¯v j i<br />

N ,<br />

i<br />

where N i is the number of molecules of A i on dV . If dV is large respect to the molecular scale and<br />

small respect to the macroscopic scale, the value ¯v i thus <strong>de</strong>fined is in<strong>de</strong>pen<strong>de</strong>nt from size and shape of dV .<br />

Velocity ¯v of the mixture is, by <strong>de</strong>finition<br />

¯v = X i<br />

j<br />

Y i¯v i .<br />

5 The coefficient of thermal diffusion must not be confused with the thermal diffusivity, Ed.


40 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

Formula (2.7) shows that there are three causes for diffusion corresponding to the three<br />

terms in the right hand si<strong>de</strong> of the equation. Namely, the differences in composition,<br />

pressure and temperature. 6<br />

first of these three causes. 7<br />

Elementary Theory of Diffusion only acknowledge the<br />

In Aerothermochemical problems it is advantageous to eliminate mole fraction<br />

X 1 from Eq. (2.7). Furthermore, by doing so a simplified expression is obtained.<br />

Elimination is immediately attained by keeping in mind the following relations 8<br />

Thus obtaining<br />

X 1 = M m<br />

Y 1 ,<br />

M 1<br />

(2.9)<br />

1<br />

= Y 1<br />

+ Y 2<br />

= 1 ( 1<br />

+ − 1 )<br />

Y 1 ,<br />

M m M 1 M 2 M 2 M 1 M 2<br />

(2.10)<br />

( )<br />

M1 − M 2<br />

¯f 1 = −ρD 12 ∇Y 1 + ρ D 12 Y 1 (1 − Y 1 ) ∇p<br />

M m p − D ∇T<br />

T 1<br />

T . (2.11)<br />

The first term in the right hand si<strong>de</strong> of this expression gives the diffusion flux corresponding<br />

to differences in composition. This term express Fick’s law. The second<br />

term gives the diffusion flux arising from differences in pressure. Since D 12 is always<br />

positive, formula (2.11) shows that pressure diffusion tends to carry the heavier<br />

molecules towards the higher pressures. The third term gives the diffusion flux arising<br />

from differences in temperature. Thermal diffusion coefficient D T 1 can be positive,<br />

negative or zero. Therefore, no general rules can be giving regarding the sens of this<br />

flux.<br />

Formula (2.11) shows that diffusion’s behaviour in a binary mixture of gases<br />

is <strong>de</strong>termined by the values of two coefficients: D 12 and D T 1 . Frequently D T 1 is<br />

substituted by<br />

k T =<br />

M 2 m<br />

M 1 M 2<br />

D T 1<br />

ρD 12<br />

. (2.12)<br />

Systematic measurements on diffusion coefficients of gas mixtures are not<br />

available. Therefore theoretical computations become necessary. They can be performed<br />

when knowing the interaction potential between molecules. If forces between<br />

molecules are in<strong>de</strong>pen<strong>de</strong>nt from their relative orientation, the interaction potential<br />

takes the form<br />

ϕ = ε 12 f<br />

( r<br />

σ 12<br />

)<br />

, (2.13)<br />

6 Where selective fields of forces exists, that is, fields acting with different strength over the two species<br />

of the mixture, they originate a new cause of diffusion. See Ref. [1], p. 143. Such occurs for example,<br />

when there exist ionized molecules and the mixture is submitted to the action of an electromagnetic field.<br />

7 See Ref. [5], p. 198.<br />

8 See chapter 1, pp. 12-13.


2.2. DIFFUSION 41<br />

where r is the distance between centers of the molecules, ε 12 is an intensity constant<br />

and σ 12 is a collision diameter. Consequently, once the form of f is known, the values<br />

of ε 12 and σ 12 <strong>de</strong>termine the potential. When forces between molecules <strong>de</strong>pend on<br />

their relative orientation, as occurs whit polar molecules, more complicated expressions<br />

than (2.13) are nee<strong>de</strong>d for ϕ. Such expressions must inclu<strong>de</strong> the influence of<br />

the relative orientation on the collision. An example of such potentials, which has<br />

been used by Hirschfel<strong>de</strong>r et al. for the computation of transport coefficients of polar<br />

molecules [6], is Stockmayer potential<br />

( (σ12 ) 12 ( σ12<br />

) ) 6<br />

ϕ = 4ε 12 − − µ2 12<br />

r r r 3 f (θ 1, θ 2 , φ 2 − φ 1 ) . (2.14)<br />

The first term in the right hand si<strong>de</strong> is Lennard-Jones’s potential and it represents<br />

the part of interaction in<strong>de</strong>pen<strong>de</strong>nt from the relative orientation of molecules.<br />

The second term gives the attraction between two dipoles whose relative orientation<br />

is <strong>de</strong>fined in Fig. 2.2. So far very few results are available on transport coefficients in<br />

mixtures of polar molecules. 9<br />

φ − φ<br />

1 2<br />

θ 1<br />

θ 2<br />

+<br />

Figure 2.2: Schematic diagram of the angles <strong>de</strong>fining the orientation of two polar<br />

molecules.<br />

The values for the coefficients of diffusion and thermal diffusion can be computed<br />

through a method of successive approximations <strong>de</strong>veloped by Chapman and<br />

Enskog.<br />

For this, the form of the interaction potential between molecules as well as its<br />

corresponding parameters must be known. Generally, a first approximation is enough<br />

for practical applications. For D 12 first approximation [D 12 ] 1<br />

, gives<br />

√<br />

[D 12 ] 1<br />

= 3<br />

( )<br />

8 √ kT M1 + M 2 1<br />

2π pσ12<br />

2 RT<br />

(2.15)<br />

M 1 M 2 Ω (1,1)∗<br />

12 (T12 ∗ ).<br />

Here T ∗ 12 is a dimensionless temperature <strong>de</strong>fined by the expression<br />

9 See Ref. [2], p. 597.<br />

+<br />

T ∗ 12 = kT<br />

ε 12<br />

, (2.16)


42 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

and Ω (1,1)∗<br />

12 is a function of T ∗ 12 whose form <strong>de</strong>pends on the potential of molecular<br />

interaction. For instance, if the molecules behave as rigid spheres Ω (1,1)∗<br />

12 = 1 for all<br />

values of T ∗ 12.<br />

For numerical computations it is more convenient to express [D 12 ] 1<br />

in the form<br />

[D 12 ] 1<br />

= 0.002628<br />

√<br />

M1 + M 2<br />

2M 1 M 2<br />

T 3<br />

(2.17)<br />

pσ12 2 Ω(1,1)∗ 12 (T12 ∗ ),<br />

where p is measured in atm, T in K, σ 12 in o A and [D 12 ] 1<br />

in cm 2 /s.<br />

The above mentioned theory does not <strong>de</strong>termine the interaction potential that<br />

should be applied in each case. To the contrary a comparison between the theoretical<br />

results corresponding to several potentials and experimental measurements, can<br />

furnish information on such potentials. In Ref. [2], pp. 589 and following, some examples<br />

of such comparisons can be found showing the agreement that can be expected<br />

between theoretical and experimental results. It is verified that for many cases the best<br />

agreement is attained with a Sutherland interaction potential or with that of Lennard-<br />

Jones. Sutherland’s potential is the one that produces between rigid spheres attracting<br />

one another with a force inversely proportional to a given power of its distance r.<br />

Therefore, it has the form<br />

This potential gives for Ω (1,1)∗<br />

12<br />

r < σ 12 : ϕ(r) = ∞,<br />

r > σ 12 : ϕ(r) = −ε 12<br />

( σ12<br />

r<br />

) δ<br />

.<br />

(2.18)<br />

Ω (1,1)∗<br />

12 = 1 + S 12<br />

T , (2.19)<br />

where S 12 = g(δ) ε 12<br />

is a constant whose value <strong>de</strong>pends on the parameters of Eq. (2.18)<br />

k<br />

and generally is <strong>de</strong>termined experimentally. Substituting Eq. (2.19) into (2.15) one<br />

obtains<br />

[D 12 ] 1<br />

= 3<br />

8 √ 2π<br />

k<br />

pσ 2 12<br />

√ ( ) 5<br />

M1 + M 2 T 2<br />

R<br />

M 1 M 2 S 12 + T . (2.20)<br />

If the interaction potential is of Lennard-Jones type, no simple expression can<br />

be given for Ω (1,1)∗<br />

12 , whose values must be calculated by numerical integration. Table<br />

2.1 gives the values calculated by Hirschfel<strong>de</strong>r et al. [7].


2.2. DIFFUSION 43<br />

T ∗ Ω (1,1)∗ Ω (2,2)∗ T ∗ Ω (1,1)∗ Ω (2,2)∗ T ∗ Ω (1,1)∗ Ω (2,2)∗<br />

0.30 2.662 2.785 1.70 1.140 1.248 4.20 0.874 0.9600<br />

0.35 2.476 2.628 1.75 1.128 1.234 4.30 0.8694 0.9553<br />

0.40 2.318 2.492 1.80 1.116 1.221 4.40 0.8652 0.9507<br />

0.45 2.184 2.368 1.85 1.105 1.209 4.50 0.8610 0.9464<br />

0.50 2.066 2.257 1.90 1.094 1.197 4.60 0.8568 0.9422<br />

0.55 1.966 2.156 1.95 1.084 1.186 4.70 0.8530 0.9382<br />

0.60 1.877 2.065 2.00 1.075 1.175 4.80 0.8492 0.9343<br />

0.65 1.798 1.982 2.10 1.057 1.156 4.90 0.8456 0.9305<br />

0.70 1.729 1.908 2.20 1.041 1.138 5.00 0.8422 0.9269<br />

0.75 1.667 1.841 2.30 1.026 1.122 6.00 0.8124 0.8963<br />

0.80 1.612 1.780 2.40 1.012 1.107 7.00 0.7896 0.8727<br />

0.85 1.562 1.725 2.50 0.9996 1.093 8.00 0.7712 0.8538<br />

0.90 1.517 1.675 2.60 0.9878 1.081 9.00 0.7556 0.8379<br />

0.95 1.476 1.629 2.70 0.9770 1.069 10.0 0.7424 0.8242<br />

1.00 1.439 1.587 2.80 0.9672 1.058 20.0 0.664 0.7432<br />

1.05 1.406 1.549 2.90 0.9576 1.048 30.0 0.6232 0.7005<br />

1.10 1.375 1.514 3.00 0.9490 1.039 40.0 0.596 0.6718<br />

1.15 1.346 1.482 3.10 0.9406 1.030 50.0 0.5756 0.6504<br />

1.20 1.320 1.452 3.20 0.9328 1.022 60.0 0.5596 0.6335<br />

1.25 1.296 1.424 3.30 0.9256 1.014 70.0 0.5464 0.6194<br />

1.30 1.273 1.399 3.40 0.9186 1.007 80.0 0.5352 0.6076<br />

1.35 1.253 1.375 3.50 0.9120 0.9999 90.0 0.5256 0.5973<br />

1.40 1.233 1.353 3.60 0.9058 0.9932 100. 0.5170 0.5882<br />

1.45 1.215 1.333 3.70 0.8998 0.9870 200. 0.4644 0.5320<br />

1.50 1.198 1.314 3.80 0.8942 0.9811 300. 0.436 0.5016<br />

1.55 1.182 1.296 3.90 0.8888 0.9755 400. 0.417 0.4811<br />

1.60 1.167 1.279 4.00 0.8836 0.9700<br />

1.65 1.153 1.264 4.10 0.8788 0.9649<br />

Table 2.1: Values of Ω (1,1)∗ and Ω (2,2)∗ as a function of T ∗ for the Lennard-Jones potential.<br />

When comparing theoretical values of the transport coefficients with those experimentally<br />

obtained, it is seen that in many cases Lennard-Jones potential gives the<br />

best agreement for a more extensive range of temperatures. 10<br />

The preceding formulae require knowing the values for ε 12 and σ 12 , which<br />

must be experimentally obtained. Generally, such values are known for potentials<br />

between equal molecules but not for those of different kind. The approximate values of<br />

the constants for the potentials between molecules of species A 1 and A 2 are obtainable<br />

as follows.<br />

10 See Ref. [2], p. 597.


44 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

Let (ε 1 , σ 1 ) and (ε 2 , σ 2 ) be the values of the constants for species A 1 and A 2 .<br />

The values of the constants for mixed potentials between both species are 11<br />

ε 12 = √ ε 1 ε 2 ,<br />

σ 12 = 1 2 (σ 1 + σ 2 ) .<br />

(2.21)<br />

Table 1.1 in chapter 1 gives the values of ε and σ for certain number of gases. If such<br />

values where not available any of the following formulas can be used for approximate<br />

values [7]:<br />

ε/k = 0.77T c , σ = 0.710(V c /N) 1 3 ,<br />

ε/k = 1.15T b , σ = 0.984(V b /N) 1 3 ,<br />

(2.22)<br />

ε/k = 1.92T m , σ = 1.031(V m /N) 1 3 ,<br />

ε/k = 0.292T B , σ = 1.030(V z /N) 1 3 .<br />

Here V is the molar volume. Subscripts c, m, b, B and z <strong>de</strong>note the critical, melting,<br />

boiling, Boyle’s 12 and absolute zero points.<br />

Table 2.2, taken from Ref. [2], gives the diffusion coefficients for some binary<br />

mixtures. Theoretical values correspond to a Lennard-Jones interaction potential.<br />

Constants for the potential have been obtained through Eq. (2.21). It is seen that<br />

agreement between theoretical and experimental values is, generally, excellent.<br />

Except for rigid spheres where Ω (1,1)∗<br />

12 is constant, this function <strong>de</strong>creases when<br />

T is increased. Hence, according to formula (2.15) [D 12 ] 1<br />

increases with T slightly<br />

faster than T 3/2 . Fig. 2.3 represents as an example the coefficient [D 12 ] 1<br />

versus T in<br />

a mixture of N 2 and CO 2 for rigid spheres and for an interaction potential of Lennard-<br />

Jones. In combustion problems, where temperature can change in a ratio of ten to one<br />

from one point to another, the influence of temperature on the values of the transport<br />

coefficients is very important. Formula (2.15) shows, as well, that diffusion coefficient<br />

is inversely proportional to the mixture’s pressure and is in<strong>de</strong>pen<strong>de</strong>nt from the mass<br />

fractions of the species in first approximation. In or<strong>de</strong>r to obtain the influence of mass<br />

fractions one must resort to higher approximations.<br />

11 See Ref. [2], pp. 589 and following<br />

12 Boyle’s temperature T B is the temperature at which curve pV vs. p has a horizontal tangent for p = 0.


2.2. DIFFUSION 45<br />

Gas Pair σ 12 ( A) o ε 12 /k (K) T (K) [D 12 ] 1<br />

(cm 2 s −1 )<br />

Calculated Experimental<br />

N 2 -H 2 3.325 55.2<br />

273.2<br />

288.2<br />

293.2<br />

0.656<br />

0.718<br />

0.739<br />

0.674<br />

0.743<br />

0.760<br />

N 2-O 2 3.557 102<br />

273.2 0.175 0.181<br />

293.2 0.199 0.220<br />

N 2 -CO 3.636 100 273.2 0.174 0.192<br />

N 2-CO 2 3.839 132<br />

273.2<br />

288.2<br />

293.2<br />

298.2<br />

0.130<br />

0.143<br />

0.147<br />

0.152<br />

0.144<br />

0.158<br />

0.160<br />

0.165<br />

N 2 -C 2 H 4 3.957 137 298.2 0.156 0.163<br />

N 2 -C 2 H 6 4.050 145 298.2 0.144 0.148<br />

N 2-nC 4H 10 4.339 194 298.2 0.0986 0.0960<br />

N 2-iso-C 4H 10 4.511 169 298.2 0.0970 0.0908<br />

H 2 -O 2 3.201 61.4 273.2 0.689 0.697<br />

H 2 -CO 3.279 60.6 273.2 0.661 0.651<br />

273.2 0.544 0.550<br />

H 2-CO 2 3.482 79.5<br />

288.2 0.597 0.619<br />

293.2 0.616 0.600<br />

298.2 0.634 0.646<br />

H 2 -N 2 O 3.424 85.6 273.2 0.552 0.535<br />

H 2 -SF 6 3.922 89.1 298.2 0.473 0.420<br />

H 2-CH 4 3.425 67.4<br />

273.2 0.607 0.625<br />

298.2 0.705 0.726<br />

H 2 -C 2 H 4 3.600 82.6 298.2 0.595 0.602<br />

H 2 -C 2 H 6 3.693 87.5 298.2 0.556 0.537<br />

CO-O 2 3.512 112 273.2 0.175 0.185<br />

CO 2 -O 2 3.715 147<br />

273.2 0.128 0.139<br />

293.2 0.146 0.160<br />

CO 2 -CO 3.793 145 273.2 0.128 0.137<br />

CO 2 -N 2 O 3.938 204 273.2 0.092 0.096<br />

CO 2-CH 4 3.939 161 273.2 0.138 0.153<br />

Table 2.2: Comparison of diffusion coefficients for some binary mixtures. Theoretical values<br />

correspond to Lennard-Jones interaction potential.<br />

Approximation of or<strong>de</strong>r j can be expressed in the form 13<br />

[D 12 ] j<br />

= [D 12 ] 1<br />

f j D 12<br />

. (2.23)<br />

For example, for the second approximation, fD 2 12<br />

<strong>de</strong>pends on the molar masses<br />

of the species, on their mass fractions, viscosity coefficients and temperature of the<br />

mixture. Its value is always close to one. For the Lennard-Jones potential, for instance,<br />

it ranges from 1 to 1.03.<br />

13 See Ref. [2], p. 539.


46 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

[D 12<br />

] 1<br />

2.0<br />

1.5<br />

1.0<br />

Rigid spheres<br />

Lennard−Jones potential<br />

0.5<br />

0.0<br />

0 200 400 600 800 1000 1200 1400 1600 1800 2000<br />

T(K)<br />

Figure 2.3: Coefficient [D 12] 1<br />

as a function of temperature in a mixture of N 2 and CO 2, for<br />

rigid spheres and for an interaction potential of Lennard-Jones.<br />

Hence, the influence of mass fractions on the value of D 12 is always small and<br />

can, generally, be neglected by adopting for it the value given by the first approximation.<br />

Frequently, the value given by elementary diffusion theory is approximate<br />

enough. According to it,<br />

D 12 ∼ T 3/2<br />

p . (2.24)<br />

Or else, an empirical formula of the form<br />

D 12 ∼ T n<br />

p , (2.25)<br />

where n is an exponent ranging from 1.5 to 2 whose value is <strong>de</strong>termined empirically.<br />

With respect to thermal diffusion ratio k T , even in the first approximation it is<br />

a complicated function of the molar masses, mass fractions of the species and of the<br />

temperature of the mixture. It is also very much influenced by the laws of molecular<br />

interaction. The expression of the first approximation of k T can be obtained from<br />

Ref. [2], p. 541, where tables for the calculation of their values are also inclu<strong>de</strong>d as<br />

well as a comparison between theoretical and experimental values for many binary<br />

mixtures. It appears that the agreement is not so good for k T as for the other transport<br />

coefficients. When comparing the successive approximations to the value of k T , it is<br />

also verified that in first approximation the error is larger here than for other transport


2.2. DIFFUSION 47<br />

coefficients. 14 The values of k T are generally small. Normally they do not exceed 0.1.<br />

Thereby thermal diffusion is of secondary importance in Aerothermochemistry and its<br />

influence is, usually, disregar<strong>de</strong>d.<br />

In Refs. [8], [9], [10] and [11] recent information on theoretical and experimental<br />

values of transport coefficients can be found.<br />

For mixtures containing more than two components, the problem has been<br />

solved by Curtiss and Hirschfel<strong>de</strong>r [2], [12]. In such case, the aforementioned three<br />

causes of diffusion, namely, differences in composition, pressure and temperature, still<br />

hold. Yet in this case diffusion flux of each species <strong>de</strong>pends not only on the gradient<br />

of its molar fraction but is a linear function of the gradients of all molar fractions.<br />

Diffusion flux ¯f i of species A i (i = 1, 2, . . . , l) is<br />

¯f i = ρY i¯v di = ρ M i<br />

M 2 m<br />

∑<br />

M j D ij<br />

′<br />

j≠i<br />

(<br />

∇X j + (X j − Y j ) ∇p<br />

p<br />

)<br />

− D T i<br />

∇T<br />

T . (2.26)<br />

Here D ′ ij is the diffusion coefficient of species A i and A j in the mixture. For a binary<br />

mixture, D ′ ij is diffusion coefficient D ij previously <strong>de</strong>fined. For the case of more than<br />

two components it differs. D T i is the thermal diffusion coefficient of species A i in<br />

the mixture.<br />

The l equations (2.26) are not in<strong>de</strong>pen<strong>de</strong>nt from each other. In fact, diffusion<br />

fluxes of l species must satisfy the condition<br />

∑<br />

¯f i = ¯0. (2.27)<br />

i<br />

Diffusion coefficients D ij ′ in system (2.26) can be expressed as a function of<br />

binary diffusion coefficients D ij of the species taken in couples. Ref. [2], pp. 541<br />

and 543, and Ref. [12] contain expressions for these coefficients as well as for those<br />

of thermal diffusion. Expressions for D ′ ij and D T i are too complicated for practical<br />

computations. Then it is advantageous to eliminate D ij ′ by substituting system (2.26)<br />

with the following that makes explicit the influence of binary diffusion coefficients 15<br />

∑ X i X j<br />

(¯v dj − ¯v di ) =∇X i + (X i − Y i ) ∇p<br />

D ij<br />

p<br />

j≠i<br />

− ∑ (<br />

X i X j DT j<br />

− D ) (2.28)<br />

T i ∇T<br />

ρD ij Y j Y i T . j≠i<br />

As was done for the case of binary diffusion, it is also convenient to eliminate<br />

molar fractions of the species by introducing their corresponding mass fractions.<br />

14 A critical study on the subject will be found in Ref. [6], p. 492 and following, as well as bibliography.<br />

15 See Ref. [2]. p. 517.


48 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

Making use of relations (1.33) and (1.36) in chapter 1, it is obtained<br />

∑<br />

Y j<br />

D ij<br />

j≠i<br />

M j<br />

[<br />

¯vdi − ¯v dj<br />

+ ∇Y i<br />

− 1 ρ<br />

− ∇Y j<br />

Y j<br />

Y i<br />

(<br />

DT j<br />

Y j<br />

+ M j − M i ∇p<br />

M m p<br />

]<br />

= ¯0, (i = 1, 2, . . . , l).<br />

− D T i<br />

Y i<br />

) ∇T<br />

T<br />

(2.29)<br />

Diffusion of a species whose mass fraction is very small is particularly important<br />

in certain problems. For example, diffusion of active particles which propagate<br />

chemical reactions through a combustion wave. Then it becomes possible to give an<br />

explicit approximate expression for the diffusion flux of such species. For diffusion<br />

arising from differences in composition this expression is<br />

(<br />

∑ Y j ∇Yj<br />

− ∇Y )<br />

i<br />

j≠iM j Y j Y i<br />

¯f i = ρY i¯v di = ρY i<br />

∑ Y j<br />

. (2.30)<br />

M j D ij<br />

j≠i<br />

System (2.29) is, generally, too complicated to be used in the solution of many<br />

of the Aerothermochemistry problems. Therefore, in some of its applications, diffusion<br />

fluxes of the species are substituted by approximate expressions un<strong>de</strong>r the assumption<br />

that diffusion flux due to differences in composition ¯f iY of each species is<br />

proportional to its gradient of molar fraction. That is, by adopting for such flux an<br />

expression of the form<br />

¯f iY = −ρ M i<br />

M m<br />

D im ∇X i . (2.31)<br />

Here D im is a binary diffusion coefficient of the species through the mixture formed<br />

by the rest. This expression is exact only when binary diffusion coefficients and molar<br />

masses of the species are all equal. If mean molar mass of the mixture is also constant,<br />

one has<br />

¯f iY = −ρD im ∇Y i , (2.32)<br />

by virtue of Eqs. (1.33) and (1.36) in chapter 1. This expressions will be used in<br />

chapter 13. In aerothermochemical problems diffusion fluxes arising from differences<br />

in pressure and temperature are, in general, negligible. C. R. Wilke [14] has given an<br />

explicit approximate expression, empirically <strong>de</strong>duced, similar to (2.31) for fluxes due<br />

to differences in composition.


2.3. VISCOSITIES 49<br />

2.3 Viscosities<br />

Mechanics of continuous media teaches that the state of stresses at each point of a<br />

medium is <strong>de</strong>termined by a symmetric tensor of second or<strong>de</strong>r<br />

⎛<br />

⎞<br />

p 11 p 12 p 13<br />

⎜<br />

⎟<br />

τ e ≡ ⎝ p 21 p 22 p 23 ⎠ , (2.33)<br />

p 31 p 32 p 33<br />

of components p ij = p ji in a rectangular cartesian system of reference. 16 The physical<br />

meaning of this tensor is the following: p ij is the component parallel to axis x j of the<br />

stress acting upon a surface element normal to axis x i , see Fig. 2.4.<br />

x<br />

p<br />

12<br />

p<br />

13<br />

p<br />

11<br />

3<br />

p<br />

23<br />

p<br />

p 22<br />

21<br />

p 1<br />

33<br />

p31<br />

x 2 p 32<br />

Figure 2.4: Schematic diagram showing the components of the stress tensor τ e.<br />

x<br />

Stress ¯f acting upon a surface element dσ, see Fig. 2.5, whose orientation is<br />

<strong>de</strong>fined by the unit vector ¯n, is given by the expression 17<br />

¯f = ¯n · τ e , (2.34)<br />

whose components are<br />

f i = ∑ j<br />

n j p ji , (i = 1, 2, 3), (2.35)<br />

where n j are the components of ¯n.<br />

Kinetic Theory of gases shows 18 that in a dilute gas the stress originates from<br />

transfer of momentum of the gas through a surface element dσ moving at velocity ¯v of<br />

16 See, i.e., Ref. [15].<br />

17 See appendix to chapter 3 for notation.<br />

18 See Ref. [1], p. 31.


50 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

d σ<br />

x<br />

3<br />

P<br />

f<br />

n<br />

f<br />

3<br />

f 1<br />

x<br />

x 2<br />

1<br />

f 2<br />

Figure 2.5: Schematic diagram showing the components of the stress ¯f = ¯n · τ e.<br />

the gas. 19 Such transfer is produced by the thermal agitation of the molecules. 20 If the<br />

state of gas is uniform, the velocity distribution of thermal agitation is a Maxwell’s<br />

distribution. In such case transfer of momentum is normal to dσ and in<strong>de</strong>pen<strong>de</strong>nt<br />

from the orientation ¯n. The corresponding state of stresses is a pure compression,<br />

<strong>de</strong>fined by a scalar: the pressure p of the gas. Tensor of Eq. (2.33) reduces as follows<br />

τ e = −pU, (2.36)<br />

where U is the unit tensor whose components are Kronecker’s δ ij<br />

U ≡<br />

⎛<br />

⎜<br />

⎝<br />

1 0 0<br />

0 1 0<br />

0 0 1<br />

Stress ¯f acting upon an element dσ in Fig. 2.5, is<br />

⎞<br />

⎟<br />

⎠ . (2.37)<br />

¯f = −p¯n. (2.38)<br />

When the state of motion of the gas is not uniform other stresses produce, which add<br />

to those of pressure, Eq. (2.38). Such stresses are given by viscous stress tensor<br />

⎛<br />

⎞<br />

τ 11 τ 12 τ 13<br />

⎜<br />

⎟<br />

τ ev ≡ ⎝ τ 21 τ 22 τ 23 ⎠ . (2.39)<br />

τ 31 τ 32 τ 33<br />

The Chapman-Enskog method <strong>de</strong>scribed in the introduction to the present chapter<br />

shows that components τ ij of the viscous stress tensor have the form 21<br />

( ) 2<br />

τ ij = 2µγ ij −<br />

3 µ − µ′ (∇ · ¯v) δ ij . (2.40)<br />

19 Within <strong>de</strong>nse gases stresses are partially due to transfer of momentum and partially to the forces of<br />

molecular interaction whose effect is not negligible.<br />

20 If ¯v is the gas velocity at a point and ¯v j the velocity of a molecule at the neighborhood of such point,<br />

velocity ¯v a of thermal agitation is, by <strong>de</strong>finition, ¯v a = ¯v j − ¯v.<br />

21 Actually Enskog and Chapman’s theory was <strong>de</strong>veloped for monotonic dilute gases, where µ ′ = 0.


2.3. VISCOSITIES 51<br />

Here µ and µ ′ are called viscosity coefficient and bulk viscosity coefficient respectively,<br />

and<br />

3<br />

8π √ 2 γ ij = 1 ( ∂vi<br />

+ ∂v )<br />

j<br />

2 ∂x j ∂x i<br />

are the components of <strong>de</strong>formation velocity tensor<br />

τ vd ≡<br />

⎛<br />

⎜<br />

⎝<br />

(2.41)<br />

⎞<br />

γ 11 γ 12 γ 13<br />

⎟<br />

γ 21 γ 22 γ 23 ⎠ , (2.42)<br />

γ 31 γ 32 γ 33<br />

which <strong>de</strong>termines the <strong>de</strong>formation suffered by a gas element as it moves. 22<br />

The method of Chapman and Enskog allows calculation of the viscosity coefficient<br />

of a gas by means of successive approximations. Expressions of the form<br />

µ j = [µ] 1<br />

f j µ (2.43)<br />

are obtained, where [µ] 1<br />

is the first approximation, j is the computed or<strong>de</strong>r of the approximation<br />

and f j µ is a function which differs slightly from unity. When comparing,<br />

for example, the first and third approximations for the Lennard-Jones potential it is<br />

verified that error in the first is smaller than 0.8%. Thereby, as for diffusion coefficients,<br />

the first approximation is usually enough.<br />

In the case of a pure gas this approximation is given by 23<br />

[µ] 1<br />

= 5<br />

16 √ N −1√ MRT<br />

π σ 2 Ω (2,2)∗ (T ∗ ) . (2.44)<br />

Here, as for diffusion, Ω (2,2)∗ is a function of reduced temperature T ∗ = kT/ε, whose<br />

law of variation <strong>de</strong>pends on the form of the molecular interaction potential. For example,<br />

for rigid spheres Ω (2,2)∗ = 1.<br />

For a Sutherland potential, Eq. (2.18), is<br />

Ω (2,2)∗ = 1 + S µ<br />

T , (2.45)<br />

where S µ = g µ (δ) ε . In practice, this value is <strong>de</strong>termined experimentally. When<br />

k<br />

Eq. (2.45) is substituted into Eq. (2.44) one obtains<br />

22 See Ref. [15].<br />

23 See Ref. [2], p. 527.<br />

[µ] 1<br />

= 5<br />

16 √ π<br />

N −1√ MR<br />

σ 2 T 3 2<br />

S µ + T . (2.46)


52 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

Values Ω (2,2)∗ as function of T ∗ for the Lennard-Jones potential have been<br />

taken into Table 2.1. It is seen that for many gases this potential represents experimental<br />

results better than others. 24<br />

For numerical computations it is more convenient to express Eq. (2.44) as<br />

√<br />

MT<br />

10 7 [µ] 1<br />

= 266.93<br />

σ 2 Ω (2,2)∗ (T ∗ ) , (2.47)<br />

where all quantities are measured in the same units than in Eq. (2.17) and [µ] 1<br />

is<br />

expressed in gr cm −1 s −1 .<br />

The preceding formulae show that the viscosity coefficient of a gas is in<strong>de</strong>pen<strong>de</strong>nt<br />

from pressure and increases with temperature slightly over √ T . Fig. 2.6 gives<br />

the law of variation with temperature of viscosity coefficients for several gases. For<br />

additional information, see Refs. [2], [8], [9], [11] and [16], where complementary<br />

bibliography is listed.<br />

7000<br />

6000<br />

O 2<br />

N 2<br />

5000<br />

10 7 µ (g cm −1 s −1 )<br />

4000<br />

3000<br />

2000<br />

CO 2<br />

CH 4<br />

H 2<br />

1000<br />

0<br />

0 200 400 600 800 1000 1200 1400 1600<br />

T(K)<br />

Figure 2.6: Viscosity coefficient as function of temperature, for several gases.<br />

Often, for practical applications an empirical formula of the form<br />

µ ∼ T n (2.48)<br />

is used, where n is an exponent ranging from 0.5 to 1.<br />

With respect to bulk viscosity coefficient µ ′ it is zero for dilute monatomic<br />

gases. For polyatomic and <strong>de</strong>nse gases µ ′ is generally different from zero, yet small.<br />

24 See Ref. [2], pp. 589 and following.


2.3. VISCOSITIES 53<br />

Its value <strong>de</strong>pends on the easiness of energy transfer between external and internal<br />

<strong>de</strong>grees of freedom of the molecules during collisions. Such facility is characterized<br />

by a relaxation time for each internal <strong>de</strong>gree of freedom. The value for µ ′ is expressed<br />

as a function of these relaxation times. Generally, it is sufficiently approximate to<br />

assume µ ′ = 0. Additional information on this problem can be found in Ref. [2],<br />

pp. 501 and 710.<br />

Viscosity coefficient of a mixture of gases is a complicated function of molar<br />

masses, mass fractions and viscosity coefficients of the species as well as of the interaction<br />

potentials between different species. 25 For binary mixture, for example, the<br />

first approximation [µ] 1<br />

is given by<br />

where<br />

1<br />

[µ] 1<br />

= X µ + Y µ<br />

1 + Z µ<br />

, (2.49)<br />

X µ = X2 1<br />

+ 2X 1X 2<br />

+ X2 2<br />

,<br />

µ 1 µ 12 µ<br />

[<br />

2<br />

]<br />

Y µ = 3 X 2<br />

5 A∗ 1 M 1<br />

12 + 2X 1X 2 (M 1 + M 2 ) 2 µ 2 12<br />

+ X2 2 M 2<br />

,<br />

µ 1 M 2 µ 12 4M 1 M 2 µ 1 µ 2 µ 2 M 1<br />

[ X<br />

2<br />

1 M 1<br />

(2.50)<br />

Z µ = 3 5 A∗ 12<br />

M 2<br />

(<br />

(M 1 + M 2 ) 2 (<br />

µ12<br />

+ 2X 1 X 2 + µ ) )<br />

12<br />

− 1<br />

4M 1 M 2 µ 1 µ 2<br />

]<br />

+ X2 2 M 2<br />

.<br />

M 1<br />

In these expressions, µ 1 and µ 2 are the first approximations of the viscosity coefficients<br />

for both species. µ 12 is given by<br />

( )<br />

µ 12 = 5<br />

N −1 M1 M 2<br />

√2RT<br />

16 √ M 1 + M 2<br />

π σ12 2 Ω(2,2)∗ 12 (T12 ∗ ) , (2.51)<br />

ε 12 and σ 12 are the constants of the interaction potential between molecules of both<br />

species. A ∗ is a function of reduced temperature T ∗ 12 = kT<br />

ε 12<br />

, whose value is given in<br />

Table 2.3 for the Lennard-Jones potential.<br />

25 See Ref. [2], p. 529.<br />

T12 ∗ 0.3 0.5 0.75 1 1.25 1.5 2<br />

A ∗ 1.046 1.093 1.105 1.103 1.099 1.097 1.094<br />

T12 ∗ 3 4 5 10 50 100 400<br />

A ∗ 1.095 1.098 1.101 1.11 1.13 1.138 1.154<br />

Table 2.3: The coefficient A ∗ as a function of T ∗ 12.


54 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

The following Table 2.4 gives, as an example, the values of the theoretical and<br />

experimental viscosity coefficients for a mixture of oxygen and carbon monoxi<strong>de</strong>.<br />

%O<br />

T (K) 0 42.01 77.33<br />

300.06<br />

Experimental 1 776 1 900 1 998<br />

Calculated 1 771 1 896 2 003<br />

500.06<br />

Experimental 2 548 2 741 2 908<br />

Calculated 2 539 2 717 2 871<br />

Table 2.4: Viscosity coefficient (10 7 µ gr cm −1 s −1 ) for a mixture of O 2 -CO.<br />

For mixtures of more than two components Bud<strong>de</strong>mberg and Wilke give the<br />

following semi-empirical expression which represents experimental results with good<br />

approximation<br />

[µ] 1<br />

= ∑ i<br />

X 2 1<br />

Xi<br />

2 + 1.385 ∑ X i X j<br />

[µ i ] 1 j≠i<br />

. (2.52)<br />

RT<br />

pM i [D ij ] 1<br />

This expression allows calculation of the viscosity coefficient of a mixture when those<br />

corresponding to the species are known as well as the diffusion coefficients of the<br />

same when taken by couples.<br />

In the preceding formulae substitution of molar fractions by mass fractions is<br />

straightforward by making use of relation<br />

X i = M m<br />

M i<br />

Y i . (2.53)<br />

Fig. 2.7 gives the law of variation as a function of the composition of the viscosity<br />

coefficients for some gas mixtures.<br />

2.4 Thermal conductivity<br />

Theory of heat conduction shows 26 that heat transfer at a point of a material medium<br />

is <strong>de</strong>fined by a vector ¯q of heat flux<br />

¯q = q 1 ī 1 + q 2 ī 2 + q 3 ī 3 . (2.54)<br />

Physical meaning of ¯q is the following: q i is heat flux per unit surface and per unit<br />

time through a surface element normal to axis x i .<br />

26 See, i.e., Ref. [16], pp. 3 and following.


2.4. THERMAL CONDUCTIVITY 55<br />

2000<br />

1800<br />

10 7 µ (gr cm −1 s −1 )<br />

1600<br />

1400<br />

1200<br />

H 2<br />

− CO<br />

H 2<br />

− N 2<br />

H 2<br />

− CO 2<br />

H 2<br />

− N 2<br />

O<br />

H 2<br />

− O 2<br />

1000<br />

800<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

X<br />

H2<br />

Figure 2.7: Viscosity for some gas mixture as function of composition.<br />

Consi<strong>de</strong>ring a surface element dσ, see Fig. 2.5, not parallel to any of the coordinate<br />

planes and whose orientation is <strong>de</strong>fined by unit vector ¯n, it is easily <strong>de</strong>monstrated<br />

that heat flux Q through dσ per unit surface and per unit time is<br />

Q = ¯q · ¯n = ∑ i<br />

q i n i . (2.55)<br />

Therefore, once the flux vector at a point is known, this <strong>de</strong>termines heat transfer<br />

through any surface element passing through it.<br />

Kinetic Theory of Gases shows that within a dilute gas heat flux originates only<br />

from molecular energy transfer when dσ moves at a velocity ¯v of the gas at the point.<br />

Such transfer originates from the motion of molecular agitation and from diffusion<br />

between species forming the gas if it is a mixture. When such transfer is calculated<br />

one obtain the expression 27<br />

¯q = −λ∇T + ρ ∑ i<br />

Y i h i¯v di + M m RT ∑ i<br />

∑<br />

j<br />

Y j D T i<br />

M i M j D ij<br />

(¯v di − ¯v dj ) , (2.56)<br />

where λ is the thermal conductivity of the gas and h i is the total specific enthalpy<br />

of species A i . Since diffusion fluxes of the species are produced by differences in<br />

composition, pressure and temperature, it results according to (2.56) that this three<br />

causes also produce heat transfer. In general, contributions from the third term in the<br />

27 See Ref. [2], p. 498.


56 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

right hand si<strong>de</strong> of (2.56) is small compared to the other two and can be neglected.<br />

Then ¯q reduces to<br />

¯q = −λ∇T + ρ ∑ Y i h i¯v di . (2.57)<br />

i<br />

In a pure gas ¯q reduces to<br />

¯q = −λ∇T. (2.58)<br />

If, moreover, the gas is monatomic, thermal conductivity coefficient can be<br />

computed through successive approximations as done for diffusion and viscosity. The<br />

first approximation [λ] 1<br />

is<br />

√<br />

T/M<br />

10 7 [λ] = 1989.1<br />

σ 2 Ω (2,2)∗ (T ∗ ) . (2.59)<br />

Here [λ] is measured in cal cm −1 s −1 K −1 and the other magnitu<strong>de</strong>s in the same units<br />

as in Eq. (2.17).<br />

When comparing (2.59) and (2.47) it is seen that for monatomic gases thermal<br />

conductivity and viscosity coefficients are proportional. Similarly to what happens for<br />

other transport coefficients the first approximation is, generally, sufficient for practical<br />

applications.<br />

Expression (2.59) does not take into account for polyatomic molecules energy<br />

transfer between internal and external <strong>de</strong>grees of freedom during collisions. Such<br />

transfer is not important in the study of viscosity and diffusion but is fundamental in<br />

thermal conductivity. 28 In polyatomic molecules [λ] must be substituted by the following<br />

approximate expressions, due to Eucken, which takes into account the influence<br />

of such energy transfer<br />

[λ] = 15<br />

4<br />

(<br />

R 4<br />

M [µ] 1<br />

15<br />

C v<br />

R + 3 )<br />

, (2.60)<br />

5<br />

where C v is molar heat of the gas at constant volume. Table 2.5 compares experimental<br />

values with those given by Eucken’s formula (2.60) for several gases. 29<br />

H 2 O 2 CO 2 CH 4 N 2<br />

Calculated 4 140 615 386 741 619<br />

Experimental 4 227 635 398 819 619<br />

Table 2.5: Values of the thermal conductivity coefficient (10 7 [λ] cal cm −1 s −1 K −1 ) for<br />

several gases at T=300 K .<br />

It is seen that agreement is fair even if not as good as for other transport coefficients.<br />

28 See Ref. [2], p. 489.<br />

29 See Ref. [2], p. 574.


2.4. THERMAL CONDUCTIVITY 57<br />

4000<br />

3500<br />

10 7 λ (cal cm −1 s −1 K −1 )<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

H 2<br />

− O 2<br />

H 2<br />

− N 2<br />

H 2<br />

− CO<br />

H 2<br />

− N 2<br />

O<br />

H 2<br />

− CO 2<br />

0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

X<br />

H2<br />

Figure 2.8: Thermal conductivity for some gas mixture as function of composition.<br />

In Ref. [2], pp. 526 and following, theoretical formulae are given for the calculation<br />

of thermal conductivity coefficients in mixtures of monatomic gases as well<br />

as empirical formulae for polyatomic gases. Here agreement between theoretical and<br />

experimental results is also not so good as for other transport coefficients. Fig. 2.8<br />

gives, as an example, the influence of the composition on the value of λ for certain<br />

binary mixtures at temperature 273.16 K.<br />

Finally, it should be noted that Aerothermochemistry <strong>de</strong>velopment is not yet<br />

sufficiently advanced so that transport coefficients are nee<strong>de</strong>d to be exact in most of its<br />

applications. Frequently a rudimentary approximation is enough, usually empirical,<br />

where the influence of composition is neglected and that of temperature is approximately<br />

taken into account. These approximations appear justified consi<strong>de</strong>ring that<br />

other features of the problems must be accounted for in a very rudimentary way, such<br />

as frequently are chemical kinetics, and, in some cases, velocity field and shape of the<br />

flame, etc.<br />

References<br />

[1] Chapman, S. and Cowling, T. G.: The Mathematical Theory on Non-Uniform<br />

Gases. Cambridge University Press, 1939.<br />

[2] Hirschfel<strong>de</strong>r, J. O., Curtiss, C. F., and Bird, R.B.: Molecular Theory of Gases<br />

and Liquids. John Wiley and Sons, Inc., New York, 1954.


58 CHAPTER 2. TRANSPORT PHENOMENA IN GAS MIXTURES<br />

[3] Enskog, D.: Kinetische Theorie <strong>de</strong>r Vorgänge in massig verdünnten Gasen. Dissertation,<br />

Upsala, 1917.<br />

[4] Chapman, S.: On the Kinetic Theory of a Gas, Part II, A Composite Monatomic<br />

Gas: Diffusion, Viscosity and Thermal Conduction. Philosophical Transactions<br />

of the Royal Society of London, A-217, 1917, p. 115.<br />

[5] Jeans, J.: An Introduction to the Kinetic Theory of Gases. Cambridge University<br />

Press, 1948.<br />

[6] Jost, W.: Diffusion in Solids, Liquids and Gases. Aca<strong>de</strong>mic Press, Inc., Publishers,<br />

New York, 1952.<br />

[7] Bird, R. B., Hirschfel<strong>de</strong>r, J. O. and Curtiss, C. F.: The Theoretical Calculation<br />

of the Equation of State and Transport Properties of Gases and Liquids. The<br />

American Society of Mechanical Engineers, Heat Transfer Division, Annual<br />

Meeting, New York. 29 Nov. to 4 Dec., 1953.<br />

[8] Johnson, E. F.: Molecular Transport Properties of Fluids. Industrial and Engineering<br />

Chemistry, Vol. 45, 1953, pp. 902-6.<br />

[9] Baron, T. and Oppenheim, A. K.: Fluid Dynamics. Industrial and Engineering<br />

Chemistry, Vol. 45, 1953, pp. 941-51.<br />

[10] Pigford, R. L.: Mass Transfer. Industrial and Engineering Chemistry, Vol. 45,<br />

1953, pp. 951-56.<br />

[11] Hirschfel<strong>de</strong>r, J. O., Bird, R. B. and Sportz, E. L.: Viscosity and Other Physical<br />

Properties of Gases and Gas Mixtures. ASME, 1949, pp. 921-37.<br />

[12] Hirschfel<strong>de</strong>r, J. O., Bird, R. B. and Sportz, E. L.: The Transport Properties<br />

for non Polar Gases. 1.- Journal of Chemical Physics, 1948, pp. 968-81.<br />

2.- Chemical Review, 1949, pp. 205-31.<br />

[13] Curtiss, C. F. and Hirschfel<strong>de</strong>r, J. O.: Transport Properties of Multicomponent<br />

Gas Mixtures. The Journal of Chemical Physics, Vol. 17, June 1949, pp. 550-<br />

55.<br />

[14] Fairbanks, D. F. and Wilke, C. R.: Diffusion Coefficients in Multicomponent<br />

Gas Mixtures. Industrial and Engineering Chemistry, March 1950, pp. 471-75.<br />

[15] Sommerfeld, A.: Mechanics of Deformable Bodies. Aca<strong>de</strong>mic Press Inc., Publishers,<br />

1950.<br />

[16] Carslaw, H. S. and Jaeger, J. C.: Conduction of Heat in Solids. Oxford at the<br />

Clarendon Press, 1950.


Chapter 3<br />

General equations<br />

3.1 Introduction<br />

In the present chapter we shall establish the general equations of Aerothermochemistry,<br />

that is, the general equations of transformation of a mixture of gases that react<br />

one to another and are in motion. It is assumed herein that the gas can assimilate to<br />

a continuous medium as previously done in the preceding chapter when analyzing the<br />

transport coefficients. This assumption is well justified except in the following cases:<br />

a) When a characteristic length of the macroscopic scale of the process, for example<br />

the size of an obstacle submerged in the gas, is of the or<strong>de</strong>r of magnitu<strong>de</strong> of the<br />

molecular mean free path.<br />

b) In the case where the phenomenological variables suffer consi<strong>de</strong>rable variations<br />

within a distance of the or<strong>de</strong>r of magnitu<strong>de</strong> of the said mean free path.<br />

The first case occurs when <strong>de</strong>aling with a very rarefied gas (Knudsen gas). The<br />

second case occurs, for example, in the analysis of the structure of a shock wave. In<br />

fact, the thickness of a sock wave is only a few times larger than the molecular free<br />

path, except when the wave is very weak. In such cases, to speak of the macroscopic<br />

functions of a point has no sense, and it is necessary to take into consi<strong>de</strong>ration the<br />

molecular structure of the gas. Such cases are exclu<strong>de</strong>d from the present study.<br />

The equations that govern the transformations of a mixture of reacting gases are<br />

similar to the general Gas Dynamics equations. Their difference is due to the fact that<br />

while Gas Dynamics <strong>de</strong>al with gases of homogeneous chemical composition invariable<br />

with time, in Aerothermochemistry the composition of the mixture varies. Such<br />

variation is due to chemical reactions taking place between the various species that<br />

form the mixture and to their mutual diffusion, and has the following consequences:<br />

59


60 CHAPTER 3. GENERAL EQUATIONS<br />

1) The need to establish a separate balance for each one of the species by means of<br />

partial continuity equations for each different species. In each of these equations<br />

the effects of the chemical reactions and of diffusion must be inclu<strong>de</strong>d.<br />

2) The modifications due to diffusion of some of the transport coefficients of the<br />

mixture. As previously seen in chapter 2, diffusion does not change the values<br />

of the viscosity coefficients but modifies the heat flux, mainly due to the transport<br />

of the enthalpy of the various species through diffusion.<br />

3) The presence of a new form of energy: the chemical energy released or absorbed<br />

in the chemical reactions.<br />

n<br />

q<br />

f<br />

Σ<br />

V<br />

F<br />

Figure 3.1: Schematic diagram of the actions upon a isolated gas element.<br />

In Gas Dynamics the equations are established by applying the laws of the conservation<br />

of mass, momentum and energy to an i<strong>de</strong>ally isolated gas element following<br />

its motion. 1<br />

The material element thus isolated (Fig. 3.1) is boun<strong>de</strong>d by a fluid surface,<br />

that is to say by a surface Σ, in which the velocity at each point is the velocity<br />

¯v of the gas particle at the same point. This element evolves in accordance with the<br />

aforementioned laws of conservation and un<strong>de</strong>r the action of the surrounding mass,<br />

which acts upon its boundary Σ. This action result in the surface forces and in the<br />

heat transport ¯q throughout each of its surface elements. Furthermore, the action of<br />

the possible force fields acting upon the element must be inclu<strong>de</strong>d; for example, the<br />

action ¯F of the gravity field. In this manner 2 we obtain:<br />

a) The continuity equation, which expresses the laws of the conservation of mass.<br />

b) The equation of motion which expresses that Newton’s Second Law of Mechanics<br />

is satisfied.<br />

c) The energy equation which expresses that the First Law of Thermodynamics is<br />

satisfied.<br />

1 See, i.e., Ref. [1], pp. 337 and following.<br />

2 The same equations could be established by starting from the molecular structure of the gas and applying<br />

the laws of the Kinetic Theory of gases. See, i.e., Ref. [2], p. 25.


3.2. EQUATION OF CONTINUITY 61<br />

The system of equations thus obtained must be completed with the state equation<br />

of the gas and with the law of variation of the viscosity coefficient and of the<br />

remaining thermodynamic functions that <strong>de</strong>al with the problem (internal energy or<br />

enthalpy, etc.) as a function of the gas state variables.<br />

In or<strong>de</strong>r to establish the general equations of Aerothermochemistry, a similar<br />

procedure can be followed, taking into account the three observations previously<br />

stated. 3 In such a case it is to be un<strong>de</strong>rstood that when speaking of the material element<br />

we refer to an element boun<strong>de</strong>d by a surface which moves with a velocity that at<br />

each point concurs with the mixture velocity. This specification is due to the need of<br />

distinguishing between the velocity of the mixture and the velocities of the different<br />

species in cases where diffusion exists. These velocities are not equal (see chapter 2).<br />

In the following, the equations thus obtained are given un<strong>de</strong>r the same notation used<br />

in the preceding chapters.<br />

Vectorial and tensorial notation is preferably used as it enables the concise<br />

writing of the equations and a simple interpretation of their terms. A summary of the<br />

symbols and formulae can be found in the Appendix.<br />

3.2 Equation of continuity<br />

Consi<strong>de</strong>r species A i , for which the partial <strong>de</strong>nsity at a point is ρ i . The variations of<br />

ρ i , following the motion of an element of the mixture, is Dρ i<br />

, where the operator<br />

Dt<br />

D (·)<br />

Dt<br />

= ∂ (·)<br />

∂t<br />

is the substantial <strong>de</strong>rivative of the motion of the mixture.<br />

The variation of ρ i is due:<br />

+ ¯v · ∇ (·) (3.1)<br />

a) To the expansion ∇ · ¯v of the volume element, which <strong>de</strong>creases ρ i in ρ i ∇ · ¯v.<br />

b) To the flux by diffusion of species A i , through the boundary of the element. If<br />

¯v di is the diffusion velocity of species A i diffusion reduces ρ i in ∇ · (ρ i¯v di ).<br />

c) To the mass of species A i produced by chemical reactions. If w i is the reaction<br />

rate of species A i 4 un<strong>de</strong>r the conditions of state and composition that prevail at<br />

the point, then the mass produced per unit volume and per unit time is w i .<br />

3 The general equations for Aerothermochemistry may also be <strong>de</strong>duced from the molecular structure of<br />

the mixture by applying the laws of Kinetic Theory.<br />

4 See chapter 1.


62 CHAPTER 3. GENERAL EQUATIONS<br />

Therefore, the continuity equation for species A i is<br />

Dρ i<br />

Dt = −ρ i∇ · ¯v − ∇ · (ρ i¯v di ) + w i . (3.2)<br />

If there are l different species A i , there is an equation similar to Eq. (3.2) for<br />

each of them. 5<br />

Adding the l equations corresponding to the l different species and making use<br />

of the obvious relation<br />

and of<br />

∑<br />

ρ i = ρ (3.3)<br />

i<br />

∑<br />

w i = 0, (3.4)<br />

i<br />

which expresses the condition that the chemical reactions do not change the mass, and<br />

of<br />

∑<br />

ρ i¯v di = ¯0, (3.5)<br />

which was <strong>de</strong>duced in chapter 2, the following expression is obtained<br />

i<br />

Dρ<br />

+ ρ∇ · ¯v = 0, (3.6)<br />

Dt<br />

which is the continuity equation for the mixture and is i<strong>de</strong>ntical to the continuity<br />

equation of Gas Dynamics.<br />

Equation (3.6) allows simplification in the form of (3.2). For this purpose it is<br />

convenient to express Eq. (3.2) as a function of the mass fraction Y i = ρ i /ρ of species<br />

5 The <strong>de</strong>tailed <strong>de</strong>duction of Eq. (3.2) is as follows.<br />

The mass m i of species A i contained in volume V is<br />

ZZZ<br />

m i = ρ i dV.<br />

V<br />

Its time variation is (see Appendix)<br />

dm i<br />

= d ZZZ ZZZ<br />

ρ i dV =<br />

dt dt V<br />

V<br />

ZZZ » –<br />

∂ρ i<br />

∂t<br />

ZZΣ<br />

dV + ∂ρi<br />

ρ i (¯v · ¯n) dσ =<br />

V ∂t + ∇ · (ρ i¯v) dV.<br />

The mass diffusing from V through Σ per unit time, is<br />

Z<br />

ZZZ<br />

ρ i¯v di · ¯n dσ = ∇ · (ρ i¯v di ) dV.<br />

Σ<br />

V<br />

The mass production in V by the chemical reaction, per unit time, is<br />

ZZZ<br />

w i dV.<br />

V<br />

Then when establishing the balance and expressing that this condition must be satisfied for all V , we obtain<br />

Eq. (3.2).<br />

In all the other formulae that follow the argumentation is i<strong>de</strong>ntical, and for this reason the expansion of<br />

the complete calculation is not consi<strong>de</strong>red necessary.


3.2. EQUATION OF CONTINUITY 63<br />

A i . By doing so and taking into account Eq. (3.6) the following simplified expression<br />

is obtained, which replaces Eq. (3.2)<br />

ρ DY i<br />

Dt + ∇ · (ρY i¯v di ) = w i , (i = 1, 2, . . . , l). (3.7)<br />

Asi<strong>de</strong> from the <strong>de</strong>nsity and velocity ¯v of the mixture, the latter implicitly contained<br />

in the substantial <strong>de</strong>rivative, this system shows the mass fractions Y i of the<br />

different species as well as their corresponding diffusion velocities ¯v di and reaction<br />

rates w i .<br />

As seen in chapter 2, diffusion velocities are linear functions of the gradients<br />

of Y i , p and T . Therein it was shown that the diffusion velocities ¯v ′ di (i = 1, 2, . . . , l)<br />

corresponding to the gradient of Y i and p are given by the system<br />

∑<br />

(<br />

Y i ¯v ′ di − ¯v ′ dj<br />

+ ∇Y i<br />

− ∇Y j<br />

+ M )<br />

j − M i ∇p<br />

= ¯0,<br />

M j D ij Y i Y j M m p<br />

j≠i<br />

∑<br />

Y j ¯v ′ dj = ¯0,<br />

j<br />

(3.8)<br />

where<br />

1<br />

= ∑ Y j<br />

(3.9)<br />

M m M<br />

j j<br />

is the mean mass of one mole of the mixture (see chapter 1).<br />

Similarly it was shown that the diffusion velocities ¯ v ′′ di corresponding to the<br />

temperature gradient are given by expressions of the form<br />

¯ v ′′ di = − D T i<br />

ρY i<br />

∇T<br />

T , (3.10)<br />

where D T i is the thermal diffusion coefficient of species A i .<br />

With respect to the reaction rate w i , it should be noted that the conclusions<br />

in chapter 1 were established un<strong>de</strong>r the assumption that both the composition and<br />

the state of the mixture are homogeneous throughout the mass. Therefore, when the<br />

composition and state of the mixture are not homogeneous, the applicability of said<br />

conclusions <strong>de</strong>pends on the assumption that neither one varies appreciably within a<br />

distance of the or<strong>de</strong>r of magnitu<strong>de</strong> of a characteristic length of the chemical transformations.<br />

This distance could be, for example, the average length of a chain when<br />

chain reactions are produced. Within such a limitation the conclusions obtained in<br />

chapter 1 can be applied here. In particular, if there are r different chemical reactions,<br />

we have seen in the aforementioned chapter that w i is a linear combination of the r<br />

reaction rates w ij corresponding to said chemical reactions<br />

∑<br />

w i = M i ν ij r j . (3.11)<br />

j


64 CHAPTER 3. GENERAL EQUATIONS<br />

3.3 Equations of motion<br />

The equations of motion are obtained by applying Newton’s Second Law of Mechanics,<br />

to the material element in Fig. 3.1.<br />

Let V be the volume of said element. Then ρV will be the mass and ρV ¯v the<br />

momentum, for which the time variation when following the element is<br />

D (ρV ¯v)<br />

Dt<br />

= ρV D¯v<br />

Dt , (3.12)<br />

since ρV does not vary. Therefore, the variation of momentum of the mass contained<br />

in the unit volume is ρ (D¯v/Dt). Such variation originates from:<br />

1) The surface forces acting upon the unit volume.<br />

2) The mass forces acting upon the unit volume.<br />

Let us calculate both, separately.<br />

Let ¯f be the force that acts upon the surface element dσ at a point of Σ, Fig. 3.1.<br />

It has been shown in chapter 1 that ¯f is <strong>de</strong>termined by the stress tensor τ e of components<br />

τ ij and that as a function of the tensor it can be expressed as follows<br />

¯f = ¯n · τ e dσ, (3.13)<br />

where ¯n is the outward normal to the surface element dσ at the point. As it can easily<br />

be proved, the resultant of all forces ¯f per unit volume is ∇ · τ e . 6<br />

Let ¯F be the field intensity of the mass forces, that is to say, the force per unit<br />

mass. The force per unit volume is, evi<strong>de</strong>ntly, ρ ¯F . 7<br />

Now, by expressing Newton’s Second Law of Mechanics, we obtain<br />

ρ D¯v<br />

Dt = ∇ · τ e + ρ ¯F . (3.14)<br />

This vectorial equation is equivalent to three scalar equations corresponding to<br />

the three components of the reference system. For example, in rectangular cartesian<br />

coordinates, the component of this equation parallel to the coordinate axis x i (i =<br />

1, 2, 3) is<br />

ρ ∂v i<br />

∂t + ρv ∂v i<br />

j = ∂τ ij<br />

+ ρF i , (3.15)<br />

∂x j ∂x j<br />

6 In fact, the resultant of the forces acting upon Σ is RR Σ ¯n · τ e dσ. Now, by applying Ostrogradsky’s<br />

theorem we obtain RR Σ ¯n · τe dσ = RRR V ∇ · τe dV. Therefore, the force per unit volume is ∇ · τe.<br />

7 If the force per unit mass <strong>de</strong>pends on the species and ¯F i is the intensity for species A i , ρ ¯F must be<br />

substituted by P ρ i ¯Fi = ρ P Y i ¯Fi .<br />

i<br />

i


3.4. ENERGY EQUATION 65<br />

where the subscripts indicate components, and the repeated subscripts indicate, according<br />

to Einstein’s rule, summation respect to them, from one to three.<br />

As shown in chapter 2, the components τ ij of the stress tensor are expressed as<br />

a function of the gas pressure p at the point, and of the velocity ¯v of the motion, in the<br />

following form<br />

( ∂vi<br />

τ ij = −pδ ij + µ + ∂v )<br />

j<br />

+<br />

(µ ′ − 2 )<br />

∂x j ∂x i 3 µ (∇ · ¯v) δ ij , (3.16)<br />

where δ ij is the Kronecker symbol and µ and µ ′ are the coefficients of viscosity and<br />

volumetric viscosity of the mixture, respectively. These coefficients <strong>de</strong>pend on the<br />

state and composition of the mixture as indicated in the above mentioned chapter. In<br />

particular, for the diluted monatomic gases µ ′ = 0 and in general, a good approximation<br />

is obtained by assuming that it is also zero for all other cases.<br />

In Eq. (3.14) the pressure and viscous stresses can be ma<strong>de</strong> explicit, by using<br />

the expression<br />

τ e = −pU + τ ev (3.17)<br />

previously given in chapter 2. By taking this expression of τ e into Eq. (3.14), we<br />

obtain<br />

ρ D¯v<br />

Dt = −∇p + ∇ · τ ev + ρ ¯F . (3.18)<br />

This equation is i<strong>de</strong>ntical to those obtained in Gas Dynamics for mixtures of uniform<br />

composition with no chemical reaction. The variations in the composition of the<br />

mixture act only through their influence on the values of the viscosity coefficients µ<br />

and µ ′ .<br />

3.4 Energy equation<br />

This equation is obtained when expressing that the variations of the energy contained<br />

in V , Fig.3.1, are due to the work done by the forces acting upon the element and the<br />

heat received by the said element. Let us calculate, separately, the different terms in<br />

this equation.<br />

a) Energy.<br />

Let u i be the internal energy per unit mass of species A i . The internal energy u<br />

per unit mass of the mixture is<br />

u = ∑ i<br />

Y i u i , (3.19)


66 CHAPTER 3. GENERAL EQUATIONS<br />

and the internal energy of the mass contained in V is ρV u. Similarly, the kinetic<br />

energy of the mass contained in V is ρV 1 2 v2 . Therefore, the total energy 8 of the<br />

mass contained in V is ρV ( u + 1 2 v2) and its time variation when following the<br />

motion of V is<br />

(<br />

D<br />

ρV (u + 1 )<br />

Dt 2 v2 ) = ρV D Dt (u + 1 2 v2 ). (3.20)<br />

Consequently, the time variation of the energy of the mass contained in the unit<br />

volume is<br />

b) Work.<br />

The work done by ¯f per unit time is<br />

ρ D Dt (u + 1 2 v2 ). (3.21)<br />

¯v · ¯f = ¯v · (¯n · τ e ) = ¯n · (¯v · τ e ) , (3.22)<br />

and the work per unit time done by the surface forces that act upon the unit<br />

volume is ∇ · (¯v · τ e ). 9 The work per unit time done by the mass forces that act<br />

upon the unit volume is ρ ¯F · ¯v. Consequently, the work per unit time done by all<br />

the forces that act upon the unit volume is<br />

∇ · (¯v · τ e ) + ρ ¯F · ¯v. (3.23)<br />

c) Heat.<br />

The heat received by the element V through dσ per unit time is −¯q · ¯n dσ, where<br />

¯q is the vector of the heat flux <strong>de</strong>fined in chapter 2. Therefore, the heat per unit<br />

time received by the unit volume is<br />

−∇ · ¯q. (3.24)<br />

In the phenomena normally studied by the Aerothermochemistry the radiation<br />

effects are of secondary importance. Furthermore, great difficulty is generally<br />

encountered for the study of the influence of these effects. Nevertheless, if they<br />

are to be taken into consi<strong>de</strong>ration, the vector ¯q r of the radiation flux must be<br />

ad<strong>de</strong>d to the vector ¯q of the heat flux. Information regarding the radiation flux<br />

can be found in the work of Hirschfel<strong>de</strong>r, Curtiss and Bird, mentioned in Ref.<br />

[4], pp. 720 and following.<br />

8 Other forms of energy different from the internal and the kinetic energies, as for example the radiation<br />

energy, are not consi<strong>de</strong>red in the present study.<br />

9 It can be proved in the same manner as for the calculation of the resultant of the surface forces.


3.4. ENERGY EQUATION 67<br />

The energy equation is now obtained by expressing that the energy variation,<br />

given by Eq. (3.21), must be equal to the work done by the exterior forces, given by<br />

Eq. (3.23), plus the heat received, given by Eq. (3.24), thus obtaining<br />

ρ D Dt (u + 1 2 v2 ) = ∇ · (¯v · τ e ) + ρ ¯F · ¯v − ∇ · ¯q. (3.25)<br />

Expanding the first term of the right hand si<strong>de</strong> of this expression, there results<br />

∇ · (¯v · τ e ) = ¯v · (∇ · τ e ) + τ e : τ v , (3.26)<br />

where τ v is the <strong>de</strong>formation velocity tensor, of components γ ij 10 <strong>de</strong>fined by<br />

γ ij = 1 2<br />

( ∂vi<br />

+ ∂v )<br />

j<br />

. (3.27)<br />

∂x j ∂x i<br />

In expression (3.26), the first term of the right hand si<strong>de</strong> measures the work<br />

done by the surface forces that act upon the unit volume when the surface moves with<br />

no <strong>de</strong>formation with velocity ¯v. The second term measures the work done by the<br />

surface forces in or<strong>de</strong>r to produce the <strong>de</strong>formation τ v .<br />

Bringing forth the pressure and the viscous stresses in τ e by using expression<br />

(3.17) for τ e , it can be verified that the <strong>de</strong>formation work (τ e : τ v ) consists of the<br />

pressure work (−p∇ · ¯v) done during the expansion of the gas, and of the work<br />

Φ = τ ev : τ v (3.28)<br />

dissipated by the viscosity to produce the <strong>de</strong>formation τ v . Function Φ, which as it can<br />

be easily proved is always positive, is the dissipation function of Lord Rayleigh. That<br />

is<br />

τ e : τ v = −p∇ · ¯v + Φ. (3.29)<br />

In or<strong>de</strong>r to bring forth the variation of the internal energy u, the energy equation<br />

(3.25) can be simplified by making use of the equation of motion (3.14). In fact, by<br />

subtracting from Eq. (3.25) the scalar product of Eq. (3.14) by ¯v, taking into account<br />

Eq. (3.26), there results<br />

or else, making use of Eq. (3.29),<br />

10 See chapter 2.<br />

ρ Du<br />

Dt<br />

ρ Du<br />

Dt = τ e : τ v − ∇ · ¯q, (3.30)<br />

= −p∇ · ¯v + Φ − ∇ · ¯q. (3.31)


68 CHAPTER 3. GENERAL EQUATIONS<br />

It has been shown in chapter 2 that ¯q is given by the expression<br />

¯q = −λ∇T + ρ ∑ i<br />

∑∑<br />

Y j D T i<br />

(¯v di − ¯v dj )<br />

i j M i M j D ij<br />

Y i h i¯v di + RT<br />

∑ Y i<br />

. (3.32)<br />

M j<br />

In this expression the first term of the right hand si<strong>de</strong> gives the heat flux through conduction<br />

due to temperature gradient. This is the only existing term in Gas Dynamics.<br />

The second term gives the enthalpy flux of the various species through diffusion. The<br />

third term gives the heat flux due to the Dufour effect of reciprocal effect of the thermal<br />

diffusion in Onsager’s sens. 11 Such an effect is zero when all the molecules have<br />

the same mass since, then, D T i is zero. In general, the Dufour effect can be neglected<br />

and thereby ¯q reduces to the expression<br />

j<br />

¯q = −λ∇T + ρ ∑ i<br />

Y i h i¯v di . (3.33)<br />

Taken this expression of ¯q into Eq. (3.30), the following final expression for<br />

the variation of internal energy of the mixture is obtained<br />

(<br />

ρ Du<br />

Dt = −p∇ · ¯v + Φ + ∇ · (λ∇T ) − ∇ · ρ ∑ )<br />

Y i h i¯v di . (3.34)<br />

i<br />

from:<br />

This equation shows that the time variation of the internal energy originates<br />

a) The work −p∇ · ¯v done by pressure to compress the gas.<br />

b) The work Φ dissipated by the viscous forces to produce the <strong>de</strong>formation.<br />

c) The heat ∇ · (λ∇T ) received through conduction.<br />

d) The enthalpy flux −∇ · (ρ ∑ i Y ih i¯v di ) of the species through diffusion.<br />

Forms a), b) and c) are the same as those that exist in Gas Dynamics when there is no<br />

change in composition.<br />

In many cases it is convenient to work with Eq. (3.25) which inclu<strong>de</strong>s kinetic<br />

energy, 12 bringing forth in this equation the enthalpy h of the mixture<br />

h = u + p ρ . (3.35)<br />

For this purpose we shall start by making the pressure explicit in the first term<br />

of the right hand si<strong>de</strong> of Eq. (3.25) by means of Eq. (3.17). Thus, the following is<br />

11 See Ref. [5], p. 118.<br />

12 See chapters 4 and 5.


3.5. GENERAL EQUATIONS 69<br />

obtained for the said term<br />

∇ · (¯v · τ e ) = −∇ · (p¯v) + ∇ · (¯v · τ ev ) . (3.36)<br />

Now, if the first term of the right hand si<strong>de</strong> of this equation is expan<strong>de</strong>d and<br />

combined with the continuity equation (3.6), the following is obtained<br />

∇ · (¯v · τ e ) = ∂p<br />

∂t − ρ D ( ) p<br />

+ ∇ · (¯v · τ ev ) . (3.37)<br />

Dt ρ<br />

Substituting this equation into Eq. (3.25), taking into account Eq. (3.35), it finally<br />

gives<br />

ρ D Dt<br />

(h + 1 2 v2 )<br />

= ∂p<br />

∂t + ∇ · (¯v · τ ev)<br />

+ ρ ¯F · ¯v + ∇ · (λ∇T ) − ∇ ·<br />

(<br />

ρ ∑ i<br />

Y i h i¯v di<br />

)<br />

,<br />

(3.38)<br />

where Eq. (3.33) has been used to eliminate ¯q.<br />

3.5 General equations<br />

As a summary of preceding paragraphs we shall once more write the general system<br />

of equations that govern the transformation of a mixture of reactant gases in motion,<br />

that is to say the general equations of Aerothermochemistry.<br />

Continuity Equation<br />

For the mixture:<br />

For the species:<br />

Dρ<br />

+ ρ∇ · ¯v = 0. (3.6)<br />

Dt<br />

ρ DY i<br />

Dt + ∇ · (ρY i¯v di ) = w i , (i = 1, 2, . . . , l). (3.7)<br />

Equation of Motion<br />

ρ D¯v<br />

Dt = −∇p + ∇ · τ ev + ρ ¯F . (3.18)


70 CHAPTER 3. GENERAL EQUATIONS<br />

Energy Equation<br />

ρ Du<br />

Dt = −p∇ · ¯v + Φ + ∇ · (λ∇T ) − ∇ · (<br />

ρ ∑ i<br />

Y i h i¯v di<br />

)<br />

. (3.34)<br />

State Equation<br />

p<br />

ρ = R mT. (3.39)<br />

Moreover the laws of variation of the thermodynamic functions, of the transport coefficients<br />

and of the reaction rates as functions of the state and composition of the<br />

mixture must be known.<br />

3.6 Entropy variation<br />

It is interesting to bring forth the entropy variation and therewith the causes of the<br />

irreversibility of the process. Moreover, the entropy variations have a great influence<br />

upon motion, for example, producing vortexes. 13<br />

Thermodynamics teaches 14 that the variation dS of the entropy of a reactant<br />

system corresponding to the variations dU, dV and dm i of its internal energy. volume<br />

and masses of the chemical species respectively, is given by the expression<br />

T dS = dU + p dV − ∑ i<br />

µ i dm i , (3.40)<br />

where µ i is the chemical potential of species A i .<br />

If one applies this expression to the mass contained in the volume element V<br />

of Fig. 3.1 following the motion, and bearing in mind that for this volume<br />

S = ρV s, U = ρV u, m i = ρV Y i ,<br />

1<br />

V<br />

DV<br />

Dt<br />

= ∇ · ¯v, (3.41)<br />

where s is the entropy of the mixture per unit mass, the following expression is obtained<br />

for the time variation of the entropy of the mass contained in the unit volume<br />

(<br />

ρ Ds<br />

Dt = 1 ρ Du<br />

T Dt + p∇ · ¯v − ρ ∑ i<br />

µ i<br />

DY i<br />

Dt<br />

)<br />

. (3.42)<br />

13 See chapter 9.<br />

14 See chapter 1.


3.6. ENTROPY VARIATION 71<br />

Taking into this expression the variations of u and Y i given by Eqs. (3.34) and<br />

(3.7) respectively, one obtains<br />

ρT Ds<br />

Dt =Φ + ∇ · (λ∇T ) − ∇ · (ρ ∑ i<br />

− ∑ i<br />

µ i w i + ∑ i<br />

Y i h i¯v di )<br />

µ i ∇ · (ρY i¯v di ).<br />

(3.43)<br />

As can easily be verified, this equation can be written in the form<br />

[<br />

]<br />

ρ Ds<br />

Dt =∇ · λ ∇T<br />

T<br />

− ρ ∑<br />

Y i (h i − µ i )¯v di + 1 (∇T )2<br />

[Φ + λ<br />

T<br />

T T<br />

− ρ ∇T<br />

T<br />

i<br />

∑<br />

Y i (h i − µ i )¯v di − ρ ∑<br />

i<br />

i<br />

Y i¯v di · ∇µ i − ∑ i<br />

µ i w i<br />

] (3.44)<br />

Then we have 15 µ i = h i − T s i , (3.45)<br />

where h i and s i are, respectively, the enthalpy and the entropy per unit mass of species<br />

A i at the partial pressure p i of the species and at the temperature T of the mixture.<br />

When taking this expression into Eq. (3.44) one obtains<br />

[<br />

ρ Ds<br />

Dt =∇ · λ ∇T<br />

T<br />

− ρ ∑ ]<br />

Y i s i¯v di + 1 (∇T )2<br />

[Φ + λ<br />

T T<br />

i<br />

− ρ∇T ∑ Y i s i¯v di − ρ ∑ Y i¯v di · ∇µ i − ∑ ]<br />

µ i w i .<br />

i<br />

i<br />

i<br />

(3.46)<br />

This equation admits a simple interpretation. In fact the first term of the right hand<br />

si<strong>de</strong> of this equation gives the reversible entropy flux across the boundary of the unit<br />

volume. This flux originates from: a) heat transport through conduction, and b) diffusion<br />

of the species. The second and third terms originate from the entropy generated<br />

in the gas per unit volume and unit time, due to irreversibilities of the process. Such<br />

irreversibilities, shown in Eq. (3.46), arise from: a) viscosity, b) heat conductivity,<br />

c) diffusion, and d) chemical reactions. Making use of Eq. (3.11) the term ∑ µ i w i<br />

i<br />

corresponding to d) may be written in the form<br />

∑<br />

µ i w i = − ∑ r j α j , (3.47)<br />

i<br />

j<br />

where α j is the chemical affinity corresponding to reaction j. 16 All contributions to<br />

the variation of the entropy from the irreversibilities of the process must be positive. 17<br />

15 See chapter 1.<br />

16 See chapter 1.<br />

17 For further information, see Refs. [4] or [5].


72 CHAPTER 3. GENERAL EQUATIONS<br />

3.7 One-dimensional motions<br />

As an application we shall see the form taken by the general equations in the case of<br />

one dimensional motions. These equations, in particular those relative to stationary<br />

motions, will be wi<strong>de</strong>ly used in the study of the <strong>de</strong>tonations and flames. 18<br />

It will be assumed that:<br />

a) No mass forces exist.<br />

b) The coefficient of volumetric viscosity µ ′ is zero.<br />

c) The effects of thermal diffusion and radiation can be neglected.<br />

We shall adopt a cartesian rectangular system with the x axis parallel to the<br />

direction of motion. Then, the only in<strong>de</strong>pen<strong>de</strong>nt variables of the motion are coordinate<br />

x and time t, and the only velocity component different from zero is that parallel to<br />

the x axis, which will be <strong>de</strong>signated as v.<br />

The substantial <strong>de</strong>rivative (3.1) reduces in this case to<br />

D (·)<br />

Dt<br />

= ∂ (·)<br />

∂t<br />

+ v ∂ (·)<br />

∂x . (3.48)<br />

Continuity equations<br />

The continuity equations for the mixture reduces to<br />

∂ρ<br />

∂t + v ∂ρ<br />

∂x + ρ ∂v = 0. (3.49)<br />

∂x<br />

Similarly, the continuity equations (3.7) for the various species, reduces to<br />

ρ ∂Y i<br />

∂t + ρv ∂Y i<br />

∂x + ∂<br />

∂x (ρY iv di ) = w i , (i = 1, 2, . . . , l). (3.50)<br />

Equations of motion<br />

In the vectorial Eq. (3.18) the only component not i<strong>de</strong>ntically zero is that parallel to<br />

x axis. Furthermore, the only component different from zero in the viscous stress<br />

tensor τ ev is<br />

τ xx = 4 3 µ ∂v<br />

∂x , (3.51)<br />

corresponding to viscous stress that acts upon the plane normal to the motion and<br />

parallel to the x axis. This stress is subtracted from the pressure. Therefore, Eq. (3.18)<br />

18 See chapters 5 and 6.


3.7. ONE-DIMENSIONAL MOTIONS 73<br />

reduces to<br />

ρ ∂v ∂v<br />

+ ρv<br />

∂t ∂x = − ∂p<br />

∂x + 4 3<br />

∂<br />

∂x<br />

(<br />

µ ∂v )<br />

. (3.52)<br />

∂x<br />

Energy equation<br />

Equation (3.34) reduces to<br />

ρ ∂u ∂u<br />

+ ρv<br />

∂t<br />

∂x = − p ∂v<br />

∂x + 4 ( ∂v<br />

3 µ ∂x<br />

− ∂ (<br />

λ ∂T<br />

∂x ∂x<br />

) 2<br />

)<br />

− ∂<br />

∂x<br />

(<br />

ρ ∑ i<br />

Y i h i v di<br />

)<br />

Or if it is <strong>de</strong>ci<strong>de</strong>d to use the total energy equation (3.38) it takes the form<br />

ρ ∂ (h + 1 )<br />

∂t 2 v2 + ρv ∂ (h + 1 )<br />

∂x 2 v2 = ∂p<br />

∂t + 4 (<br />

∂<br />

µv ∂v )<br />

3 ∂x ∂x<br />

+ ∂ (<br />

λ ∂T ) (<br />

− ∂ ρ ∑ )<br />

Y i h i v di .<br />

∂x ∂x ∂x<br />

i<br />

.<br />

(3.53)<br />

(3.54)<br />

The system of equations (3.49), (3.50), (3.52) and (3.53), or (3.54), must be<br />

completed with the following equations.<br />

Total mass fraction equation<br />

∑<br />

Y i = 1. (3.55)<br />

i<br />

Diffusion equations<br />

The system of equations (3.8) which gives the diffusion velocities reduces to<br />

∑<br />

[ (<br />

Y j vdi − v dj 1 ∂Y i<br />

+<br />

M j D ij Y i ∂x − 1 )<br />

∂Y j<br />

Y j ∂x<br />

j<br />

+ M j − M i<br />

M m<br />

( )] 1 ∂p<br />

= 0, (i = 1, 2, . . . , l),<br />

p ∂x<br />

∑<br />

Y i v di = 0.<br />

i<br />

(3.56)<br />

State equation<br />

The state equation is, by assumption, that of i<strong>de</strong>al gases<br />

p<br />

ρ = R mT. (3.39)


74 CHAPTER 3. GENERAL EQUATIONS<br />

State functions<br />

The internal energy u i of species A i is 19<br />

The internal energy of the mixture is<br />

where<br />

u = ∑ i<br />

∫ T<br />

u i = u i0 + c vi dT. (3.57)<br />

T 0<br />

Y i u i = ∑ i<br />

∫ T<br />

Y i u i0 + c v dT, (3.58)<br />

T 0<br />

c v = ∑ Y i c vi (3.59)<br />

i<br />

is the mixture heat capacity at constant volume.<br />

Similarly, for the enthalpy we have<br />

and for the enthalpy of the mixture<br />

with<br />

h = ∑ i<br />

h i = h i0 +<br />

Y i h i = ∑ i<br />

c p = ∑ i<br />

∫ T<br />

T 0<br />

c pi dT, (3.60)<br />

∫ T<br />

Y i h i0 + c p dT, (3.61)<br />

T 0<br />

Y i c pi . (3.62)<br />

Transport coefficients<br />

In system (3.56), D ij are the binary diffusion coefficients between species A i and<br />

A j . For each pair of these species the coefficients <strong>de</strong>pend on the state variables, as<br />

indicated in chapter 2.<br />

λ and µ are the coefficients of thermal conductivity and viscosity of the mixture,<br />

respectively. These coefficients <strong>de</strong>pend on the state variables and on the composition<br />

of the mixture as indicated in chapter 2.<br />

Reaction rates<br />

The reaction rate of species A i is given in Eq. (3.11). In this equation the coefficients<br />

ν ij and the velocities r j corresponding to the different elementary reactions of the<br />

mixture are given by the formulae of chapter 1.<br />

19 See chapter 1.


3.8. STATIONARY, ONE-DIMENSIONAL MOTIONS 75<br />

3.8 Stationary, one-dimensional motions<br />

If the motion is not only one-dimensional but stationary, the preceding equations are<br />

consi<strong>de</strong>rably simplified. This type of motion, as previously stated, is of special interest<br />

for the study of the structure of combustion waves.<br />

The stationary motions are characterized by the condition<br />

∂ (·)<br />

= 0. (3.63)<br />

∂t<br />

Therefore, the only in<strong>de</strong>pen<strong>de</strong>nt variable is x.<br />

Let us see how the equations in the preceding paragraph can be simplified when<br />

condition (3.63) is introduced therein.<br />

Continuity Equations<br />

The continuity equation (3.49) reduces to<br />

v dρ<br />

dx + ρ dv<br />

dx ≡<br />

This equation can be integrated, thus obtaining<br />

d (ρv) = 0. (3.64)<br />

dx<br />

ρv = m, (3.65)<br />

where m is an integration constant which gives the mass flux per unit surface normal<br />

to the direction of motion.<br />

Similarly, equations (3.50) reduces to<br />

d<br />

dx (ρY iv + ρY i v di ) = w i , (i = 1, 2, . . . , l) . (3.66)<br />

By introducing the flux of the different species, these equations reduce to a<br />

very simple form. In fact, let m i = mε i be the flux of species A i through the unit<br />

surface normal to the direction of motion, that is to say that ε i is the fraction of the<br />

total flux corresponding to species A i . Since the velocity v i of species A i is<br />

v i = v + v di , (3.67)<br />

evi<strong>de</strong>ntly the said flux is<br />

mε i = ρY i (v + v di ) . (3.68)<br />

Now, when this expression is compared with the left hand si<strong>de</strong> of Eq. (3.66) it<br />

is seen that this system can be written in the form<br />

m dε i<br />

dx = w i, (i = 1, 2, . . . , l). (3.69)


76 CHAPTER 3. GENERAL EQUATIONS<br />

The physical interpretation of this equations is obvious, m dε i is the difference<br />

between the fluxes of species A i through two surfaces normal to the direction of motion,<br />

separated one from another by the distance dx. But this difference has its origin<br />

in the amount of the said species produced per unit time by the chemical reactions that<br />

take place within the space that separates both surfaces, which is evi<strong>de</strong>ntly w i dx.<br />

The l equations of system (3.69) are not in<strong>de</strong>pen<strong>de</strong>nt. In fact:<br />

1) Formulae (3.11) shows that w i are linear combinations of the r reaction rates r j<br />

corresponding to the different reactions that take place between the species of the<br />

mixture. Consequently, the maximum number of in<strong>de</strong>pen<strong>de</strong>nt w i is at most r.<br />

2) The chemical reactions do not change the total number of atoms of the elements<br />

that form the species. Let g be the number of different chemical elements of the<br />

species and let a ij (j = 1, 2, ...., g) be the number of atoms of the element j in<br />

species i. The conservation of the element in the chemical reactions imposes the<br />

following system of conditions between w i<br />

∑ w i<br />

a ij = 0, (j = 1, 2, . . . , g). (3.70)<br />

M i<br />

i<br />

It might occur however that not all these conditions are in<strong>de</strong>pen<strong>de</strong>nt. In fact<br />

the number of in<strong>de</strong>pen<strong>de</strong>nt equations in (3.70) is the rank of the matrix of coefficients<br />

a ij . This rank is the minimum number of components nee<strong>de</strong>d to form the l species<br />

A i in the sense of the phase rule. Therefore, if g ′ ≤ g is the number of components of<br />

the mixture, there exist g ′ in<strong>de</strong>pen<strong>de</strong>nt linear relations between w i , due to Eq. (3.70),<br />

and the number of in<strong>de</strong>pen<strong>de</strong>nt w i is, at most, l − g ′ .<br />

Thereby, the number l ′ < l of the in<strong>de</strong>pen<strong>de</strong>nt w i is the smallest one of the<br />

numbers r and l ′ − g ′ . 20<br />

System (3.69) shows that there are as many ε i in<strong>de</strong>pen<strong>de</strong>nt from each other as<br />

there are w i , that is to say, l ′ < l, which makes it possible to reduce the number of<br />

variables of the problem in l − l ′ .<br />

Equation of Motion<br />

Equation (3.52) reduces to<br />

m dv<br />

dx = − dp<br />

dx + 4 3<br />

(<br />

d<br />

µ dv )<br />

, (3.71)<br />

dx dx<br />

20 As it can easily be proved, the difference l − g ′ between the number of species and the number<br />

of components is the maximum number of in<strong>de</strong>pen<strong>de</strong>nt reactions equations which can exist among the l<br />

species. This property simplifies the interpretation of the conclusion concerning the number of in<strong>de</strong>pen<strong>de</strong>nt<br />

w i . In fact, if l − g ′ < r, the r chemical reactions are not linearly in<strong>de</strong>pen<strong>de</strong>nt from each other.


3.9. THE CASE OF ONLY TWO CHEMICAL SPECIES 77<br />

which by integration gives<br />

where i is an integration constant.<br />

p + mv − 4 3 µ dv = i, (3.72)<br />

dx<br />

Energy Equation<br />

In this case it is convenient to make use of the total energy equation (3.54) which by<br />

integration gives<br />

mh + ρ ∑ i<br />

Y i h i v di + 1 2 mv2 − λ dT<br />

dx − 4 dv<br />

µv = e, (3.73)<br />

3 dx<br />

where e is an integration constant.<br />

The two first terms of this equation give the enthalpy flux m ∑ ε i h i of the<br />

i<br />

mixture per unit surface normal to x, as can easily be proved by keeping in mind<br />

(3.61) and (3.68). Therefore Eq. (3.73) may also be written in the form<br />

(<br />

1<br />

m<br />

2 v2 + ∑ )<br />

ε i h i − λ dT<br />

dx − 4 dv<br />

µv = e, (3.74)<br />

3 dx<br />

i<br />

whose physical interpretation is obvious.<br />

Diffusion Equations<br />

Equations (3.56) remain valid<br />

∑<br />

j<br />

Y j<br />

M j<br />

[<br />

vdi − v dj<br />

D ij<br />

+<br />

( 1 dY i<br />

Y i dx − 1 dY j<br />

Y j dx<br />

( 1 dp<br />

p dx<br />

+ M j − M i<br />

M m<br />

)<br />

)]<br />

= 0, (i = 1, 2, . . . , l),<br />

∑<br />

Y i v di = 0.<br />

i<br />

(3.75)<br />

3.9 The case of only two chemical species<br />

Let us now assume that in the preceding case the number of chemical species of the<br />

mixture reduces to two, A 1 and A 2 . Then, a single chemical variable Y 1 , <strong>de</strong>noted Y ,<br />

is enough to <strong>de</strong>fine the composition of the mixture, since Y 2 = 1 − Y . Similarly the<br />

flux fraction ε i of A 1 will be <strong>de</strong>noted ε and one has ε 2 = 1 − ε.


78 CHAPTER 3. GENERAL EQUATIONS<br />

reduces to<br />

Y v d1<br />

D 12<br />

System (3.69) reduces in this case to the single equation<br />

m = dε = w. (3.76)<br />

dx<br />

Equation of motion (3.72) remains the same and the energy equation (3.74)<br />

m<br />

(h 2 + (h 1 − h 2 ) ε + 1 )<br />

2 v2 − λ dT<br />

dx − 4 dv<br />

µv = e. (3.77)<br />

3 dx<br />

Finally, the system of diffusion equations (3.75) reduces to the single equation<br />

(<br />

+ dY<br />

dx + (M2 − M 1 ) [ ])<br />

(M 2 − M 1 ) Y + M 1 Y (1 − Y ) dp<br />

= 0, (3.78)<br />

M 1 M 2 p dx<br />

which making use of Eq. (3.68) that in this case reduces to<br />

can be written as<br />

mε = ρY (v + v d1 ) = mY + ρY v d1 , (3.79)<br />

dY<br />

dx + (M 2 − M 1 ) [ ]<br />

(M 2 − M 1 ) Y + M 1 Y (1 − Y ) dp<br />

M 1 M 2 p dx = m (Y − ε). (3.80)<br />

ρD 12<br />

The unknown of the problem are in this case v, p, ρ, T , Y and ε. The six equation<br />

which <strong>de</strong>termine their values are the following: continuity (3.65) and (3.76), motion<br />

(3.72), energy (3.77), diffusion (3.80) and state (3.39). The system thus formed<br />

will be studied in <strong>de</strong>tail in chapters 5 and 6 in discussing the structure of the combustion<br />

waves, specially un<strong>de</strong>r the simplified form studied in the following. Let us<br />

assume, in particular, the following two conditions:<br />

1) The heat capacities at constant pressure of both species are equal<br />

c p1 = c p2 = c p . (3.81)<br />

2) The molar masses of both species are equal, or else the gradient of pressure is<br />

very small.<br />

By virtue of the first assumption, and taking into account (3.60), one has<br />

h 2 − h 1 = h 20 − h 10 = q, (3.82)<br />

where q is the heat of reaction per unit mass of the mixture.<br />

Due to (3.82), equation (3.77) reduces to<br />

m<br />

(h 2 − qε + 1 )<br />

2 v2 − λ dT<br />

dx − 4 dv<br />

µv = e. (3.83)<br />

3 dx


3.10. STATIONARY, ONE-DIMENSIONAL MOTION OF IDEAL GASES WITH HEAT ADDITION 79<br />

By virtue of the second assumption, the diffusion equation (3.80) reduces to<br />

which is the expression of Fick’s Law.<br />

dY<br />

dx = m (Y − ε), (3.84)<br />

ρD 12<br />

If, furthermore, the heat capacity c p is in<strong>de</strong>pen<strong>de</strong>nt from temperature, then<br />

equation (3.83), consi<strong>de</strong>ring (3.60), takes the form<br />

m<br />

(c p T − qε + 1 )<br />

2 v2 − λ dT<br />

dx − 4 dv<br />

µv = e, (3.85)<br />

3 dx<br />

where e must now inclu<strong>de</strong> the constant terms that come from (3.60).<br />

3.10 Stationary, one-dimensional motion of i<strong>de</strong>al gases<br />

with heat addition<br />

Let us assume in the above case that the action of viscosity and thermal conductivity<br />

can be neglected in the equation of motion (3.72) and energy (3.85). This is justified if<br />

the gradients of velocity and temperature are not very ”large”. 21 Then (3.72) reduces<br />

to<br />

p + mv = i. (3.86)<br />

Likewise (3.85) reduces to<br />

m<br />

(c p T − qε + 1 )<br />

2 v2 = e. (3.87)<br />

These two equations together with (3.65)<br />

ρv = m (3.88)<br />

and the equation of state (3.39)<br />

p<br />

ρ = R mT, (3.89)<br />

<strong>de</strong>termine the values for p, ρ, T and v corresponding to each value of ε, if the values of<br />

such variables corresponding to a given value of ε are known, for example, to ε = 0.<br />

This enables the <strong>de</strong>termination of the constants of Eqs. (3.86), (3.87) and (3.88). In<br />

such case, in Eq. (3.87) the term Q = qε represents the heat ad<strong>de</strong>d to the gas per unit<br />

mass from the initial state ε = 0. The result is in<strong>de</strong>pen<strong>de</strong>nt from the way in which the<br />

heat is ad<strong>de</strong>d; be it by chemical reactions or transmitted from external sources.<br />

21 See chapter 5 for the exact meaning of this expression.


80 CHAPTER 3. GENERAL EQUATIONS<br />

In the following, R m is assumed to be constant as occurs if the molar masses<br />

of both species are equal.<br />

Taking in abscissas the heat ad<strong>de</strong>d to the gas and in ordinates the values for<br />

v and T 22 a diagram can be drawn which clearly shows the transformations taking<br />

place within the gas. This has been done in Fig. 3.3 for the case where the relation<br />

γ of heat capacities is 1.4. The dimensionless variables which will be <strong>de</strong>fined in the<br />

following are introduce in or<strong>de</strong>r to have an universal diagram.<br />

1) Let<br />

b = γ + 1<br />

γ<br />

ma s0<br />

, (3.90)<br />

i<br />

where a s0 = √ γR m T s0 is the velocity of sound at the stagnation condition T s0<br />

corresponding to the initial state Q = 0.<br />

A simple calculation shows that b can be expressed as a function of the Mach<br />

number M 0 in the form:<br />

b = (γ + 1) M 0<br />

1 + γM 2 0<br />

√<br />

1 + γ − 1 M0 2 2<br />

. (3.91)<br />

Similarly, let M ∗ 0 be ratio of v 0 to the critical velocity v cr of the gas at the initial<br />

state. b can also be expressed as a function of M ∗ 0 in the form<br />

b = √ 2(γ + 1)<br />

M ∗ 0<br />

1 + M ∗2<br />

0<br />

. (3.92)<br />

b and M 0 are represented in Fig. 3.2 as functions of M ∗ 0 for γ = 1.4.<br />

2) Let T s be the stagnation temperature of the gas when the ad<strong>de</strong>d heat is Q, and<br />

let n be the ratio of T s to T s0 ,<br />

n = T s<br />

T s0<br />

. (3.93)<br />

The relation between T s and T s0 is obtained when expressing e in Eq. (3.87) by<br />

means of T s and T s0 . Thus resulting<br />

n = 1 +<br />

Q<br />

c p T s0<br />

. (3.94)<br />

The ad<strong>de</strong>d heat is expressed by means of the variable x <strong>de</strong>fined by<br />

3) The velocity is expressed in dimensionless form by<br />

22 See von Kármán, Ref. [7].<br />

x = b 2 n. (3.95)<br />

v ∗ = b v<br />

a s0<br />

. (3.96)


3.10. STATIONARY, ONE-DIMENSIONAL MOTION OF IDEAL GASES WITH HEAT ADDITION 81<br />

1.2<br />

6<br />

1.0<br />

0.8<br />

b<br />

5<br />

4<br />

b<br />

0.6<br />

3<br />

M 0<br />

0.4<br />

2<br />

0.2<br />

M 0<br />

1<br />

0<br />

0<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

*<br />

M 0<br />

Figure 3.2: M 0 and b (Eq. (3.92)) as a function of M0 ∗ .<br />

4) Likewise, the temperature is expressed in dimensionless form by<br />

θ = b 2 T<br />

T s0<br />

. (3.97)<br />

If these variables are taken into the previous system and this system is solved<br />

for v ∗ and θ, one obtains<br />

(v ∗ − 1) 2 = 1 − 2x<br />

γ + 1 , (3.98)<br />

θ = x − γ − 1 v ∗2 . (3.99)<br />

2<br />

The values of v ∗ and θ are represented in the diagram Fig. 3.3 where the law<br />

of variation of the Mach number M of the motion is also inclu<strong>de</strong>d. M is given by the<br />

expression<br />

M = v∗<br />

√<br />

θ<br />

. (3.100)<br />

It can easily be verified that for v ∗ and θ to be real, x cannot exceed the value<br />

x max = γ + 1 , (3.101)<br />

2<br />

to which corresponds a maximum value n max of n given by the expression<br />

n max = γ + 1<br />

2b 2 . (3.102)


82 CHAPTER 3. GENERAL EQUATIONS<br />

2.5<br />

1.25<br />

v * , M<br />

2.0<br />

1.5<br />

1.0<br />

θ −<br />

B<br />

v +<br />

*<br />

θ +<br />

M +<br />

A<br />

1.00<br />

0.75<br />

0.50<br />

θ<br />

0.5<br />

M − v−<br />

*<br />

0.25<br />

O<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

0.00<br />

1.4<br />

x<br />

Figure 3.3: Mach number (M), nondimensional velocity (v ∗ ) and temperature (θ) as a function<br />

of the nondimensional heat ad<strong>de</strong>d to the gas (x).<br />

To this value corresponds a maximum value Q max of Q given by the expression<br />

( ) γ + 1<br />

Q max =<br />

2b 2 − 1 c p T s0 . (3.103)<br />

This value is <strong>de</strong>termined by the initial condition through b 2 and c p T s0 . Q max<br />

is the maximum heat that can be ad<strong>de</strong>d to the flow consistent with the given initial<br />

conditions.<br />

The values of v ∗ 1, θ 1 and M 1 of v ∗ , θ and M corresponding to x = x max are<br />

v ∗ 1 = 1, θ 1 = 1, M 1 = 1. (3.104)<br />

Therefore the velocity of the gas at this point is the velocity of sound. In the<br />

curve of velocities point A, corresponding to x = x max , divi<strong>de</strong>s this curve into two<br />

branches OA subsonic and AB supersonic. Fig. 3.4 shows that if the initial velocity<br />

is subsonic the gas accelerates as the heat is ad<strong>de</strong>d whilst if the initial velocity is<br />

supersonic the gas <strong>de</strong>celerates. In both cases the maximum heat that can be ad<strong>de</strong>d<br />

is given by (3.102) or (3.103). Conversely, the maximum initial velocity M0,max ∗ for<br />

each value of n is obtained when Eq. (3.102) is solved for M 0 . Thus results<br />

M0,max ∗ = √ n ± √ n − 1 . (3.105)<br />

Sign plus corresponds to supersonic velocities since M ∗ 0,max > 1. Sign minus


3.10. STATIONARY, ONE-DIMENSIONAL MOTION OF IDEAL GASES WITH HEAT ADDITION 83<br />

corresponds to subsonic velocities. This case has great importance in technical applications<br />

as it <strong>de</strong>termines, for example, the maximum velocity permissible at the inlet<br />

of a constant cross-section combustion chamber as a function of the heat released in<br />

the same. 23 The value of the corresponding Mach number M 0,max is given by<br />

√<br />

2M0,max<br />

∗2<br />

M 0,max =<br />

(γ + 1) − (γ − 1) M0,max<br />

∗2 . (3.106)<br />

The values of M 0,max and M ∗ 0,max as a function of n are represented in Fig. 3.4.<br />

1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 2.1<br />

14<br />

subsonic<br />

supersonic<br />

n (supersonic)<br />

12<br />

M * 0,max , M 0,max (subsonic)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

M 0,max<br />

M * 0,max<br />

10<br />

8<br />

6<br />

4<br />

M * 0,max , M 0,max (supersonic)<br />

0.2<br />

2<br />

M 0,max<br />

0.0<br />

0<br />

1 3 5 7 9 11 13 15 17 19 21 23<br />

n (subsonic)<br />

Figure 3.4: Values of M 0,max and M ∗ 0,max as a function of n.<br />

It is easily seen that this choking effect imposes an important limitation to the Mach<br />

number at the inlet of the combustion chamber.<br />

With respect to the law of variation of temperature as a function of Q it is<br />

interesting to remark that the maximum temperature has the value<br />

θ max =<br />

(γ + 1)2<br />

, (3.107)<br />

4γ<br />

and it is reached for a value of n slightly smaller that its maximum value as well as<br />

for a subsonic velocity. This is due to the fact that for velocities slightly subsonic the<br />

acceleration produced by the ad<strong>de</strong>d heat is so large that the heat received is unable to<br />

balance the cooling produced by the expansion of the accelerating gas. In chapter 5,<br />

where these results are applied, it is seen that this effect produces in <strong>de</strong>tonation.<br />

23 See chapter 10.


84 CHAPTER 3. GENERAL EQUATIONS<br />

3.11 Appendix: Notation and repertoire of vectorial<br />

and tensorial formulae used in the present<br />

chapter<br />

Scalar: a<br />

Vector: ā (components a 1 , a 2 and a 3 )<br />

Scalar product: ā · ¯b (scalar)<br />

Vectorial product: ā × ¯b (vector)<br />

The Hamilton ”nabla”:<br />

Gradient of a scalar:<br />

Divergence of a vector:<br />

∂ (·) ∂ (·) ∂ (·)<br />

∇ (·) = ī 1 + ī 2 + ī 3 (vectorial operator)<br />

∂x 1 ∂x 2 ∂x 3<br />

ī 1 , ī 2 and ī 3 are unit vectors parallel to the coordinate axes.<br />

∇a = ī 1<br />

∂a<br />

∂x 1<br />

+ ī 2<br />

∂a<br />

∂x 2<br />

+ ī 3<br />

∂a<br />

∂x 3<br />

(vector)<br />

∇ · ā = ∂a 1<br />

∂x 1<br />

+ ∂a 2<br />

∂x 2<br />

+ ∂a 3<br />

∂x 3<br />

(scalar)<br />

Curl of a vector:<br />

( ∂a3<br />

∇ × ā = ī 1 − ∂a ) (<br />

2 ∂a1<br />

+ ī 2 − ∂a ) (<br />

3 ∂a2<br />

+ ī 3 − ∂a )<br />

1<br />

∂x 2 ∂x 3 ∂x 3 ∂x 1 ∂x 1 ∂x 2<br />

(vector)<br />

Convective <strong>de</strong>rivative:<br />

¯v · ∇ (·) = v 1<br />

∂ (·)<br />

∂x 1<br />

+ v 2<br />

∂ (·)<br />

∂x 2<br />

+ v 3<br />

∂ (·)<br />

∂x 3<br />

∇ · (a¯b ) = ¯b · ∇a + a ( ∇ · ¯b ) .<br />

(scalar operator)<br />

Tensor:<br />

τ =<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

τ 11 τ 12 τ 13<br />

⎟<br />

τ 21 τ 22 τ 23 ⎠<br />

τ 31 τ 32 τ 33<br />

Transposed ¯τ of tensor τ (permutation of rows by columns):<br />

⎛<br />

⎞<br />

τ 11 τ 21 τ 31<br />

⎜<br />

⎟<br />

¯τ = ⎝ τ 12 τ 22 τ 32 ⎠<br />

τ 13 τ 23 τ 33


3.11. APPENDIX: NOTATION AND REPERTOIRE OF VECTORIAL AND TENSORIAL FORMULAE 85<br />

Symmetric tensor: τ = ¯τ .<br />

If τ is symmetric:<br />

τ ij = τ ji (i, j = 1, 2, 3)<br />

Product of a vector by a tensor:<br />

ā · τ = ī 1 (a 1 τ 11 + a 2 τ 21 + a 3 τ 31 )<br />

+ ī 2 (a 1 τ 12 + a 2 τ 22 + a 3 τ 32 )<br />

+ ī 3 (a 1 τ 13 + a 2 τ 23 + a 3 τ 33 ) (vector)<br />

Divergence of a tensor:<br />

∇ · τ =<br />

(<br />

∂τ11<br />

ī 1 + ∂τ 21<br />

+ ∂τ )<br />

31<br />

∂x 1 ∂x 2 ∂x 3<br />

(<br />

∂τ12<br />

+ ī 2 + ∂τ 22<br />

+ ∂τ )<br />

32<br />

∂x 1 ∂x 2 ∂x 3<br />

(<br />

∂τ13<br />

+ ī 3 + ∂τ 23<br />

+ ∂τ )<br />

33<br />

∂x 1 ∂x 2 ∂x 3<br />

(vector)<br />

Gradient of a vector:<br />

⎛<br />

∇ ā =<br />

⎜<br />

⎝<br />

∂a 1<br />

∂x 1<br />

∂a 2<br />

∂x 1<br />

∂a 3<br />

∂x 1<br />

∂a 1<br />

∂x 2<br />

∂a 2<br />

∂x 2<br />

∂a 3<br />

∂x 2<br />

∂a 1<br />

∂x 3<br />

∂a 2<br />

∂x 3<br />

∂a 3<br />

∂x 3<br />

⎞<br />

⎟<br />

⎠<br />

(tensor)<br />

Double product:<br />

ā · (¯b · τ<br />

)<br />

(scalar)<br />

If τ is symmetric, then:<br />

ā · (¯b · τ<br />

)<br />

= ¯b ·<br />

(ā<br />

· τ<br />

)<br />

Contraction of two tensors:<br />

3∑ 3∑<br />

τ : τ ′ = τ ij τ ji<br />

′<br />

i=1 j=1<br />

(scalar)<br />

Therefore<br />

and<br />

τ : τ ′ = τ ′ : τ<br />

∇ · (ā · τ) = ā · (∇ · τ) + (∇ā) : τ (scalar)


86 CHAPTER 3. GENERAL EQUATIONS<br />

If τ is symmetric<br />

∇ · (ā · τ) = ā · (∇ · τ) + (∇ā) : τ = ā · (∇ · τ) + 1 (∇ā + ∇ā) : τ<br />

2<br />

The Gauss formula:<br />

In a domain V enclosed by surface S:<br />

∫∫<br />

∫∫∫<br />

(¯n ◦ ϕ) dσ =<br />

S<br />

V<br />

(∇ ◦ ϕ) dV<br />

ϕ can be a scalar, vector or tensor. Symbol ◦ represents any of the several types of<br />

products previously <strong>de</strong>fined.<br />

Derivative of an integral with changing boundary:<br />

∫∫∫<br />

∫∫∫<br />

∫∫<br />

d<br />

∂ϕ (t, ¯x)<br />

ϕ (t, ¯x) dV =<br />

dV +<br />

dt<br />

∂t<br />

V (t)<br />

V (t)<br />

Σ(t)<br />

(¯n · ¯v) ϕ dσ<br />

where ¯v is the velocity of Σ at dσ and ¯n is the outwards normal to Σ at dσ.<br />

References<br />

[1] Kuethe, A. N. and Schetzer, J. D.: Foundations of Aerodynamics. John Wiley<br />

and Sons, Inc. New York. 1950.<br />

[2] Green, R. S.: The Molecular Theory of Fluids. Interscience Publishers, Inc,<br />

New York, 1952.<br />

[3] Richardson, J. M. and Brinkley, S. R.: Mechanics of Reacting Continua. Combustion<br />

Processes, Sec. F, Vol. II of High Speed Aerodynamics and Jet Propulsion,<br />

Princeton University Press, 1956, p. 203.<br />

[4] Hirschfel<strong>de</strong>r, J. O., Curtiss, C. F. and Bird, R. B.: Molecular Theory of Liquid<br />

and Gases. John Wiley and Sons, Inc., New York, 1954.<br />

[5] De Groot, S. R.: Thermodynamics of Irreversible Processes. North Holland<br />

Publishing Comp., Amsterdam, 1951.<br />

[6] Prigogine, I. and Defay, R.: Chemical Thermodynamics. Longmans Green and<br />

C., New York, 1954.<br />

[7] von Kármán, Th.: The Theory of Shock Waves and the Second Law of Thermodynamics.<br />

L’Aerotecnica, Febr. 15, 1951, pp. 82-3.<br />

[8] Chambré, P. and Lin, C. C.: On the Steady Flow of a Gas through a Tube with<br />

Heat Exchange or Chemical Reaction. The Journal of Aeronautical Sciences,<br />

Oct. 1946, pp. 537-42.


3.11. APPENDIX: NOTATION AND REPERTOIRE OF VECTORIAL AND TENSORIAL FORMULAE 87<br />

[9] Shapiro, A. H. and Hawthorne, W. R.: The Mechanics and Thermodynamics<br />

of Steady One-Dimensional Gas Flows. Journal of Applied Mechanics, Trans.<br />

A.S.M.E., Vol. 14, No. 4, 1947, p. 317.<br />

[10] Foa, J. V. and Rudinger, G.: On the Addition of Heat to a Gas Flowing in a<br />

Pipe at Subsonic Speed. The Journal of Aeronautical Sciences, Febr. 1949,<br />

pp. 84-94.


88 CHAPTER 3. GENERAL EQUATIONS


Chapter 4<br />

Combustion Waves<br />

4.1 Detonation and <strong>de</strong>flagration<br />

When an explosive mixture is ignited, a wave or reaction front originates which propagates<br />

the combustion throughout the mixture. The wave thickness is normally very<br />

small. By assuming this thickness to be zero, that is, assuming that gases burn instantaneously<br />

as they cross the wave, the state of the mixture and the state of the burn<br />

gases can be compared, in a way similar to that used when studying shock waves; 1<br />

that is, by applying the laws of the conservation mass, momentum and energy across<br />

the front, without taking into consi<strong>de</strong>ration the intermediate transformations between<br />

the initial and final states Thus, the following system of three equations is obtained<br />

ρ 1 v 1 = ρ 2 v 2 , (4.1)<br />

ρ 1 v1 2 + p 1 = ρ 2 v2 2 + p 2 , (4.2)<br />

1<br />

2 v2 1 + h 1 = 1 2 v2 2 + h 2 , (4.3)<br />

where ρ, p and h are the <strong>de</strong>nsity, pressure and total enthalpy per unit mass and v the<br />

normal velocity relative to the reaction front. Subscripts 1 and 2 refer to the unburnt<br />

gases and to the combustion products respectively. In particular, v 1 is the propagation<br />

velocity of the wave, that is, the velocity at which the wave moves through the unburnt<br />

gases, Fig. 4.1.<br />

For a given composition, the specific enthalpy h 1 of the unburnt gases is a<br />

function of the pressure p 1 and the <strong>de</strong>nsity ρ 1 of the mixture<br />

h 1 = h 1 (p 1 , ρ 1 ) . (4.4)<br />

1 See i.e. Courant-Friedrichs: Supersonic Flow and Shock Waves. Intersciences Pub. Inc., New York,<br />

1948, pp. 121 and following.<br />

89


function of p 2 and ρ 2<br />

h 2 = h 2 (p 2 , ρ 2 ) . (4.5)<br />

90 CHAPTER 4. COMBUSTION WAVES<br />

Unburned gases<br />

Burned gases<br />

V<br />

1<br />

V 2<br />

p ρ 1 T p ρ T<br />

1 1<br />

2 2 2<br />

Combustion wave<br />

Figure 4.1: Schematic diagram of a combustion wave to obtain the relations between initial<br />

and final states.<br />

Similarly, if the burnt gases are in thermodynamic equilibrium, h 2 is only a<br />

When equations (4.4) and (4.5) are substituted into (4.3), this equation together<br />

with Eq. (4.1) and (4.2) form a system of three equations with six unknowns: p 1 , ρ 1 ,<br />

v 1 , p 2 , ρ 2 and v 2 . If three of these values are known, the system <strong>de</strong>termines the values<br />

for the other three. As said before, the state of the unburnt gases is <strong>de</strong>fined by the<br />

values p 1 and ρ 1 . Therefore, in or<strong>de</strong>r to <strong>de</strong>termine the propagation, another of these<br />

values must be known, for instance that of the propagation velocity v 1 . The analysis<br />

of the propagation velocity shows the existence of two types of essentially different<br />

waves. In fact, the elimination of v 2 between Eq. (4.1) and (4.2), gives for v 1 ,<br />

v 1 = 1 ρ 1<br />

√ √√√ p 2 − p 1<br />

1<br />

ρ 1<br />

− 1 ρ 2<br />

. (4.6)<br />

The study is simplified by introducing in the above system the specific volume<br />

τ, which is related to the <strong>de</strong>nsity ρ by<br />

τ = 1 ρ , (4.7)<br />

thus obtaining for v 1<br />

v 1 = τ 1<br />

√<br />

p2 − p 1<br />

τ 1 − τ 2<br />

. (4.8)<br />

satisfied<br />

Since the propagation velocity must be real, the following condition must be<br />

p 2 − p 1<br />

τ 1 − τ 2<br />

> 0. (4.9)


4.2. KINDS OF DETONATIONS AND DEFLAGRATIONS 91<br />

Consequently, if p 2 > p 1 , then τ 1 > τ 2 , and if p 2 < p 1 , then τ 1 < τ 2 .<br />

Therefore, when the pressure increases across the wave the gases contract, and if the<br />

pressure <strong>de</strong>creases, the gases expand. Both types of waves are physically observed.<br />

The first type is called <strong>de</strong>tonation and the second <strong>de</strong>flagration. The propagation velocity<br />

of a <strong>de</strong>tonation wave in an explosive mixture is of the or<strong>de</strong>r of several thousand<br />

meters per second. The propagation velocity of a <strong>de</strong>flagration wave is normally of the<br />

or<strong>de</strong>r of 1 m/s.<br />

4.2 Kinds of <strong>de</strong>tonations and <strong>de</strong>flagrations<br />

The study can be simplified by representing the state of unburnt and burnt gases in the<br />

diagram (p, τ), Fig. 4.2, proposed by Grussard [1].<br />

p<br />

III<br />

H<br />

I<br />

p 2<br />

A<br />

p 1<br />

II<br />

P<br />

B<br />

IV<br />

0 τ0 τ1 τ2 τ3<br />

Figure 4.2: Sketch of the Hugoniot curve for the final states.<br />

D<br />

τ<br />

In this diagram, let P (p 1 , τ 1 ) be the representative point of the initial state.<br />

Condition (4.9) exclu<strong>de</strong>s regions I and II from this diagram, since in these regions<br />

(p 2 − p 1 ) / (τ 1 − τ 2 ) is smaller than zero. The representative points of the final states<br />

must be, consequently, within regions III and IV. The points in region III correspond to<br />

<strong>de</strong>tonations, since, in this region p 2 > p 1 and τ 2 < τ 1 . Points in region IV correspond<br />

to <strong>de</strong>flagrations.<br />

The possible final states corresponding to the initial state (p 1 , τ 1 ) and compatible<br />

with conditions (4.1), (4.2) and (4.3), lie on a curve H, called the Hugoniot curve.


92 CHAPTER 4. COMBUSTION WAVES<br />

This curve is obtained from the elimination of v 1 and v 2 between Eqs. (4.1), (4.2) and<br />

(4.3). There results for H, the following equation<br />

h 2 − h 1 − 1 2 (τ 1 + τ 2 ) (p 2 − p 1 ) = 0. (4.10)<br />

This curve has a <strong>de</strong>tonation branch and a <strong>de</strong>flagration branch.<br />

Let p 2 be the final pressure corresponding to an adiabatic combustion at constant<br />

volume τ 1 . Since the reaction is exothermic p 2 > p 1 and the representative point<br />

of the state (p 2 , τ 1 ), will be, for instance, point A. The <strong>de</strong>tonation branch starts at this<br />

point. The propagation velocity corresponding to this <strong>de</strong>tonation at constant volume<br />

is, according to Eq. (4.8), infinite. Hence, at this point the combustion propagates<br />

instantaneously throughout the mass.<br />

Similarly, let τ 2 be the specific volume corresponding to an adiabatic combustion<br />

at constant pressure p 1 . Here τ 2 > τ 1 , and point B, corresponding to the state p 1 ,<br />

τ 2 , is the starting point of the <strong>de</strong>flagration branch. The propagation velocity of this<br />

limit constant pressure <strong>de</strong>flagration is, according to (4.8), zero.<br />

Hereinafter, it is assumed that the Hugoniot curve satisfies the following conditions<br />

( ) ( ∂p<br />

∂ 2 p<br />

< 0 ,<br />

∂τ<br />

H<br />

∂τ<br />

)H<br />

2 > 0, (4.11)<br />

where subscript H indicates differentiation along the Hugoniot curve. These conditions<br />

mean that H is monotonically <strong>de</strong>creasing and turns its concavity to the axis<br />

p > 0. Both conditions correspond to the curves generally observed in practice. In<br />

general, curve H has an asymptote, parallel to the pressure axis, for τ = τ 0 > 0, and<br />

cuts the specific volume axis at point τ 3 . Therefore, the general form of H is the one<br />

shown in Fig. 4.2.<br />

Let us consi<strong>de</strong>r a straight line that starts from point P , corresponding to the<br />

initial state, and enters in region III of <strong>de</strong>tonations. Let α, Fig. 4.3(a), be the angle<br />

between this line and the negative direction of the axis. Depending on the values of α,<br />

the three following cases are possible:<br />

1) If α is smaller than a certain limit value α min which corresponds to the tangent<br />

from P to H, the straight line corresponding to α cannot cut the Hugoniot curve.<br />

2) If α = α min , the corresponding line is, as aforesaid, tangent to H at point J.<br />

3) If α > α min , the corresponding line cuts the Hugoniot curve at two points E<br />

and E ′ at different si<strong>de</strong>s of point J.


4.2. KINDS OF DETONATIONS AND DEFLAGRATIONS 93<br />

p<br />

C<br />

E’<br />

p<br />

1<br />

p<br />

p 1<br />

P<br />

B<br />

J<br />

α<br />

α min<br />

E<br />

A<br />

P<br />

τ 0 1<br />

(a) Detonation branch<br />

τ<br />

τ<br />

J’<br />

D<br />

τ 1 τ 3<br />

τ<br />

(b) Deflagration branch<br />

Figure 4.3: Chapman-Jouguet points in the <strong>de</strong>tonation and <strong>de</strong>flagration branches of the<br />

Hugoniot curve.<br />

These properties can immediately be translated into properties for the propagation<br />

velocity of the <strong>de</strong>tonation. In fact, we have 2<br />

tan α = p 2 − p 1<br />

τ 1 − τ 2<br />

. (4.12)<br />

Taking this value into Eq. (4.8), the following expression for the propagation velocity<br />

is obtained<br />

v 1 = τ 1<br />

√<br />

tan α. (4.13)<br />

Therefore, it results that the propagation velocity of a <strong>de</strong>tonation is minimum<br />

at point J, where the straight line P J is tangent to H. This property was observed<br />

by Chapman in 1899 [2].<br />

Point J is called the Chapman-Jouguet point and the<br />

corresponding <strong>de</strong>tonation the Chapman-Jouguet <strong>de</strong>tonation. The Chapman-Jouguet<br />

<strong>de</strong>tonation is important since it is the one usually observed. Hence, one conclu<strong>de</strong>s<br />

that of all the possible <strong>de</strong>tonations, compatible with the given initial conditions, the<br />

Chapman-Jouguet <strong>de</strong>tonation is the one which propagates at a minimum velocity.<br />

2 To account for dimensions a constant should be inclu<strong>de</strong>d. Hereinafter it is omitted for simplicity.


94 CHAPTER 4. COMBUSTION WAVES<br />

The <strong>de</strong>tonations corresponding to the upper part JC of H, for which the pressure<br />

of the burnt gases is greater than the pressure of the Chapman-Jouguet <strong>de</strong>tonation,<br />

have been called by Courant and Friedrichs strong <strong>de</strong>tonations and those corresponding<br />

to the lower part AJ, weak <strong>de</strong>tonations.<br />

In the same manner it can be proven, Fig. 4.3(b), that in the <strong>de</strong>flagration branch<br />

there exists a point J ′ in which H and P J ′ are tangent and the corresponding <strong>de</strong>flagration,<br />

also called Chapman-Jouguet, propagates with a maximum velocity. Point J ′<br />

divi<strong>de</strong>s the <strong>de</strong>flagration branch in two parts, one J ′ D of strong <strong>de</strong>flagrations and the<br />

other BJ ′ of weak <strong>de</strong>flagrations.<br />

4.3 Velocity of the burnt gases<br />

It is interesting to <strong>de</strong>termine the subsonic or supersonic character of the velocity v 2<br />

of the burnt gases with respect to the wave front, since the stability of the <strong>de</strong>tonation<br />

wave <strong>de</strong>pends on the character of this velocity, that is, the possibility for this wave to<br />

propagate unchanged through the gas. In or<strong>de</strong>r to <strong>de</strong>termine the character of v 2 it is<br />

necessary to compare this velocity with the sound velocity a 2 in the burnt gases. Let<br />

us see how this can be attained.<br />

Gas Dynamics shows 3 that the sound velocity a in a gas is given by the expression<br />

a = τ<br />

√<br />

−<br />

( ) ∂p<br />

, (4.14)<br />

∂τ<br />

S<br />

where subscripts S means differentiation along the isentropic that passes through the<br />

representative point of the gas state.<br />

Let α S , Fig. 4.4, be the angle between the tangent to the isentropic that passes<br />

through the point (p 2 , τ 2 ) of the <strong>de</strong>tonation branch, and the negative direction of x<br />

axis. From Eq. (4.14) we have<br />

a 2 = τ 2<br />

√ tan αS . (4.15)<br />

On the other hand, the velocity v 2 of the burnt gases with respect to the flame front,<br />

in virtue of Eq. (4.1), is v 2 = τ 2<br />

τ 1<br />

v 1 . Then, because of Eq. (4.13), this velocity can be<br />

expressed as<br />

v 2 = τ 2<br />

√<br />

tan α , (4.16)<br />

where α is the previously <strong>de</strong>fined angle. If α is larger, equal to or smaller than α s ,<br />

then v 2 will be larger, equal to or smaller than a 2 ; and the flow of the burnt gases<br />

3 See Courant-Friedrichs: Supersonic Flow and Shock Waves. Intersciences Pub. Inc., New York, 1948.


4.3. VELOCITY OF THE BURNT GASES 95<br />

α S<br />

α > : supersonic α = : sonic α < : subsonic<br />

α S<br />

α S<br />

τ ,<br />

α<br />

α S<br />

( τ , )<br />

( τ , )<br />

( τ , )<br />

2 p 2<br />

S<br />

α<br />

2 p 2<br />

α<br />

α S<br />

S<br />

2 p 2<br />

αS<br />

P( τ 1,<br />

p 1 )<br />

P( τ 1,<br />

p 1 )<br />

P( 1 p 1 )<br />

S<br />

Figure 4.4: Schematic diagram showing the line connecting the initial and final states and<br />

the corresponding isentropic curve.<br />

relative to the wave will be supersonic, sonic or subsonic. Hence, the problem reduces<br />

to a comparison between the slope of the isentropic that passes through each point of<br />

H, and that of the radius vector that joins this point with point P of the initial state.<br />

The three possible cases appear schematically in Fig. 4.4.<br />

To perform this comparison let us start by studying the variation of the entropy<br />

s, along the Hugoniot curve. The elemental entropy variation ds corresponding to the<br />

variations dh and dp of enthalpy h and pressure p, is<br />

T ds = dh − τ dp. (4.17)<br />

Now, by differentiating (4.10), keeping p 1 and τ 1 constant and taking the result into<br />

(4.17), the following expression for the entropy variation along H, is obtained<br />

( ∂s2<br />

T 2 =<br />

∂τ 2<br />

)H<br />

τ [ ( ]<br />

1 − τ 2 p2 − p 1 ∂p2<br />

+<br />

2 τ 1 − τ 2 ∂τ 2<br />

)H<br />

(4.18)<br />

= (τ 1 − τ 2 ) (tan α − tan α H )<br />

,<br />

2<br />

where α H is the angle between the tangent to H at point p 2 , τ 2 , and the negative<br />

direction of τ axis.<br />

Fig. 4.3 shows that in the strong <strong>de</strong>tonation branch α H<br />

virtue of (4.18), in this branch<br />

( ∂s2<br />

∂τ 2<br />

)H<br />

> α. Therefore, in<br />

< 0. (4.19)<br />

On the contrary, in the weak <strong>de</strong>tonation branch α H < α, that is<br />

( ∂s2<br />

> 0. (4.20)<br />

∂τ 2<br />

)H


96 CHAPTER 4. COMBUSTION WAVES<br />

Finally, at the Chapman-Jouguet point α H = α, that is<br />

( ∂s2<br />

= 0. (4.21)<br />

∂τ 2<br />

)H<br />

From these three inequalities it results that the entropy of the burnt gases is<br />

minimum at the Chapman-Jouguet point and increases from there on, both in strong<br />

and weak <strong>de</strong>tonation branches.<br />

Since at point J condition (4.21) is satisfied, the isentropic passing through this<br />

point is tangent to the Hugoniot curve. Therefore, at point J the following conditions<br />

are satisfied<br />

α S = α H = α = α min . (4.22)<br />

These values, when taken into Eqs. (4.15) and (4.16), give at point J<br />

a 2 = v 2 , (4.23)<br />

thus resulting the following important property: in a Chapman-Jouguet <strong>de</strong>tonation<br />

the velocity of the burnt gases with respect to the wave is sonic. This property is<br />

the one that best characterizes a Chapman-Jouguet <strong>de</strong>tonation. As will presently be<br />

seen, owing to this property, the <strong>de</strong>tonations physically observed are of this same type.<br />

Jouguet stated this property in 1905 [3].<br />

For the purpose of <strong>de</strong>termining the character of v 2 in strong and weak <strong>de</strong>tonations,<br />

it is necessary, as previously seen in Fig. 4.3, to compare α and α s . This<br />

means that it is necessary to <strong>de</strong>termine the relative position of the isentropic that passes<br />

through each point of H, with respect to the straight line that joins P to that point. For<br />

the <strong>de</strong>termination of this relative position it is not sufficient to know the entropy variation<br />

along H, but it is also necessary to know the entropy variation along another<br />

direction, for instance, that of the radius vector with its origin at point P . This can<br />

easily be attained by consi<strong>de</strong>ring the function<br />

F (p 1 , τ 1 ; p 2 , τ 2 ) ≡ h 2 − h 1 − 1 2 (τ 1 + τ 2 ) (p 2 − p 1 ) . (4.24)<br />

On the Hugoniot curve H the value of this function is zero, as results from (4.10).<br />

This curve divi<strong>de</strong>s the plane in two regions: the lower region, that contains point P ,<br />

representative of the initial state, and the upper region. F takes opposite signs in both<br />

regions. As can be easily be verified, in the lower region F < 0 and in the upper<br />

region F > 0. For this purpose, it is enough to study, for instance, the sign of F in P .<br />

In B we have (Fig. 4.2)<br />

F (p 1 , τ 1 ; p 1 , τ 2 ) = h 2 (p 1 , τ 2 ) − h 1 (p 1 , τ 1 ) = 0. (4.25)


4.3. VELOCITY OF THE BURNT GASES 97<br />

Similarly, in P<br />

F (p 1 , τ 1 ; p 1 , τ 1 ) = h 2 (p 1 , τ 1 ) − h 1 (p 1 , τ 1 ) . (4.26)<br />

Now subtracting (4.25) from (4.26), results<br />

F (p 1 , τ 1 ; p 1 , τ 1 ) = h 2 (p 1 , τ 1 ) − h 2 (p 1 , τ 2 ) . (4.27)<br />

But, τ 1 is smaller than τ 2 , and, since to reduce the volume of the burnt gases from τ 2<br />

to τ 1 without varying their pressure p 1 it is necessary to cool the burnt gases, that is to<br />

reduce their enthalpy, then<br />

h 2 (p 1 , τ 1 ) < h 1 (p 1 , τ 2 ) , (4.28)<br />

that is<br />

F (p 1 , τ 1 ; p 1 , τ 1 ) < 0. (4.29)<br />

Therefore F is negative in the lower region and positive in the upper region.<br />

By means of the function F , the entropy variation can be expressed in a simple<br />

manner by differentiating Eq. (4.24), keeping p 1 and τ 1 constant and then combining<br />

the result with Eq. (4.17), thus, obtaining<br />

T 2 ds 2 = dF + 1 2 (p 2 − p 1 ) dτ 2 + 1 2 (τ 1 − τ 2 ) dp 2 . (4.30)<br />

But the variations dp 2 and dτ 2 corresponding to the straight line that joins P with<br />

(p 2 , τ 2 ) must satisfy condition<br />

dp 2<br />

dτ 2<br />

= p 2 − p 1<br />

τ 2 − τ 1<br />

, (4.31)<br />

this expression, when taken into Eq. (4.30), gives for entropy variation along a radius<br />

vector<br />

T 2 ds 2 = dF. (4.32)<br />

Therefore, along a radius vector, F and s 2 increase and <strong>de</strong>crease in the same direction.<br />

Now, at point E (Fig. 4.5), F increases from E to E ′ . Therefore, s 2 also<br />

increases from E to E ′ . In the same manner, at point E ′ , F and s 2 increase in the<br />

direction E ′ E.<br />

Furthermore, s 2 <strong>de</strong>crease along H in the directions E ′ J and EJ. Thereby it is<br />

<strong>de</strong>duced that the isentropic curves that pass by E and E ′ must, necessarily, have the<br />

positions shown in Fig. 4.5. There results that:<br />

a) In E, α S < α and the velocity of the burnt gases is supersonic.<br />

b) In E ′ , α S > α and the velocity of the burnt gases is subsonic.


98 CHAPTER 4. COMBUSTION WAVES<br />

p<br />

C<br />

E’<br />

S 2<br />

S 2<br />

J<br />

E<br />

A<br />

p<br />

1<br />

P<br />

τ 0 1<br />

Figure 4.5: Position of the isentropic curves in the <strong>de</strong>tonation branch of the Hugoniot curve.<br />

τ<br />

τ<br />

Hence the velocity of the burnt gases is supersonic in the weak <strong>de</strong>tonations and subsonic<br />

in the strong <strong>de</strong>tonations.<br />

Similarly, it can be <strong>de</strong>monstrated that the entropy of the burn gases in the <strong>de</strong>flagration<br />

branch is maximum at J ′ . It can also be <strong>de</strong>monstrated that the velocity of<br />

the unburnt gases is subsonic in the weak <strong>de</strong>flagrations, sonic in the Chapman-Jouguet<br />

<strong>de</strong>flagration and supersonic in the strong <strong>de</strong>flagrations.<br />

4.4 Propagation velocity<br />

In or<strong>de</strong>r to complete the study, we shall analyze the character of the propagation velocity<br />

v 1 , by comparing its values with the value of the velocity a 1 of the sound propagation<br />

in the state (p 1 , τ 1 ). The study can be ma<strong>de</strong> in an i<strong>de</strong>ntical manner to that<br />

previously used for v 2 , that is by consi<strong>de</strong>ring all the states (p 1 , τ 1 ) of the unburnt gases<br />

that are compatible with a given state (p 2 , τ 2 ) of the burnt gases. All these states lie<br />

on a Hugoniot curve H ′ which is obtained by fixing in Eq. (4.10) the values of p 2 and<br />

τ 2 and varying p 1 and τ 1 . This curve, like curve H, has two branches (Fig. 4.6): a<br />

lower <strong>de</strong>tonation branch and an upper <strong>de</strong>flagration branch. It can easily be proved that<br />

in the former, the entropy of the unburnt gases <strong>de</strong>creases starting from A, and in the<br />

latter increases starting from B. Furthermore by studying the variation of function F ,<br />

which is <strong>de</strong>fined by (4.24) fixing the values of p 2 and τ 2 and varying p 1 and τ 1 , we<br />

obtain that H ′ divi<strong>de</strong>s the plane (p, τ) in two regions. The upper region containing


4.4. PROPAGATION VELOCITY 99<br />

p H’<br />

S 2<br />

B ( τ 2<br />

, p 2<br />

)<br />

p 2<br />

S 2<br />

A<br />

H’<br />

Figure 4.6: Position of the isentropic curves in the Hugoniot curve of the admissible initial<br />

states.<br />

τ 2<br />

τ<br />

point (p 2 , τ 2 ), in which F < 0, and the lower region, in which F > 0. Finally, it<br />

is also verified that along the radius vectors that pass through the point (p 2 , τ 2 ) the<br />

variations of F and s 1 are opposite. Now, by combining these properties as previously<br />

done for v 2 , we can <strong>de</strong>termine the relative position of the isentropic that passes<br />

through each point H ′ with respect to the straight line that joins this point to the point<br />

(p 2 , τ 2 ). It results that this position is the one indicated in Fig. 4.6. Therefore, in the<br />

<strong>de</strong>tonation branch α < α S and in the <strong>de</strong>flagration branch α > α S . Therefore, one<br />

conclu<strong>de</strong>s that the propagation velocity of a <strong>de</strong>tonation is always supersonic, and the<br />

propagation velocity of a <strong>de</strong>flagration is always subsonic.<br />

As a result of the preceding study, the Table 4.1 can be constructed. This<br />

table contains the results of Jouguet’s studies and <strong>de</strong>fines the characteristics of the<br />

combustion waves.<br />

Detonations<br />

Weak Chapman-Jouguet Strong<br />

Velocity before<br />

Supersonic<br />

Velocity after Supersonic Sonic Subsonic<br />

Deflagrations<br />

Weak Chapman-Jouguet Strong<br />

Velocity before<br />

Subsonic<br />

Velocity after Subsonic Sonic Supersonic<br />

Table 4.1: Characteristics of the combustion waves.


100 CHAPTER 4. COMBUSTION WAVES<br />

4.5 Applications<br />

As an application of the previous study we shall consi<strong>de</strong>r the propagation of a combustion<br />

wave, assuming that both the unburnt and burnt gases are perfect gases and<br />

that the chemical composition of the burnt gases is in<strong>de</strong>pen<strong>de</strong>nt from their state. In<br />

this case, the enthalpy of the burnt gases can be expressed in the form<br />

∫ T2<br />

h 2 = h 21 + c p2 dT, (4.33)<br />

T 1<br />

where h 21 is their enthalpy at the temperature T 1 . Moreover, it is assumed that the<br />

heat capacity c p2 is in<strong>de</strong>pen<strong>de</strong>nt from the temperature in the interval (T 1 , T 2 ). Then<br />

Eq. (4.33) reduces to<br />

h 2 = h 21 + c p2 (T 2 − T 1 ) . (4.34)<br />

Now, by taking this expression into the Hugoniot equation (4.10), we have<br />

c p2 T 2 = (h 11 − h 21 ) + c p2 T 1 + 1 2 (τ 1 + τ 2 ) (p 2 − p 1 ) , (4.35)<br />

where h 11 − h 21 is the difference between the formation enthalpies of the burnt and<br />

unburnt gases at the temperature T 1 of the unburnt gases. Let Q be this value. Then<br />

Eq. (4.35) can be written in the form<br />

c p2 T 2 = Q + c p2 T 1 + 1 2 (τ 1 + τ 2 ) (p 2 − p 1 ) . (4.36)<br />

Let<br />

p 2 τ 2 = R 2 T 2 (4.37)<br />

be the state equation of the burn gases, and γ 2 = c p2 /c v2 the ratio of their capacities.<br />

Furthermore, let x = τ 2 /τ 1 and y = p 2 /p 1 the specific volume and pressure of the<br />

burnt gases, referred to the corresponding values for the unburnt gases, and<br />

q = 2<br />

( ) ( )<br />

γ2 − 1 Q<br />

+ 2γ 2ν<br />

γ 2 + 1 p 1 τ 1 γ 2 + 1 − γ 2 − 1<br />

γ 2 + 1 , (4.38)<br />

where ν = τ ′ 1/τ 1 and τ ′ 1 is the specific volume that would correspond to the burnt<br />

gases at the pressure p 1 and temperature T 1 of the unburnt gases, that is,<br />

ν = R 2<br />

R 1<br />

= M u<br />

M b<br />

, (4.39)<br />

where M u and M b are the average mole masses for the unburnt gases and for the<br />

combustion products respectively.


4.5. APPLICATIONS 101<br />

By substituting these values and that of the temperature given by Eq. (4.37)<br />

into Eq. (4.36), the following expression, for the Hugoniot curve, is obtained<br />

xy = q + γ 2 − 1<br />

(y − x) . (4.40)<br />

γ 2 + 1<br />

Therefore, the Hugoniot curve is an equilateral hyperbola, with an asymptote parallel<br />

to the axis y for x 0 = (γ 2 − 1) / (γ 2 + 1). This curve cuts axis x at the point x 1 =<br />

q (γ 2 + 1) / (γ 2 − 1).<br />

Let us see how the Chapman-Jouguet points are <strong>de</strong>termined. The equation of<br />

an isentropic for the burnt gases is<br />

yx γ2 = const. (4.41)<br />

Therefore, the Chapman-Jouguet points are obtained by solving the system formed<br />

by Eq. (4.40) and by the equations that express the tangency condition of the curves<br />

Eq. (4.40) and Eq. (4.41), namely,<br />

(γ 2 + 1) y + (γ 2 − 1)<br />

(γ 2 + 1) x − (γ 2 − 1) = γ y<br />

2<br />

x . (4.42)<br />

The solution of Eq. (4.40) and Eq. (4.42) gives for x and y the following equations<br />

y 2 − ( (γ 2 + 1)q − (γ 2 − 1) ) y + q = 0, (4.43)<br />

x 2 − 1 γ 2<br />

(<br />

(γ2 + 1)q + (γ 2 − 1) ) x + q = 0. (4.44)<br />

When solving these equations the root x 1 < 1 must be assign to the root y 1 > 1 and<br />

root x 2 > 1 to the root y 2 < 1.<br />

In virtue of Eq. (4.8), the propagation velocity is<br />

v 1<br />

= 1<br />

√<br />

y − 1<br />

√<br />

a 1 γ1 1 − x , (4.45)<br />

where a 1 is the sound velocity of the unburnt gases, and γ 1 the ratio of their heat<br />

capacities.<br />

Similarly, the velocity of the burnt gases with respect to the wave front is<br />

v 2<br />

=<br />

x<br />

√<br />

y − 1<br />

√<br />

a 1 γ1 1 − x . (4.46)<br />

The corresponding Mach number M 2 is<br />

√<br />

M 2 = √ 1 ( )<br />

x y − 1<br />

. (4.47)<br />

γ2 y 1 − x


102 CHAPTER 4. COMBUSTION WAVES<br />

By subtracting from the left hand si<strong>de</strong> of Eq. (4.42) the right hand si<strong>de</strong> multiplied<br />

by two, we obtain<br />

y<br />

γ 2<br />

x = y − 1<br />

1 − x , (4.48)<br />

which, together with Eq. (4.47), show that at the Chapman-Jouguet points, the velocity<br />

of the burnt gases with respect to the combustion point is equal to the sound velocity.<br />

For example, by taking the typical values p 1 = 1 atm, τ 1 = 1000 cm 3 /gr,<br />

T 1 = 300 K, γ 1 = γ 2 = 1.4, Q = 1000 cal gr −1 , c p2 = 0.46 cal gr −1 K −1 and ν = 1<br />

the following is obtained for the Hugoniot curve<br />

xy = 15.23 + 1 (y − x) . (4.49)<br />

6<br />

This curve has been represented in Fig. 4.7. The branch of <strong>de</strong>tonations (Fig. 4.7(a))<br />

corresponds to the values of x in the interval<br />

1<br />

6 ≤ x ≤ 1 . (4.50)<br />

The branch of <strong>de</strong>flagrations (Fig. 4.7(b)) corresponds to the interval<br />

13.2 ≤ x ≤ 91.4 . (4.51)<br />

The Chapman-Jouguet points for <strong>de</strong>tonations and <strong>de</strong>flagrations are, respectively,<br />

Detonations: x 1 = 0.58, y 1 = 35.7 .<br />

Deflagrations: x 2 = 24.82, y 2 = 0.45 .<br />

The corresponding propagation velocities are<br />

Chapman-Jouguet <strong>de</strong>tonation:<br />

Chapman-Jouguet <strong>de</strong>flagration:<br />

v 1<br />

= 7.68 .<br />

a 1<br />

v 1<br />

= 0.126 .<br />

a 1<br />

4.6 Remarks<br />

If the variation of the chemical composition along the Hugoniot curve is taken into<br />

account, that is the influence of the dissociation of the combustion products on the<br />

characteristics of the wave, the problem is far more complicated. To attain a solution<br />

one must resort to the equations of thermodynamic equilibrium, which must be combined<br />

with the state equations and the invariants across the wave. Then, the Hugoniot<br />

curve must be drawn point by point. For additional information see Ref. [4].


4.7. INDETERMINACY OF THE SOLUTION 103<br />

y<br />

125<br />

100<br />

x =1/6 0<br />

(a) Detonation branch<br />

75<br />

50<br />

J<br />

25<br />

0<br />

P A<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

y<br />

1.0<br />

0.8<br />

P<br />

B<br />

(b) Deflagration branch<br />

x<br />

0.6<br />

0.4<br />

J’<br />

0.2<br />

0.0<br />

0<br />

20<br />

40<br />

60<br />

80<br />

100<br />

x<br />

Figure 4.7: Hugoniot curve (x = τ/τ 1, y = p/p 1) corresponding to p 1 = 1 atm, τ 1 =<br />

1000 cm 3 /gr, T 1 = 300 K, γ 1 = γ 2 = 1.4, Q = 1000 cal gr −1 , c p2 = 0.46<br />

cal gr −1 K −1 , ν = 1.<br />

4.7 In<strong>de</strong>terminacy of the solution<br />

The study performed in this chapter allows the establishment of the different types of<br />

combustion waves that are compatible with the principles of the conservation of mass,<br />

momentum and energy. This has been attained by comparison between the initial and<br />

final state of the gases. In or<strong>de</strong>r to obtain this result it has not been necessary to take<br />

into account the intermediate states un<strong>de</strong>rgone by the gas, within the wave. Thus, the<br />

problem is essentially simplified. However, if the influence of the intermediate states<br />

is neglected the result obtained is incomplete, since the propagation velocity of the<br />

combustion wave remains un<strong>de</strong>termined, whilst experience shows that it takes a well<br />

<strong>de</strong>fined value for each different case. In or<strong>de</strong>r to eliminate the said in<strong>de</strong>terminacy and<br />

select among the combustion waves obtained herein those that will actually occur, it<br />

is necessary to analyze the internal structure of the wave and see the mechanism that


104 CHAPTER 4. COMBUSTION WAVES<br />

propagates the reaction to the unburnt gases. This study will be the subject of the<br />

following chapter. Therein, it will be seen that:<br />

1) The weak <strong>de</strong>tonations and strong <strong>de</strong>flagrations are impossible.<br />

2) The <strong>de</strong>tonations physically observed must be of the Chapman-Jouguet type.<br />

3) Deflagrations propagate with a well <strong>de</strong>fined velocity.<br />

References<br />

[1] Crussard, L.: On<strong>de</strong>s <strong>de</strong> choc et on<strong>de</strong> explosive. Bulletin <strong>de</strong> la Société <strong>de</strong><br />

l’Industrie Minérale, Vol. VI, 1907.<br />

[2] Chapman, D. L.: On the rate of explosion in gases. Phylosophical Magazine,<br />

Jan. 1899.<br />

[3] Jouguet, E.: Sur la propagation <strong>de</strong>s réactions chimiques dans les gaz. Journal<br />

<strong>de</strong> Mathématiques Pures et Appliquées, 1905-6.<br />

[4] Taylor, J.: Detonation in Con<strong>de</strong>nsed Explosives. Oxford University Press, 1952.


Chapter 5<br />

Structure of the combustion<br />

waves<br />

5.1 Introduction<br />

The different types of combustion waves compatible with the principles of the conservation<br />

of mass, momentum and energy throughout the wave have been previously<br />

established in chapter 4. Table 4.1 of that chapter summarizes the results of the said<br />

analysis. These results where attained simply by comparing the initial and final state<br />

of the gases. The conclusions are correct provi<strong>de</strong>d that the wave thickness is small<br />

enough, so that its geometrical shape has no influence upon the transformations that<br />

occur within the wave.<br />

Owing to its nature, the aforementioned analysis is necessary incomplete, due<br />

to the fact that the possible influence of the intermediate states is ignored. Therefore,<br />

it is not to be expected that all the types of combustion waves compatible with the said<br />

principles of conservation will actually occur. On the contrary, when analyzing the<br />

internal mechanism of the propagation of the process, the existence of new limitations<br />

is to be expected. It should also be expected that such limitations will enable us to<br />

eliminate the in<strong>de</strong>terminacy, mentioned at the end of the preceding chapter, and to<br />

select from all the waves compatible with the laws of conservation the one which will<br />

actually occur in each case.<br />

To make this point clear, it is necessary to take into consi<strong>de</strong>ration the internal<br />

structure of the wave. This is done in the present chapter, following a similar procedure<br />

to the one used in Gas Dynamics [1] when studying the internal structure of<br />

shock waves.<br />

105


106 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

For simplicity, the study herein will be restricted to the case of a perfect gas<br />

with heat capacity and number of moles constant throughout the wave. The limiting<br />

solutions, obtained by assuming that a characteristic time of the thermodynamic<br />

transformations is very small compared to a characteristic time of the chemical transformations<br />

(as explained in §3), will be carefully analyzed. This plausible assumption<br />

enables an essential simplification of the equations, favouring the study of the properties<br />

of the corresponding solutions. This analysis will follow, mainly, the line of von<br />

Kármán’s reasoning [2]. More <strong>de</strong>tailed studies, for example those of Friedrichs [3] and<br />

Hirschfel<strong>de</strong>r [4], which take into account the influence of the terms neglected herein,<br />

show that the discrepancy between both cases is very small. Therefore, the method<br />

followed in this study is well justified. Furthermore, these studies allow one to <strong>de</strong>rive<br />

conclusions for the cases in which, due to the existence of ”abnormal” reaction rates,<br />

the simplified treatment <strong>de</strong>veloped herein is not applicable.<br />

In the following analysis special attention is given to the study of <strong>de</strong>tonations<br />

(see §5), since <strong>de</strong>flagrations are the subject of a <strong>de</strong>tailed analysis in the following<br />

chapter.<br />

5.2 Wave equations<br />

Let us consi<strong>de</strong>r an in<strong>de</strong>finite plane wave which propagates in undisturbed uniform<br />

combustible mixture. We shall adopt a reference system fixed to the wave were the x<br />

axis is parallel to the propagation direction and positive towards the burnt gases. With<br />

respect to this reference system, the process is stationary and x is the only in<strong>de</strong>pen<strong>de</strong>nt<br />

variable. The values of x vary from x = −∞ for the unburnt gases to x = +∞ for<br />

the burnt gases.<br />

For simplification it will be assumed that the following consi<strong>de</strong>rations are satisfied<br />

throughout the wave:<br />

1) The mixture behaves as a perfect gas.<br />

2) The heat capacity at constant pressure c p is in<strong>de</strong>pen<strong>de</strong>nt from the temperature<br />

and composition of the mixture.<br />

3) The ratio γ of the heat capacities is in<strong>de</strong>pen<strong>de</strong>nt from the composition of the<br />

mixture.<br />

4) The chemical composition of the mixture at each point of the wave is <strong>de</strong>termined<br />

by only one chemical parameter. We shall adopt as chemical parameter the <strong>de</strong>gree<br />

of advancement of the combustion, measured by the mass fraction ε(x) of<br />

the gas that has burnt when the point x of the wave is reached. Due to the action


5.2. WAVE EQUATIONS 107<br />

of diffusion, this <strong>de</strong>gree of advancement of the combustion differs from the mass<br />

fraction Y (x) of the burnt gases that <strong>de</strong>fines the mixture composition at point x,<br />

that is, the fraction Y (x) of burnt gases that would be obtained by analyzing the<br />

composition of a gas sample taken from the said point.<br />

The simplifying assumptions previously stated do not limit the extent of the<br />

study that follows, which is of a qualitative nature. On the other hand, these assumptions<br />

simplify calculations.<br />

The wave equations are obtained by particularizing the general equations of<br />

continuity, momentum and energy 1 to the case of a one-dimensional stationary motion<br />

(∂/∂y = ∂/∂z = ∂/∂t = 0). When this is done, and in addition to the preceding<br />

assumptions are taken into account, the following system of equations is obtained:<br />

a) Continuity equation.<br />

that is<br />

d (ρv)<br />

dx<br />

ρv = m,<br />

= 0, (5.1)<br />

(5.1.a)<br />

where m is the mass flow normal to the wave, per unit surface.<br />

b) Continuity equation for the burnt gases. Since the mass flow per unit surface is<br />

ρv and the burnt fraction is ε, the mass flow of the burnt gases at section x is<br />

ρvε. Its variation with x is due to chemical reactions. The equation that gives<br />

this variation is<br />

d (ρvε)<br />

= w, (5.2)<br />

dx<br />

where w is the reaction rate. In virtue of (5.1.a), equation (5.2) can also be<br />

written in the form<br />

c) Momentum equation.<br />

m dε<br />

dx = w.<br />

ρv dv<br />

dx = − dp<br />

dx + 4 3<br />

(5.2.a)<br />

(<br />

d<br />

µ dv )<br />

. (5.3)<br />

dx dx<br />

This equation can be immediately integrated by taking into account (5.1.a), giving<br />

p + mv − 4 3 µ dv<br />

dx = i,<br />

(5.3.a)<br />

where i is an integration constant that must be <strong>de</strong>termined by the boundary conditions.<br />

2<br />

1 See chapter 3.<br />

2 In equations (5.3) and (5.4) it has been assumed that the second viscosity coefficient of the mixture<br />

(see chapter 2) is zero. When not, it is sufficient to substitute µ for `µ + 3 4 µ′´ in (5.3) and (5.4).


108 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

d) Energy equation.<br />

ρv d ( 1<br />

dx 2 v2 + c p T − qε)<br />

= d (<br />

λ dT )<br />

+ 4 dx dx 3<br />

(<br />

d<br />

µv dv )<br />

, (5.4)<br />

dx dx<br />

which can be immediately integrated, giving<br />

( )<br />

1<br />

m<br />

2 v2 + c p T − qε − λ dT<br />

dx − 4 dv<br />

µv<br />

3 dx = me,<br />

(5.4.a)<br />

where e is an integration constant which, like i, must be <strong>de</strong>termined by the<br />

boundary conditions.<br />

e) Diffusion equations.<br />

which can be integrated and gives<br />

(<br />

d<br />

ρD dY ) ( dY<br />

= ρv<br />

dx dx dx − dε )<br />

, (5.5)<br />

dx<br />

ρD<br />

m<br />

dY<br />

dx − Y + ε = f,<br />

(5.5.a)<br />

where f is an integration constant.<br />

The previous system must be completed with the state equation. It can easily be<br />

proven that assumptions 2) and 3) imply that the average mole masses of the unburnt<br />

and burnt gases are equal. Therefore, the particular constant R m = R/M m of the gas<br />

is in<strong>de</strong>pen<strong>de</strong>nt from its composition, and the state equation valid throughout the wave<br />

is<br />

where R m is in<strong>de</strong>pen<strong>de</strong>nt from Y .<br />

p<br />

ρ = R mT, (5.6)<br />

Equations (5.1.a), (5.2.a), (5.3.a), (5.4.a), (5.5.a) and (5.6) form a system of<br />

six equations for the six unknowns p, ρ, T , v, ε and Y . With the a<strong>de</strong>quate boundary<br />

conditions, this system <strong>de</strong>termines the possible solutions of the problem.<br />

Let p 0 , ρ 0 , T 0 and v 0 , be the values for p, ρ, T and v in the unburnt gases, and<br />

p f , ρ f , T f and v f the corresponding values for the burnt gases. Furthermore, since in<br />

the unburnt gases there are no combustion products, in them Y = ε = 0. Similarly,<br />

in the burn gases, Y = ε = 1. Therefore, the looked-for solutions must satisfy the<br />

following boundary conditions:<br />

1) Unburnt gases,<br />

x → −∞ : p → p 0 , ρ → ρ 0 , T → T 0 , v → v 0 , Y → 0, ε → 0. (5.7)


5.2. WAVE EQUATIONS 109<br />

2) Burnt gases,<br />

x → +∞ : p → p f , ρ → ρ f , T → T f , v → v f , Y → 1, ε → 1. (5.8)<br />

Such conditions imply, naturally, the following ones,<br />

x → ±∞ :<br />

dp<br />

dx → 0, dρ<br />

dx → 0, dT<br />

dx → 0, dY<br />

dx → 0,<br />

dε<br />

→ 0. (5.9)<br />

dx<br />

Herein, the possibility of satisfying the previous boundary conditions, will not<br />

be discussed since it is subjected to a careful study in the chapter <strong>de</strong>dicated to <strong>de</strong>flagrations.<br />

Therein, it will be seen that the boundary conditions relative to the cold<br />

boundary of the flame give rise to a problem which, so far, has not been satisfactorily<br />

solved.<br />

In the present study it will be assumed that the boundary conditions can be<br />

satisfied and, therefore, that the problem is completely <strong>de</strong>termined.<br />

By taking the boundary conditions (5.7), (5.8) and (5.9) into the system of<br />

equations (5.1.a), (5.4.a)) and (5.4.a), the following laws of conservation are obtained,<br />

which agree with the invariants <strong>de</strong>duced in the preceding chapter<br />

ρ 0 v 0 = ρ f v f , (5.10)<br />

p 0 + ρ 0 v 2 0 = p f + ρ f v 2 f , (5.11)<br />

1<br />

2 v2 0 + c p T 0 + q = 1 2 v2 f + c p T f . (5.12)<br />

The boundary conditions relative, for instance, to the burnt gases, make possible<br />

the elimination of constants i, e and f from equations (5.3.a), (5.4.a) and (5.5.a),<br />

thus obtaining<br />

p + mv − 4 3 µ dv<br />

dx = p f + mv f , (5.13)<br />

1<br />

)<br />

m(<br />

2 v2 + c p T − qε − λ dT<br />

dx − 4 dv<br />

( 1<br />

)<br />

µv<br />

3 dx = m 2 v2 f + c p T f − q , (5.14)<br />

ρD dY<br />

− Y + ε = 0. (5.15)<br />

m dx<br />

The discussion and analysis of the possible solutions of the previous system are<br />

rather complicated. An example of a discussion for a similar system (but in which the<br />

diffusion neglected) can be found in Friedrichs’s work [3]. As a result of his analysis,<br />

Friedrichs conclu<strong>de</strong>s that except in the case where the reaction rate w is exceptionally<br />

high (and takes a well <strong>de</strong>fined value) weak <strong>de</strong>tonations are impossible. Consequently,<br />

the only possible <strong>de</strong>tonations with a normal reaction rate are strong <strong>de</strong>tonations and


110 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

Chapman-Jouguet <strong>de</strong>tonations. As will presently be seen, the type of <strong>de</strong>tonation that<br />

occurs in each particular case <strong>de</strong>pends on the boundary conditions on the burnt si<strong>de</strong> of<br />

the <strong>de</strong>tonation wave. As for the <strong>de</strong>flagrations, Friedrichs obtained the conclusion that<br />

strong <strong>de</strong>flagrations are impossible and that weak <strong>de</strong>flagrations are possible only for a<br />

well <strong>de</strong>fined value of the propagation velocity. This value <strong>de</strong>pends on the state of the<br />

unburnt gases, on the reaction rate and on the transport coefficients of the mixture.<br />

The discussion of the differential system of combustion waves is consi<strong>de</strong>rably<br />

simplified by analyzing, as done by von Kármán [2], the limiting system <strong>de</strong>duced from<br />

the previous one by assuming that a characteristic time of the thermodynamic transformations,<br />

for example the average time between two molecular collisions, is very small<br />

compared to a characteristic time of the chemical transformations, for example the average<br />

time between two collisions of molecules of different species un<strong>de</strong>r the required<br />

circumstances for these molecules to react to one another. This assumption appears<br />

justified if one consi<strong>de</strong>rs that, in a mixture of reactants, of all the molecular collisions<br />

only a few are accompanied by chemical transformations. In fact, for this to happen<br />

a number of favourable circumstances must concur: for example, the molecules must<br />

be of the a<strong>de</strong>quate species, the orientation of the collision and the energy in certain<br />

<strong>de</strong>grees of freedom must be the proper ones, etc. 3<br />

In the following, the solutions of the aforementioned differential system will<br />

be discussed mainly from this stand-point, following von Kármán’s reasoning.<br />

5.3 Characteristic times<br />

As a characteristic τ t of the thermodynamic transformations the following ratio is<br />

adopted<br />

τ t = µ p . (5.16)<br />

This ratio is proportional to the average time between molecular collisions, which is<br />

given by ratio l/¯v of the mean free path l of the molecules to the average velocity ¯v<br />

of the molecular motion. In fact, the Kinetic Theory of Gases shows that µ is of the<br />

form<br />

µ ∼ ρ¯vl, (5.17)<br />

and that ¯v is of the form<br />

√ p<br />

¯v ∼<br />

ρ . (5.18)<br />

3 See chapter 1.


5.4. LIMITING FORM OF THE WAVE EQUATIONS 111<br />

Now, by substituting the expressions of ¯v and l <strong>de</strong>duced from (5.17) and (5.18) into<br />

the ratio l/¯v, it results that this ratio is proportional to τ t as <strong>de</strong>fined by (5.16).<br />

As a characteristic time τ c of the chemical transformations, the following ratio<br />

is adopted<br />

τ c = ρ w , (5.19)<br />

that measures the average time between those molecular collisions successful in producing<br />

chemical reaction. 4<br />

Let<br />

α = τ t<br />

= µw<br />

τ c ρp<br />

(5.20)<br />

be the ratio of the thermodynamic time τ t to the chemical time τ c . The assumption<br />

that chemical transformations are very slow when compared to the thermodynamic<br />

transformations is expressed by the condition that parameter α is much smaller than<br />

unity<br />

α ≪ 1. (5.21)<br />

The limiting form of the wave equations, obtained when taking assumption (5.21) into<br />

the system <strong>de</strong>duced in the preceding paragraph, is given in the paragraph that follows.<br />

5.4 Limiting form of the wave equations<br />

In or<strong>de</strong>r to discuss the relative values of the various terms of the differential system<br />

of the wave, we shall start by showing the influence of α and the flow Mach number<br />

M = v/a = v/ √ γp/ρ.<br />

For this purpose, let us first eliminate x from (5.3.a), (5.4.a) and (5.5.a), making<br />

use of equation (5.2.a), then<br />

p + mv − 4 µw dv<br />

= i,<br />

3 m dε<br />

(5.22)<br />

1<br />

2 v2 + c p T − qε − λw dT<br />

m 2 dε − 4 µw<br />

3 m 2 v dv = e,<br />

dε<br />

(5.23)<br />

ρDw<br />

m 2<br />

dY<br />

dε<br />

− Y + ε = 0. (5.24)<br />

4 See chapter 1.


112 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

Now, the system of equations (5.22), (5.23) and (5.24) can be written in the<br />

following way, which evi<strong>de</strong>nces the influence of α and M,<br />

(<br />

p 1 + γM 2 − 4 3 α 1 )<br />

dv<br />

= i, (5.25)<br />

v dε<br />

(<br />

c p T 1 + γ − 1 M 2 −<br />

q<br />

2 c p T ε − 1 α 1 dT<br />

γP r M 2 T dε − 4 γ − 1<br />

α 1 )<br />

dv<br />

= e, (5.26)<br />

3 γ v dε<br />

1 α dY<br />

γS c M 2 − Y + ε = 0, (5.27)<br />

dε<br />

where P r = µc p /λ and S c = µ/ (ρD) are the Prandtl and Schmidt numbers, respectively,<br />

for the mixture.<br />

Since α ≪ 1, the last term of the left hand si<strong>de</strong> of equation (5.25) can be<br />

neglected when compared to unity, thus obtaining, instead of equation (5.25)<br />

p ( 1 + γM 2) = i. (5.28)<br />

Similarly, in equation (5.26) one can neglect the last term of the left hand si<strong>de</strong>. Since<br />

the Prandtl P r and the Schmidt S c numbers are of or<strong>de</strong>r unity, 5 the or<strong>de</strong>r of magnitu<strong>de</strong><br />

of terms 1 α 1 dT<br />

P r M 2 T dε of equation (5.26), and 1 α dY<br />

S c M 2 of equation (5.27) <strong>de</strong>pends<br />

dε<br />

on the values of the ratio α/M 2 . Here, there are two possible cases:<br />

a) If the Mach number M of the flow is of the or<strong>de</strong>r of magnitu<strong>de</strong> of unity or larger,<br />

then α/M 2 ≪ 1, and the said terms can also be neglected.<br />

b) Whereas if α/M 2 are of the or<strong>de</strong>r one, then said terms must be preserved. In<br />

this case, however, M 2 ≪ 1, and γM 2 can be neglected in equation (5.28)<br />

and γ − 1 M 2 can be neglected in equation (5.26). Therefore the two following<br />

2<br />

cases are obtained:<br />

1) α ≪ 1, M 2 ∼ 1.<br />

Will be <strong>de</strong>signated case A. For this case, Eq. (5.28) is valid and will be<br />

written in the form<br />

p + ρv 2 = i.<br />

(5.28.a)<br />

In Eq. (5.26) the fourth and fifth terms of the left hand si<strong>de</strong> disappear,<br />

obtaining the following simplified equation<br />

(<br />

c p T 1 + γ − 1 M 2 −<br />

q )<br />

2 c p T ε = e, (5.29)<br />

that can be written<br />

c p T + 1 2 v2 − qε = e.<br />

(5.29.a)<br />

5 See Hirschfel<strong>de</strong>r, Curtiss & Bird: Molecular Theory of Gases and Liquids, John Wiley & Sons Inc.,<br />

New York, 1954, p. 16.


5.4. LIMITING FORM OF THE WAVE EQUATIONS 113<br />

Equation (5.27) reduces to<br />

Y = ε. (5.30)<br />

Therefore in this case, the effects of viscosity, thermal conductivity and<br />

diffusion can be neglected, and the mixture can be consi<strong>de</strong>red as an i<strong>de</strong>al<br />

gas whose composition varies due to chemical reactions.<br />

2) α ≪ 1, α/M 2 ∼ 1.<br />

Will be <strong>de</strong>signated case B. As aforesaid, in this case M 2<br />

(5.28) is reduced to<br />

≪ 1, and Eq<br />

p = i, (5.31)<br />

that is to say, the combustion occurs in first approximation at constant pressure.<br />

In Eq. (5.26) the second and last terms of the left hand si<strong>de</strong> can be<br />

neglected. Then, the following simplified equation is obtained<br />

(<br />

c p T 1 − q<br />

c p T ε − 1 )<br />

α 1 dT<br />

P r M 2 = e, (5.32)<br />

T dε<br />

which can be written in the following form, returning to the old variable x<br />

by means of equation (5.2.a)<br />

c p T − qε − λ m<br />

dT<br />

dx = e.<br />

(5.32.a)<br />

All the terms of equation (5.27) must be preserved. By returning to variable<br />

x, equation (5.15) is obtained.<br />

Of the two limiting cases <strong>de</strong>termined herein, the former is represented by the<br />

system of equations (5.1.a), (5.2.a), (5.28.a) and (5.30), that is, by<br />

Case A: α ≪ 1, M 2 ∼ 1 .<br />

ρv = m,<br />

m dε<br />

dx = w,<br />

p + ρv 2 = i,<br />

c p T + 1 2 v2 − qε = e,<br />

(5.1.a)<br />

(5.2.a)<br />

(5.28.a)<br />

(5.29.a)<br />

Y = ε. (5.30)<br />

As aforesaid, in this system, the influence of viscosity, thermal conductivity<br />

and diffusion has disappeared. As will presently be seen, the solutions corresponding<br />

to this system represent <strong>de</strong>tonation waves.


114 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

The second case is represented by the system of equations (5.1a), (5.2a), (5.31)<br />

and (5.32a), that is, by<br />

Case B: α ≪ 1, α/M 2 ∼ 1 .<br />

ρD<br />

m<br />

dY<br />

dx<br />

c p T − qε − λ m<br />

ρv = m,<br />

m dε<br />

dx = w,<br />

(5.1.a)<br />

(5.2.a)<br />

− Y + ε = 0, (5.15)<br />

p = i, (5.31)<br />

dT<br />

dx = e,<br />

(5.32.a)<br />

in which the influence of the kinetic energy and viscosity has disappeared. This system<br />

represents a combustion at constant pressure, as limiting case of the weak <strong>de</strong>flagrations<br />

that are physically observed.<br />

In the following paragraphs both cases will be studied separately.<br />

5.5 Detonations<br />

Let us first discuss the solutions represented by system A).<br />

The three equations (5.1.a), (5.28.a) and (5.29.a), together with the state equation<br />

(5.6), allows one to express the variation laws of p, T and v as function of the<br />

<strong>de</strong>gree of advancement ε of the combustion, or of the mass fraction Y = ε of the<br />

burnt gases in the mixture. This is the problem of heat addition in the i<strong>de</strong>al gas in<br />

one-dimensional and stationary motion. 6 In particular, the following equation for v is<br />

obtained<br />

v 2 γ i − 1)<br />

− 2 v + 2(γ (e + qε) = 0, (5.33)<br />

(γ + 1) m γ + 1<br />

which shows that to each value of ε satisfying the condition<br />

(<br />

0 ≤ ε ≤ 1 γ 2 ( ) 2 i<br />

q 2(γ 2 − e)<br />

, (5.34)<br />

− 1) m<br />

corresponds two different real values of v, given by the expression<br />

( √<br />

)<br />

γ i<br />

v =<br />

1 ± 1 − 2 (γ2 − 1)<br />

( m<br />

) 2<br />

(e + qε)<br />

(γ + 1) m<br />

γ 2<br />

. (5.35)<br />

i<br />

6 See chapter 3, paragraph 9.


5.5. DETONATIONS 115<br />

Of these two velocities, one is subsonic and the other supersonic. In fact, the<br />

critical velocity v cr , corresponding to the point <strong>de</strong>fined by the value ε of the <strong>de</strong>gree of<br />

advancement of the combustion, is given, in virtue of (5.29a), by<br />

v 2 cr =<br />

2 (γ − 1)<br />

γ + 1<br />

(e + qε) . (5.36)<br />

This value, however, is exactly the product of roots v 1 and v 2 of equation (5.33).<br />

Consequently<br />

v 1 v 2 = v 2 cr, (5.37)<br />

that is, of both velocities one is subcritical, that is to say subsonic, and the other supercritical,<br />

that is to say, supersonic. Furthermore, the two values v 1 and v 2 correspond to<br />

the velocities before and after a normal shock wave, since (5.37) is the Prandtl relation<br />

for shock waves. 7<br />

A<br />

F<br />

v<br />

C<br />

E<br />

B<br />

ε<br />

D<br />

Figure 5.1: Schematic diagram showing the two possible velocities resulting from the heat<br />

addition.<br />

Figure 5.1 represents the pair of values corresponding to Eq. (5.33) for variable<br />

ε. Their corresponding curve is a parabola. The upper branch of this parabola, AC,<br />

corresponds to the supersonic velocities, and the lower branch, BC, corresponds to the<br />

subsonic velocities. At point C, where both branches join, the velocity of the gases<br />

with respect to the wave is equal to the sound velocity.<br />

7 See R. Courant and K. O. Friedrichs: Supersonic Flow and Shock Waves. Interscience Pub. Inc., New<br />

York, 1948, p. 147.


116 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

Let A and B be the two representative points of the two possible initial states<br />

(ε = 0), compatible with the assumed values for m, i and e. Velocity is supersonic<br />

in A and subsonic in B. The jump from A to B occurs through a shock wave, as<br />

aforesaid. Since, as previously seen in the preceding chapter, the propagation velocity<br />

of a <strong>de</strong>tonation is supersonic, the representative point for the initial state of the mixture<br />

is A. When ε increases, that is, when the reaction occurs, two alternatives are possible<br />

either the representative point of the intermediate states of the mixture throughout the<br />

wave moves along the upper branch AC (see Fig. 5.1), or else a shock wave produces<br />

first which changes the velocity from A to B and then, as the combustion progresses,<br />

the representative point moves along the lower branch BC. A <strong>de</strong>cision between both<br />

alternatives can not be taken from purely hydrodynamic consi<strong>de</strong>rations. Let us analyze<br />

both cases separately.<br />

Suppose that the point moves on the upper branch, starting from A. This means<br />

that the combustion initiates in the state of the unburnt gases. But in this state, the<br />

temperature is small and the reaction velocity cannot be sufficiently large to burn<br />

the gases with the speed required by the <strong>de</strong>tonation wave. Therefore, the <strong>de</strong>tonation<br />

must be initiated by a shock wave which, by making the gases pass from the state<br />

represented by point A to the one represented by point B, compresses and heats the<br />

gases, taking them to a state in which the reaction velocity can be sufficiently large to<br />

burn them with the required speed.<br />

The ratio of the thickness of the shock wave to the thickness of the <strong>de</strong>tonation<br />

wave is measured by the ratio of the thermodynamic characteristic time τ t to the chemical<br />

characteristic time τ c . Therefore, the thickness of the shock wave is very small<br />

with respect to the thickness of the <strong>de</strong>tonation wave. 8 This means that while the gases<br />

pass through the shock wave, the burnt fraction is insignificant. Thus the <strong>de</strong>tonation<br />

wave appears as formed by the shock wave, followed by a combustion wave.<br />

The need of a shock wave to initiate the chemical reaction in the <strong>de</strong>tonation<br />

wave has always been acknowledged. This has been the stand-point, for instance, for<br />

Vieille [5] and Jouguet [6] in 1900. 9 These authors, however, have consi<strong>de</strong>r the <strong>de</strong>tonation<br />

always as a discontinuity, in which not only the shock wave is instantaneously<br />

produced, but also the subsequent reaction. The i<strong>de</strong>as <strong>de</strong>veloped in the present study,<br />

concerning the structure of the wave, belong to the mo<strong>de</strong>rn theories on <strong>de</strong>tonation<br />

8 The study of the dynamic transformations within the shock wave cannot be performed with the system<br />

obtained by neglecting the action of viscosity and thermal conductivity, whose action is essential within the<br />

wave. See for example M. Roy: Structure <strong>de</strong> l’on<strong>de</strong> <strong>de</strong> choc et <strong>de</strong>s flammes <strong>de</strong>flagrantes. ONERA, Paris,<br />

1952.<br />

9 In the fundamental work of Jouguet [6] data can be found concerning the historical evolution of the<br />

i<strong>de</strong>as relative to the classical theory of the <strong>de</strong>tonation waves.


5.5. DETONATIONS 117<br />

waves. Such theories were initiated by the works of Taylor in England, von Neumann<br />

[7] in the U.S.A., Zeldovich [8] in the U.R.S.S. and Döring [9] in Germany.<br />

Within the combustion zone that follows the initial shock wave, the gases move<br />

along the lower branch of the curve shown in Fig. 5.1, starting from point B. The point<br />

ε = 1, at which the combustion ends and the thermodynamic equilibrium is reached<br />

after the wave, cannot lie at the right hand si<strong>de</strong> of D. Therefore, the three following<br />

cases can occur:<br />

1) The combustion ends at a point such as E of the lower branch, in which the<br />

velocity is subsonic. This case represents a strong <strong>de</strong>tonation.<br />

2) The combustion ends at the point C that separates the subsonic from the supersonic<br />

branch. In C the velocity of the burnt gases with respect to the <strong>de</strong>tonation<br />

wave is sonic. The corresponding <strong>de</strong>tonation for this case is of the Chapman-<br />

Jouguet type.<br />

3) The representative point of the final state is point F in the supersonic branch.<br />

Consequently, the final velocity is supersonic and the corresponding <strong>de</strong>tonation<br />

is weak.<br />

We shall presently see that the last case cannot possibly occur. In fact, in or<strong>de</strong>r<br />

to reach point F either one has to pass by point C, in which case the reaction between<br />

C and F would be endothermic and therefore in contradiction with what happen in the<br />

combustion, or else one must pass from B to D and from D to F by means of a jump<br />

or expansion shock which is also impossible. Consequently, weak <strong>de</strong>tonations are not<br />

possible.<br />

On the other hand, nothing opposes a strong <strong>de</strong>tonation or a <strong>de</strong>tonation of the<br />

Chapman-Jouguet type. For a given case the occurrence of one or the other <strong>de</strong>pends<br />

on the boundary conditions after the wave. Let us see the influence of these conditions.<br />

When the <strong>de</strong>tonation is strong, the velocity of the burnt gases, with respect to<br />

the wave is subsonic. Any perturbation produced in the burnt gases will propagate<br />

throughout them with sound velocity and can therefore reach the wave and alter its<br />

nature. For example, an expansion of the burnt gases will reach the wave, reducing<br />

its strength down to the point where the velocity of the burnt gases with respect to the<br />

wave equals the sound velocity. In such a case the wave becomes insensible to the<br />

perturbations that occur in the burnt gases. Such is the situation that normally occurs<br />

in practice; for example, when the <strong>de</strong>tonation propagates along a tube, starting either<br />

at an open or closed end. Thus the <strong>de</strong>tonations normally observed in practice are of the<br />

Chapman-Jouguet type, due to the fact that these are the only stable ones with respect<br />

to perturbations coming from the burnt gases.


118 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

The study of the aerodynamic field subsequent to a <strong>de</strong>tonation wave can be performed<br />

by applying the Gas Dynamic methods. 10 For instance, Zeldovich [10] has calculated<br />

the aerodynamic field following a <strong>de</strong>tonation wave that propagates unchanged<br />

from the closed end of a tube, initially full of <strong>de</strong>tonant gas at rest. By neglecting the<br />

friction and the heat lost through the walls of the tube, a solution is obtained which<br />

propagates with the Chapman-Jouguet velocity. The <strong>de</strong>tonation wave is followed by<br />

an expansion region with an approximate length of 50% of the total length covered<br />

by the <strong>de</strong>tonation wave. The remaining burnt gases (approximately 40% of the burnt<br />

mass) are at rest. By including the effect of friction and heat losses through the walls<br />

of the tube, a stable Chapman-Jouguet wave is obtained, followed by an expansion<br />

in which the direction of the motion of the burnt gases in the tube is reversed. Following<br />

the expansion there also exists in this case a region at rest in which the burnt<br />

gases have cooled down to ambient temperature. There is photographic evi<strong>de</strong>nce of<br />

the existence of this reversal motion in the expansion region.<br />

The aerodynamic field that follows a spherical <strong>de</strong>tonation wave has been studied<br />

by G.J. Taylor [11] in England, and by Zeldovich in U.S.S.R. [12]. The calculations<br />

also <strong>de</strong>monstrate the existence of a solution which propagates unchanged with<br />

the Chapman-Jouguet velocity. Immediately after the <strong>de</strong>tonation wave there is a strong<br />

expansion region, followed by a central nucleus at rest. Within the expansion region<br />

more than 90% of the burnt mass is in motion. The spherical <strong>de</strong>tonation waves have<br />

been experimentally observed, for example, by Manson and Ferrié [13]. A strong <strong>de</strong>tonation<br />

can occur, for example, when the <strong>de</strong>tonation propagates insi<strong>de</strong> a tube, starting<br />

from an end closed by a piston, that moves after the wave with a subsonic velocity<br />

with respect to the wave. In such a case the intensity of the strong <strong>de</strong>tonation would<br />

be <strong>de</strong>termined by the compatibility condition obtained by expressing that the velocity<br />

of the burnt gases with respect to the wave must equal the velocity of the piston.<br />

Then, when friction and heat losses through the walls of the tube are neglected, it<br />

results that the strong <strong>de</strong>tonation wave propagates unchanged throughout the mass of<br />

unburnt gases.<br />

The relation between pressure and <strong>de</strong>nsity within the wave can be obtained<br />

from Eqs. (5.1.a) and (5.28.a) by elimination of v. Thus resulting<br />

p + m2<br />

ρ<br />

= i, (5.38)<br />

or else, introducing the state (p 1 , ρ 1 ) of the unburnt gases<br />

( 1<br />

p − p 1 = m 2 − 1 )<br />

. (5.39)<br />

ρ 1 ρ<br />

10 See Courant and Friedrichs, ib., pp. 218 and 416 for plane and spherical waves, respectively.


5.5. DETONATIONS 119<br />

This relation shows that the pressure at each point of the wave is <strong>de</strong>termined<br />

only by the corresponding value of the <strong>de</strong>nsity. This relation plays here the same part<br />

as, for example, in Gas Dynamics the relation p = Cρ γ for the isentropic motions of<br />

non-reacting gases.<br />

In or<strong>de</strong>r to represent the result in the diagram (p, τ), as done in chapter 3, the<br />

specific volume τ = 1/ρ must be introduced in place of the <strong>de</strong>nsity in Eq. (5.39),<br />

which then takes the form<br />

p − p 1 = m 2 (τ 1 − τ). (5.40)<br />

This equation is represented in diagram (p, τ), see Fig. 5.2, by a straight line joining<br />

the representative points of the initial and final states. On this line also lie all the<br />

representative points of the intermediate states corresponding to the reaction zone. 11<br />

70<br />

60<br />

D<br />

p/p 1<br />

E<br />

50<br />

C<br />

Trayectory through shock wave<br />

40<br />

30<br />

ε=1.0<br />

J<br />

20<br />

E‘<br />

ε=0.3 ε=0.7<br />

10<br />

ε=0<br />

1<br />

P<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

τ /τ 1<br />

Figure 5.2: Hugoniot curves from different values of ε.<br />

Bringing forth in Eqs. (5.1.a), (5.28.a) and (5.29.a) the specific volume and the<br />

initial conditions corresponding to the unburnt gases, there results<br />

v<br />

τ = v 1<br />

τ 1<br />

, (5.41)<br />

p + v2<br />

τ = p 1 + v2 1<br />

τ 1<br />

, (5.42)<br />

c p T + 1 2 v2 − qε = c p T 1 + 1 2 v2 1. (5.43)<br />

11 W. Michelson was the first to postulate that within the reaction zone of the <strong>de</strong>tonation wave the linear<br />

relation (40) is verified.


120 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

The elimination of T , v, T 1 and v 1 between these equations and the state equations,<br />

pτ = R m T and p 1 τ 1 = R m T 1 , (5.44)<br />

gives the following Hugoniot relation for each value of the <strong>de</strong>gree of advancement ε<br />

of the combustion<br />

γ + 1<br />

2(γ − 1) (pτ − p 1τ 1 ) + 1 2 (p 1τ − pτ 1 ) = qε. (5.45)<br />

This equation represents a family of hyperbolas. For each value of ε a hyperbola<br />

is obtained, which represents the locus of the possible states of the gas when<br />

burnt fraction is ε and the initial state is (p 1 , τ 1 ). The Hugoniot curve E’JE is obtained<br />

for the particular case ε = 1. This curve corresponds to all the possible states of the<br />

burnt gases, see Fig. 5.2. Similarly, for ε = 0 the Hugoniot curve PD is obtained,<br />

which corresponds to all the shock waves that can form in the unburnt gases at the<br />

initial state represented by point P . Between both curves lie those corresponding to<br />

the intermediate states. Two of them are shown in Fig. 5.2, corresponding to ε = 0.3<br />

and ε = 0.7.<br />

Now, let us consi<strong>de</strong>r, for example, a Chapman-Jouguet <strong>de</strong>tonation. Equation<br />

(5.40) corresponding to this <strong>de</strong>tonation is represented by the straight line PJC. The<br />

transformations that occur throughout the shock wave, preceding the combustion,<br />

carry the gas from the initial state P to the state C after the shock wave with no combustion.<br />

However, the trajectory of the gases through the shock wave is not represented<br />

by the straight line PC. In fact, we have seen that the thickness of the shock wave<br />

is very small and, as aforesaid, the thermal conductivity and viscosity therein cannot<br />

be neglected. In particular, Eq. (5.40) must be substituted by the following relation,<br />

<strong>de</strong>duced from (5.3.a), which takes into account the viscosity, 12<br />

p − p 1 = m 2 (τ 1 − τ) + 4 3 µ dv<br />

dx . (5.46)<br />

But since, through the shock wave the velocity <strong>de</strong>creases, dv/ dx is negative. Therefore,<br />

the representative points of Eq. (5.46) lie below the straight line PC, as indicated<br />

in Fig. 5.2. 13<br />

The representative states of the combustion that follow the shock wave, that<br />

is, those corresponding to section BE, Fig. 5.1, are obtained by traversal segment CJ,<br />

starting from C. The points at which this segment intersects the successive Hugoniot<br />

12 Assuming that the transformations through the shock wave can be <strong>de</strong>scribed by variables corresponding<br />

to a continuum. See chapter 3, §1.<br />

13 Hirschfel<strong>de</strong>r et al. ib. page 810.


5.5. DETONATIONS 121<br />

curves, represent the successive states of the mixture for the corresponding values of<br />

the <strong>de</strong>gree of advancement of the combustion.<br />

Now, let us consi<strong>de</strong>r a strong <strong>de</strong>tonation, which in Fig. 5.2 would be represented<br />

by a straight line such as PED. The initial shock wave produces a jump from<br />

state P to state D. From D to E the reaction takes place and is accompanied by an<br />

expansion and acceleration of the gases. The intermediate states are represented by<br />

the points of segment DE.<br />

Figure 5.2 also shows that weak <strong>de</strong>tonations are not possible. In fact, once<br />

point E is reached, point E’ corresponding to a weak <strong>de</strong>tonation can only be reached<br />

by means of an expansion wave between E and E’, which is impossible, or else by<br />

means of an endothermic reaction, corresponding to segment EE’.<br />

The temperature variation throughout the wave can be analyzed in the same<br />

manner as that used for the pressure variations. For this purpose it is sufficient to<br />

eliminate p and p 1 from Eq. (5.40), by making use of (5.44). Thus, we obtain<br />

T<br />

= (1 + γM1 2 ) τ ( ) 2 τ<br />

− γM1<br />

2 , (5.47)<br />

T 1 τ 1 τ 1<br />

in which the value of the Mach number M 1 of the unburnt gases has been ma<strong>de</strong> explicit.<br />

To each value of the Mach number M 1 corresponds a different parabola, on<br />

which lie the representative points of the successive states of the mixture within the<br />

combustion zone of the <strong>de</strong>tonation wave. In Fig. 5.3 two parabolas are shown. One<br />

corresponds to the Chapman-Jouguet <strong>de</strong>tonation and the other to a strong <strong>de</strong>tonation.<br />

The same letters are used to <strong>de</strong>signate homologous points in Figs. 5.1 and 5.2.<br />

The Hugoniot curves are obtained from Eq. (5.45) eliminating p and p 1 by<br />

means of Eq. (5.44), as has been done to obtain Eq. (5.47). There results<br />

( ) 2 τ<br />

+ γ + 1 T τ<br />

− T ( 2γ qε<br />

−<br />

+ γ + 1 ) τ<br />

= 0. (5.48)<br />

τ 1 γ − 1 T 1 τ 1 T 1 γ − 1 c p T 1 γ − 1 τ 1<br />

These curves are a family of hyperbolas. A hyperbola is obtained for each value of<br />

ε. In particular, for ε = 1 one obtains the Hugoniot curve of the burnt gases and for<br />

ε = 0 the curve of the shock waves of the unburnt gases. Both have been taken into<br />

Fig. 5.3. Their intersection with curve of Eq. (5.47), representative of the intermediate<br />

states, <strong>de</strong>termines section CJ or DE, corresponding to the said intermediate states. The<br />

state corresponding to a given fraction of burnt gases is given by the intersection of<br />

curve of Eq. (5.48), corresponding to this fraction, with section CJ, or DE, as shown<br />

in Fig. 5.3 for the values ε = 0.3 and ε = 0.7.<br />

The preceding consi<strong>de</strong>rations give an i<strong>de</strong>a of the structure of a <strong>de</strong>tonation wave.<br />

The initial shock wave compresses, <strong>de</strong>celerates and heats sud<strong>de</strong>nly the gasses, within a


122 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

zone of negligible thickness compared to that of the <strong>de</strong>tonation wave, where no appreciable<br />

chemical reaction occurs. After the shock wave the combustion is initiated. The<br />

gases expand and accelerate and the pressure <strong>de</strong>creases. Temperature rises due to the<br />

reaction. For Chapman-Jouguet <strong>de</strong>tonation and slightly strong <strong>de</strong>tonations, the maxi-<br />

20<br />

18<br />

T/T 1<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

D<br />

C<br />

ε =0.3<br />

E<br />

ε =1<br />

J<br />

K<br />

ε =0.7<br />

E‘<br />

4<br />

2<br />

0<br />

ε =0<br />

Trayectory through shock wave<br />

P<br />

0.2 0.4 0.6 0.8 1.0<br />

τ /τ 1<br />

Figure 5.3: Hugoniot diagram in variables (T, τ).<br />

15<br />

50<br />

1.0<br />

13<br />

40<br />

C<br />

P/P 1<br />

K<br />

J<br />

0.8<br />

11<br />

30<br />

J<br />

0.6<br />

T/T 1<br />

9<br />

P/P 1<br />

20<br />

T/T 1<br />

J<br />

0.4<br />

τ/τ 1<br />

7<br />

10<br />

C<br />

C<br />

τ/τ 1<br />

0.2<br />

5<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Figure 5.4: Typical values of T , p and τ in a Chapman-Jouguet <strong>de</strong>tonation as a function of<br />

the fraction of burnt gases.<br />

ε


5.6. DEFLAGRATIONS 123<br />

mum temperature is reached shortly before the combustion ends (as is clearly shown<br />

in Fig. 5.3), after which a slight temperature drop (100 to 200 ◦ C for typical cases)<br />

takes place due to the fast expansion of the combustion products. For the very strong<br />

<strong>de</strong>tonations, like DE represented in Fig. 5.3, the maximum temperature is reached at<br />

the end of the combustion. These results appear in Fig. 5.4, which shows the values<br />

corresponding to a typical case for a Chapman-Jouguet <strong>de</strong>tonation.<br />

5.6 Deflagrations<br />

We have said that the <strong>de</strong>flagrations, also known as flames, will be the subject of a<br />

careful study in the following chapter. Consequently, herein we shall only inclu<strong>de</strong> a<br />

few brief consi<strong>de</strong>rations concerning their possible existence, structure and propagation<br />

velocity.<br />

In the preceding chapter we have seen that the propagation velocity of a <strong>de</strong>flagration<br />

wave is always subsonic. Thereby it cannot be ascertained that in the <strong>de</strong>flagration<br />

waves the condition α ≪ M 2 will be satisfied, and the system that must be used<br />

for the study of its structure is B, at least within the region of the flame close to the<br />

cold boundary. 14 Let us consi<strong>de</strong>r separately the weak and strong <strong>de</strong>flagrations, and let<br />

us start by <strong>de</strong>monstrating that the latter cannot exist.<br />

In the strong <strong>de</strong>flagration waves, the final velocity of the gases is supersonic.<br />

Consequently, at least in the final zone of such waves, the condition M 2 ∼ 1 is satisfied,<br />

and system A can be used. In particular in this zone, the variation law for the<br />

velocity of the gases with respect to the wave is given in Fig. 5.1. However, due to<br />

the influence of thermal conductivity and diffusion, this law can differ within the zone<br />

close to the cold boundary. Hence, in a strong <strong>de</strong>flagration wave, the variation of the<br />

velocity of the gases through the wave must follow a law as the one represented in<br />

Fig. 5.5, where the curve in Fig. 5.1 representing the limiting solution of system A has<br />

also been represented by a dash line. The velocity of the unburnt gases is represented<br />

by point P . Starting from this point, the gases accelerate. When their Mach number<br />

is such that condition α ≪ M 2 is satisfied, the curve overlaps the dash line of the<br />

limiting solution (in the figure this happens at point B). If now one must reach point<br />

D, corresponding to the final state where the velocity is supersonic, this can only be<br />

achieved by passing through point C where the velocity is sonic. But the jump from C<br />

14 It has been shown in the preceding chapter that the propagation velocity of a <strong>de</strong>flagration is always equal<br />

to or smaller than that corresponding to the Chapman-Jouguet <strong>de</strong>flagration. In this case the Mach number<br />

corresponding to the propagation velocity is always much smaller than unity. Thereby, the propagation<br />

Mach number of the <strong>de</strong>flagrations observed is always much smaller than unity. In practice, of the or<strong>de</strong>r of<br />

10 −3 .


124 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

to D could only be attained through an endothermic reaction, and this kind of reaction<br />

is impossible in a combustion process. As a consequence, the non-existence of the<br />

strong <strong>de</strong>flagrations is conclu<strong>de</strong>d.<br />

D<br />

v<br />

C<br />

B<br />

P<br />

ε<br />

Figure 5.5: Schematic diagram showing the variation of the velocity of gases in a <strong>de</strong>flagration.<br />

On the other hand, since for a weak <strong>de</strong>flagration the final velocity is subsonic,<br />

nothing is opposed to its existence.<br />

In the weak <strong>de</strong>flagrations diffusion and heat conduction are important and the<br />

system to be used is system B. Diffusion and heat conduction are responsible for<br />

the propagation of the combustion to the unburnt gases, activating their reaction by<br />

diffusion of the active centers (atoms, radicals, etc.) and by heating. As they heat, the<br />

gases expand and accelerate, producing a slight pressure drop throughout the wave.<br />

Fig. 5.6 shows qualitatively the structure of a <strong>de</strong>flagration wave bringing forth, by<br />

comparison with Fig. 5.3, the difference in the mechanism that propagate the process<br />

in each case. The system B of differential equations that must be integrated in or<strong>de</strong>r<br />

to obtain the structure of the <strong>de</strong>flagration wave can be simplified by eliminating x, as<br />

done in §4. Thus obtaining<br />

p Dw<br />

R m T m 2<br />

dY<br />

dε<br />

= Y − ε, (5.49)<br />

λw dT<br />

m 2 dε = c pT − qε + e. (5.50)<br />

In the <strong>de</strong>duction of this system, the state equation p = ρR m T has been used in or<strong>de</strong>r<br />

to eliminate the <strong>de</strong>nsity from the diffusion equation.


5.6. DEFLAGRATIONS 125<br />

8<br />

7<br />

v/v 1<br />

=T/T 1<br />

6<br />

5<br />

4<br />

3<br />

−∆ P/(1/2)ρ 1<br />

v 1<br />

2<br />

T 2<br />

/T 1<br />

=8<br />

ρ D c p<br />

/λ=1.5<br />

2<br />

1<br />

Y<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

ε<br />

Figure 5.6: Typical values of T , v, Y and ∆p in a <strong>de</strong>flagration as a function of the fraction<br />

of burnt gases .<br />

In this system the reaction rate w is a known function of Y and T . The pressure<br />

p is constant in virtue of Eq. (5.31). One must look for a solution of this system, in<br />

the interval 0 ≤ ε ≤ 1, satisfying the following boundary conditions<br />

ε = 0 : Y = 0, T = T 0 , (5.51)<br />

ε = 1 : Y = 1, T = T f . (5.52)<br />

Since this is a system of two first or<strong>de</strong>r equations, the four conditions (5.51)<br />

and (5.52) cannot be satisfied unless parameters e and m take well <strong>de</strong>fined values.<br />

Parameter e is given by the expression: e = q − c p T f , which results from expressing<br />

the condition w = 0 for ε = 1, that is, at the end of the combustion. As for m it must<br />

taken well <strong>de</strong>fined value, so that the previous differential system to be compatible.<br />

This means that weak <strong>de</strong>flagrations can only propagate with a well <strong>de</strong>fined velocity<br />

which <strong>de</strong>pends on the state of the mixture, its reaction rate and its coefficients of thermal<br />

conductivity and diffusion. The problem of <strong>de</strong>termining the propagation velocity<br />

of the flame through a combustible mixture appears, thus, as an “eigenvalue” problem.<br />

The solution to this problems will be studied in the following chapter.<br />

The same conclusions are reached when taking into account the influence of all<br />

the terms in the differential equations of the combustion wave, instead of consi<strong>de</strong>ring<br />

a limiting solution as done herein.


126 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

5.7 Transition from <strong>de</strong>flagration to <strong>de</strong>tonation<br />

The preceding study presents <strong>de</strong>flagrations and <strong>de</strong>tonations as in<strong>de</strong>pen<strong>de</strong>nt phenomena.<br />

Both, <strong>de</strong>flagrations and <strong>de</strong>tonations can be initiated by means of various procedures.<br />

For example, <strong>de</strong>flagrations can be initiated by means of an electric spark within<br />

a combustible mixture, and <strong>de</strong>tonations by making a shock wave cross the <strong>de</strong>tonant<br />

mixture. Such procedure was applied, for example, by Fay [14] using a shock tube.<br />

However, a <strong>de</strong>tonation can also be produced starting from a <strong>de</strong>flagration by compression<br />

of the unburnt gases and acceleration of the combustion wave. This can occur, for<br />

example, when a <strong>de</strong>tonant mixture filling a tube is ignited at the closed end. In such<br />

a case a <strong>de</strong>flagration wave initiates at the ignition point. This <strong>de</strong>flagration accelerates<br />

as it propagates along the tube, producing a succession of intermediate states, nonstationary,<br />

which end with the establishment of a stable Chapman-Jouguet <strong>de</strong>tonation,<br />

Hereinafter, we shall restrict ourselves to a qualitative <strong>de</strong>scription of the process, following<br />

the lines of Zeldovich’s work [8], who carefully studied this phenomenon. For<br />

this purpose the flame front will be consi<strong>de</strong>red as a discontinuity that travels along the<br />

tube, as done in the preceding chapter.<br />

Let ϕ be the propagation velocity of the flame through the unburnt gases, and<br />

S the area of the cross-section of the tube. If the flame front is plane and normal to the<br />

tube axis, the mass of the gases burnt per unit time is ρ 1 ϕS, where ρ 1 is the <strong>de</strong>nsity<br />

of the unburnt gases before the flame. Before burning, this mass occupies a volume<br />

ϕS. Since the gases expand as they burn, the volume that must occupy the said mass<br />

at the same pressure 15 is nϕS. Here n is the ratio of the temperature of the burnt gases<br />

to temperature of the unburnt gases. Consequently, the increase in volume due to the<br />

combustion is (n − 1)ϕS. Such an increase in volume sets the unburnt gases into<br />

motion in front of the combustion wave to empty the necessary space. The total mass<br />

of unburnt gases is not set into motion simultaneously, but progressively by a pressure<br />

wave that travels through the mass with a velocity close to the velocity of sound. The<br />

situation is shown schematically in Fig. 5.7. The gases are at rest within the region<br />

between O and the flame front A. Between the front A and the pressure wave B, lie the<br />

unburnt gases, compressed and moving towards the right-hand si<strong>de</strong>. Starting from the<br />

pressure wave B and towards tho right-hand si<strong>de</strong>, lie the unburnt gases in the initial<br />

state and at rest. The intensity of the pressure jump across the pressure wave is such<br />

that the burnt gases are at rest. Let v p1 and v p2 be the velocities of the unburnt gases<br />

before and after the pressure wave, measured with respect to this wave. The velocity<br />

at which the pressure wave travels along the tube is v p1 . The velocity of the unburnt<br />

15 The pressure drop across the flame is negligible.


5.7. TRANSITION FROM DEFLAGRATION TO DETONATION 127<br />

O<br />

A<br />

B<br />

P P<br />

n ϕ<br />

2<br />

1<br />

P v p − v p = (n−1)ϕ v p1<br />

3 1 2<br />

P 3<br />

P 2<br />

P, v<br />

(n−1)ϕ<br />

P 1<br />

Figure 5.7: Schematic diagram of the flame propagation in a closed tube.<br />

gases after the wave, with respect to the tube, is v p1 −v p2 . The condition that the burnt<br />

gases must be at rest is<br />

v p1 − v p2 = (n − 1)ϕ, (5.53)<br />

as it can easily be verified. The velocity v a at which the flame travels along the tube is<br />

v a = nϕ, (5.54)<br />

where ϕ is the propagation velocity of the flame throughout the gases in their state<br />

after the pressure wave. Such propagation velocity differs from that corresponding<br />

to the state of the gases before the flame. This state of affairs can be maintained<br />

unchanged along the tube. For the <strong>de</strong>tonation to occur it is necessary to activate the<br />

combustion, increasing the burnt mass per second. Such an increase is necessary in<br />

or<strong>de</strong>r to increase the intensity of the pressure jump across the pressure wave, up to the<br />

point of self-ignition of the mixture after the wave, nee<strong>de</strong>d to maintain the <strong>de</strong>tonation.<br />

The slight compression of the mixture, produced by the compression wave, is not<br />

sufficient to increase, appreciably, the propagation velocity of the flame. Therefore,<br />

the increase of burnt mass per second must be due to other effects. The explanation<br />

can be found in the influence of the tube walls. In fact, due to friction, the distribution<br />

of velocities in the cross-section of a tube is not uniform but maximum at the axis<br />

and zero on the walls. Due to this non-uniformity of the velocity distribution, the<br />

flame front curves increasing its surface. Due to this surface increase, the burnt mass<br />

per second increases and therewith the velocity of the gases and the intensity of the<br />

pressure jump, producing an interaction of both effects such that when the tube is<br />

sufficiently long it leads to the establishment of the <strong>de</strong>tonation wave. This standpoint<br />

is corroborated by the fact that for tubes with rough walls or small diameter the


128 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES<br />

<strong>de</strong>tonation is soon establish because both effects increase the influence of friction. On<br />

the other hand, if the mixture is ignited at the open end of the tube, the <strong>de</strong>tonation is<br />

established with greater difficulty as the burnt gases can escape through the open end<br />

and the pressure wave that prece<strong>de</strong>s the combustion wave is much weaker. In this case,<br />

a longer tube is necessary so that the pressure wave may reach the required intensity<br />

for the <strong>de</strong>tonation to occur.<br />

For the case previously analyzed, A.K. Oppenhein [15] has studied the succession<br />

of intermediate states that must be produced in the transition from <strong>de</strong>flagration to<br />

<strong>de</strong>tonation by acceleration of the <strong>de</strong>flagration wave up to the point where it overtakes<br />

the shock wave. He <strong>de</strong>monstrated his results by means of a hydraulic analogy.<br />

References<br />

[1] Backer, R.: Stosswelle und Detonation. Zeitschrift für Physic, Vol. 8, 1922.<br />

[2] von Kármán, Th.: Aerothermodynamics and Combustion Theory. L’Aerotecnica,<br />

Vol. XXXIII, Fasc. 1st., 1953, pp. 80-86.<br />

[3] Friedrichs, K. O.: On the Mathematical Theory of Deflagrations and Detonations.<br />

NAVORD Report 79-46, Institute for Mathematics and Mechanics, New<br />

York University, 1946.<br />

[4] Hirschfel<strong>de</strong>r, J. O., Curtiss, C. F. and Campbell, D.E.: The Theory of Flames<br />

and Detonations. Fourth Symposium (International) on Combustion, Williams<br />

and Wilkins Co., Baltimore, 1953, pp. 190-211.<br />

[5] Vieille, P.: Rôle <strong>de</strong>s Discontinuites dans la Propagation <strong>de</strong>s Phénomènès Explosifs.<br />

Comptes Rendus <strong>de</strong>s Séances <strong>de</strong> 1’Académie <strong>de</strong>s Sciences, Vol. CXXXI,<br />

1900, p. 413.<br />

[6] Jouguet, E.: Mecanique <strong>de</strong>s Explosifs, Paris, 1917, p. 325.<br />

[7] Neumann, J. von: Progress Report on the Theory of Detonation Waves. N D R<br />

C, Division B, 0 S R D, No. 549, 1942.<br />

[8] Zeldovich, Y. B.: On the Theory of Propagation of Detonation. Journal of<br />

Experimental and Theoretical Physics, Vol. 10, 1940, p. 542, Translated as<br />

NACA Technical Memorandum No. 1261, 1950.<br />

[9] Döring, W.: Annalen <strong>de</strong>r Physic, Vol. 43, 1943, p. 421.<br />

[10] Zeldovich, Y. B.: Theory of Combustion and Detonation of Gases. A9-T-45,<br />

Air Documents Division, Air Material Command, U S A F .


5.7. TRANSITION FROM DEFLAGRATION TO DETONATION 129<br />

[11] Taylor, G. I.: Detonation Waves. Ministry of Supply, Explosives Research Committee,<br />

R.C. 178, A.C. 639, 1941.<br />

[12] Zeldovich, Y. B.: The Distribution of Pressure and Velocity in the Products of<br />

a Detonation Explosion. Journal of Experimental and Theoretical Physics, Vol.<br />

12, 1942, p. 389. Also ref. [8].<br />

[13] Manson, N. and Ferrié, P.: Contribution to the Study of Spherical Detonation<br />

Waves. Fourth Symposium (International) on Combustion, Williams and<br />

Wilkins Co., Baltimore, 1953, pp. 486-94.<br />

[14] Fay, J. A.: Some Experiments on the Initiation of Detonation in 2H 2 -O 2 Mixtures<br />

by Uniform Shock Waves. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953, pp. 501-507.<br />

[15] Oppenheim, A. K.: Gasdynamics Analysis of the Development of Gaseous Detonation<br />

and its Hydraulic Analogy. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953, pp. 471-480.


130 CHAPTER 5. STRUCTURE OF THE COMBUSTION WAVES


Chapter 6<br />

Laminar flames<br />

6.1 Introduction<br />

It was <strong>de</strong>monstrated in chapter 4 that through a combustible mixture of gases two<br />

types of combustion waves may propagate: <strong>de</strong>tonation waves and <strong>de</strong>flagration waves<br />

or flames. Chapter 5 inclu<strong>de</strong>s the analysis of the nature of such waves, their structure<br />

and the conditions which <strong>de</strong>termine their propagation velocity. This analysis proved<br />

that while in <strong>de</strong>tonation waves the propagation velocity is <strong>de</strong>termined only by a condition<br />

of mechanical stability within the zone of burnt gases, which simplifies extremely<br />

its <strong>de</strong>termination, when <strong>de</strong>aling with flames their propagation velocity is <strong>de</strong>termined<br />

by an internal equilibrium between the processes of chemical reaction, heat transfer<br />

(conductivity) and transport of the chemical species (diffusion), which makes the<br />

study specially difficult.<br />

The initial attempt to establish a theory on these bases, and in particular the<br />

<strong>de</strong>duction of the first formula for the computation of the propagation velocity of a<br />

flame through a given mixture, which is the fundamental problem of combustion, is<br />

own to Mallard and Le Chatelier [1]. Their theory is purely thermal and it establishes<br />

a primitive balance between the heat received by the gases through conduction, when<br />

an assumed ignition temperature is reached, and the heat content of these gases. The<br />

concept of reaction velocity is not applied in this theory.<br />

The work by Mallard and Le Chatelier was the starting point for the <strong>de</strong>velopment<br />

of the “thermal theories of the flame”, so called because they did not inclu<strong>de</strong> the<br />

diffusion of species.<br />

The i<strong>de</strong>a of chemical reaction velocity, incorporated by Crussard [2] to this<br />

thermal mo<strong>de</strong>l, was a <strong>de</strong>finite step forward in the progress of the theory. From there<br />

131


132 CHAPTER 6. LAMINAR FLAMES<br />

on, the interest of the later works was centered both on the use of more realistic expressions<br />

for the reaction velocity which consi<strong>de</strong>red the influence of the concentration<br />

of species and of the temperature of the mixture, as well as on the <strong>de</strong>velopment of<br />

approximate methods, analytical and numerical, for the integration of the flame equations.<br />

A critical discussion on thermal theories, whose <strong>de</strong>velopment and application<br />

have been pursued up to our days, may be found in the work by Goward and Payman<br />

[3]. Frank-Kamenetskii-Semenov [4], Boys-Corner [5] and von Kármán [6] have specially<br />

contributed to the <strong>de</strong>velopment of analytical methods and Hirschfel<strong>de</strong>r and his<br />

collaborators [7] to the expansion of numerical ones.<br />

Initially these studies were performed with a very simplified kinetic mo<strong>de</strong>l<br />

which consi<strong>de</strong>red only two different chemical species (reactants and products) and<br />

they applied a law of reaction which reflected approximately the relationship between<br />

the reaction velocity of the mixture and its temperature and composition at each point.<br />

Later on these mo<strong>de</strong>ls were improved by taking into account the influence of the different<br />

species and chemical radicals which act in the process.<br />

The first attempt to inclu<strong>de</strong> the influence of both diffusion and chemical radicals<br />

on the velocity of the flame is own to Lewis and von Elbe [8] in their study of<br />

ozone <strong>de</strong>composition flame, which will be given further consi<strong>de</strong>ration in §13. This<br />

study established the bases for the <strong>de</strong>velopment of a complete theory on flames. Other<br />

authors followed their steps, trying to find a more accurate and complete formulation<br />

for the problem. At the same time the analytical and numerical methods of the thermal<br />

theory were exten<strong>de</strong>d to the new formulation. A very complete analysis of the different<br />

theories and methods available, as well as of the assumptions used in expanding<br />

them, may be found in the work by Evans [9].<br />

In the last years a great effort has been applied to the <strong>de</strong>velopment of the theory<br />

of the flame, which has resulted in the obtaining of a complete formulation of the<br />

problem and its boundary conditions. At the same time a great amount of attention was<br />

given to the <strong>de</strong>velopment of analytical and numerical methods, very approximated,<br />

for the solution of the resulting equations, while the theory was applied to a certain<br />

number of practical cases, although very few due to the fact that it is very difficult to<br />

know the physico-chemical constants of the process and of the chemical scheme of<br />

the reactions. It is practically impossible to try to give a complete bibliography of the<br />

work achieved in these last years since it is very abundant and scattered, specially in<br />

the United States, United Kingdom, and Russia. Therefore, it is far more practical<br />

to refer the rea<strong>de</strong>r to the Proceedings of the International Symposia on Combustion,<br />

in special starting from the third one which took place in 1948, as well as to the<br />

volumes of the Selected Combustion Problems published by AGARD, containing the


6.1. INTRODUCTION 133<br />

papers presented at the two International Colloquia sponsored by this organization in<br />

1954 and 1956. The work by von Kármán [10] is a complete review of the state of<br />

knowledge on this problem up to 1956.<br />

At the same time that this theory which consi<strong>de</strong>rs the flame as a wave propagating<br />

through a continuous medium was <strong>de</strong>veloped, other theories have attempted to<br />

emphasize only partial aspects of the phenomenon. From them the most representative<br />

is the one proposed by Tanford and Pease on the diffusion of radicals from the hot<br />

boundary [11]. It has also been objected that flames cannot be treated by the methods<br />

of the Mechanics of Continua. This objection has originated such theories as the<br />

one proposed by van Tiggelen [12], who applies a method analogous to that used by<br />

Semenov in studying chain reactions.<br />

At present, however, it seems clearly established and wi<strong>de</strong>ly accepted that the<br />

correct method which should be applied to this problem is the one expan<strong>de</strong>d in the following<br />

paragraphs, and that its difficulty reduces to the use of the a<strong>de</strong>quate chemical<br />

kinetic scheme for each case; the knowledge of the corresponding physico-chemical<br />

constants and of the transport coefficients of the mixtures; and finally to the obtaining<br />

of satisfactory methods of solution of the difficult mathematical problem of the<br />

integration of the flame equations.<br />

In addition to this progress of theory, a substantial improvement has been<br />

achieved in the field of experimental techniques to measure the propagation velocity<br />

of the flame. An increasing number of data is now available on the influence of<br />

several parameters such as the composition, pressure, initial temperature of the mixture,<br />

etc., on the flame propagation velocity. As a bibliography for the study of these<br />

techniques, their limitations and applicability we recommend the works by Jost [13],<br />

Gaydon and Wolfhard [14] and specially the one of Ref. [15]. Linnet [16] has published<br />

a review on these methods, in which he classifies the experimental techniques<br />

most commonly used into the following five groups:<br />

1) Method of the bunsen burner (Gouy and Michelson). The most universal, of<br />

which a great number of variations are used.<br />

2) Method of the tube (Coward-Hartwell, Gerstein, etc). Not consi<strong>de</strong>red reliable<br />

enough.<br />

3) Method of the spherical flame in a constant volume bomb (Fick, Lewis-von<br />

Elbe). Not frequently used and difficult to observe.<br />

4) Method of the soap film or spherical flame at constant pressure. Not frequently<br />

used and limited to mixtures which are not sensitive to the influence of water<br />

vapor.


134 CHAPTER 6. LAMINAR FLAMES<br />

5) Method of the flat burner (Egerton-Pauling, Spalding-Botha). Indicated for mixtures<br />

with a low flame velocity.<br />

At the same time, several attempts have been ma<strong>de</strong>, with limited success, to<br />

measure the distribution of certain variables through the flame in or<strong>de</strong>r to analyze<br />

its structure. Information of this matter may be found in the above mentioned work<br />

by Linnet and in the Proceedings of the 6th International Symposium on Combustion.<br />

Initially, attempts were ma<strong>de</strong> to obtain temperature profiles whose results show<br />

a fair agreement with theoretical predictions. Later on, attempts were ma<strong>de</strong> to obtain<br />

distributions of the concentration of the main species, and finally, of the radicals.<br />

These observations are difficult due to the small thickness of the flame. Although it<br />

may be increased by reducing pressure, then the perturbation effects of the measuring<br />

instruments and of radiation become more important, and it is difficult to know the<br />

magnitu<strong>de</strong> of the corrections which must be introduced into the results obtained so<br />

that the values be exact.<br />

In the following paragraphs of this chapter, we shall <strong>de</strong>duce and discuss the<br />

flame equations, its boundary conditions and the methods of integration most commonly<br />

used. The application of the theory to some practical cases will be performed<br />

in §13, 14 and 15.<br />

6.2 Equation for the combustion wave in the case of<br />

two chemical species<br />

The equations for the propagation of a plane stationary combustion wave were <strong>de</strong>duced<br />

in the preceding chapter, system B, un<strong>de</strong>r the following assumptions:<br />

1) There are only two different chemical species, reactants and products.<br />

2) The mixture behaves like a perfect gas.<br />

3) The specific heat of the mixture is in<strong>de</strong>pen<strong>de</strong>nt from its composition and temperature.<br />

The system obtained un<strong>de</strong>r these conditions, referred to the axis advancing with the<br />

wave, and preserving the notation previously used, is as follows:<br />

a) Continuity equation.<br />

ρv = m (6.1)<br />

b) Chemical reaction equation.<br />

m dε = w(Y, p, T ) (6.2)<br />

dx


6.3. BOUNDARY CONDITIONS 135<br />

c) Diffusion equation.<br />

d) Momentum equation.<br />

e) Energy equation.<br />

f) State equation.<br />

ρD dY<br />

dx<br />

= m(Y − ε) (6.3)<br />

p = const. (6.4)<br />

c p T − qε − λ m<br />

dT<br />

dx = e (6.5)<br />

p<br />

ρ = R gT (6.6)<br />

In this system, m and e are two constants, whose values will result from the<br />

boundary conditions. The elimination of ρ through (6.6) reduces the system to three<br />

first or<strong>de</strong>r differential equations (6.2), (6.3) and (6.5) with three unknown ε, Y and T .<br />

6.3 Boundary conditions<br />

In or<strong>de</strong>r for the solution of this system to represent a stationary wave, it is necessary<br />

that it satisfies the following boundary conditions, which insures the transition from an<br />

uniform state of the unburnt gases before the wave to an uniform state of the products<br />

in chemical equilibrium after it,<br />

Unburnt gases, x → −∞ : Y → 0, ε → 0, T → T 0 . (6.7)<br />

Products, x → +∞ : Y → 1, ε → 1, T → T f . (6.8)<br />

These conditions imply, as well, the following<br />

x → ±∞ :<br />

dY<br />

dx → 0,<br />

dε<br />

dx → 0,<br />

dT<br />

dx<br />

→ 0. (6.9)<br />

The first of conditions (6.9) is satisfied by virtue of (6.3), since ε and Y take<br />

the same values both in x = +∞ and in x = −∞, by virtue of (6.7) and (6.8).<br />

The third condition (6.9) is also satisfied by virtue of (6.5) when the following<br />

values are assigned to q and e<br />

q = c p (T f − T 0 ),<br />

e = c p T 0 ,<br />

(6.10)


136 CHAPTER 6. LAMINAR FLAMES<br />

which, furthermore, enables (6.5) to be written as<br />

λ dT<br />

dx = mc (<br />

p (T − Tf ) + (T f − T 0 )(1 − ε) ) , (6.5.a)<br />

in which form it will be used further on.<br />

As for the second condition (6.9) it is also satisfied for x = +∞, since the<br />

condition of chemical equilibrium for the products may be expressed as<br />

w(1, ρ f , T f ) = 0. (6.11)<br />

To the contrary, condition dε → 0 for x → −∞, will not be satisfied unless<br />

dx<br />

w(0, ρ 0 , T 0 ) = 0, (6.12)<br />

which, generally, will be in contradiction with the laws of Chemical Kinetics.<br />

This circumstance implies a fundamental difficulty, whose origin lies upon the<br />

fact that it is impossible to maintain invariable the composition of the unburnt gases<br />

located before the wave, as imposed by boundary conditions (6.7) and (6.9), when the<br />

reaction velocity of such gases is not zero.<br />

The existence of stationary combustion waves observed in practice may be<br />

explained by the fact that such waves do not correspond exactly to the one-dimensional<br />

mo<strong>de</strong>l proposed in this work, since the mixing of reactant species forms, in general,<br />

shortly before they reach the wave, which always looses a certain amount of heat<br />

through the walls of the flame hol<strong>de</strong>r, for example, through the walls of the burner.<br />

Furthermore, these walls act as chain breakers for the chemical reactions of the wave.<br />

There is a possibility of eluding the above mentioned difficulty by incorporating<br />

to the one-dimensional mo<strong>de</strong>l the effect of all these complex circumstances, which<br />

prevent the reaction of the unburnt gases in a real flame, through a slight modification<br />

of the boundary conditions at the neighborhood of the “cold limit” of the flame, which<br />

may be attained in several different ways. The practical interest of these solutions<br />

is based on the fact that for reactions with an appreciable activation energy, as those<br />

expected in combustion, the structure and propagation velocity of the flame are insensible<br />

to a modification of the said boundary conditions, which makes it unnecessary to<br />

<strong>de</strong>fine them exactly, within a wi<strong>de</strong> range of values of the <strong>de</strong>fining parameters. Moreover,<br />

the proposed solutions are equivalent since they lead to the same fundamental<br />

results. The following paragraph is <strong>de</strong>voted to a discussion of these solutions.


6.4. MODIFICATION OF THE CONDITIONS AT THE “COLD BOUNDARY” 137<br />

6.4 Modification of the conditions at the “cold<br />

boundary”<br />

There are several solutions available in or<strong>de</strong>r to elu<strong>de</strong> the difficulty of the cold boundary,<br />

which has passed unnoticed until recently. The history of the evolution of thought<br />

on this subject may be followed by consulting references [17] through [19]. Herein<br />

we will only study the two solutions currently used, which consist in introducing an<br />

ignition temperature T i and a flame hol<strong>de</strong>r.<br />

Ignition temperature<br />

By adopting the same assumption used in the classical theories of Combustion, that<br />

is, by assuming that there is an ignition temperature T i , such that reaction velocity of<br />

the mixture is zero for T < T i , this temperature divi<strong>de</strong>s the combustion wave into<br />

two zones: a “heating and diffusion zone”, corresponding to temperatures un<strong>de</strong>r T i , at<br />

which no chemical reaction takes place, and a “reaction zone”, which corresponds to<br />

temperatures over T i .<br />

The differential equations for the reaction zone are the same given in the preceding<br />

paragraph, which are summarized in the following, as well as the boundary<br />

conditions, assuming that the origin of coordinates x = 0 is located at ignition point<br />

T = T i .<br />

Reaction zone, x > 0.<br />

m dε = w(Y, ρ, T ), (6.2)<br />

dx<br />

ρD dY<br />

dx<br />

= m(Y − ε), (6.3)<br />

λ dT<br />

dx = mc p<br />

The boundary conditions for this system will be<br />

(<br />

(T − Tf ) + (T f − T 0 )(1 − ε) ) . (6.5.a)<br />

x = 0 : ε = 0, T = T i , (6.13)<br />

x → +∞ : ε = 1, T = T f , Y = 1. (6.14)


138 CHAPTER 6. LAMINAR FLAMES<br />

The equations for the heating and diffusion zone, x < 0, may be obtained<br />

from those given §2 by introducing in them the conditions expressing the absence of<br />

chemical reaction, that is: w = ε = 0. There resulting<br />

Heating and diffusion zone, x < 0.<br />

ρD dY<br />

dx<br />

The corresponding boundary conditions are<br />

= mY, (6.15)<br />

λ dT<br />

dx = mc p(T − T 0 ). (6.16)<br />

x = −∞ : T → T 0 , Y → 0, (6.17)<br />

x = 0 : T = T i . (6.18)<br />

Since for x → −∞ both dY / dx and dT / dx tend to zero, as it can be verified<br />

by taking (6.17) into (6.15) and (6.16), the introduction of an ignition temperature<br />

eliminates the difficulty at the cold boundary. Furthermore by comparing the equations<br />

for the heating and diffusion zone with those for the reaction zone, given in the<br />

preceding paragraph, and the boundary conditions (6.13) and (6.18), it may be readily<br />

verified that the transition from one zone to another at point x = 0 is continuous for<br />

all variables when adopting the additional condition that the value for Y at x = 0,<br />

which is un<strong>de</strong>termined, be the same for both solutions<br />

x = 0 : Y (0 − ) = Y (0 + ). (6.19)<br />

The objection to this way of <strong>de</strong>aling with the difficulty at the cold boundary<br />

states that the solution obtained, and in special the propagation velocity for the flame,<br />

will <strong>de</strong>pend on the value adopted for T i , which does not actually exist. Nevertheless, it<br />

will be proved further on, when studying the solutions for the combustion wave, that as<br />

long as the velocity of the chemical reaction <strong>de</strong>pends substantially on the temperature<br />

of the mixture, it will happen that the solution obtained will be in<strong>de</strong>pen<strong>de</strong>nt from the<br />

values of T i , except for values of this temperature very close to T 0 or to T f . Un<strong>de</strong>r<br />

such conditions the influence of the ignition temperature will vanish completely from<br />

the result, when calculating the propagation velocity of the flame.<br />

The presence of a flame hol<strong>de</strong>r<br />

Another solution proposed to the problem of the cold boundary consists in imagining<br />

the presence of a flame hol<strong>de</strong>r placed in front of the wave which has the double mission


6.4. MODIFICATION OF THE CONDITIONS AT THE “COLD BOUNDARY” 139<br />

of absorbing a certain amount of heat Q from the flame and of acting as a filter which<br />

allows the unburnt gases to reach the flame but prevents the combustion products from<br />

diffusing towards the unburnt gases. By assuming that the hol<strong>de</strong>r is located at point<br />

x = 0, the system of Eqs. (6.2), (6.3) and (6.5a) will still hold, but conditions (6.13)<br />

at the cold boundary will have to be substituted by the following<br />

x = 0 : T = T 0 , ε = 0, λ dT<br />

dx<br />

≠ 0. (6.20)<br />

To the contrary, for x < 0 the composition and state of the mixture are uniform<br />

x < 0 : Y = ε = 0, T = T 0 . (6.21)<br />

The value of Y at x = 0 remains un<strong>de</strong>termined, but different from zero. Therefore<br />

through x = 0 there exists a discontinuity in the composition of the mixture which<br />

becomes possible due to the existence of the filter.<br />

It happens here, as in the case of ignition temperature, that the structure and<br />

propagation velocity of the flame are in<strong>de</strong>pen<strong>de</strong>nt from the value of parameter Q,<br />

provi<strong>de</strong>d it is not close to heat of reaction q or very close to zero, which justifies the<br />

practical value of the proposed mo<strong>de</strong>l.<br />

T f<br />

T f<br />

T T 0 i 1−ε T 0<br />

1−Y<br />

Karman ´ ´ boundary condition<br />

1−Y<br />

Hirschfel<strong>de</strong>r boundary condition<br />

1−ε<br />

Figure 6.1: Boundary conditions for ignition temperature and flame hol<strong>de</strong>r mo<strong>de</strong>ls.<br />

Figure 6.1 summarizes and compares the conditions at the cold boundary for<br />

the two solutions proposed. Since both lead to the same results, in the following we<br />

will use only the assumption that an ignition temperature exists.<br />

Figure 6.2 shows in a qualitatively way the form of the solutions obtained for<br />

the propagation velocity of the flame in a given mixture when the ignition temperature,<br />

assumed for it, is changed. It is seen that there is a wi<strong>de</strong> interval AB of ignition<br />

temperatures within which the propagation velocity of the flame is in<strong>de</strong>pen<strong>de</strong>nt from<br />

T i . The quantitative solutions for some typical cases will be inclu<strong>de</strong>d further on.


140 CHAPTER 6. LAMINAR FLAMES<br />

1<br />

0.8<br />

0.6<br />

S b<br />

0.4<br />

A<br />

B<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

T 0<br />

/T f T /T i f<br />

Figure 6.2: Schematic diagram showing the flame propagation velocity vs ignition temperature.<br />

6.5 Propagation velocity of the flame<br />

A simple check of the boundary conditions at the “reaction zone” shows that they<br />

are superabundant. In fact, since we are <strong>de</strong>aling with a system of three equations of<br />

first-or<strong>de</strong>r, the solution of the system is <strong>de</strong>termined for each value of m by boundary<br />

conditions (6.14), except for a translation, which is <strong>de</strong>termined by the additional condition<br />

that for x = 0 be T = T i as imposed by (6.13). In or<strong>de</strong>r to satisfy as well the<br />

additional condition x = 0 : ε = 0, it is necessary that m takes a particular value: the<br />

“eigenvalue” of the system which makes compatible boundary conditions (6.13) and<br />

(6.14).<br />

This eigenvalue will <strong>de</strong>termine the propagation velocity of the flame, which by<br />

virtue of (6.1) is given by<br />

u 0 = m . (6.22)<br />

ρ 0<br />

6.6 Example<br />

Even in the case of only two chemical species as consi<strong>de</strong>red in the preceding paragraphs,<br />

the non-linear character of the flame equations and in particular the shape of<br />

the reaction velocity add difficulties to the problem to such an extent that it becomes


6.6. EXAMPLE 141<br />

generally impossible to obtain explicit solutions for the equations. Hence, it is necessary<br />

to resort to cumbersome numerical integrations or to analytical methods which<br />

will be discussed further on. However, in or<strong>de</strong>r to illustrate the nature of the solutions<br />

and the way in which the eigenvalue appears <strong>de</strong>termined, the present paragraph offers<br />

an example in which through an a<strong>de</strong>quate simplification of both transport coefficients<br />

and chemical reaction velocity, it is possible to obtain an explicit solution of the problem<br />

giving correct qualitative predictions. A possible objection to this solution could<br />

be that the results obtained <strong>de</strong>pend on the ignition temperature of the mixture, which,<br />

as before said, does not exist. But this is due to the fact that the simplification assumed<br />

for the reaction velocity is excessive and consequently the inconvenient will vanish for<br />

more realistic cases, as will be proven in the following.<br />

For this example the following assumptions will be adopted:<br />

1) The coefficient of thermal conductivity λ and the product ρD are constant.<br />

2) The reaction is of first-or<strong>de</strong>r and its velocity is in<strong>de</strong>pen<strong>de</strong>nt from temperature, in<br />

which case it has the form 1 w = ρ 0 k(1 − Y ). (6.23)<br />

It is precisely this simplified form of the reaction velocity and, in special, the fact<br />

that in it the activation energy is assumed to be zero, an indispensable condition<br />

for the integrability of the system, which makes the solution <strong>de</strong>pends on the<br />

ignition temperature, as aforesaid.<br />

When expression (6.23) is taken into (6.2), it gives<br />

The diffusion equation subsists in the form (6.3)<br />

m dε<br />

dx = ρ 0k(1 − Y ). (6.24)<br />

ρD dY<br />

dx<br />

= m(Y − ε). (6.25)<br />

Finally, when the dimensionless temperature θ = T/T f is introduced, the energy<br />

equation (6.5.a) may be written<br />

where is θ 0 = T 0 /T f .<br />

λ dθ<br />

mc p dx = θ − 1 + (1 − θ 0)(1 − ε), (6.26)<br />

1 See §8 of Chap. 1.


142 CHAPTER 6. LAMINAR FLAMES<br />

In or<strong>de</strong>r to find the solution, it is convenient to eliminate x from the preceding<br />

system. This is done by dividing (6.24) and (6.25) by (6.26), thus resulting<br />

In this system<br />

dY<br />

dθ = L Y − ε<br />

θ − 1 + (1 − θ 0 )(1 − ε) , (6.27)<br />

dε<br />

dθ = Λ 1 − Y<br />

θ − 1 + (1 − θ 0 )(1 − ε) . (6.28)<br />

L =<br />

λ<br />

ρDc p<br />

(6.29)<br />

is the Lewis-Semenov number of the mixture, which is constant.<br />

Λ is a dimensionless parameter, also constant, <strong>de</strong>fined by<br />

Λ = λρ 0k<br />

m 2 c p<br />

. (6.30)<br />

The problem lies, precisely, in <strong>de</strong>termining the eigenvalue of this parameter. Once it is<br />

known, the velocity u 0 of the flame propagation may be <strong>de</strong>rived from it and Eq. (6.22),<br />

thus obtaining<br />

√<br />

λk<br />

u 0 = Λ −1/2 . (6.31)<br />

ρ 0 c p<br />

The system of equations (6.27) and (6.26) holds only within the reaction zone, where<br />

boundary conditions are<br />

where we have written θ i = T i /T f .<br />

θ = θ i : ε = 0, (6.32)<br />

θ = 1 : ε = θ = 1, (6.33)<br />

Precisely because these three conditions exist for a system of second-or<strong>de</strong>r, it<br />

is necessary to look for the value of Λ which makes them compatible.<br />

The system is readily integrated by testing lineal solutions of the form<br />

1 − ε = α(1 − θ), (6.34)<br />

1 − Y = β(1 − θ), (6.35)<br />

which satisfy boundary conditions (6.33) at the hot boundary.<br />

Condition (6.32) at the cold boundary gives the following relation<br />

1 = α(1 − θ i ). (6.36)


6.6. EXAMPLE 143<br />

Two more relations may be obtained by taking (6.34) and (6.35) into (6.27) and<br />

(6.28), there resulting<br />

α =<br />

β =<br />

Λβ<br />

−1 + (1 − θ 0 )α , (6.37)<br />

α − β<br />

−1 + (1 − θ 0 )α . (6.38)<br />

The three Eqs. (6.36), (6.37) and (6.38) become a <strong>de</strong>termined system for the<br />

computation of α, β and Λ. The solution is<br />

α = 1<br />

1 − θ i<br />

, (6.39)<br />

β = 1 (1 + 1 ) −1<br />

θ i − θ 0<br />

, (6.40)<br />

1 − θ i L 1 − θ i<br />

Λ = θ i − θ 0<br />

(1 + 1 )<br />

θ i − θ 0<br />

. (6.41)<br />

1 − θ i L 1 − θ i<br />

Once Λ is known, then Eq. (6.31) provi<strong>de</strong>s the following dimensionless expression<br />

for the propagation velocity of the flame<br />

θ = θ i ,<br />

u 0<br />

√<br />

ρ0 c p<br />

λk = Λ−1/2 . (6.42)<br />

Finally, since we know β, Eq. (6.35) gives the value Y i for Y corresponding to<br />

Y i = 1 − β(1 − θ i ). (6.43)<br />

The solution for the heating and diffusion zone, x < 0, will be obtained by<br />

integrating the following equation<br />

dY<br />

dθ = L Y , (6.44)<br />

θ − θ 0<br />

which results from (6.27) when making ε = 0. The solution to this equation is<br />

Y = Y i<br />

( θ − θ0<br />

θ i − θ 0<br />

) L<br />

. (6.45)<br />

Once the preceding solutions are known, one may pass to the physical space<br />

by making use of solutions (6.34) and (6.35) and Eqs. (6.24), (6.25) and (6.26), as<br />

follows. 2<br />

2 This far, the problem could have been solved un<strong>de</strong>r less drastic assumptions, i.e., not being constant λ<br />

nor ρD, but the Lewis-Semenov number.


144 CHAPTER 6. LAMINAR FLAMES<br />

a) Reaction zone.<br />

After substituting (6.34) into (6.26), we have<br />

dθ<br />

dξ = θ i − θ 0<br />

1 − θ 0<br />

(1 − θ), (6.46)<br />

where the following dimensionless distance was introduced<br />

being<br />

ξ = x l , (6.47)<br />

l =<br />

λ<br />

mc p<br />

(6.48)<br />

a characteristic length of the wave. The integration of (6.46), once it is assumed that<br />

reaction starts at point x = 0, gives<br />

Likewise, one obtains for ε and Y<br />

1 − θ = (1 − θ i ) e −θ i − θ 0<br />

1 − θ i<br />

ξ<br />

. (6.49)<br />

1 − ε = e −θ i − θ 0<br />

1 − θ i<br />

ξ<br />

, (6.50)<br />

1 − Y =<br />

(<br />

1 + 1 ) −1<br />

θ i − θ 0<br />

e −θ i − θ 0<br />

ξ<br />

1 − θ i . (6.51)<br />

L 1 − θ i<br />

1.0<br />

0.8<br />

Y(L=0.5)<br />

0.6<br />

ε, Y<br />

0.4<br />

Y(L=1)<br />

ε<br />

0.2<br />

Y(L=2)<br />

0.2 θ 0.6 0.8 1.0<br />

i<br />

=0.4<br />

θ =0.125 0 θ<br />

Figure 6.3: Distributions of Y and ε as a function of θ when the activation energy is zero.


6.7. REACTION VELOCITY 145<br />

1.0<br />

0.8<br />

Y(L=2)<br />

θ, ε, Y<br />

0.6<br />

0.4<br />

0.2<br />

θ 0<br />

θ<br />

Y(L=0.5)<br />

Y(L=1)<br />

ε<br />

0.0<br />

−10 −8 −6 −4 −2 0 2 4 6 8 10<br />

ξ<br />

Figure 6.4: Distributions of Y , ε and θ as functions of ξ, when the activation energy is zero.<br />

b) Heating zone.<br />

Likewise, in this zone one obtains<br />

θ = θ 0 + (θ i − θ 0 ) e ξ , (6.52)<br />

Y = 1 L<br />

θ i − θ 0<br />

(1 + 1 ) −1<br />

θ i − θ 0<br />

e Lξ , (6.53)<br />

1 − θ i L 1 − θ i<br />

ε = 0. (6.54)<br />

Figures 6.3 and 6.4 represent the solutions computed for the following typical<br />

values: θ 0 = 0.125, θ i = 0.4 and L = 0.5, 1 and 2.<br />

6.7 Reaction velocity<br />

The correct expression for the reaction velocity w is, in accordance with the laws of<br />

Chemical Kinetics, 3 w = ρ n k(1 − Y ) n . (6.55)<br />

3 See Chap. 1, §8.


146 CHAPTER 6. LAMINAR FLAMES<br />

In this expression, n is the or<strong>de</strong>r of reaction and k a function of temperature, which,<br />

generally, is of the form<br />

k = A<br />

( T<br />

T f<br />

) δ<br />

e −E/RT , (6.56)<br />

where A is a constant and E is the activation energy of the reaction.<br />

It is convenient to eliminate the <strong>de</strong>nsity from Eq. (6.55) so that only two variables<br />

appear explicitly: the temperature and the mass fraction of products.<br />

Let M r and M p be the molar masses of reactants and products, respectively,<br />

and a the relation<br />

Then, the mean molar mass of the mixture is 4<br />

a = M r<br />

M p<br />

− 1. (6.57)<br />

M =<br />

M r<br />

1 + aY , (6.58)<br />

and the <strong>de</strong>nsity of the mixture may be expressed as a function of pressure and temperature<br />

in the form<br />

By taking now Eq. (6.56) and (6.59) into (6.55), we finally obtain for the reaction<br />

velocity<br />

ρ = pM r<br />

RT<br />

1<br />

1 + aY . (6.59)<br />

w = A<br />

( ) n ( ) n ( ) δ pMr 1 − Y T<br />

e −E/RT . (6.60)<br />

RT 1 + aY T f<br />

It is advantageous to bring forth in Eq. (6.60) the dimensionless temperature θ =<br />

T/T f as <strong>de</strong>fined in the preceding paragraph, and to introduce dimensionless activation<br />

temperature θ a as <strong>de</strong>fined by<br />

Now, Eq. (6.60) may be written<br />

θ a =<br />

E<br />

RT f<br />

. (6.61)<br />

( ) n 1 − Y<br />

w = Bθ δ−n e −θ a/θ , (6.62)<br />

1 + aY<br />

where<br />

( ) n pMr<br />

B = A<br />

(6.63)<br />

RT f<br />

is a constant of the process. Eq. (6.62) is the form of the reaction velocity that will be<br />

used in the following.<br />

4 See Chap. 1, Eq. (1.36).


6.8. FLAME EQUATIONS 147<br />

6.8 Flame equations<br />

When (6.62) is taken into (6.2) , the later takes the form<br />

m dε ( ) n 1 − Y<br />

dx = Bθδ−n e −θ a/θ . (6.64)<br />

1 + aY<br />

Equation (6.3) is still written<br />

ρD dY<br />

dx<br />

= m(Y − ε). (6.65)<br />

If we substitute into Eq. (6.5.a) the dimensionless temperature, we have for it<br />

λ dθ<br />

dx = mc (<br />

p θ − 1 + (1 − θ0 )(1 − ε) ) . (6.66)<br />

In or<strong>de</strong>r to solve the above system it is convenient to divi<strong>de</strong> (6.64) and (6.65)<br />

by (6.66), as was done with the example treated in §6, thus obtaining<br />

In this system,<br />

(<br />

λ 1 − Y<br />

θ δ−n 1 − θ<br />

dε<br />

dθ = Λ(1 − θ λ f 1 + aY<br />

0)<br />

θ − 1 + (1 − ε)(1 − θ 0 ) e−θ a<br />

θ , (6.67)<br />

dY<br />

dθ = L Y − ε<br />

θ − 1 + (1 − θ 0 )(1 − ε) . (6.68)<br />

Λ =<br />

Bλ f e −θ a<br />

m 2 c p (1 − θ 0 )<br />

) n<br />

(6.69)<br />

is an un<strong>de</strong>termined parameter of the problem and L is the previously <strong>de</strong>fined Lewis-<br />

Semenov number of the mixture. 5<br />

The boundary conditions of system (6.67), (6.68) are the same given in (6.32)<br />

and (6.33), namely,<br />

Hot boundary: θ = 1, Y = 1, (6.70)<br />

Cold boundary: θ = θ i , ε = 0. (6.71)<br />

The problem lies in <strong>de</strong>termining the value of Λ which makes compatible these<br />

three conditions in the system formed by the two first-or<strong>de</strong>r equations (6.67) and (6.68).<br />

5 In the numerator of (6.67) we bring forth factor e −θ a 1−θ<br />

θ , in lieu of e − θa θ in or<strong>de</strong>r to simplify<br />

calculations since θ a is generally much larger than unity and otherwise, we would have to <strong>de</strong>al with very<br />

small numbers instead of doing it with numbers of the or<strong>de</strong>r of unity.


148 CHAPTER 6. LAMINAR FLAMES<br />

Once Λ is known then the velocity u 0 of the flame is given by<br />

√<br />

Bλ f e<br />

u 0 =<br />

−θa<br />

ρ 2 0 c p(1 − θ 0 ) Λ−1/2 . (6.72)<br />

6.9 Solution of the flame equations<br />

As before said, the preceding system cannot by integrated analytically. Hence, it<br />

becomes necessary to resort either to numerical solutions or else to semi-analytical<br />

approximation methods.<br />

The numerical methods are very cumbersome since we are <strong>de</strong>aling with an<br />

eigenvalue problem with given boundary conditions at both extremes of the interval,<br />

and, therefore, we would have to make numerous tentatives before reaching the exact<br />

solution. Generally, the use of electronic computers is required and in particular they<br />

were wi<strong>de</strong>ly utilized by Hirschfel<strong>de</strong>r and his collaborators.<br />

In recent years, several approximate methods have been proposed generally<br />

based in the unusual behavior of the solutions due to the presence of factor e −θa/θ in<br />

the reaction velocity. In fact, when the reduced activation temperature θ a is large, as<br />

should be expected in combustion reactions, due to this factor the result is <strong>de</strong>termined,<br />

essentially, by the form of the solution close to the hot boundary and this enables<br />

the <strong>de</strong>velopment of methods, as those proposed by Zeldovich-Frank-Kamenetskii-<br />

Semenov [4] and the one by Boys-Corner [5], Adams [20], Wil<strong>de</strong> [21] and von Kármán<br />

[6], with different variations and approximations. Only very recently a comparative<br />

study of these methods has been performed. Information on the subject will be found<br />

in the work by von Kármán [10] and, specially, in the more complete study carried<br />

out by Millán, Sendagorta and Da Riva [22]. Figures 6.5 and 6.6, taken from this<br />

study show a comparison between the approximate values for Λ −1/2 ( to which the<br />

flame velocity is proportional by virtue of (6.72)), obtained through several of those<br />

methods, with the exact value, given by a numerical integration. This comparison was<br />

performed for several values of the temperature of the unburnt gases and of the activation<br />

energy of the reaction. All the results shown in these figures correspond to a<br />

Lewis-Semenov number equal to unity and to a constant mean molar mass of the mixture.<br />

However, the unpublished results of the calculations performed for more general<br />

cases indicate that the same conclusions are still valid.<br />

Figures 6.5 and 6.6 show that the Boys-Corner method in its first iteration<br />

gives <strong>de</strong>fect values with an important <strong>de</strong>viation. The same happens with the method<br />

by Zeldovich et al., but giving excess values. The approximation improves with the


6.9. SOLUTION OF THE FLAME EQUATIONS 149<br />

second iteration of the Boys-Corner method but it is very arduous to perform and,<br />

furthermore, like the first one it does not converge towards the correct solution for<br />

increasing activation temperatures.<br />

1.5<br />

1.4<br />

1.3<br />

ZELDOVICH<br />

1.2<br />

Λ −1/2 / Λ −1/2<br />

exact<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

WILDE<br />

KARMAN 3<br />

SENDAGORTA<br />

BOYS−CORNER 2nd IT.<br />

KARMAN 2<br />

KARMAN 1<br />

0.7<br />

BOYS−CORNER<br />

0.6<br />

2 4 6 8 10 12 14 16<br />

θ a<br />

Figure 6.5: Comparison between the flame propagation problem eigenvalue obtained by<br />

different approximated methods and the exact value for θ 0 = 0.125.<br />

1.3<br />

Λ −1/2 / Λ −1/2<br />

exact<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

ZELDOVICH<br />

WILDE SENDAGORTA<br />

KARMAN 3<br />

BOYS−CORNER 2nd IT.<br />

KARMAN 2<br />

KARMAN 1<br />

BOYS−CORNER<br />

0.6<br />

2 4 6 8 10 12 14 16<br />

θ a<br />

Figure 6.6: Comparison between the flame propagation problem eigenvalue obtained by<br />

different approximated methods and the exact value for θ 0 = 0.250.


150 CHAPTER 6. LAMINAR FLAMES<br />

The other methods are consi<strong>de</strong>rably better mainly when applied to normal activation<br />

temperatures. This is specially true about Wil<strong>de</strong>’s method, the second and third<br />

alternatives of von Kármán’s one and that proposed by Sendagorta [23], which is the<br />

best and does not require more work than the others.<br />

1.0<br />

0.8<br />

BOYS−CORNER<br />

0.6<br />

EXACT<br />

θ<br />

0.4<br />

0.2<br />

KARMAN 3<br />

0.0<br />

−0.4 −0.2 0.0 0.2 0.4 0.6 0.8 1.0<br />

ε<br />

Figure 6.7: Comparison of the distributions of ɛ vs θ obtained by two approximate methods<br />

with the exact result.<br />

Figure 6.7 shows, for one of the case calculated, the law of the variation of ε<br />

with θ of the Boys-Corner approximation, Kármán’s third alternative and the exact<br />

solutions. It is evi<strong>de</strong>nt that the approximation reached with Kármán’s method is very<br />

satisfactory. The following studies the nature of this method after referring the rea<strong>de</strong>r<br />

to the above references for a more <strong>de</strong>tailed analysis of the other methods mentioned.<br />

If in Eq. (6.67) we bring the <strong>de</strong>nominator of the right hand si<strong>de</strong> into the left one,<br />

and this equation is integrated between the cold boundary (6.71) and the hot boundary<br />

(6.70), we obtain<br />

1 − θ 0<br />

2<br />

−<br />

∫ 1<br />

0<br />

∫ 1<br />

( ) n<br />

λ 1 − Y<br />

(1 − θ) dε = Λ(1 − θ 0 ) θ δ−n e −θ 1 − θ<br />

a<br />

θ dθ.<br />

θ i<br />

λ f 1 + aY<br />

(6.73)<br />

Consequently, in or<strong>de</strong>r to <strong>de</strong>termine Λ, the problem reduces now to obtaining<br />

an approximation of θ vs ε on the integral of the left si<strong>de</strong> of Eq. (6.73) and an approximation<br />

of Y vs θ on the integral of the right si<strong>de</strong> of the same equation. These integrals


6.9. SOLUTION OF THE FLAME EQUATIONS 151<br />

will be written for shortness as<br />

∫ 1<br />

( ) n<br />

λ 1 − Y<br />

I =(1 − θ 0 ) θ δ−n e −θ 1 − θ<br />

a<br />

θ dθ, (6.74)<br />

λ f 1 + aY<br />

J =<br />

∫ 1<br />

0<br />

θ i<br />

(1 − θ) dε. (6.75)<br />

With this notation, the fundamental Eq. (6.73) may be written<br />

1 − θ 0<br />

2<br />

− J = ΛI.<br />

The following is a separate study of both approximations.<br />

(6.73.a)<br />

Approximation of integral I<br />

Heat transfer λ is a function of temperature θ of the mixture and its composition<br />

<strong>de</strong>fined by the value of Y . 6 Consequently, in or<strong>de</strong>r to calculate I the problem reduces<br />

to finding an approximation for Y as a function of θ. The solution <strong>de</strong>pends on the<br />

value of the Lewis-Semenov number.<br />

Lewis-Semenov number equal to unity<br />

If L = 1, then Y is a lineal function of θ in the following form<br />

In fact, when condition<br />

Y = θ − θ 0<br />

1 − θ 0<br />

. (6.76)<br />

L ≡<br />

is satisfied, the diffusion Eq. (6.65) may be written in the form<br />

λ<br />

ρDc p<br />

= 1 (6.77)<br />

λ dY<br />

dx = mc p(Y − ε). (6.78)<br />

By adding to it the energy Eq. (6.66), multiplied by an arbitrary constant C,<br />

λ d<br />

(<br />

dx (Y + Cθ) = mc p Y + C(θ − θ 0 ) − ( C(1 − θ 0 ) + 1 ) )<br />

ε . (6.79)<br />

This equation is satisfied i<strong>de</strong>ntically with the following two conditions<br />

6 See Chap. 2, §4.<br />

C(1 − θ 0 ) + 1 = 0, (6.80)<br />

Y + C(θ − θ 0 ) = 0, (6.81)


152 CHAPTER 6. LAMINAR FLAMES<br />

the first is satisfied by adopting the following value for C<br />

C = − 1<br />

1 − θ 0<br />

. (6.82)<br />

This value when taken into Eq. (6.41) gives for Y the expression (6.76) which is, in<br />

fact, the <strong>de</strong>sired solution valid throughout the interval of temperatures as the boundary<br />

conditions are met at both extremes.<br />

θ = θ 0 : Y = 0, θ = 1 : Y = 1. (6.83)<br />

0.40<br />

θ a<br />

=4<br />

0.35<br />

θ 0<br />

=0.125<br />

Λ −1/2<br />

0.30<br />

0.25<br />

0.20<br />

0.5 1.0 1.5 2.0<br />

L<br />

Figure 6.8: Dimensionless flame velocity as a function of the Lewis number for θ a = 4 and<br />

θ 0 = 0.125.<br />

Lewis-Semenov number different from unity<br />

Actually the Lewis-Semenov number is only i<strong>de</strong>ntical to unity when <strong>de</strong>aling with pure<br />

mono-atomic gases (in which case D is the coefficient of autodiffusion). Generally<br />

in more complex cases L differs from unity and it varies with the composition and<br />

temperature of the mixture. However, it is frequently very close to unity and the<br />

preceding approximation is satisfactory. Moreover, the influence of the value of the<br />

Lewis-Semenov number on the flame velocity is, generally speaking, quite small. This<br />

is verified in Fig. 6.8 taken from Ref. [22] which shows, as an example, that when<br />

L doubles its value, the flame velocity increases less than 20 per cent, in this case<br />

corresponding to a first-or<strong>de</strong>r reaction with an activation temperature θ a = 4, which


6.9. SOLUTION OF THE FLAME EQUATIONS 153<br />

is small, and therefore the influence of L is more obvious than it would be for larger<br />

values of θ a .<br />

On the other hand, due to the presence of factor e −θ 1 − θ<br />

a<br />

θ<br />

in the integral of<br />

the right hand si<strong>de</strong> of Eq. (6.73) it occurs that for large values of θ a , as being those<br />

normally appearing in combustion, the value of the integral is influenced only by the<br />

value taken at the neighborhood of θ = 1 by the quantity un<strong>de</strong>r the integral.<br />

1.0<br />

0.8<br />

θ a<br />

=4<br />

L=0.5<br />

ε<br />

0.6<br />

θ<br />

Y<br />

0.4<br />

0.2<br />

θ 0<br />

=0.125<br />

0.0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

ε, Y<br />

Figure 6.9: Solutions of the flame equations 6.67 and 6.68 for L = 0.5, θ a = 4, θ 0 =<br />

0.125, θ i = 0.4, n = 1, a = 0 and δ = 1.<br />

In or<strong>de</strong>r to obtain the <strong>de</strong>sired approximation Y vs θ it is sufficient to observe<br />

Fig. 6.3 corresponding to the example treated in §7, since this figure shows that even<br />

when the activation energy θ a is zero it happens that near θ = 1 is 1 − ε ≫ 1 − Y ,<br />

and also 1 − ε ≫ 1 − θ. This fact appears more evi<strong>de</strong>nt for reactions with a high<br />

activation temperature as shown in Fig. 6.9 which was calculated for a more realistic<br />

case. Eq. (6.68) may be written<br />

dY<br />

dθ = L (1 − ε) − (1 − Y )<br />

1 − θ 0<br />

(1 − ε) − 1 − θ , (6.84)<br />

1 − θ 0<br />

and, in accordance with the preceding consi<strong>de</strong>rations, both 1 − Y and 1 − θ<br />

1 − θ 0<br />

are negligible<br />

close to θ = 1 and the following approximation may be obtained for Eq. (6.84)<br />

θ ≃ 1 :<br />

dY<br />

dθ ≃ L , (6.85)<br />

1 − θ 0


154 CHAPTER 6. LAMINAR FLAMES<br />

whose solution is<br />

θ ≃ 1 : 1 − Y ≃ L<br />

1 − θ 0<br />

(1 − θ), (6.86)<br />

this approximation must replace Eq. (6.76) when L is different from unity and furthermore<br />

the value taken for L must be that corresponding to the temperature and<br />

composition of the hot boundary conditions of the flame, T = T f .<br />

If the activation energy has a small value and an approximation of Y vs θ that<br />

will be valid within a wi<strong>de</strong>r range of temperatures is <strong>de</strong>sired, then the linear approximation<br />

(6.86) may be substituted by the following parabolic<br />

Y ≃<br />

( θ − θ0<br />

1 − θ 0<br />

) L<br />

, (6.87)<br />

which for θ = 1 coinci<strong>de</strong>s with (6.86) and for θ ≃ θ 0 it behaves like the solution<br />

corresponding to the heating zone (ε ≡ 0). This can readily be verified. Fig. (6.10)<br />

shows some of the curves (6.86) and (6.87).<br />

1.0<br />

0.8<br />

L=0.5<br />

L=0.75<br />

0.6<br />

L=1.5<br />

L=2.0<br />

Y<br />

0.4<br />

L=1.0<br />

0.2<br />

0.0<br />

0.125 0.2 0.4 0.6 0.8 1.0<br />

θ<br />

Figure 6.10: Linear and parabolic approximations of Y vs θ given by Eqs. 6.86 and 6.87<br />

for θ 0 = 0.125.<br />

The preceding consi<strong>de</strong>rations prove that it is generally sufficient, when calculating<br />

the integral to take for λ/λ f the value unity which corresponds to the hot<br />

boundary, or some other approximation at the neighborhood of this point. Frequently<br />

the following approximations are used<br />

λ<br />

λ f<br />

≃ θ,<br />

λ<br />

λ f<br />

≃ √ θ. (6.88)


6.9. SOLUTION OF THE FLAME EQUATIONS 155<br />

1.4<br />

1.3<br />

Λ −1/2 / Λ −1/2<br />

λ=λ λ=λ θ<br />

f f<br />

1.2<br />

1.1<br />

θ 0<br />

=0.250<br />

θ 0<br />

=0.125<br />

1.0<br />

2 4 6 8 10 12 14 16<br />

θ a<br />

Figure 6.11: Ratio of flame velocities corresponding to λ = λ f and λ = λ f θ as a function<br />

of the activation temperature θ a for two different values of the initial temperature<br />

θ 0.<br />

Figure 6.11, taken from Ref. [22], sets forth the scarce influence of the variation<br />

of λ with temperature on the flame velocity, specially for high activation temperature.<br />

Two different solution are compared in this figure, one assuming that λ is<br />

constant through the flame λ = λ f ; and another where λ varies linearly with temperature<br />

λ = λ f θ, in accordance with (6.88). The solution was computed for two different<br />

values of the temperature of unburnt gases The <strong>de</strong>viations are un<strong>de</strong>r 10 per cent.<br />

Once established that the preceding consi<strong>de</strong>rations solve the problem of the<br />

calculation of integral I we shall proceed with the study of J.<br />

Approximation of integral J<br />

In or<strong>de</strong>r to approximate integral J, we must first consi<strong>de</strong>r that in the normal cases, this<br />

is to say when θ a ≫ 1, this integral is much smaller than the other term of the left<br />

hand si<strong>de</strong> of Eq. (6.73), that is<br />

J ≪ 1 − θ 0<br />

. (6.89)<br />

2<br />

Such a property is illustrated in Fig. 6.12 which corresponds to a typical case<br />

where the value of J is shown by a dot-area while the value for (1 − θ 0 )/2 is represented<br />

by the area of triangle ABC.


156 CHAPTER 6. LAMINAR FLAMES<br />

1.0<br />

A<br />

1<br />

∫ (1−θ) dθ<br />

θ0<br />

0.8<br />

0.6<br />

θ<br />

0.4<br />

(1−θ 0<br />

)/2<br />

0.2<br />

θ 0<br />

=0.125<br />

C<br />

B<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

ε<br />

Figure 6.12: Areas representing the values of J = R 1<br />

θ 0<br />

(1 − θ)dε and (1 − θ 0)/2 in a typical<br />

case.<br />

Hence, it results that a first approximation is obtained when J is neglected<br />

respect to (1 − θ 0 )/2, in which case the value of Λ is given by expression<br />

Λ = 1 − θ 0<br />

. (6.90)<br />

2I<br />

Actually, this approximation was introduced by Zeldovich et al., although by<br />

means of a more complicated justification. Furthermore, the authors applied an ina<strong>de</strong>quate<br />

approximation for the value of integral I which is the main reason for the<br />

error resulting from their method as it was shown in Figs. 6.5 and 6.6. Professor<br />

von Kármán, after a more accurate calculation of integral I, as before said, has improved<br />

the approximation obtained by Zeldovich et al. This can be seen in Figs. 6.5<br />

and 6.6 where the curves named Kármán correspond to (6.90) when using for I the<br />

approximation <strong>de</strong>veloped in the preceding paragraph.<br />

Von Kármán improved the approximation for J, through an approximation of<br />

1 − θ vs 1 − ε, which is obtained by studying the behavior of the differential equation<br />

(6.67) at the neighborhod of the hot boundary in the following way.<br />

Close to point θ = 1, difference 1 − ε is an infinitesimal, whose or<strong>de</strong>r <strong>de</strong>pends<br />

only on the or<strong>de</strong>r of 1 − θ and it can be easily <strong>de</strong>rived from (6.67) by introducing into<br />

it the following approximations:


6.9. SOLUTION OF THE FLAME EQUATIONS 157<br />

1) We neglect 1 − θ respect to (1 − θ 0 ) (1 − ε), in agreement with the procedure<br />

used in obtaining (6.85).<br />

2) We substitute (1 − Y ) by its approximation as a function of (1 − θ), given by<br />

Eq. (6.86).<br />

3) The remaining terms of (6.67) are substituted by the corresponding values at the<br />

hot boundary.<br />

Once these approximation are introduced into (6.67) one obtains the following<br />

approximation for this equations, which is valid at the neighborhood of the hot<br />

boundary<br />

1 − θ ≪ (1 − ε)(1 − θ 0 ) :<br />

dε<br />

dθ ≃<br />

Λ (<br />

) n<br />

L (1 − θ) n<br />

1 − θ 0 (1 − θ 0 )(1 + a) 1 − ε . (6.91)<br />

After integration the following approximation 1 − θ vs 1 − ε, is obtained<br />

( n + 1<br />

1 − θ ≃<br />

2<br />

) 1<br />

1 − θ 0<br />

Λ<br />

n + 1 ( (1 − θ 0 )(1 + a)<br />

L<br />

which when taken into (6.75) and integrated, gives for J<br />

J = n + 1 ( n + 1<br />

n + 3 2<br />

) 1<br />

1 − θ 0<br />

Λ<br />

n + 1 ( (1 − θ 0 )(1 + a)<br />

L<br />

) n<br />

n + 1 (1 − ε)<br />

2<br />

n + 1 , (6.92)<br />

) n<br />

n + 1 . (6.93)<br />

By taking this value of J and the one for I given by Eq. (6.74) into Eq. (6.73), the<br />

following equation results for the calculation of Λ<br />

1 − θ 0<br />

2<br />

− n + 1 ( n + 1<br />

n + 3 (1 − θ 0)<br />

2<br />

) 1<br />

n + 1 ( 1 + a<br />

L<br />

) n<br />

n + 1 Λ<br />

− 1<br />

n + 1 = ΛI, (6.94)<br />

which may be readily solved, either with a graphical method or by iteration, using for<br />

this last case as a first approximation the one given by (6.90) and as the correcting<br />

term for successive iterations the right hand si<strong>de</strong> of Eq. (6.94).<br />

The result reached with (6.94) are represented in Figs. 6.5 and 6.6 where the<br />

corresponding curves are named Kármán 2.<br />

The approximation for J may still be improved by writing it in the form<br />

J =<br />

∫ 1<br />

θ<br />

(1 − θ) dε dθ, (6.95)<br />

dθ<br />

and using for dε an approximation of (6.67) <strong>de</strong>rived as follows<br />

dθ<br />

(1 − ε) dε<br />

dθ ≃<br />

Λ (<br />

L<br />

1 − θ 0 (1 − θ 0 )(1 + a)<br />

) n<br />

λ<br />

λ f<br />

θ δ−n e −θ a<br />

1 − θ<br />

θ (1 − θ) n , (6.96)


158 CHAPTER 6. LAMINAR FLAMES<br />

which through integration, gives<br />

√ ( 2Λ L<br />

1 − ε ≃<br />

1 − θ 0 (1 − θ 0 )(1 + a)<br />

) n 2<br />

√ ∫ 1<br />

λ (1 − θ) n<br />

λ f θ n−δ e −θ 1 − θ<br />

a<br />

θ dθ. (6.97)<br />

This expression, when taken into the <strong>de</strong>nominator of (6.67) after we have neglected<br />

1 − θ, supplies an approximation for dε/ dθ which introduced in (6.95) gives<br />

the <strong>de</strong>sired value for J. The result obtained from this approximation corresponds to<br />

the curves named Kármán 3 of Fig. 6.5 and 6.6.<br />

Tho approximation for J is even better when, after Sendagorta, we calculate<br />

dε/ dθ in (6.95) by using for ε vs θ an approximation of the form<br />

where P is a polynomial of the form<br />

θ<br />

ε = P e −θ 1 − θ<br />

a<br />

θ , (6.98)<br />

P = 1 + a 1 (1 − θ) + a 2 (1 − θ) 2 + · · · + a m (1 − θ) m , (6.99)<br />

whose coefficients must be <strong>de</strong>termined by studying the behavior ε close to θ = 1 in<br />

the differential equation (6.67).<br />

For example, in the case of a first-or<strong>de</strong>r reaction (n = 1), it gives for J<br />

J = H 0 + a 1 H 1 + a 2 H 2 + · · · + a m H m , (6.100)<br />

where<br />

H j =<br />

∫ 1<br />

θ i<br />

(1 − θ) j e −θ a<br />

1 − θ<br />

θ dθ, (j = 1, 2, . . . m). (6.101)<br />

This integral may be expressed by a law of recurrences starting from H 0 , which<br />

in turn is expressed through the exponential integral in the following way<br />

being<br />

H 0 =<br />

∫ 1<br />

θ i<br />

e −θ 1 − θ<br />

a<br />

θ dθ<br />

= 1 − θ a e θ a<br />

E 1 (θ a ) − θ i e −θ a<br />

E 1 (θ) =<br />

∫ ∞<br />

θ<br />

1 − θ i<br />

θ i<br />

e −z<br />

z<br />

+ θ a e θ a<br />

E 1<br />

(<br />

θa<br />

θ i<br />

)<br />

,<br />

(6.102)<br />

dz. (6.103)<br />

For example, if in the preceding case one limits approximation (6.99) to the first two<br />

terms,<br />

P = 1 + a 1 (1 − θ), (6.104)


6.10. STRUCTURE OF THE COMBUSTION WAVE 159<br />

in which case the only un<strong>de</strong>termined parameter would be a 1 to be obtained when<br />

comparing the value for ( dε/ dθ) θ=1 given by (6.98),<br />

( ) dε<br />

= θ a − a 1 , (6.105)<br />

dθ<br />

θ=1<br />

with the one resulting from (6.67)<br />

( ) dε<br />

(1 − θ 0 = ( )<br />

dθ<br />

θ=1<br />

dε<br />

(1 − θ 0 )<br />

dθ<br />

θ=1<br />

The following second-<strong>de</strong>gree equation will be obtained for a 1<br />

. (6.106)<br />

− 1<br />

1 − θ 0<br />

θ a − a 1 = Λ<br />

(1 − θ 0 )(θ a − a 1 ) − 1 , (6.107)<br />

which when solved gives the value of the un<strong>de</strong>termined parameter, which in turn,<br />

when taken into (6.100) provi<strong>de</strong>s the <strong>de</strong>sired approximation for J.<br />

Therefore, this approximation of J <strong>de</strong>pends on Λ, and when introduced into<br />

(6.73.a), it supplies an equation for Λ, whose solution gives the eigenvalue which<br />

solves the problem.<br />

6.10 Structure of the combustion wave<br />

Once <strong>de</strong>termined the value of Λ which makes compatible the boundary conditions of<br />

the flame equations, a simple numerical integration of such equations, through any of<br />

the methods available, will give the distribution of temperature, concentrations, etc.<br />

through the combustion wave. If the integration is performed by using system (6.67),<br />

(6.68), the curves obtained will be as those shown in Fig. 6.9. When the structure is<br />

<strong>de</strong>sired on a physical plane it is sufficient to integrate the system of equations (6.64),<br />

(6.65) and (6.66). However for this case one must select the position of origin of coordinates<br />

which is un<strong>de</strong>termined. Fig. 6.13 shows the solution corresponding to the case<br />

represented in Fig. 6.9. It is seen that the structure is analogous to the one represented<br />

in Fig. 6.4, corresponding to the simplified case discussed in §6. Fig. 6.13, shows a<br />

dot-line representation of the distribution of temperatures corresponding to the case of<br />

pure thermal conduction, where chemical reaction is absent, which practically coinci<strong>de</strong>s<br />

with the wave up to a θ = 0.5, proving that up this point the action of chemical<br />

reaction is extremely small.<br />

the same<br />

As a characteristic length of the phenomenon it was adopted, for its solution,<br />

l = λ f<br />

mc p<br />

(6.108)


160 CHAPTER 6. LAMINAR FLAMES<br />

applied before in §6. This is the characteristic length of the process and it is a measure<br />

of the thickness of the flame, since although theoretically it is infinite, almost the<br />

entire variation of the temperature and composition of the gases takes place within a<br />

narrow zone which is a few times the value of l, as shown in the solution computed in<br />

Fig. 6.13.<br />

By calculating the value of l corresponding to a typical case it may be verified<br />

that it is a small fraction of a millimetre and, consequently, the thickness of the flame,<br />

un<strong>de</strong>r normal conditions, has the same or<strong>de</strong>r of magnitu<strong>de</strong>, as proven by experimental<br />

observation.<br />

This circumstance increases consi<strong>de</strong>rably the difficulties of the experimental<br />

observation of the structure of the flame, since the recording of the distribution of<br />

temperatures, for instance, must be performed within a very narrow zone and when<br />

attempting to measure the composition the difficulties encountered are even greater.<br />

Several attempts have been ma<strong>de</strong> of measurements of this nature, mainly by the research<br />

team of the Applied-Physics Laboratory of the Johns Hopkins University. A<br />

<strong>de</strong>scription of the present state of this problem may be found in Ref. [24]. The measurements<br />

performed agree completely with theoretical predictions.<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

Y<br />

ε<br />

Y, θ, ε<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

θ<br />

L=0.5<br />

θ =4 a<br />

θ 0<br />

=0.125<br />

θ =0.4 i<br />

0<br />

−2.0 −1.5 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0 2.5<br />

ξ=mc p<br />

x/λ<br />

Figure 6.13: Structure of the combustion wave represented in Fig. 6.9. The origin of coordinates<br />

corresponds to the ignition temperature, where chemical reaction starts.


6.11. IGNITION TEMPERATURE 161<br />

6.11 Ignition temperature<br />

We have established that, except when activation temperature θ a is zero or very small,<br />

the ignition temperature θ i bears no influence on the flame velocity unless it has a<br />

value close to the temperature of the burnt gases θ = 1, or very close to initial temperature<br />

θ 0 .<br />

1.6<br />

1.2<br />

θ a<br />

=0<br />

Λ −1/2<br />

0.8<br />

θ a<br />

=2<br />

0.4<br />

θ a<br />

=4<br />

θ a<br />

=8<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

θ =0.125 0<br />

θ i<br />

Figure 6.14: Flame velocity as a function of the ignition temperature θ i for different values<br />

of the activation temperature θ a.<br />

This property is illustrated by Fig. 6.14, where the law of variation of Λ −1/2<br />

(to which u 0 is proportional as before said), as a function of θ i is shown for a typical<br />

case, corresponding to a first-or<strong>de</strong>r reaction, with a Lewis-Semenov number equal to<br />

unity and a temperature of the unburnt gases θ 0 = 0.125. The figure represents the<br />

solutions corresponding to values of θ a comprised between 0 and 8 and it is seen that<br />

except for the first case (θ a = 0), there exists a wi<strong>de</strong> range on values of θ i in which<br />

Λ −1/2 is practically constant, as it was previously announced. 7<br />

This property is of a general nature and it justifies the assumption of a ignition<br />

temperature which eliminate the difficulty of the cold boundary.<br />

7 For θ a ≫ 1 and θ i very close to θ 0 , it can be shown by means of asymptotic techniques that Λ −1/2 =<br />

„<br />

(1 − θ 0 ) ln` 1 ´ 1 − θ 0 ´«1/2<br />

exp`−θa , so that when θ i → θ 0 there exists an exponentially<br />

θ a(θ i − θ 0 )<br />

θ 0<br />

thin layer in which the transition for Λ −1/2 = constant to Λ −1/2 = ∞ occurs, Ed.


162 CHAPTER 6. LAMINAR FLAMES<br />

0.05<br />

0.04<br />

θ a<br />

=8<br />

(1−θ)e −θ a (1−θ)/θ<br />

0.03<br />

0.02<br />

0.01<br />

θ l<br />

0.00<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

θ<br />

Figure 6.15: Curve showing the typical values of the integrand of I as a function of θ.<br />

The reason for the insensibility of the solutions to θ i appears clearly in Fig. 6.15<br />

where it is seen that the values of θ i comprised between θ 0 and θ l do not influence in<br />

the value of integral I, since in this zone the integrand is practically zero. To the<br />

contrary, if θ i is comprised between θ l and 1, then one disregards in the solution the<br />

contribution of the part corresponding to θ l < θ < θ i . Since Λ −1/2 is proportional to<br />

√<br />

I, this explains the fact that Λ −1/2 should <strong>de</strong>crease and tend to zero when is θ i → 1.<br />

The fact that the combustion velocity tends to ∞ when θ i → θ 0 may be physically<br />

explained since then all the mass of gas burns simultaneously giving infinite<br />

wave velocity. In or<strong>de</strong>r to give a theoretical explanation it would be necessary to<br />

analyze in <strong>de</strong>tail the differential system.<br />

The fact that Λ −1/2 is in<strong>de</strong>pen<strong>de</strong>nt from θ i enables the elimination of this value<br />

in all the equations of the preceding paragraphs, in special in the lower limit of integral<br />

I (Eq. 6.74), by substituting it for zero without changing the result.<br />

6.12 General equations for the combustion wave in the<br />

case of more than two chemical species<br />

Hirschfel<strong>de</strong>r and his collaborators were the first to give the general equations for the<br />

combustion wave in the case of more than two chemical species [7]. A <strong>de</strong>rivation of<br />

these equations may be found in a paper by Kármán and Penner [6].


6.12. GENERAL EQUATIONS FOR THE COMBUSTION WAVE 163<br />

Such equations are a particular case of those corresponding to the one-dimensional<br />

stationary flow of a mixture of gases which were <strong>de</strong>duced in §8, chapter 3 of this book<br />

and will be used as a starting point for the present study, although certain modifications<br />

and simplifications will be required in or<strong>de</strong>r to adapt these equations to the<br />

problem in which we are now interested.<br />

Reaction equations<br />

The reaction equation which corresponds to species i is the Eq. (3.63) of the above<br />

mentioned chapter<br />

m dε i<br />

dx = w i, (i = 1, 2, . . . , l). (6.109)<br />

where l is the number of different chemical species. In this expression w i is the<br />

reaction velocity of species i, which, in the case of more than one chemical reaction,<br />

is the resultant of the contributions of each one of them<br />

r∑<br />

w i = w ij . (6.110)<br />

j=1<br />

Here, w ij is the mass of species i produced per volume unit and time unit due<br />

to reaction j, and r is the number of different reactions, counting as such, the two<br />

opposite ones which take place in each reaction.<br />

Velocity w ij corresponding to reaction j is given by expression (1.119) of<br />

chapter 1 which will be written herein in the following way by bringing forth in it<br />

the molar fractions X of the species in lieu of the mass fractions which were applied<br />

in the case of two chemical species (this change is done for the reasons stated further<br />

on the diffusion equations)<br />

w ij = M i (ν ′′<br />

ij − ν ′ ij)k j<br />

( p<br />

RT<br />

) nj<br />

l∏<br />

s=1<br />

X ν′ sj<br />

s . (6.111)<br />

In this expression M i is the molar mass of species i, ν ij ′ and ν′′ ij are the stoichiometric<br />

coefficients of this species in the left and right si<strong>de</strong>s of reaction equation j, n j =<br />

l∑<br />

ν sj ′ is the or<strong>de</strong>r of the reaction and k j is a function of the temperature of the mixture<br />

s=1<br />

which, as in the case of two species, is generally of the form<br />

k j = A j T δ j<br />

e −E j/RT . (6.112)<br />

When (6.112) is taken into (6.111), the equation may be written for shortness<br />

w ij = B j T δ j−n j<br />

e −E j/RT<br />

l ∏<br />

s=1<br />

X ν′ sj<br />

s , (6.111.a)


164 CHAPTER 6. LAMINAR FLAMES<br />

which is the form that will be used is the following. In the equation it has been taken<br />

B j = A j<br />

( p<br />

R<br />

) nj<br />

. (6.113)<br />

As aforesaid said in chapter 3, §8, even when there exist l reaction equations (6.109)<br />

corresponding to i chemical species they are not all in<strong>de</strong>pen<strong>de</strong>nt from one another.<br />

In fact, the number of in<strong>de</strong>pen<strong>de</strong>nt equations is the smallest of the following two: 1)<br />

number of chemical reactions; 2) number of in<strong>de</strong>pen<strong>de</strong>nt components of the mixture,<br />

in the sense of the rule of phases.<br />

If we take (6.111.a) into (6.109), the latter may be written in the form<br />

m dε i<br />

dx =<br />

r∑<br />

j=1<br />

M i B j (ν ′′<br />

ij − ν ′ ij)T δj−nj e −E j/RT<br />

which is the form that will be used in the following.<br />

l ∏<br />

s=1<br />

X ν′ sj<br />

s , (6.109.a)<br />

Diffusion equations<br />

Since it is evi<strong>de</strong>nt that<br />

ρv = m, (6.114)<br />

equation (3.68) of Chap. 3 may be written<br />

ρY j v dj = m(ε j − Y j ), (j = 1, 2, . . . , l). (6.115)<br />

If we now take this expression of the diffusion velocities into the equations of system<br />

(3.75) (which in this case still holds if we nullify the term corresponding to pressure<br />

diffusion since it is neglectable), we shall obtain the <strong>de</strong>sired system of diffusion equations.<br />

However, when written in this form, the said system has the inconvenient that<br />

the <strong>de</strong>rivatives of the mass fractions dY j / dx do not appear explicit as it would be<br />

necessary in or<strong>de</strong>r to obtain the system of equations for the flame in the canonical<br />

form. Such an inconvenience can be easily avoi<strong>de</strong>d by simply expressing the result<br />

as a function of molar fractions X i in lieu of the mass fractions. In fact, in this case<br />

the system of diffusion equations to be used is the one given by Eq. (2.28), which<br />

after applied to the one-dimensional flow and neglecting the pressure and temperature<br />

diffusions takes the form<br />

dX i<br />

dx =<br />

l∑<br />

j=1<br />

X i X j<br />

D ij<br />

(v dj − v di ). (6.116)


6.12. GENERAL EQUATIONS FOR THE COMBUSTION WAVE 165<br />

On the other hand, consi<strong>de</strong>ring that 8<br />

and<br />

Y j = M j<br />

M m<br />

X j (6.117)<br />

ρ = pM m<br />

RT , (6.118)<br />

the elimination of v di , v dj , Y j and ρ between these three last equations and (6.115),<br />

finally gives the following system of diffusion equations in the form which will be<br />

applied to the solution of the flame equations<br />

dX i<br />

dx = mRT<br />

p<br />

l∑<br />

j=1<br />

(<br />

)<br />

1 ε j ε i<br />

X i − X j , (i = 1, 2, . . . , l). (6.119)<br />

D ij M j M i<br />

Energy equation<br />

The energy equation is the one give in Eq. (3.74), when we disregard in it the terms<br />

due to kinetic energy and viscosity, thus reducing it to the following<br />

m<br />

l∑<br />

j=1<br />

ε j h j − λ dT<br />

dx<br />

where e is a constant <strong>de</strong>fined by the initial or final conditions.<br />

= e, (6.120)<br />

Before transforming this equations into the form in which it will be applied we<br />

shall first consi<strong>de</strong>r the boundary conditions.<br />

Boundary conditions<br />

These conditions are analogous to those applied in the case of two chemical species,<br />

namely,<br />

Cold boundary, x → −∞:<br />

T → T 0 , ε i → ε i0 , X i → X i0 ,<br />

dT<br />

dx → 0, dε i<br />

dx → 0, dX i<br />

→ 0. (6.121)<br />

dx<br />

Hot boundary, x → +∞:<br />

T → T f , ε i → ε if , X i → X if ,<br />

dT<br />

dx → 0, dε i<br />

dx → 0, dX i<br />

→ 0. (6.122)<br />

dx<br />

8 See Chap. 1, Eq.1.34


166 CHAPTER 6. LAMINAR FLAMES<br />

In (6.121), T 0 , ε i0 = Y i0 and X i0 = Y i0 M m0 /M i are the values of T , ε i and<br />

X i corresponding to the temperature and composition of the unburnt gases. M m0 is<br />

the value of M m un<strong>de</strong>r these conditions.<br />

Likewise, in (6.122), T f , ε if = Y if and X if = Y if M mf /M i are the temperature<br />

and composition corresponding to the adiabatic combustion products in chemical<br />

equilibrium.<br />

A recount of conditions, similar to the one performed in §6 for the case of two<br />

chemical species, shows that conditions are superabundant and that their compatibility<br />

imposes that m takes an “eigenvalue” which <strong>de</strong>termines the propagation velocity of<br />

the flame.<br />

Boundary conditions (6.121) and (6.122) <strong>de</strong>termine the value of constant e in<br />

(6.120), when it is particularized for both extremes, thus obtaining<br />

l∑<br />

l∑<br />

e = m ε j0 h j0 = m ε jf h jf . (6.123)<br />

j=1<br />

j=1<br />

In this expression the equality of the last two terms expresses simply that the combustion<br />

is adiabatic and at constant pressure.<br />

form<br />

Substituting (6.123) into (6.120) the latter may be written as<br />

λ dT<br />

dx = m<br />

l∑<br />

(ε j h j − ε jf h jf ). (6.124)<br />

j=1<br />

In Chap. 1, §3, it was established that the enthalpy of a diluted gas is of the<br />

∫ T<br />

h j = h 0 j + c pj dT, (6.125)<br />

T 0<br />

where h 0 j is the formation enthalpy9 of species j at temperature T 0 and c pj is the<br />

specific heat of the same at constant pressure.<br />

When taking (6.125) into (6.124) the latter is written<br />

⎛<br />

⎞<br />

λ dT<br />

dx = m ⎝ ∑ ∫ T ∑<br />

h 0 jε j + c pj ε j dT ⎠<br />

j<br />

T 0 j<br />

⎛<br />

⎞<br />

− m ⎝ ∑ ∫ Tf ∑<br />

h 0 jε jf + c pj ε jf dT ⎠ .<br />

j<br />

T 0<br />

j<br />

(6.126)<br />

9 To avoid confusion with h j0 , total enthalpy of species j at T 0 , the notation is slightly different from<br />

that used in chapter 1. Ed.


6.12. GENERAL EQUATIONS FOR THE COMBUSTION WAVE 167<br />

In this equation, expression ∑ c pj ε j <strong>de</strong>pends on temperature T and on the values for<br />

j<br />

mass fluxes ε j of the species. In or<strong>de</strong>r to simplify this expression we adopt a mean<br />

value c p in<strong>de</strong>pen<strong>de</strong>nt from temperature and the composition of the mixture so that<br />

Eq. (6.126) may be written in the form<br />

λ dT (∑<br />

)<br />

dx = m h 0 j(ε j − ε jf ) + c p (T − T f ) . (6.127)<br />

j<br />

The value for c p is <strong>de</strong>rived from (6.127) when expressing the condition that<br />

(6.127) vanishes at the cold boundary in agreement with (6.121) obtaining<br />

∑<br />

j<br />

c p =<br />

h0 j (ε j0 − ε jf )<br />

, (6.128)<br />

T f − T 0<br />

or else since, as before said, ε jf = Y jf and ε j0 = Y j0 due to the fact that diffusion is<br />

absent in both limits<br />

c p =<br />

∑<br />

j h0 j (Y j0 − Y jf )<br />

T f − T 0<br />

. (6.129)<br />

If, as before done, we introduce dimensionless temperature θ = T/T f into Eq. (6.127),<br />

this may be written<br />

λ dθ (<br />

dx = mc p θ − 1 +<br />

l∑<br />

j=1<br />

h 0 )<br />

j<br />

(ε j − ε jf ) . (6.130)<br />

c p T f<br />

It is also convenient here to eliminate variable x, by dividing the reaction equation<br />

(6.109.a) and diffusion Eq. (6.119) by (6.130), as it was done for the case of two<br />

chemical species. Thus the system of the flame equations reduces to the following<br />

Reaction equations<br />

dε i<br />

dθ = λ λ f<br />

r∑<br />

j=1<br />

(ν ′′<br />

ij − ν ′ ij)Λ ij θ δ j−n j<br />

e −θ aj/θ<br />

θ − 1 +<br />

l∑<br />

q j (ε j − ε jf )<br />

j=1<br />

l ∏<br />

s=1<br />

X s ν ′ sj<br />

, (i = 1, 2, . . . , l), (6.131)<br />

where, by analogy with the case of two chemical species, we have taken<br />

Λ ij = M iλ f B j T δ j−n j<br />

f<br />

m 2 , (6.132)<br />

c p<br />

θ aj =<br />

E j<br />

RT f<br />

, (6.133)


168 CHAPTER 6. LAMINAR FLAMES<br />

and<br />

q j =<br />

h j<br />

c p T f<br />

. (6.134)<br />

In comparison with the case of only two chemical species, we observe here the appearing<br />

of a set of parameters Λ ij . However, all these parameters have only one unknown<br />

quantity, the value of m, this is to say the flame velocity, therefore the set Λ ij may be<br />

expressed as a function of just one of the parameters.<br />

Diffusion equations<br />

l∑<br />

( )<br />

Mi<br />

L ij X i ε j − X j ε i<br />

dX<br />

M<br />

i<br />

dθ = j=1<br />

j<br />

, (i = 1, 2, . . . , l), (6.135)<br />

l∑<br />

θ − 1 + q j (ε j − ε jf )<br />

j=1<br />

where<br />

L ij =<br />

λRT<br />

M i c p pD ij<br />

, (6.136)<br />

Boundary conditions<br />

The boundary conditions reduce to the following<br />

θ → θ 0 : ε i → ε i0 , X i → X i0 , (6.137)<br />

θ → 1 : ε i → ε if , X i → X if . (6.138)<br />

Like in the case of two chemical species the difficulty at the cold boundary may be<br />

eliminated by introducing an ignition temperature θ i whose value does not influence<br />

the result when the reduced activation energies θ aj , or at least some of them, are large<br />

as normally happens in practical cases for the same reasons stated in <strong>de</strong>tail in the two<br />

species case.<br />

A recount of the number of boundary conditions when compared to the number<br />

of equations proves that the conditions are superabundant and hence, an eigenvalue of<br />

m, this is, one of the values of Λ ij , must exist which makes compatible the said<br />

conditions and therefore the problem is essentially the same than in the case of two<br />

species only.


6.13. OZONE DECOMPOSITION FLAME 169<br />

Be Λ ij the parameter chosen to express the rest of them as functions of it. Once<br />

its value is known, the value for velocity u 0 of the flame, by virtue of Eqs. (6.22) and<br />

(6.132), is<br />

√<br />

u 0 =<br />

M iλ f B j T δj−nj<br />

f<br />

ρ 2 0 c p<br />

Λ −1/2<br />

ij . (6.139)<br />

6.13 Ozone <strong>de</strong>composition flame<br />

The ozone <strong>de</strong>composition flame in a mixture of this gas with oxygen is the first example<br />

we have chosen among the few cases of flame propagation which have been<br />

calculated. Its first theoretical study was due to Lewis and von Elbe [8] who thus became<br />

the first to analyze flame propagation consi<strong>de</strong>ring, in addition to the effects of<br />

heat transfer, those of the diffusion of reactants and products, as well as the influence<br />

of the radicals (oxygen atoms) in the process. For this purpose they applied the law of<br />

chemical reaction which inclu<strong>de</strong>s the influence of temperature and the concentrations<br />

of the various species on the reaction rate. The three chemical species in the process<br />

are ozone, molecular oxygen and atomic oxygen. Lewis and von Elbe simplified the<br />

problem by adopting the following assumptions:<br />

1) The Lewis-Semenov number for the mixture O 2 - O 3 is equal to unity. This<br />

assumption enables the expression of the concentrations of ozone and molecular<br />

oxygen as functions of temperature.<br />

2) The chemical processes take place in accordance with the following kinetic<br />

scheme<br />

O 3 ⇆ O 2 + O, (6.140)<br />

O + O 3 → 2 O 2 . (6.141)<br />

The first of these two reaction equations expresses that the concentration of<br />

atoms of oxygen at each point of the flame is the one that would correspond<br />

to a mixture of O, O 2 and O 3 in chemical equilibrium at the temperature and<br />

composition of the point.<br />

Having stated the problem, and if, furthermore, we take into account that the<br />

molar fraction of atomic oxygen is much smaller than unity, its solution is straightforward.<br />

In fact, the first of the above assumptions allows the expression of the molar<br />

fractions of O 2 and O 3 as functions of temperature, as shown in §9 of this chapter,<br />

and equation (6.140) allows the same for the molar fraction of O.


170 CHAPTER 6. LAMINAR FLAMES<br />

Thus, the problem reduces to the integration of only one reaction equation<br />

(that corresponding to O 2 or to O 3 ), which may be performed by means of one of the<br />

methods enumerated in the aforesaid.<br />

Lewis and von Elbe carried out a more arduous calculation, through the numerical<br />

integration of the resulting equation. A <strong>de</strong>tailed information on their calculations<br />

may be found in the above reference or in [9]. Even when the values of the flame velocity<br />

obtained from this calculation are several times larger than those experimentally<br />

observed, consi<strong>de</strong>ring the state of knowledge at the time, their work was published,<br />

the fact that predicted and experimental values were of the same or<strong>de</strong>r of magnitu<strong>de</strong><br />

was consi<strong>de</strong>red an important success.<br />

Recently, von Kármán and Penner [10] have recomputed the flame velocity for<br />

a set of mixtures of O 2 and O 3 , using the kinetic scheme proposed by Lewis and von<br />

Elbe, but applying the semi-analytical methods <strong>de</strong>veloped in §9. The results agree<br />

with those obtained by Lewis and von Elbe and they shown that the influence of the<br />

composition of the mixture on velocity of the flame obtained through this procedure,<br />

differ essentially from those observed by experimentation. This discrepancy is basically<br />

due to the fact that the set of chemical reactions proposed by Lewis and von Elbe<br />

and, in particular, the distribution of oxygen atoms within the flame, are different from<br />

the actual ones.<br />

Hirschfel<strong>de</strong>r and his associates [25] have recently calculated the same flame,<br />

applying the following complete kinetic scheme<br />

O 3 + G → O + O 2 + G 1)<br />

O + O 2 + G → O 3 + G 2)<br />

O + O 3 → 2 O 2 3)<br />

2 O 2 + G → O + O 3 4)<br />

O 2 + G → 2 O + G 5)<br />

(6.142)<br />

2 O + G → O 2 + G 6)<br />

This computation was carried out by means of an arduous numerical integration<br />

for which electronic computers were used.<br />

Later on, von Kármán and Penner [6] performed a simplified analysis of the<br />

problem, utilizing the same kinetic scheme and physico-chemical constants as Hirschfel<strong>de</strong>r<br />

but applying Kármán’s analytical method of integration, and postulating a distribution<br />

of oxygen atoms within the flame <strong>de</strong>termined through the extension of the<br />

steady state assumption which is wi<strong>de</strong>ly used in classical Chemical Kinetics.


6.13. OZONE DECOMPOSITION FLAME 171<br />

The following is a summary of the von Kármán-Penner study which can be<br />

found fully <strong>de</strong>scribed in the papers of Refs. [6] and [10].<br />

If we assign subscripts 1, 2 and 3 to species O, O 2 and O 3 respectively, and<br />

making use of the flame equations <strong>de</strong>rived in §11 of this chapter, we obtain the following<br />

system.<br />

Reaction equation<br />

It is sufficient to write two equations corresponding, for example, to species O 2 and<br />

O 3 thus obtaining after Eq. (6.131)<br />

dε 1<br />

dθ = λ λ f<br />

[<br />

Λ 11 θ −3/2 e −θ a1/θ X 3 − Λ 12 θ −5/2 e −θ a2/θ X 1 X 2<br />

− Λ 13 θ −3/2 e −θa3/θ X 1 X 3 + Λ 14 θ −3/2 e −θa4/θ X2<br />

2 ]<br />

(6.143)<br />

+ 2Λ 15 θ −3/2 e −θa5/θ X 2 − 2Λ 16 θ −5/2 e −θa6/θ X1<br />

2<br />

[<br />

] −1,<br />

· θ − 1 + q 1 (ε 1 − ε 1f ) + q 2 (ε 2 − ε 2f ) + q 3 (ε 3 − ε 3f )<br />

dε 3<br />

dθ = λ [<br />

−Λ 31 θ −3/2 e −θa1/θ X 3 + Λ 32 θ −5/2 e −θa2/θ X 1 X 2<br />

λ f<br />

]<br />

− Λ 33 θ −3/3 e −θa3/θ X 1 X 3 + Λ 34 θ −3/2 e −θa4/θ X2<br />

2 (6.144)<br />

[<br />

] −1.<br />

· θ − 1 + q 1 (ε 1 − ε 1f ) + q 2 (ε 2 − ε 2f ) + q 3 (ε 3 − ε 3f )<br />

Diffusion equations<br />

Likewise, the diffusion equations corresponding to O and O 3 are, according to (6.135)<br />

dX 1<br />

dθ = L ( 1<br />

12 2 X ) (<br />

1ε 1 − X 2 ε 1 + 1 L13<br />

3 X )<br />

1ε 1 − X 3 ε 1<br />

θ − 1 + q 1 (ε 1 − ε 1f ) + q 2 (ε 2 − ε 2f ) + q 3 (ε 3 − ε 3f ) , (6.145)<br />

dX 3<br />

dθ = L 31 (3X 1 ε 3 − X 3 ε 1 ) + L 32<br />

( 3<br />

2 X 3ε 2 − X 2 ε 3<br />

)<br />

θ − 1 + q 1 (ε 1 − ε 1f ) + q 2 (ε 2 − ε 2f ) + q 3 (ε 3 − ε 3f ) . (6.146)<br />

The following Table 6.1shows the values of the physico-chemical constants<br />

corresponding to the six reactions of Eq. (6.142).


172 CHAPTER 6. LAMINAR FLAMES<br />

Reaction<br />

Or<strong>de</strong>r<br />

n j<br />

δ j<br />

E j<br />

cal/mol<br />

A j<br />

1 2 1/2 24.14 10.56 × 10 12<br />

2 3 1/2 0.00 0.230 × 10 12<br />

3 2 1/2 6.00 7.15 × 10 12<br />

4 2 1/2 99.21 2.93 × 10 12<br />

5 2 1/2 117.34 8.92 × 10 12<br />

6 3 1/2 0.00 0.482 × 10 12<br />

Table 6.1: Physico-chemical parameters for reactions 6.142.<br />

The boundary conditions for this system are as follows:<br />

a) Hot boundary.<br />

At the hot boundary (θ = 1) the concentration of ozone is zero and the concentration<br />

of atoms of oxygen is very small. Therefore it can be written<br />

θ = 1 : ε 1 ≡ ε 1f ≃ 0, ε 3 ≡ ε 3f = 0, (6.147)<br />

X 1 ≡ X 1f ≃ 0, X 3 ≡ X 3f = 0. (6.148)<br />

b) Cold boundary.<br />

At the cold boundary since no chemical reactions has been yet produced, the<br />

mass flux of ozone is the one corresponding to the initial mixture and the one of<br />

atomic oxygen is zero<br />

θ = θ i : ε 1 ≡ ε 10 = 0, ε 3 = Y 30 (datum). (6.149)<br />

Simplification of the equations<br />

Before proceeding to integrate the system, we will simplify it, following von Kármán<br />

and Penner. For the purpose, we shall begin by analyzing the values of the physicochemical<br />

constants of the reactions which appear summarized in the above Table 6.1,<br />

starting by the activation energies.<br />

We see in this table that for reactions 4 and 5 in (6.142), this is, those producing<br />

atomic oxygen from molecular oxygen, the activation energies are much larger than<br />

for the rest, and consequently the reaction velocities in which they participate will be<br />

very small and may be neglected. Hence, we can eliminate the terms corresponding<br />

to these reactions from equations (6.143) and (6.144).<br />

Let us now consi<strong>de</strong>r, coefficients Λ ij which are the ones that may also influence<br />

the or<strong>de</strong>r of magnitu<strong>de</strong> of the reaction velocities. In or<strong>de</strong>r to simplify this comparison,


6.13. OZONE DECOMPOSITION FLAME 173<br />

in the following table we give their values, referred to the value of Λ 31 which we are<br />

adopting as a reference according to the reasons stated in the foregoing.<br />

Λ 11<br />

Λ 31<br />

= 0.333<br />

Λ 14<br />

Λ 31<br />

= 0.925<br />

Λ 32<br />

= 3.486 × 10 −3 T −1<br />

f<br />

Λ 31<br />

Λ 12<br />

= 1.162 × 10 −3 T −1<br />

f<br />

Λ 31<br />

Λ 15<br />

Λ 31<br />

= 0.282<br />

Λ 33<br />

Λ 31<br />

= 0.677<br />

Λ 13<br />

Λ 31<br />

= 0.226<br />

Λ 16<br />

= 2.435 × 10 −3 T −1<br />

f<br />

Λ 31<br />

Λ 34<br />

Λ 31<br />

= 0.277<br />

Table 6.2: Values of the ratios Λ ij/Λ 31 appearing in Eqs. (6.145) and (6.146).<br />

It appears here that Λ 12 , Λ 16 and Λ 32 , which correspond to reactions 2 and 6<br />

in (6.142), are much smaller than the others, due to the following two reasons. First,<br />

because the corresponding coefficients A j are smaller, as shown by Table (6.1), and<br />

second because they are reactions requiring triple collision, as shown by Eq. (6.142),<br />

which explains the presence of term T −1<br />

f<br />

in these coefficients.<br />

Hence, it results that we can also eliminate the corresponding terms in reaction<br />

equations (6.143) and (6.144).<br />

Finally, with reference to the <strong>de</strong>nominator of these equations, it can also be<br />

simplified, keeping in mind that:<br />

1) In accordance with boundary conditions (6.147) it is ε 1f = ε 3f = 0.<br />

2) Moreover, ε 1 is very small compared to unity throughout the flame, and therefore,<br />

the corresponding term may be neglected.<br />

3) Finally, in accordance with Eq. (6.124) and consi<strong>de</strong>ring that h 0 2 is zero, it is<br />

q 2 = 0.<br />

following<br />

Taking into account these consi<strong>de</strong>rations, the said <strong>de</strong>nominator reduces to the<br />

θ − 1 + q 1 (ε 1 − ε 1f ) + q 2 (ε 2 − ε 2f ) + q 3 (ε 3 − ε 3f )1 ≃ θ − 1 + q 3 ε 3 . (6.150)<br />

Consequently if we introduce the above simplifications into the system of reaction<br />

equations (6.143) and (6.144), this system reduces to<br />

dε 1<br />

dθ = Λ λ θ ( )<br />

−3/2 e −θa1/θ X 3 − e −θa3/θ X 1 X 3<br />

, (6.151)<br />

λ f θ − 1 + q 3 ε 3<br />

dε 3<br />

dθ = −Λ λ λ f<br />

θ −3/2 ( e −θ a1/θ X 2 + e −θ a3/θ X 1 X 3<br />

)<br />

θ − 1 + q 3 ε 3<br />

. (6.152)


174 CHAPTER 6. LAMINAR FLAMES<br />

Steady state assumption<br />

In spite of the simplification introduced into the system, it is still very difficult to<br />

perform its integration by semi-analytical methods.<br />

In an effort to find further simplifications, von Kármán and Penner checked the<br />

applicability of the steady state assumption for the distribution of atomic oxygen. This<br />

assumption is expressed through nullifying the reaction velocity of the oxygen atoms,<br />

thus obtaining<br />

e −θ a1/θ X 3 − e −θ a3/θ X 1 X 3 = 0, (6.153)<br />

from which we <strong>de</strong>rive the following distribution of atomic oxygen throughout the<br />

flame<br />

X 1 = e − θ a1 − θ a3<br />

θ . (6.154)<br />

10<br />

9<br />

8<br />

7<br />

Hirschfel<strong>de</strong>r<br />

Karman−Penner<br />

6<br />

10 4 X 1<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

θ<br />

Figure 6.16: Mole fraction of atomic oxygen as a function of the temperature for the ozone<br />

<strong>de</strong>composition flame.<br />

Figure 6.16 taken from [6] compares the distribution of X 1 corresponding to<br />

this assumption, with that obtained by Hirschfel<strong>de</strong>r through numerical integration of<br />

the flame equations, without making use of the assumption. It is seen that the result is<br />

completely satisfactory and it plainly justifies the assumption of Kármán and Penner,<br />

which is not applicable just within a narrow zone very close to the maximum temperature,<br />

at which, since the concentration of ozone is practically reduced to zero there<br />

are other reactions, different from 1 and 3 in (6.142) controlling the formation of O.


6.13. OZONE DECOMPOSITION FLAME 175<br />

But this zone is unimportant when studying the flame structure and in particular its<br />

propagation velocity.<br />

The applicability of the steady state assumption is an important contribution<br />

to the theory of flames which essentially simplifies. The success reached with ozone<br />

flames has encouraged the i<strong>de</strong>a of applying this assumption to the study of similar<br />

practical cases, as will be seen later on.<br />

Recently, however, the applicability of such an assumption has been discussed,<br />

among others, by Campbell [26], Giddings and Hirschfel<strong>de</strong>r [27] and Spalding [28].<br />

They based their objections first on the fact that the crossing of the flame is so rapid<br />

that there is not sufficient time for the formation of radicals to reach the level required<br />

by the condition of equilibrium established by the said assumption. On the other hand,<br />

even if such level were reached the diffusion of radicals would change substantially<br />

the distribution. Although other studies by Gilbert and Altman [29] and by Millán<br />

and Sanz [30] tend to confirm the applicability of the steady state assumption for the<br />

cases studied in the foregoing paragraphs, however, the problem cannot be consi<strong>de</strong>red<br />

as solved, until a general study enables the establishment of the conditions that must<br />

be satisfied in a flame so that this assumption be valid.<br />

Let us proceed with the study of the ozone flame. The application of the steady<br />

state assumption eliminates Eq. (6.151) and reduces (6.152) to the following<br />

where, to simplify notation, we have written<br />

dε 3<br />

dθ = −2Λ λ λ f<br />

θ −3/2 e −θa1/θ X 3<br />

θ − 1 + q 3 ε 3<br />

, (6.155)<br />

Λ 13 = Λ = M 1λ f B 3<br />

. (6.156)<br />

m 2 c p T 3/2<br />

f<br />

Taking into account the preceding consi<strong>de</strong>rations, the diffusion equation reduces to<br />

where also, for shortness, we write<br />

dX 3<br />

dθ = L 3<br />

2 X 3 − ε 3 − 1 2 X 3ε 3<br />

θ − 1 + q 3 ε 3<br />

, (6.157)<br />

L 13 = L =<br />

λRT<br />

M 1 c p pD 13<br />

. (6.158)<br />

The problem reduces now to the integration of these two equations with the following<br />

boundary conditions<br />

θ = θ i : ε 3 = 0, (6.159)<br />

θ = 1 : ε 3 = X 3 = 0. (6.160)


176 CHAPTER 6. LAMINAR FLAMES<br />

The compatibility of these three conditions <strong>de</strong>termines, the “eigenvalue” of Λ, as established<br />

in §9, and once it is <strong>de</strong>termined,we obtain with it the value for the flame<br />

velocity by virtue of Eq. (6.139).<br />

The integration of the system is very simple using Kármán’s method, in the<br />

following way. In the zone of interest, this is for θ close to 1, it is ε 3 ≫ X 3 and<br />

ε 3 ≫ 1 − θ, therefore a first approximation of Eq. (6.157) is the following<br />

where consi<strong>de</strong>ring that for θ = 1 it is X 3 = 0, we obtain<br />

dX 3<br />

dθ ≃ − L q 3<br />

, (6.161)<br />

X 3 ≃ − L q 3<br />

(1 − θ). (6.162)<br />

If this value is taken into (6.155) and integrating between θ = 1 and θ = θ i ≃ 0 as<br />

expressed in §9 of this chapter, we obtain the <strong>de</strong>sired value of Λ.<br />

Von Kármán and Penner have calculated the result for X 30 = 0.25, 0.40, 0.50,<br />

0.75, and 1.00. Furthermore, these authors calculated as well the flame velocity for<br />

the same values by means of the Lewis-von Elbe chemical scheme.<br />

A comparison of results with those obtained through experimentation by Grosse<br />

[31] proves that the approximation of the theoretical values reached by Kármán and<br />

Penner is satisfactory since the <strong>de</strong>viations, in both cases, are un<strong>de</strong>r 25 per cent.<br />

6.14 Hydrazine <strong>de</strong>composition flame<br />

Hydrazine combustion has been experimentally studied by Murray and Hall [32].<br />

Tests were performed by burning mixtures of hydrazine and water vapours in a Bunsen<br />

burner at ambient pressure. The flame propagation velocity was obtained by measuring<br />

the area of the luminous cone of the flame.<br />

The analysis of the combustion products shown that they correspond to the<br />

overall reaction<br />

N 2 H 4 ⇄ 1 2 H 2 + 1 2 N 2 + NH 3 . (6.163)<br />

Murray and Hall also calculated the theoretical propagation velocity of the<br />

flame for one of the mixtures experimentally analyzed (97.2% hydrazine). They assumed<br />

that such velocity was <strong>de</strong>termined by the unimolecular <strong>de</strong>composition reaction<br />

N 2 H 4 → 2NH 2 , (6.164)


6.14. HYDRAZINE DECOMPOSITION FLAME 177<br />

which is the initial reaction in the complex process that is stoichiometrically represented<br />

by (6.163).<br />

Making use of the thermal theory of Zeldovich and Frank-Kamenetskii [33]<br />

and using the specific reaction rate<br />

k = 4 × 10 12 e −60 000/RT , (6.165)<br />

proposed by Szwarc [34], they obtained a theoretical propagation velocity of the flame<br />

of 110 cm/s compared to the 185 cm/s obtained from experiments.<br />

More precise calculations were performed by Hirschfel<strong>de</strong>r [35] and by Kármán<br />

and Penner [6], making use of Murray and Hall’s kinetic scheme with results that<br />

confirm those obtained by Murray and Hall. Hirschfel<strong>de</strong>r and Kármán and Penner<br />

also studied the influence over the propagation velocity of the flame of the diffusion<br />

of the different species. In this case they obtain values consi<strong>de</strong>rably smaller than<br />

the experimental ones since diffusion reduces substantially the value of the flame’s<br />

propagation velocity.<br />

The following table summarizes the results of the works by Murray-Hall and<br />

Kármán-Penner.<br />

Experimental value<br />

Theoretical values<br />

Refs. Murray-Hall Murray-Hall Kármán-Penner<br />

Zero diffusion Zero diffusion Non-zero diffusion 10<br />

u 0 185 cm/s 110 cm/s 95 cm/s 42 cm/s<br />

Table 6.3: Theoretical and experimental propagation velocities of the flame in a mixture of<br />

vapour of hydrazine and water. Mass fraction of hydrazine = 0.972. Pressure =<br />

1 atm. Initial temperature of the mixture = 150 ◦ C.<br />

If the reaction controlling the process is the one given by Eq. (6.164) with the<br />

specific reaction rate (6.165), the following two conditions must be satisfied:<br />

1) Since chemical reaction (6.164) is of the first or<strong>de</strong>r, the propagation velocity of<br />

the flame must be inversely proportional to the square root of pressure.<br />

2) Since the activation energy E of reaction (6.164) is 60 000 cal/mol and the flame<br />

propagation velocity is approximately proportional to exp (−E/2RT f ) according<br />

to Eq. (6.72) where T f is the flame temperature, when such velocity is represented<br />

as a function of 1/T f in the logarithmic scale one must evi<strong>de</strong>ntly obtain<br />

a slope equal to −E/2R.<br />

10 For a Lewis number equal to unity.


178 CHAPTER 6. LAMINAR FLAMES<br />

With the experiments ma<strong>de</strong> by Murray and Hall it is not possible to verify the<br />

first conclusion since they were all performed at ambient pressure, and the second<br />

conclusion was not checked by the authors. In or<strong>de</strong>r to verify both, Adams and Stock<br />

[36] measured the combustion velocity of hydrazine at different pressures. For this<br />

purpose, they stabilized the flame over a column of liquid hydrazine contained in a<br />

capillary tube and measured the recession velocity of the liquid level as its vapour<br />

burnt. This method is not practical to measure the absolute propagation velocity of<br />

the flame due to the irregularity of its front but is very convenient for the comparative<br />

study inten<strong>de</strong>d by these authors. The results of their experiments show that the influence<br />

of pressure is approximately that corresponding to a first or<strong>de</strong>r reaction. The<br />

discrepancies could be due to the experimental method used.<br />

As for the apparent activation energy it <strong>de</strong>creases with the drop in flame temperature<br />

and it is in all cases way un<strong>de</strong>r the 60 000 cal/mol that correspond to the<br />

process proposed by Szwarc and applied to the theoretical studies mentioned herein.<br />

Such result led Adams and Stocks to conclu<strong>de</strong> that the <strong>de</strong>composition of hydrazine<br />

should occur through a complicated system of chain reactions of which (6.164)<br />

is only the initial one. As a result of a <strong>de</strong>tailed discussion of all possible reactions the<br />

above mentioned authors proposed a simplified kinetic scheme [36], which summarizes<br />

the actual process. This proposed scheme consi<strong>de</strong>rs only three chemical species:<br />

reacting species A (hydrazine), stable products C of the combustion (mixture of NH 3 ,<br />

N 2 , H 2 , etc.), and one radical B, that propagates the chain (which could be NH 2 , H,<br />

etc). The Adams and Stock simplified scheme is a follows<br />

A → 2 B, (6.166)<br />

B + A → B + 2 C, (6.167)<br />

B + B + X → 2 C + X. (6.168)<br />

Of the three reactions, the first initiates the chain, the second is the propagation reaction<br />

and the third the chain breaking reaction.<br />

Lately, Spalding [37] has calculated the propagation velocity of the flame with<br />

the scheme proposed by Adams and Stocks and he applied an interesting method of<br />

numerical integration <strong>de</strong>veloped by him. Spalding calculates two cases corresponding,<br />

respectively, to a cold flame and to one of the hot flames experimented by Murray<br />

and Hall. In both cases he gets results in close agreement with those experimentally<br />

obtained by Murray and Hall. These results are summarized in Table 6.4.


6.14. HYDRAZINE DECOMPOSITION FLAME 179<br />

Theoretical<br />

(Spalding)<br />

Experimental<br />

Hot flame 190 cm/s 185 cm/s (Murray-Hall)<br />

Cold flame 12 cm/s 10 cm/s (Adams-Stocks)<br />

Table 6.4: Propagation velocity of the flame in cm/s.<br />

As Spalding points out the excellent numerical agreement is casual to a certain<br />

extent since there are doubts respect to the correct values of the transport coefficients.<br />

Moreover, Spalding did not inclu<strong>de</strong> the influence of the variation of such coefficients<br />

with the temperature across the flame. More interesting is the fact that Spalding’s<br />

results predict correctly the influence of the combustion temperature on the velocity<br />

of the flame.<br />

Spalding also conclu<strong>de</strong>s that the propagation velocity of the flame that would<br />

be obtained when calculating the radical distributions by means of the steady state<br />

assumption would be much too large. He bases this conclusion upon the fact that<br />

the distribution of radicals obtained with the steady state assumption is consi<strong>de</strong>rably<br />

larger, at temperatures close to combustion temperature, than the distribution given by<br />

correct calculation.<br />

Millán and Sanz [30] calculated the propagation velocity of hydrazine <strong>de</strong>composition<br />

flame applying the same simplified mo<strong>de</strong>l proposed by Adams and Stocks.<br />

Two cases were consi<strong>de</strong>red, the first assuming steady state for concentration of radicals<br />

and solving the flame equations through Kármán’s method; the second, for which<br />

an approximate calculation method was <strong>de</strong>veloped to obtain flame velocity without<br />

consi<strong>de</strong>ring the steady state assumption. The authors conclu<strong>de</strong>d that the results obtained<br />

applying the steady state assumption to the radical concentrations are satisfactory<br />

when compared to those corresponding to a more correct distribution for the<br />

same.<br />

Simultaneously, Gilbert and Altman [38] obtained the flame propagation velocity<br />

corresponding to complete kinetic mo<strong>de</strong>l proposed by Adams and Stocks [36]<br />

through the application of the Boys-Corner method in finding the solution of the flame<br />

equations and calculating the concentration of radicals throughout the flame un<strong>de</strong>r the<br />

steady state assumption introduced by Kármán and Penner [6].<br />

Recently [39] a new analysis of the problem was conducted using the complete<br />

kinetic mo<strong>de</strong>l of Adams and Stocks, and adopting the steady state assumption for the<br />

<strong>de</strong>termination of the concentration of radicals, with Kármán’s method in calculating<br />

the solution.


180 CHAPTER 6. LAMINAR FLAMES<br />

Kinetic mo<strong>de</strong>l of the reaction of hydrazine <strong>de</strong>composition<br />

The mechanism of <strong>de</strong>composition proposed by Adams and Stocks is as follows<br />

N 2 H 4 → 2NH 2 1)<br />

N 2 H 4 + NH 2 → NH 3 + N 2 H 3 2)<br />

N 2 H 3 → N 2 + H 2 + H 3)<br />

H + N 2 H 4 → 2NH 3 + NH 2 4)<br />

H + NH 2 + X → NH 3 + G 5 c)<br />

(6.169)<br />

H + H + X → H 2 + G 5 d)<br />

In these equations, G represents any one of the particles of the mixture.<br />

In or<strong>de</strong>r to i<strong>de</strong>ntify the molecules of the components of the mixture the following<br />

subscripts will be used<br />

1 ↔ N 2 H 4 reactant<br />

2 ↔ NH 3 product<br />

3 ↔ N 2 product<br />

4 ↔ H 2 product<br />

(6.169.a)<br />

5 ↔ NH 2 radical<br />

6 ↔ N 2 H 3 radical<br />

7 ↔ H radical<br />

According to the law of the mass action, the velocity reaction w i of the radicals<br />

must be expressed as a function of the mole concentration in the following way.<br />

For NH 2 :<br />

For N 2 H 3 :<br />

For H :<br />

w 5<br />

M 5<br />

= 2k 1 cX 1 − k 2 c 2 X 1 X 5 + k 4 c 2 X 1 X 7 − k 5c c 3 X 5 X 7 , (6.170)<br />

w 6<br />

M 6<br />

= k 2 c 2 X 1 X 5 − k 3 cX 6 , (6.171)<br />

w 7<br />

M 7<br />

= k 3 cX 6 − k 4 c 2 X 1 X 7 − k 5c c 3 X 5 X 7 − 2k 5d c 3 X 2 7 . (6.172)<br />

Likewise, the hydrazine <strong>de</strong>composition velocity is<br />

− w 1<br />

M 1<br />

= k 1 cX 1 + k 2 c 2 X 1 X 5 + k 4 c 2 X 1 X 7 . (6.173)


6.14. HYDRAZINE DECOMPOSITION FLAME 181<br />

Here, X i are the mole fractions of the different species, k i the specific reaction velocities<br />

given by Eq. (6.112) (subscript of k indicates the corresponding reaction in accordance<br />

with the number assigned to them in the preceding paragraph) and c = ρ/M<br />

is the mole concentration per cm 3 .<br />

The following values, also applied by Gilbert and Altman [38], are adopted for<br />

specific velocities<br />

k 1 = 4 × 10 12 e −60 000/RT s −1 (6.174)<br />

k 2 = 10 13 e −4 600/RT cm 3 mol −1 s −1 (6.175)<br />

k 4 = 10 13 e −7 000/RT cm 3 mol −1 s −1 (6.176)<br />

k 5c = 5 × 10 15 cm 6 mol −2 s −1 (6.177)<br />

k 5d = 10 16 cm 6 mol −2 s −1 (6.178)<br />

Reaction velocity of hydrazine un<strong>de</strong>r to steady state assumption for<br />

radicals<br />

The steady state assumption is expressed by making to reaction velocities of the radicals<br />

referred by 5, 6 and 7 in (6.187.a) equal to zero. Thereby<br />

k 2 cX 1 X 5 = k 3 X 6 , (6.179)<br />

cX 1 (k 2 X 5 − k 4 X 7 ) = 2k 1 X 1 − k 5c c 2 X 5 X 7 , (6.180)<br />

cX 1 (k 2 X 5 − k 4 X 7 ) = k 5c c 2 X 5 X 7 + 2k 5d c 2 X7 2 . (6.181)<br />

Since mole fraction X 7 of radical H is small throughout the reaction with<br />

respect to reactant X 1 , it occurs that k 5c cX 7<br />

is negligible when compared to 2k 1<br />

.<br />

k 2 X 1 k 2 cX 5<br />

Hence Eqs. (6.180) and (6.181) may be written<br />

k 2 X 5 − k 4 X 7 = 2k 1<br />

c , (6.182)<br />

k 2 X 5 − k 4 X 7 = 2k 5dcX 2 7<br />

X 1<br />

. (6.183)<br />

Consequently, the concentration of radicals H is supplied by<br />

√<br />

k1 √<br />

X 7 = X1<br />

k 5d c 2 . (6.184)


182 CHAPTER 6. LAMINAR FLAMES<br />

From (6.182) and (6.179) the following expressions may be <strong>de</strong>duced for mole<br />

fractions X 5 of NH 2 and X 6 of N 2 H 3 :<br />

X 5 = 2k 1 + k 4 cX 7<br />

, (6.185)<br />

k 2 c<br />

X 6 = 2k 1 + k 4 cX 7<br />

k 3<br />

X 1 . (6.186)<br />

If the values for X 7 and X 5 given by (6.184) and (6.185) are substituted into<br />

the hydrazine reaction velocity given by (6.173), the following is obtained<br />

− w 1<br />

M 1<br />

= cX 1 (3k 1 + kX 1/2<br />

1 ), (6.187)<br />

where<br />

( ) 4k<br />

2 1/2<br />

k = 4 k 1<br />

= 4 × 10 11 e −37 000/RT . (6.188)<br />

k 5d<br />

It is easily seen that the first term in the parenthesis of Eq. (6.187) may be neglected<br />

respect to the second. Hence it finally results<br />

− w 1<br />

M 1<br />

= kcX 3/2<br />

1 . (6.189)<br />

Let us assume that the proportion between concentrations of products NH 3 , N 2 and<br />

H 2 is the one corresponding to the complete reaction, and the concentrations of radicals<br />

are very small, then the following relations will be obtained between the mole<br />

fractions and the mass fractions of the main species<br />

Furthermore, the mean mole mass of the mixture will be<br />

X 1 + 2X 2 = 1, (6.190)<br />

Y 1<br />

= X 1<br />

M 1 M , (6.191)<br />

Y 2<br />

= X 2<br />

M 2 M . (6.192)<br />

M = M 1 X 1 + X 2<br />

(M 2 + 1 2 M 3 + 1 2 M 4<br />

From the system of equations (6.190) to (6.193) it results<br />

)<br />

. (6.193)<br />

1 − 32<br />

X 1 = 17 Y 2<br />

1 + 32 . (6.194)<br />

17 Y 2<br />

The reaction velocity of ammonia results to be<br />

⎛<br />

w 2<br />

= − w 1 − 32 ⎞3/2<br />

1 ⎜<br />

= kc ⎝<br />

17 Y 2<br />

⎟<br />

M 2 M 1<br />

1 + 32 ⎠ . (6.195)<br />

17 Y 2


6.14. HYDRAZINE DECOMPOSITION FLAME 183<br />

Energy equation<br />

When mean specific heat is assumed to be constant and the influence of radicals on<br />

the energy of the mixture is neglected, the energy equation may be written<br />

( )<br />

λ dT 17<br />

m dx = q r<br />

32 − ε 2 − c p (T f − T ), (6.196)<br />

where q r is the reaction heat per gram of ammonia produced. If reduced temperatures<br />

θ = T T f<br />

and θ 0 = T 0<br />

T f<br />

(6.197)<br />

are introduced into Eq. (6.196) and taking into consi<strong>de</strong>ration the conditions at the cold<br />

boundary, θ = θ 0 , ε 2 = 0, this equation may be written<br />

where<br />

λ<br />

mc p<br />

Chemical reaction and diffusion equations<br />

The reaction equation for ammonia is<br />

dθ<br />

dx = (1 − θ 0)(1 − ε) − (1 − θ), (6.198)<br />

ε = 32<br />

17 ε 2. (6.199)<br />

m dε 2<br />

dx = w 2. (6.200)<br />

The diffusion equation corresponding to the same is given by Fick’s law and written<br />

dY 2<br />

dx = m (Y 2 − ε 2 ). (6.201)<br />

ρD 2m<br />

Coordinate x is eliminated from the above two equations through Eq. (6.198), and the<br />

following system results<br />

where<br />

and<br />

dε<br />

dθ = 32<br />

17<br />

is the Lewis-Semenov number.<br />

λ<br />

m 2 c p<br />

w 2<br />

(1 − θ 0 )(1 − ε) − (1 − θ) , (6.202)<br />

dY<br />

dθ = L Y − ε<br />

(1 − θ 0 )(1 − ε) − (1 − θ) , (6.203)<br />

Y = 32<br />

17 Y 2, (6.204)<br />

L =<br />

λ<br />

ρDc p<br />

(6.205)


184 CHAPTER 6. LAMINAR FLAMES<br />

where<br />

Taking (6.195) into (6.202)<br />

( ) 3/2 1 − Y<br />

e −θ 1 − θ<br />

a<br />

θ<br />

dε<br />

dθ = Λ 1 + Y<br />

(1 − θ 0 )(1 − ε) − (1 − θ) , (6.206)<br />

Λ = 128 × 10 11 e −θ a<br />

If a linear variation of λ with temperature is assumed<br />

we obtain from Eq. (6.207).<br />

Λ = 128 × 10 11 e −θ a<br />

( p<br />

) λ<br />

RT m 2 . (6.207)<br />

c p<br />

λ = λ f θ, (6.208)<br />

( ) p λf<br />

RT f m 2 . (6.209)<br />

c p<br />

Then the problem reduces to the integration of the system of Eqs. (6.203) and (6.206),<br />

with the following boundary conditions<br />

θ = θ 0 : ε = 0, (6.210)<br />

θ = 1 : ε = 1, Y = 1. (6.211)<br />

The value of Λ which makes compatible both boundary condition will be the<br />

eigenvalue for the system, and it <strong>de</strong>termines the flame propagation velocity.<br />

Flame velocity<br />

The flame velocity is given by<br />

u 0 = m ρ 0<br />

= 1 ρ 0<br />

(<br />

128 × 10 11 e −θ a<br />

p<br />

RT f<br />

where Λ is the eigenvalue referred to.<br />

When Eq. (6.206) is written<br />

1 − θ 0<br />

2<br />

−<br />

∫ 1<br />

0<br />

λ f<br />

c p<br />

) 1/2<br />

Λ −1/2 , (6.212)<br />

∫ 1<br />

( ) 3/2 1 − Y<br />

(1 − θ) dε = Λ<br />

e −θ 1 − θ<br />

a<br />

θ dθ, (6.213)<br />

θ 0<br />

1 + Y<br />

the problem reduces (following the i<strong>de</strong>a of Karman’s method) to finding an approximation<br />

of θ vs ε for the calculation of integral<br />

∫ 1<br />

Y vs θ for the computation of the integral of the right hand si<strong>de</strong>.<br />

0<br />

(1 − θ) dε, and an approximation of


6.14. HYDRAZINE DECOMPOSITION FLAME 185<br />

The parabolic approximation given by Eq. (6.87) is adopted for Y vs θ<br />

( ) L θ − θ0<br />

Y =<br />

, (6.214)<br />

1 − θ 0<br />

which satisfies Eq. (6.203) for ε = 0, and for ε = 1 is tangent to the straight line<br />

1 − Y = L 1 − θ<br />

1 − θ 0<br />

, (6.215)<br />

which, in turn, is also tangent to the solution of Eq. (6.216) as it can easily be verified<br />

when consi<strong>de</strong>ring that for θ ≃ 1 is<br />

and<br />

1 − ε ≫ 1 − Y, (6.216)<br />

1 − ε ≫ 1 − θ . (6.217)<br />

1 − θ 0<br />

When approximation (6.214) is taken into Eq. (6.213), it results<br />

where<br />

1 − θ 0<br />

2<br />

−<br />

∫ 1<br />

∫ ⎡ ( ) 1 θ − L ⎤3/2<br />

θ0<br />

1 −<br />

⎢ 1 − θ<br />

I K = ⎣<br />

0 ⎥<br />

( ) θ − L ⎦<br />

θ0<br />

θ 0 1 +<br />

1 − θ 0<br />

The approximation of θ vs ε for the computation of<br />

0<br />

(1 − θ) dε = ΛI K , (6.213.a)<br />

e −θ 1 − θ<br />

a<br />

θ dθ. (6.218)<br />

∫ 1<br />

θ<br />

(1 − θ) dε can be obtained<br />

from the solution of Eq. (6.206) for values of ε close to 1. Through the procedure<br />

applied in Ref. [30] , it results 11<br />

Therefore<br />

∫ 1<br />

0<br />

[ ( ) ] 3/2 −2/5<br />

1 − θ<br />

= (1 − ε) 4/5 4 L<br />

1 − θ 0 5 Λ . (6.219)<br />

2<br />

[<br />

(1 − θ) dε ≃ 5 ( ) ] 3/2 −2/5<br />

9 (1 − θ 4 L<br />

0)<br />

Λ −2/5 . (6.220)<br />

5 2<br />

If Eqs. (6.218) and (6.220) are taken into Eq. (6.213) the following equation will be<br />

obtained for the eigenvalue Λ<br />

⎛<br />

Λ = 1 − θ 0<br />

2I K<br />

⎝1 − 10<br />

9<br />

which is solved through iteration or else graphically.<br />

[ ( ) ] ⎞<br />

3/2<br />

−2/5<br />

4 L<br />

Λ −2/5 ⎠ , (6.221)<br />

5 2<br />

11 Essentially, it reduces to neglecting (1 − θ) respect to (1 − ε) in the <strong>de</strong>nominator of Eq. (6.206) and to<br />

apply the linear approximation given by Eq. (6.215) in or<strong>de</strong>r to eliminate Y in the numerator of the same.


186 CHAPTER 6. LAMINAR FLAMES<br />

In or<strong>de</strong>r to perform the numerical calculation of the flame velocity the following<br />

values given by Hirschfel<strong>de</strong>r were applied. These same values were also used by<br />

Gilbert and Altman in his work:<br />

ρ 0 = 9.17 × 10 −4 gr/cm 3 ,<br />

T 0 = 423 K,<br />

T f = 1933 K,<br />

λ = 0.00067<br />

cal<br />

cm s K ,<br />

c p = 0.6623 cal<br />

gr K ,<br />

L = 1.33.<br />

(6.222)<br />

With them, the reduced activation temperature results to be<br />

θ a =<br />

The numerical integration of Eq. (6.218) gives<br />

When the prece<strong>de</strong>nt values are taken into (6.221)<br />

which, when taken into (6.212), finally results<br />

37 000<br />

= 9 632. (6.223)<br />

1.933 × 1.9872<br />

I K<br />

1 − θ 0<br />

= 270.9 × 10 −5 . (6.224)<br />

Λ −1/2 = 8.2962 × 10 −2 , (6.225)<br />

u 0 = Λ−1/2<br />

3.97<br />

≃ 209 cm/s. (6.226)<br />

The agreement between this value and the 200 cm/s given by experimental results is<br />

better than it could be expected consi<strong>de</strong>ring the dubiousness of many of the numerical<br />

values used.<br />

Influence of pressure<br />

As shown in Eq. (6.212) the flame velocity varies with pressure in accordance with<br />

the following law<br />

u 0 ∼ p1/2<br />

ρ 0<br />

∼ p −1/2 , (6.227)<br />

in agreement with the predictions of theory for first-or<strong>de</strong>r reaction flames Eq. (6.75)<br />

and as verified by the experiments carried-out by Adams and Stocks [36].


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 187<br />

6.15 Flame propagation in Hydrogen-Bromine mixtures<br />

The present example contains a calculation of the velocity of the hydrogen-bromine<br />

flame performed with a variation of the method <strong>de</strong>veloped by von Kármán and Penner<br />

[40]. The method presented here differs from earlier calculations because the<br />

influence of dissociation of bromine on the flame velocity is evaluated properly. Von<br />

Kármán and Penner inclu<strong>de</strong>d the influence of the atoms of bromine on the reaction<br />

rate but when calculating the distribution of the mole fraction of Br 2 , H 2 and HBr,<br />

they neglected the mole fraction of Br atoms. However, for temperatures close to the<br />

maximum flame temperature the dissociation of Br 2 reduces substantially the value<br />

for the mole fraction of Br 2 , which has an important influence on the flame velocity.<br />

Hence it seemed <strong>de</strong>sirable to compute the burning velocity correctly, especially<br />

in hydrogen-rich flames where the influence of dissociation is most important. Von<br />

Kármán and Penner also neglect the influence of the energy of dissociation of Br 2 ,<br />

which should be taken into account since it may be of the same or<strong>de</strong>r of magnitu<strong>de</strong> as<br />

the energy transferred by convection through the flame.<br />

In the following, a method is <strong>de</strong>veloped for the correct calculation of the mole<br />

fraction of Br 2 and of the influence of its dissociation energy on the velocity of the<br />

flame, within the limitations of the steady-state assumption for the distribution of<br />

bromine and hydrogen atoms. This method will be applied to computation for the<br />

four hydrogen-rich flames previously consi<strong>de</strong>red by von Kármán and Penner. The results<br />

are compared with those obtained by these authors and are shown to differ by not<br />

more than 25% even for the hottest flame.<br />

A <strong>de</strong>tailed study of the structure of the flame is also presented. It will be shown<br />

that dissociation of Br 2 reduces the flame velocity because of a <strong>de</strong>crease in the number<br />

of molecules of bromine which exist near the hot flame boundary. On the other hand,<br />

it will be seen that the influence of the dissociation energy can be neglected if thermal<br />

convection is ignored because these two effects tend to cancel.<br />

The procedure <strong>de</strong>veloped here follows closely the work performed in Ref. [41]<br />

and could also be useful for computations on other types of flames.<br />

Flame Equations<br />

Since there are only five different chemical components, namely, HBr, Br 2 , H 2 , Br<br />

and H, the unknowns are:<br />

1) Five flux fractions, ε i .


188 CHAPTER 6. LAMINAR FLAMES<br />

2)


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 189<br />

Expressions for the reaction rates w i are obtained from chemical kinetics. Between<br />

the five chemical components, the following in<strong>de</strong>pen<strong>de</strong>nt chemical reactions<br />

exist [42]<br />

Br 2 + X −→ k1<br />

2Br + X, (6.233)<br />

k<br />

Br + H<br />

2<br />

2 −→ HBr + H, (6.234)<br />

k<br />

H + Br<br />

3<br />

2 −→ HBr + Br, (6.235)<br />

HBr + H k 4<br />

−→ H 2 + Br, (6.236)<br />

2Br + X k 5<br />

−→ Br 2 + X. (6.237)<br />

These equations give the following expressions for the reaction rates w 1 , w 4<br />

and w 5<br />

( ) 2 p<br />

w 1 = M 1 θ −2 (k 2 X 3 X 4 + k 3 X 2 X 5 − k 4 X 1 X 5 ), (6.238)<br />

RT f<br />

( ) p<br />

w 4 = M 4 θ −1 p<br />

(2k 1 X 2 − 2k 5 θ −1 X4 2 ) − M 4<br />

w 5 , (6.239)<br />

RT f RT f M 5<br />

( ) 2 p<br />

w 5 = M 5 θ −2 (k 2 X 3 X 4 − k 3 X 2 X 5 − k 4 X 1 X 5 ). (6.240)<br />

RT f<br />

Energy Equation<br />

The energy equation is<br />

λ dθ (<br />

dx = mc p (θ − 1) +<br />

5∑<br />

i=1<br />

)<br />

(ε i − ε if ) h0 i<br />

. (6.241)<br />

c p T f<br />

Diffusion Equations<br />

Between mole fractions X i the following relation exists<br />

X 1 + X 2 + X 3 + X 4 + X 5 = 1. (6.242)<br />

Consequently, four additional equations are nee<strong>de</strong>d in or<strong>de</strong>r to complete the <strong>de</strong>termination<br />

of all of the mole fractions. The remaining equations are those for the diffusion<br />

of four of the components, and they are obtained from Eq. (6.119)<br />

dX i<br />

5∑<br />

(<br />

)<br />

dx = mRT f<br />

p<br />

θ 1 ε j ε i<br />

X i − X j , (i = 1, 2, 3, 4). (6.243)<br />

D ij M j M i<br />

j=1


190 CHAPTER 6. LAMINAR FLAMES<br />

Equations (6.228) through (6.232), (6.241), (6.242) and the four expressions<br />

contained in Eq. (6.243) form a system of eleven equations for the computation of<br />

the eleven unknowns of the flame. In the following section we simplify the problem<br />

by introducing the steady-state approximation for the concentration of bromine and<br />

hydrogen atoms.<br />

The steady-state approximation<br />

The following is the expression of the steady-state assumption for bromine and hydrogen<br />

atoms, respectively<br />

w 4 = 0, (6.244)<br />

w 5 = 0. (6.245)<br />

If Eqs. (6.239) and (6.240) are combined with Eqs. (6.244) and (6.245), we obtain<br />

for the mole fractions of Br and H the following relations in terms of the principal<br />

components and of the temperature<br />

X 4 =<br />

√<br />

k1<br />

k 5<br />

√<br />

RT f<br />

√<br />

θX2 , (6.246)<br />

p<br />

and<br />

X 5 = k √ √<br />

2 k1 RT f<br />

k 3 k 5 p<br />

X 3<br />

√ θX2<br />

X 2 + k 4<br />

k 3<br />

X 1<br />

. (6.247)<br />

From Eqs. (6.238), (6.242) and (6.246) the following expression is <strong>de</strong>duced for w 1<br />

( ) 2 p<br />

w 1 = 2M 1 θ −2 k 3 X 2 X 5 . (6.248)<br />

RT f<br />

Using in Eq. (6.248) the value for X 5 given in Eq. (6.247), we obtain<br />

( ) 3/2<br />

√<br />

p<br />

k1<br />

w 1 = 2M 1 k 2 θ −3/2 X 3 X 3/2<br />

2<br />

RT f k 5<br />

X 2 + k . (6.249)<br />

4<br />

X 1<br />

k 3<br />

Equation (6.249) expresses the rate of formation of HBr as a function of the mole<br />

fractions of the principal components and of temperature. Introduction of Eq. (6.249)<br />

into Eq. (6.230) leads to the result<br />

m dε ( ) 3/2<br />

√<br />

1<br />

p<br />

dx = 2M k1<br />

1<br />

k 2<br />

RT f<br />

θ −3/2 X 3 X 3/2<br />

2<br />

k 5<br />

X 2 + k 4<br />

k 3<br />

X 1<br />

. (6.250)


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 191<br />

The steady-state assumption implies the hypothesis that the diffusion of radicals may<br />

be neglected. Therefore, the flux fractions of Br and H must be proportional to their<br />

corresponding mole fractions, viz.,<br />

ε 4 = M 4<br />

M m<br />

X 4 , (6.251)<br />

ε 5 = M 5<br />

M m<br />

X 5 . (6.252)<br />

The mole fraction of hydrogen, X 5 , is always negligibly small. Therefore, in<br />

view of Eq. (6.252), the corresponding flux fraction may also be neglected.<br />

The preceding assumption and consi<strong>de</strong>rations allow a consi<strong>de</strong>rable simplification<br />

of the flame equations. Since<br />

Eqs. (6.228) and (6.229) reduce to<br />

X 5 ≪ 1 and ε 5 ≪ 1, (6.253)<br />

ε 1 + ε 2 + ε 3 + ε 4 = 1, (6.254)<br />

X 1 + X 2 + X 3 + X 4 = 1. (6.255)<br />

The mole fraction of Br, X 4 , is expressed as a function of X 2 through. Eq. (6.246).<br />

The rate of formation of HBr is given by Eq. (6.250).<br />

Eq. (6.254), now reduces to<br />

where<br />

and<br />

The energy equation, see<br />

λ dθ (<br />

)<br />

dx = mc p θ − 1 + q 1 (ε 1f − ε 1 ) − q 4 (ε 4f − ε 4 ) , (6.256)<br />

q 1 =<br />

(<br />

M4<br />

h 0 2 + M )<br />

5<br />

1<br />

h 0 3 − h 0 1<br />

(6.257)<br />

M 1 M 1 c p T f<br />

q 4 = (h 0 4 − h 0 1<br />

2)<br />

(6.258)<br />

c p T f<br />

are, respectively, the reduced reaction heats per unit mass corresponding to reactions<br />

1<br />

2 H 2 + 1 2 Br 2 −→ HBr + M 1 q 1 c p T f , (6.259)<br />

and<br />

Br + Br −→ Br 2 + M 2 q 4 c p T f . (6.260)<br />

Finally, by virtue of Eqs. (6.251) and (6.252), only two of the four diffusion equations,<br />

see Eqs. (6.243), remain. Moreover, the terms involving X 5 and ε 5 may be ignored.


192 CHAPTER 6. LAMINAR FLAMES<br />

Selection is ma<strong>de</strong> of the two equations corresponding to X 1 and X 3 because that<br />

corresponding to X 2 presents some difficulties, as will be seen later on. Thus<br />

and<br />

dX 1<br />

dx<br />

dX 3<br />

dx<br />

= RT 4∑<br />

(<br />

)<br />

f<br />

p<br />

θ 1 ε j ε 1<br />

X 1 − X j , (6.261)<br />

D<br />

j=1 1j M j M 1<br />

= RT 4∑<br />

(<br />

)<br />

f<br />

p<br />

θ 1 ε j ε 3<br />

X 3 − X j . (6.262)<br />

D<br />

j=1 3j M j M 3<br />

Hence the system of the flame equations has been reduced to the five relations<br />

given in Eqs. (6.229), (6.246), (6.251), (6.254) and (6.255) and to the four<br />

differential equations (6.250), (6.256), (6.261) and (6.262) for the nine unknowns ε i<br />

(i = 1, 2, 3, 4), X i (i = 1, 2, 3, 4) and θ. If, furthermore, the differential equations<br />

(6.250), (6.261) and (6.262) are divi<strong>de</strong>d by Eq. (6.256), the coordinate x is eliminated<br />

and θ is introduced as in<strong>de</strong>pen<strong>de</strong>nt variable. The results are<br />

dε 1<br />

dθ = λ ( ) 3/2<br />

√<br />

p<br />

m 2 2M 1 θ −3/2 k1<br />

k 2<br />

c p RT f k 5<br />

(<br />

X 3 X 3/2<br />

2 X 2 + k )<br />

4<br />

X 1 − 1<br />

k 3<br />

·<br />

θ − 1 + q 1 (ε 1f − ε 1 ) − q 4 (ε 4f − ε 4 ) . (6.263)<br />

dX 1<br />

dθ<br />

dX 3<br />

dθ<br />

(<br />

)<br />

4∑ 1 ε j ε 1<br />

= λ<br />

X 1 − X j<br />

RT f<br />

mc p p θ j=1D 1j M j M 1<br />

θ − 1 + q 1 (ε 1f − ε 1 ) − q 4 (ε 4f − ε 4 ) , (6.264)<br />

(<br />

)<br />

4∑ 1 ε j ε 3<br />

= λ<br />

X 3 − X j<br />

RT f<br />

mc p p θ j=1D 3j M j M 3<br />

θ − 1 + q 1 (ε 1f − ε 1 ) − q 4 (ε 4f − ε 4 ) . (6.265)<br />

Solution of the reaction equations<br />

Von Kármán and Penner [40] give the following expressions for k 2<br />

√<br />

k1<br />

k 5<br />

and k 4<br />

k 3<br />

and<br />

where<br />

k 2<br />

√<br />

k1<br />

k 5<br />

= 0.8 × 10 14 e −θ 1/θ , (6.266)<br />

k 4<br />

= 1 = 0.119, (6.267)<br />

k 3 8.4<br />

θ a =<br />

40 200<br />

RT f<br />

(6.268)


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 193<br />

is the reduced activation temperature for the formation of HBr. These authors assume<br />

also that the coefficient of thermal conductivity λ is in<strong>de</strong>pen<strong>de</strong>nt of the composition<br />

of the mixture and that it varies proportionally with the square root of the temperature,<br />

viz.,<br />

λ = λ f<br />

√<br />

θ. (6.269)<br />

If Eqs. (6.266), (6.267), and (6.269) are introduced into Eq. (6.263), we obtain<br />

where<br />

dε 1<br />

dθ = X 3 X 3/2<br />

Λθ−1 2 (X 2 + 0.119X 1 ) −1 e −θ 1 − θ<br />

a<br />

θ<br />

, (6.270)<br />

θ − 1 + q 1 (ε 1f − ε 1 ) − q 4 (ε 4f − ε 4 )<br />

Λ = 1.6 × 10 14 λ ( ) 3/2<br />

f p<br />

m 2 M 1 e −θ a (6.271)<br />

c p RT f<br />

is the eigenvalue which <strong>de</strong>termines the propagation velocity u 0 of the flame through<br />

the relation<br />

u 0 = 1.265 × 10 7 1 ρ 0<br />

√<br />

λ f M 1<br />

c p<br />

( p<br />

RT f<br />

) 3/4<br />

e −θ a/2 Λ −1/2 . (6.272)<br />

The eigenvalue Λ can be obtained by applying the method of von Kármán<br />

and Penner [6] which involves integration of Eq. (6.270) between the cold and hot<br />

boundaries of the flame in the following form<br />

ε 2 ∫ ε1f (<br />

) ∫ 1<br />

1f<br />

q 1<br />

2 − X 3 X 3/2<br />

2 e −θ 1 − θ<br />

a<br />

θ<br />

1 − θ + q 4 (ε 4f − ε 4 ) dε 1 = Λ<br />

0<br />

θ 0<br />

θ(X 2 + 0.119X 1 )<br />

The problem lies now in obtaining:<br />

dθ. (6.273)<br />

1) Approximate expressions for (1 − θ) and (ε 4f − ε 4 ) as function of (ε 1f − ε 1 )<br />

for the computation on the integral on the left hand of Eq. (6.273).<br />

2) Approximate expressions for X 1 , X 2 , and X 3 as functions of θ, for the computation<br />

of the integral appearing on the right hand si<strong>de</strong>.<br />

Von Kármán and Penner [40] performed this computation by assuming that<br />

X 4 as well as ε 4 may be neglected (compared with unity). In this case, the diffusion<br />

equations (for Lewis number close to unity) show that X 1 , X 2 , and X 3 are linear<br />

functions of 1 − θ near θ = 1; a study of Eq. (6.270) shows that ε 1f − ε 1 is a function<br />

of θ of the form<br />

ε 1f − ε 1 = (1 − θ) n (6.274)


194 CHAPTER 6. LAMINAR FLAMES<br />

near θ = 1, where the exponent n takes the following values<br />

X 3,0 < 0.5 : n = 1,<br />

X 3,0 = 0.5 : n = 7 4 ,<br />

(6.275)<br />

X 3,0 > 0.5 : n = 5 4 .<br />

The approximation given in Eq. (6.274) holds only for values of θ very close<br />

to unity. In fact, the values for ε 1f − ε 1 given by such an approximation <strong>de</strong>crease very<br />

rapidly with 1 − θ, reaching the value zero for values of θ very close to unity. A better<br />

approximation for ε 1f − ε 1 can be obtained as follows. In Eq. (6.270), (1 − θ) as well<br />

as q 4 (ε 4f − ε 4 ) are much smaller than q 1 (ε 1f − ε 1 ) for the interesting range of values<br />

of θ. 13 Therefore, a good approximation for Eq. (6.270) can be obtained if these terms<br />

are neglected, thus yielding<br />

(ε 1f − ε 1 ) dε 1<br />

dθ = Λ X 3 X 3/2<br />

2 e −θ 1 − θ<br />

a<br />

θ<br />

q 1 θ(X 2 + 0.119X 1 ) . (6.276)<br />

Equation (6.276) may be integrated from the hot boundary forward. The result is<br />

√ ∫ 1<br />

ε 1f − ε 1 = √ 2Λ q 1<br />

θ<br />

f(θ ′ ) dθ ′ , (6.277)<br />

where<br />

f(θ) = X 3X 3/2<br />

2 e −θ 1 − θ<br />

a<br />

θ<br />

θ(X 2 + 0.119X 1 ) . (6.278)<br />

It can be seen that Eq. (6.277) is an approximation for (ε 1f −ε 1 ) which is valid<br />

for a range of temperatures far more extensive than that covered by Eq. (6.274).<br />

The improved approximation is necessary when the velocity of the flame is<br />

computed, as it is done here, without neglecting the influence of ε 4 and X 4 . In fact, if<br />

an approximation similar to Eq. (6.274) is used in this case, when computing the integral<br />

on the left-hand si<strong>de</strong> of Eq. (6.273), the contribution of this integral would be consi<strong>de</strong>rably<br />

overestimated, as can be verified easily. The approximation of Eq. (6.277)<br />

gives<br />

√<br />

dε 1<br />

dθ = Λ<br />

2q 1<br />

13 However a very small interval near θ = 1 must be exclu<strong>de</strong>d.<br />

f(θ)<br />

√ ∫ . (6.279)<br />

1<br />

θ f(θ′ ) dθ ′


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 195<br />

This expression permits the integral of the left-hand si<strong>de</strong> of Eq. (6.273) to be written<br />

as follows<br />

∫ ε1f<br />

(<br />

1 − θ+q4 (ε 4f − ε 4 ) ) dε 1<br />

0<br />

√<br />

∫<br />

Λ 1 (<br />

)<br />

f(θ)<br />

=<br />

1 − θ + q 4 (ε 4f − ε 4 ) √<br />

2q 1 ∫ dθ.<br />

θ 1<br />

0<br />

θ f(θ′ ) dθ ′<br />

(6.280)<br />

By using these expressions in Eq. (6.273), the following equation for Λ is obtained<br />

ε 2 ∫ 1<br />

1f<br />

q 1<br />

2 − Λ f(θ) dθ<br />

θ<br />

√<br />

0<br />

∫<br />

Λ 1 (<br />

M<br />

)<br />

4<br />

f(θ)<br />

−<br />

1 − θ + q 4 (X 4f − X 4 ) √<br />

2q 1 θ 0<br />

M m ∫ dθ = 0.<br />

1<br />

θ f(θ′ ) dθ ′<br />

Here Eq. (6.251) was used in or<strong>de</strong>r to eliminate ε 4 . 14<br />

notation will be used<br />

and<br />

I =<br />

∫ 1<br />

θ 0<br />

f(θ) dθ =<br />

∫ 1<br />

(6.281)<br />

For simplicity the following<br />

X 3 X 3/2<br />

2 e −θ 1 − θ<br />

a<br />

θ<br />

dθ, (6.282)<br />

θ 0<br />

θ(X 2 + 0.119X 1 )<br />

∫ 1 (<br />

M<br />

)<br />

4<br />

f(θ)<br />

J = (1 − θ) + q 4 (X 4f + X 4 ) √<br />

θ 0<br />

M m ∫ dθ. (6.283)<br />

1<br />

θ f(θ′ ) dθ ′<br />

If these expressions are used in Eq. (6.281) and the resulting equation is solved, the<br />

following is obtained for √ Λ<br />

√<br />

Λ =<br />

√q 1 ε 2 1f<br />

2I<br />

+ 1 ( ) 2 J<br />

+ 1<br />

8q 1 I 2 √ J<br />

2q 1 I . (6.284)<br />

This value, when substituted in Eq. (6.272), gives the corresponding flame velocity.<br />

Computation of X 2 and X 4<br />

Equation (6.284) shows that when computing Λ it is only necessary to know the values<br />

for I and J. In or<strong>de</strong>r to compute I, one must know f(θ) whereas for the calculation<br />

of J one must know f(θ) and X 4 as functions of θ. Finally, Eq. (6.278) shows that<br />

to obtain f(θ), one must know X 1 , X 2 , and X 3 as function of θ. Consequently, the<br />

problem reduces now to the computation of X 1 , X 2 , X 3 , and X 4 as functions of θ.<br />

14 In Eq. (6.281) when changing from ε 4 to X 4 , it has also been assumed that M m is constant through<br />

the flame. Actually its variation is very small.


196 CHAPTER 6. LAMINAR FLAMES<br />

This can be done by means of the diffusion relations given in Eqs. (6.264) and (6.265),<br />

the steady-state Eq. (6.246), and Eq. (6.255).<br />

expressed as<br />

Let X 1f and X 3f be the final values for X 1 and X 3 . These variables can be<br />

X 1 = X 1f − α 1 , (6.285)<br />

X 3 = X 3f + α 3 , (6.286)<br />

where α 1 and α 3 are functions of θ which approach zero when θ approaches one.<br />

When Eqs. (6.246), (6.285) and (6.286) are combined with Eq. (6.255), the<br />

following equation is obtained for the <strong>de</strong>termination of X 2<br />

√<br />

k 1 RT f<br />

X 1f − α 1 + X 3f + α 3 + X 2 +<br />

k 5 p θX 2 = 1. (6.287)<br />

The computation that follows holds for hydrogen-rich flames where X 2f =<br />

X 4f = 0 and the following condition is consequently satisfied<br />

X 1f + X 3f = 1. (6.288)<br />

Equation (6.287) can now be written as<br />

√<br />

k 1 RT f<br />

X 2 +<br />

k 5 p θX 2 − (α 1 − α 3 ) = 0. (6.289)<br />

On solving Eq. (6.289), one obtains for X 2<br />

(√<br />

√ )2<br />

k 1<br />

X 2 = θ RT f<br />

+ α 1 − α 3 − 1 k 1<br />

θ RT f<br />

. (6.290)<br />

4k 5 p<br />

2 k 5 p<br />

Von Kármán and Penner [40] give for<br />

where<br />

√<br />

k1<br />

k 5<br />

the expression<br />

√<br />

k1<br />

k 5<br />

= 1.676 e − θ r<br />

θ , (6.291)<br />

θ r =<br />

Finally, combining Eqs. (6.291) and (6.290) it is obtained<br />

22 605<br />

RT f<br />

. (6.292)<br />

X 2 =<br />

(√<br />

g(θ)2 + α 1 − α 3 − g(θ)) 2<br />

, (6.293)


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 197<br />

where<br />

θ √<br />

g(θ) = 0.838 e − r<br />

θ θ RT f<br />

p . (6.294)<br />

Likewise, when Eq. (6.291) is substituted into Eq. (6.246), one finds<br />

√<br />

X 4 = 1.676 e − θr<br />

θ θ RT √<br />

f X2 . (6.295)<br />

p<br />

Equations (6.293) and (6.295) permit the computation of X 2 and X 4 as soon as α 1<br />

and α 3 are known. Nevertheless, before initiating the calculation of these quantities<br />

let us study the behavior of X 2 as a function of θ, since this relation influences more<br />

than any other the value of the flame velocity, as shown by Eq. (6.282).<br />

For θ → 1, the following conditions are satisfied<br />

√<br />

g(θ) → 0.838 e −θ RT<br />

r f<br />

≠ 0,<br />

p<br />

α 1 − α 3 → 0.<br />

(6.296)<br />

Therefore, for θ near 1<br />

α 1 − α 3<br />

g(θ) 2 ≪ 1, (6.297)<br />

it is now simple to show from Eq. (6.293) that, near θ = 1, X 2 behaves as follows<br />

X 2 ≃ 1<br />

2θ r<br />

p<br />

θ −1 e θ (α 1 − α 3 ) 2 . (6.298)<br />

2.808 RT f<br />

√<br />

On the other hand, when the difference 1−θ increases, the term 0.838 e −θr/θ θ RT f<br />

p<br />

<strong>de</strong>creases very rapidly since θ r ≫ 1, whilst the term α 1 − α 3 increases. Therefore,<br />

when the difference 1 − θ is not very small compared to unity, X 2 behaves as follows<br />

X 2 ≃ α 1 − α 3 . (6.299)<br />

In the different behaviors of Eqs. (6.298) and (6.299) lie essentially the influence<br />

of the dissociation of bromine on the propagation velocity of the flame. In fact,<br />

if the term X 4 is neglected in Eq. (6.255) when computing X 2 , as was done in [40],<br />

then the approximation given in Eq. (6.299) holds for all temperatures. On the contrary,<br />

when the influence of X 4 is taken into account, the values for X 2 near θ = 1<br />

are far smaller for than those given by Eq. (6.299), since then X 2 varies as the square<br />

of (α 1 − α 3 ), as shown by Eq. (6.298). Since the values for X 2 are smaller, the value<br />

for the integral I is also smaller, as shown by Eq. (6.28). Finally, if I <strong>de</strong>creases, √ Λ<br />

increases as shown by Eq. (6.284). Therefore, according to Eq. (6.272), u 0 <strong>de</strong>creases.


198 CHAPTER 6. LAMINAR FLAMES<br />

Hence, dissociation of bromine reduces the flame propagation velocity basically because<br />

such dissociation reduces the molar fractions of Br 2 at temperatures close to<br />

T f . Dissociation also influences the value for J, as shown by Eq. (6.283). However,<br />

this influence is small and it acts opposite to the influence of thermal conductivity; the<br />

value for J is very small and its influence is negligible. This last conclusion will be<br />

verified when numerical calculations are performed later on.<br />

Introducing Eq. (6.298) into Eq. (6.295) we find, for θ ≃ 1, the following<br />

behavior of X 4<br />

X 4 ≃ 1.676 e −θ r<br />

√<br />

RT f<br />

p (α 1 − α 3 ), (6.300)<br />

thus X 4 is a linear function of (α 1 − α 3 ) and, for θ ≃ 1, X 4 ≫ X 2 .<br />

Evaluation of X 1 and X 3<br />

The evaluation of X 1 and X 3 can be easily performed from the diffusion relations<br />

Eqs. (6.264) and (6.265).<br />

Equations (6.229) and (6.254) permit us to express ε 2 and ε 3 as functions of ε 1<br />

and ε 4 as follows<br />

and<br />

ε 2 = M 4<br />

M 1<br />

(ε 1f − ε 1 ) − ε 4 , (6.301)<br />

ε 3 = ε 3f + M 1 − M 4<br />

M 1<br />

(ε 1f − ε 1 ). (6.302)<br />

By introducing these expressions, together with Eqs. (6.285) and (6.286) , into<br />

Eq. (6.264) and (6.265), and if ε 4 , X 2 and X 4 are expressed as functions of (α 1 − α 3 )<br />

by use of the relations given in Eqs. (6.251), (6.298) and (6.300), two equations are<br />

obtained which <strong>de</strong>pend only on θ, (ε 1f − ε 1 ), α 1 and α 3 . These expressions can be<br />

<strong>de</strong>veloped in series of these variables near θ = 1. The results are<br />

dX 1<br />

dθ<br />

dX 3<br />

dθ<br />

[( =λ f RT f M4 X 1f<br />

+ M 1 − M 4<br />

pc p q 1 M 1 M 2 D 12 M 1 M 3<br />

(<br />

α1 α 3<br />

+O , ,<br />

ε 1f − ε 1 ε 1f − ε 1<br />

[( =λ f RT f M4 X 3f<br />

− M 1 − M 4<br />

pc p q 1 M 1 M 2 D 23 M 1 M 3<br />

(<br />

α1 α 3<br />

+O , ,<br />

ε 1f − ε 1 ε 1f − ε 1<br />

X 1f<br />

D 13<br />

+<br />

1 )<br />

X 3f<br />

M 1 D 13<br />

)<br />

1 − θ<br />

+ higher or<strong>de</strong>r terms<br />

ε 1f − ε 1<br />

X 1f<br />

D 13<br />

−<br />

1 )<br />

X 3f<br />

M 1 D 13<br />

)<br />

1 − θ<br />

+ higher or<strong>de</strong>r terms<br />

ε 1f − ε 1<br />

]<br />

, (6.303)<br />

]<br />

. (6.304)


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 199<br />

dX 1<br />

dθ<br />

Hence, for θ ≃ 1, we find<br />

≃ λ (<br />

f RT f M4 X 1f<br />

+ M 1 − M 4 X 1f<br />

+<br />

pc p q 1 M 1 M 2 D 12 M 1 M 3 D 13<br />

1 )<br />

X 3f<br />

≡ β 1 , (6.305)<br />

M 1 D 13<br />

dX 3<br />

dθ<br />

≃ λ (<br />

f RT f M4 X 3f<br />

− M 1 − M 4 X 1f<br />

−<br />

pc p q 1 M 1 M 2 D 23 M 1 M 3 D 13<br />

1 )<br />

X 3f<br />

≡ −β 3 , (6.306)<br />

M 1 D 13<br />

which leads to the following approximations for X 1 and X 3<br />

X 1 ≃ X 1f − β 1 (1 − θ), (6.307)<br />

X 3 ≃ X 3f + β 3 (1 − θ). (6.308)<br />

Thus<br />

where<br />

α 1 − α 3 = (β 1 − β 3 )(1 − θ) = β 2 (1 − θ), (6.309)<br />

β 2 = β 1 − β 3 . (6.310)<br />

near θ = 1<br />

In short, the following approximations for X 1 , X 2 , X 3 and X 4 are obtained<br />

X 1 = X 1f − β 1 (1 − θ), (6.311)<br />

X 2 =<br />

(√<br />

g(θ)2 + β 2 (1 − θ) − g(θ)) 2<br />

, (6.312)<br />

X 3 = X 3f − β 3 (1 − θ), (6.313)<br />

(√ )<br />

X 4 = 2g(θ) g(θ)2 + β 2 (1 − θ) − g(θ) . (6.314)<br />

Evaluation of f(θ), I and J<br />

When computing f(θ), the only values of X 1 , X 2 and X 3 nee<strong>de</strong>d are those for θ near<br />

1, since the reduced activation temperature θ a is much larger than unity. The relevant<br />

expressions are given in Eqs. (6.311) through (6.313). Once f(θ) is known, the<br />

integral I can be computed by numerical of graphical integration. Likewise, knowing<br />

f(θ) and X 4 , the integral J can be obtained either numerically or graphically.<br />

Computation of the flame velocity<br />

The preceding results may be summarized by the following method for the computation<br />

of the velocity of the flame.


200 CHAPTER 6. LAMINAR FLAMES<br />

1) Values for β 1 , β 3 and β 2 are obtained from Eqs. (6.305), (6.306) and (6.310).<br />

2) Introducing these values into Eqs. (6.307), (6.308) and (6.312) one obtains X 1 ,<br />

X 2 and X 3 as functions of θ.<br />

3) Substituting the results into Eq. (6.278), the value of f(θ) is obtained.<br />

4) From the numerical or graphical integration of f(θ) between θ 0 and 1, the value<br />

of the integral I, <strong>de</strong>fined in Eq. (6.282), is obtained.<br />

5) The numerical or graphical integration of f(θ) between θ and 1 gives the value<br />

of<br />

∫ 1<br />

θ<br />

f(θ ′ ) dθ ′ .<br />

6) We introduce this value of<br />

∫ 1<br />

θ<br />

f(θ) dθ, the value of f(θ), computed in (6.230),<br />

and the value for X 4 , given by Eq. (6.314), into Eq. (6.283). Then we integrate<br />

Eq. (6.283), either numerically or graphically, between θ 0 and 1 in or<strong>de</strong>r to<br />

obtain the value of J.<br />

7) When both the values for I and J are substituted into Eq. (6.284) the value for<br />

√<br />

Λ is obtained.<br />

8) Finally, u 0 is computed from Eq. (6.272).<br />

It has been stated that the influence of J on the value of u 0 is negligible. This<br />

fact allows a consi<strong>de</strong>rable simplification of the method since it eliminates steps 5) and<br />

6). By neglecting J in Eq. (6.284), the following simplified equation results<br />

√<br />

Λ =<br />

√q 1 ε 2 1f<br />

2I . (6.315)<br />

Results<br />

The method <strong>de</strong>veloped herein has been applied to the computation of the four cases<br />

calculated by von Kármán and Penner for hydrogen-rich flames. The results obtained<br />

are given in Table (6.5), in which are also listed the values obtained by von Kármán<br />

and Penner.<br />

X 3,0 0.55 0.60 0.65 0.70<br />

Von Kármán and Penner 30.6 42.3 34.2 23.6<br />

u 0 calculated from Eq. (6.315) 22.9 33.7 30.5 21.9<br />

u 0 calculated from Eq. (6.284) 23.3 34.2 – –<br />

Table 6.5: Flame velocity (cm/s) for hydrogen-rich H 2-Br 2 mixtures.


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 201<br />

The physico-chemical constants given by von Kármán and Penner were used<br />

in our calculations. The values of these constants, as well as those for the integrals<br />

and parameters nee<strong>de</strong>d for evaluation of u 0 , are summarized in Table 6.6.<br />

In or<strong>de</strong>r to analyze the influence of convection and of dissociation of bromine<br />

upon the value of u 0 , the flame velocity was computed for two cases, using the correct<br />

relation given in Eq. (6.284) for √ Λ. The results are listed in Table 6.5 and 6.6, which<br />

show that the correction is negligible small due to the cancellation of convection and<br />

dissociation energy effects.<br />

X 3,0 0.55 0.60 0.65 0.70<br />

X 2,0 0.45 0.40 0.35 0.30<br />

X 1,f 0.90 0.80 0.70 0.60<br />

X 3,f 0.10 0.20 0.30 0.40<br />

ε 1f 0.997 0.994 0.990 0.984<br />

T f K 1660 1585 1460 1324<br />

10 5 λ f (cal/cm-s-K) 19.8 25.0 29.4 33.1<br />

c p (cal/g-K) 0.105 0.116 0.132 0.151<br />

M (g/mole) 73.1 65.2 57.3 49.4<br />

θ a 12.2 12.75 13.9 15.28<br />

θ r 6.81 7.13 7.83 8.55<br />

θ 0 0.1945 0.204 0.221 0.244<br />

β 1 1.532 1.745 1.762 1.684<br />

β 2 1.380 1.560 1.560 1.470<br />

β 3 0.152 0.185 0.202 0.214<br />

q 1 0.807 0.800 0.787 0.768<br />

q 4 1.63 1.55 1.475 1.42<br />

10 3 I 0.667 1.601 2.639 3.395<br />

√<br />

Λ 24.54 13.82 12.09 10.47<br />

u 0<br />

√<br />

Λ (cm/s) 551 529 369 228<br />

u 0 (cm/s) 22.9 33.7 30.5 21.8<br />

10 3 J -0.87 -0.94 — —<br />

√<br />

Λ from Eq. (6.284) 24.03 15.49 — —<br />

u 0 (cm/s) from Eq. (6.284) 23.34 34.15 —<br />

Table 6.6: Parameters for the computation of flame velocities.<br />

The results given in Table 6.5 show that dissociation of bromine reduces the<br />

flame velocity consi<strong>de</strong>rably. Values for u 0 have been plotted in Fig. 6.17 which also<br />

shows the results calculated by von Kármán and Penner [40] and by Gilbert and Altman<br />

[29] . Also listed are experimental data obtained by An<strong>de</strong>rson and his collaborators<br />

[43]. It will be seen that the influence of dissociation of bromine on u 0 predicted<br />

by Gilbert and Altman is exaggerated. This conclusion is not surprising if one consi<strong>de</strong>rs<br />

that these authors estimated the influence of dissociation from the results obtained<br />

through a thermal theory. Then they applied the reduction factor obtained from a ther-


202 CHAPTER 6. LAMINAR FLAMES<br />

S b<br />

(cm/s)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

von Kármán − Penner<br />

von Kármán − Millán<br />

Gilbert − Altman (Dissoc. Negl.)<br />

Gilbert − Altman (Dissoc. Incl.)<br />

Boys − Corner (Diss. Negl.)<br />

(1) Flame cone area method<br />

(2) Flame cone angle method<br />

(3) Tube method<br />

(1)<br />

(2)<br />

(3)<br />

10<br />

0<br />

0.50 0.55 0.60 0.65 0.70<br />

X 3,0<br />

Figure 6.17: The quantity S b as a function of X 3,0 for hydrogen-rich H 2 − Br 2 mixtures.<br />

mal theory to a diffusion theory in which the influence of dissociation was neglected.<br />

On the other hand, the excellent agreement between their results and the experimental<br />

ones is not too relevant since the former were obtained by applying the method of Boys<br />

and Corner, which gives consi<strong>de</strong>rably smaller values for u 0 than those obtained from<br />

the application of a more correct method, as will now be verified. With the purpose of<br />

estimating the uncertainty due to the use of a poor method of computation, the value<br />

of u 0 for X 3,0 = 0.70 was computed with the method of Boys and Corner, as applied<br />

by Gilbert and Altman, and with the method of von Kármán and Penner. In both cases<br />

the influence of the dissociation of bromine was neglected and the physico-chemical<br />

constants of von Kármán and Penner were used, so that the difference in numbers was<br />

only due to the difference in method. The following are the results obtained.<br />

Boys-Corner method, as applied by Gilbert and Altman:<br />

Von Kármán and Penner method:<br />

u 0 = 13.8 cm/s,<br />

u 0 = 23.6 cm/s.<br />

Therefore, if Gilbert and Altman had applied a better method than that of Boys and<br />

Corner, they would have obtained consi<strong>de</strong>rably larger values for u 0 . The agreement<br />

between their values and those obtained from experiments must be consi<strong>de</strong>red to be<br />

largely fortuitous.


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 203<br />

1.00<br />

0.75<br />

X 1<br />

X 2<br />

0.50<br />

X 2<br />

(dissociation neglected)<br />

X 3<br />

ε 1f<br />

− ε 1<br />

X 4<br />

0.25<br />

0.0 0.1 0.2 0.3<br />

1 − θ<br />

Figure 6.18: Composition and flux profiles for X 3,0 = 0.65.<br />

Structure of the Flame<br />

The distributions of X 1 , X 2 , X 3 , X 4 and ε 1f − ε 1 corresponding to one of the cases<br />

consi<strong>de</strong>red are shown in Fig. 6.18.<br />

In or<strong>de</strong>r to show the influence of X 4 on the distribution of X 2 , the figure also<br />

indicates the values for X 2 obtained when X 4 is neglected. Obviously, the influence<br />

of X 4 is very important, especially for values of θ near 1. On the other hand, this is<br />

the most important influence of neglecting X 4 . It may be seen that small <strong>de</strong>viations of<br />

X 1 and X 3 from the linear law have little influence upon the value of u 0 , since X 1f<br />

and X 3f are different from zero,<br />

References<br />

[1] Mallard, E. and Le Chatelier, H.: Recherches sur la Combustion <strong>de</strong>s Melanges<br />

Gateaux Explosives Ann. Mines, Vol. 8 series 4, 1883, p. 274.<br />

[2] Crussard, L.: Les Deflagrations on Régime Permanent dans les Mileux Conducteurs.<br />

Compt. Rend., Vol. 158, 1914, pp. 125, 340.<br />

[3] Coward H. F. and Payman W.: Chem. Rev., Vol. 21, 1937, p. 359.<br />

[4] Zeldovich, Y. B. and Semenov, N.: Kinetics of Chemical Reactions in Flames.<br />

NACA Tech. Memo. No. 1084, 1946.


204 CHAPTER 6. LAMINAR FLAMES<br />

[5] Boys, S. F. and Corner, J.: The Structure of the Reaction Zone in a Flame. J.<br />

Proc. Roy Soc. London, 1949, A1-97,90.<br />

[6] von Kármán, Th. and Penner, S. S.: Fundamental Approach to Laminar Flame<br />

Propagation. Selected Combustion Problems, Vol. I, AGARD, 1954, pp. 5-41.<br />

[7] Hirschfel<strong>de</strong>r, J. O., Curtiss, C. F. and Campbell, E.: The Theory of Flames and<br />

Detonations. Fourth Symposium (International) on Combustion, Williams and<br />

Wilkins Co., Baltimore, 1953.<br />

[8] Lewis, B. and von Elbe, G.: On the Theory of Flame Propagation. Phys., Vol.<br />

2, 1934, pp. 537-546.<br />

[9] Evans, N. W.: Current Theoretical Concepts of Steady-State Flame Propagation.<br />

Chem. Rev., Vol. 51, 1952.<br />

[10] von Kármán, Th.: The Present Status of the Theory of Laminar Flame Propagation.<br />

Sixth Symposium (International) on Combustion, Reinhold Publishing<br />

Corp., New York, 1957.<br />

[11] Tanford, C. and Pease, R.: Equilibrium Atom and Free Radical Concentrations<br />

in Carbon Monoxi<strong>de</strong> Flames and Correlation with Burning Velocities. J. Chem.<br />

Phys., Vol. 15, 1947, pp. 431-433.<br />

[12] van Tiggelen, A.: Chemical Theory of the Speed of Flame Propagation. Bull.<br />

Soc. Chim. Belg., Vol. 58, 1949, p. 259.<br />

[13] Jost, W.: Explosion and Combustion Processes in Gases. McGraw-Hill, New-<br />

York-London, 1946.<br />

[14] Gaydon A. G. and Wolfhard, H. G.: Flames, Their Structure, Radiation and<br />

Temperature. Chapman Hall, London, 1953.<br />

[15] La<strong>de</strong>nburg, R. W., Lewis, B., Pease, R. N. and Taylor, H. S.: Physical Measurements<br />

in Gas Dynamics and Combustion. Vol. IX of High Speed Aerodynamics<br />

and Jet Propulsion, Princeton University Press, 1954.<br />

[16] Linnet, J. W.: Methods of Measuring Burning Velocities. Fourth Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore, 1953,<br />

pp. 20-35.<br />

[17] Emmons, H. W., Harr, J. A. and Strong, P.: Harvard University, Contract No.<br />

AT (30-1)-497, 1950.<br />

[18] Adamson, T. C.: On the Theory of One Dimensional Flame Propagation. Jet<br />

Propulsion, January-February 1952.<br />

[19] von Kármán Th. and Millán, G.: The Thermal Theory of Constant Pressure<br />

Deflagration. Biezeno’s Anniversary Volume, Delft, Holland, 1953.


6.15. FLAME PROPAGATION IN HYDROGEN-BROMINE MIXTURES 205<br />

[20] Adams E. N., quoted by Henkel, M. J., Spaulding, W. P. and Hirschfel<strong>de</strong>r J. O.:<br />

Theory of Propagation of Flames, Part II Approximate Solutions. Third Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1949, pp. 127-135.<br />

[21] Wil<strong>de</strong>, K. A.: J. Chem. Phys., Vol. 22, 1954, p. 1788.<br />

[22] Millán G., Sendagorta, J. M. and Da Riva, I.: Comparison of Analytical Methods<br />

for the Calculation of Laminar Flame Velocity. ARDC Contract AF 61<br />

(514), 997, 1957.<br />

[23] Sendagorta, J. M.: Method for the Computation of the Propagation Velocity of<br />

a Plane Laminar Flame. ARDC Contract AF 61 (514), 997, 1957.<br />

[24] Fristrom, R. M.: The Structure of Laminar Flame Fronts. Sixth Symposium<br />

(International) on Combustion, Reinhold Publishing Corp., New York, 1957.<br />

[25] Hirschfel<strong>de</strong>r, J. O., Curtis, J. P. and Campbell, D. E.: The Theory of Flame<br />

Propagation. IV Rep. No. OM-756. Univ. Wisconsin, 1952.<br />

[26] Campbell, E. S.: Theoretical Study of the Hydrogen Bromine Flame. Sixth<br />

Symposium (International) on Combustion, Reinhold Publishing Corp., New<br />

York, 1957.<br />

[27] Giddings, J. C. and Hirschfel<strong>de</strong>r J. O.: Flame Properties and the Kinetics of<br />

Chain-Branching Reactions. Sixth Symposium (International) on Combustion,<br />

Reinhold Publishing Corp., New York, 1957.<br />

[28] Spalding, D. B.: The Theory of Flame Phenomena with a Chain Reaction. Phil.<br />

Trans. Roy. Soc., London 1956.<br />

[29] Gilbert, M. and Altman, D.: The Chemical Steady State in HBr Flames. Sixth<br />

Symposium (International) on Combustion, Reinhold Publishing Corp., New<br />

York, 1957.<br />

[30] Millán, G. and Sanz, S.: Hydrazine Decomposition Flame. ARDC Contract No.<br />

AF 61 (514)-734-O, 1956.<br />

[31] Strong, A. G. and Grosse, A. V.: The Ozone to Oxygen Flame. Sixth Symposium<br />

(International) on Combustion, Reinhold Publishing Corp., New York,<br />

1957.<br />

[32] Murray, R. O. and Hall. A. R.: Flame Speeds in Hydrazine Vapour and in<br />

Mixtures of Hydrazine and Ammonia with Oxygen. Trans. Faraday Soc., Vol.<br />

57, 1951, pp. 743-751.<br />

[33] Zeldovich, Y. B.: Theory of Flame Propagation. NACA Tech. Memo. No.<br />

1282, 1951.<br />

[34] Szwarc M. J.: Chem. Phys. Vol. 17, 1949, p. 505.


206 CHAPTER 6. LAMINAR FLAMES<br />

[35] Hirschfel<strong>de</strong>r, J. O. and Curtiss O. F.: J. Phys Chem. Vol. 57.<br />

[36] Adams, G. K. and Stocks, G. W.: The Combustion of Hydrazine. Fourth Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, pp. 239-248.<br />

[37] Spalding, D. B.: The Theory of Flame Phenomenon with a Chain Reaction.<br />

Phil. Trans. Roy. Soc. London, 1956.<br />

[38] Gilbert M. and Altman D.: Hydrazine Decomposition Flame. ORDCIT Project<br />

Contract No. DA-04-495, Ord. 18 Progress Report No. 20-278, 1956.<br />

[39] Millán, G. and Sendagorta J. M.: Hydrazine Decomposition Flame. ARDC<br />

Contract No. AF 61-(514)-997, 1957.<br />

[40] von Kármán, Th. and Penner S. S.: The Theory of One-Dimensional Laminar<br />

Flame Propagation for Hydrogen-Bromine Mixtures. Part I. Dissociation<br />

Neglected. Tech. Rep. No. 16, Contract No. DA 04-495-0rd. 446, 1956.<br />

[41] von Kármán Th. and Millán G.: The Theory of One-Dimensional Laminar<br />

Flame Propagation for H 2 -Br 2 Mixtures. Part II. Dissociation Inclu<strong>de</strong>d. Tech.<br />

Rep. No. 16, Contract No. DA-04-495-Ord.-446, 1956.<br />

[42] Mileson, D. F.: The Thermal Theory of Laminar Flame Propagation for Hydrogen-<br />

Bromine Mixtures. Tech. Rep. No. 6, Contract No. DA 04-495-Ord.-446, 1954.<br />

[43] An<strong>de</strong>rson, R. C.: Hydrogen-Halogen Flames. AGARD Combustion Panel Meeting,<br />

Oslo, 1956.


Chapter 7<br />

Turbulent flames<br />

7.1 Introduction<br />

There is experimental evi<strong>de</strong>nce that turbulence increases the propagation velocity of<br />

the flame through a combustible mixture. This is a very important fact in technical<br />

applications since it allows a consi<strong>de</strong>rable reduction of the space and time required to<br />

burn a given mass. The influence of turbulence over combustion was first recognized<br />

by Mallard and Le Chatelier in 1883 [1]. However the attempts ma<strong>de</strong> to <strong>de</strong>termine the<br />

causes of this influence as well as to estimate quantitatively its value are quite recent.<br />

The basic problem lies in <strong>de</strong>termining the propagation velocity of the flame through a<br />

combustible mixture in turbulent motion knowing the characteristics of the turbulence<br />

and the state and composition of the mixture.<br />

From the experimental stand-point, several techniques are available for the <strong>de</strong>termination<br />

of the flame velocity. However the basis for such techniques are not as<br />

solid as those applied to the case of laminar flames and the measurements taken are<br />

not as numerous or systematic. Among them, the technique more commonly used<br />

consists in photographing a flame, obtained for example in a bunsen burner, then measuring<br />

the area of the combustion front and dividing it by the flow rate of the burner<br />

as is done for the case of laminar flames. One of the difficulties of such techniques<br />

is the fact that the combustion front of a turbulence flame is not well <strong>de</strong>fined since<br />

the long exposure photographs show a thick luminous zone (see Fig. 7.1a) whereas<br />

a short exposure “Schlieren” photograph reveals a very irregular and wrinkled structure<br />

(see Fig. 7.1b). These circumstances impose the introduction of an arbitrariness<br />

in estimating the flame area. Damköhler [2], for instance, adopted as surface of the<br />

turbulent combustion front the one that limits internally the luminous zone. Later on,<br />

207


208 CHAPTER 7. TURBULENT FLAMES<br />

(a) Time exposure photograph<br />

(b) Instantaneous Schlieren photograph<br />

Figure 7.1: Stoichiometric natural gas-air flame, Re = 25 000, burner tube diameter =<br />

5.08 cm (by courtesy of Bureau of Mines).<br />

preference was given to the use of the mean surface between those internally and externally<br />

limiting the luminous zone [3] or else to the surface where luminosity reaches<br />

its maximum intensity [4].<br />

Additional difficulty lies in the fact that the paths of the gas particles are not<br />

known. Therefore it is risky to i<strong>de</strong>ntify the fraction of unburnt mixture corresponding<br />

to each element of the flame front. Such difficulty has often been elu<strong>de</strong>d by measuring<br />

the mean velocity obtained when the total flow rate of the mixture is divi<strong>de</strong>d by the<br />

area of the flame front. The propagation velocity of turbulent flames has been measured<br />

with these or similar techniques by Damköhler [2] , Bollinger and Williams [3],<br />

Scurlock [4], Williams, Hottel and Scurlock [5], Karlovitz, Denniston and Wells [6],<br />

Wohl, Shore, von Rosemberg and Weil [7], Leason [8], Bowditch [9], Wohl and Shore<br />

[10], Mickelsein and Ernstein [11], etc.<br />

Figure 7.2 taken from Ref. [3] shows, as an example, the results of the measurements<br />

performed by Bollinger and Williams in mixtures of several hydrocarbons<br />

and air. For each case the mixture giving the maximum velocity was used. This figure<br />

also shows the laminar velocities corresponding to the same mixtures. It is seen that<br />

the effect is maximum in the acetylene flame where the value of the laminar flame ve-


7.1. INTRODUCTION 209<br />

300<br />

250<br />

TURBULENT FLAME SPEED, cm/s<br />

200<br />

150<br />

100<br />

50<br />

0<br />

0<br />

10 20 30 40x10 3<br />

REYNOLDS NUMBER OF PIPE FLOW<br />

Figure 7.2: Variation of flame speed with Reynolds number of flow.<br />

locity is half of that for a Reynolds number 3 × 10 4 . Other measurements have given<br />

turbulent velocities consi<strong>de</strong>rably higher than these observed here.<br />

From the theoretical stand-point several attempts have been ma<strong>de</strong> to explain the<br />

activation of combustion due to turbulence. The first attempt was ma<strong>de</strong> by Damköhler<br />

[2] who pointed-out two causes for the action of turbulence. One is the increase in the<br />

transport coefficients (conductivity and diffusion) due to turbulent diffusivity. The second<br />

cause is the distortion of the laminar flame front due to the turbulent oscillations<br />

of the velocity which would increase, consi<strong>de</strong>rably, the effective surface of the combustion<br />

front. The importance of either one of these two factors would <strong>de</strong>pend in each<br />

case on the relation between the scale of turbulence and the thickness of the laminar


210 CHAPTER 7. TURBULENT FLAMES<br />

flame. In technical applications this ratio is generally very large, hence the main cause<br />

for the activation of combustion would be in this case the distortion of the flame front.<br />

Figure 7.1b seems to confirm this point of view. The major part of the later works on<br />

the subject have attempted to estimate the importance of either one of these causes,<br />

preserving Damköhler’s mo<strong>de</strong>l. In 1952 von Kármán and Marble [12] proposed a different<br />

mo<strong>de</strong>l which consi<strong>de</strong>red the turbulent flame as a zone whose structure should<br />

be <strong>de</strong>fined by mean values of the characteristic variables (temperature, reaction rate,<br />

etc.) to which were superimposed static oscillations of turbulent nature. Later, this<br />

mo<strong>de</strong>l was adopted and <strong>de</strong>veloped by Sommerfield and his collaborators [13], who<br />

obtained some experimental evi<strong>de</strong>nce (not sufficient) that the approximation is correct.<br />

The comparison between theoretical and experimental results is not conclusive<br />

enough in any case to establish either one of the proposed theories, which are actually<br />

in a preliminary phase.<br />

Consequently the present study will only be a brief exposure of the basic principles<br />

and fundamental results of these theories.<br />

7.2 Turbulent combustion theories<br />

So far, the only type of turbulence whose influence on combustion has been studied is<br />

the isotropic. This turbulence is characterized by two magnitu<strong>de</strong>s: its scale l and its<br />

intensity v ′ . 1 Damköhler studies the influence of turbulence on the flame by comparing<br />

l and v ′ with the corresponding magnitu<strong>de</strong>s of the laminar combustion wave of the<br />

mixture; these being thickness d l of the flame and the laminar propagation velocity<br />

u l . When performing this comparison the following cases arise.<br />

flame<br />

1) The turbulence scale is small when compared to the thickness of the laminar<br />

l<br />

d l<br />

≪ l (7.1)<br />

In this case the action of turbulence is reduced to an activation of the transport coefficients<br />

conductivity and diffusion at the flame). Let<br />

x =<br />

λ<br />

ρc p<br />

(7.2)<br />

and<br />

ε ∼ lv ′ , (7.3)<br />

1 For an exposure of the principles of turbulence, see, i.e., H.L. Dry<strong>de</strong>n: A review of the Statistical Theory<br />

of Turbulence. Quart. Applied Math., Vol. 1, 1943, pp. 7-42.


7.2. TURBULENT COMBUSTION THEORIES 211<br />

be the laminar and turbulent thermal diffusivities, respectively. The turbulent propagation<br />

velocity u t of the flame is obtained from u l when x is substituted by ε. Since<br />

u l is proportional to √ x, it results for u t<br />

u t<br />

u l<br />

=<br />

√ ε<br />

x<br />

(7.4)<br />

Shelkin [14] also consi<strong>de</strong>rs the influence of laminar diffusivity on the turbulent flame.<br />

In this case, ε must be substituted in (7.3) by x + ε, thus resulting<br />

√<br />

u t<br />

= 1 + ε u l x<br />

(7.5)<br />

This expression has the advantage over (7.4) that when turbulence approaches zero,<br />

u t approaches u l .<br />

Case 1) very seldom arises since un<strong>de</strong>r normal conditions the thickness of the<br />

laminar flame is only of some tenths of a millimeter.<br />

2) Scale of turbulence and flame thickness are of the same or<strong>de</strong>r of magnitu<strong>de</strong><br />

l<br />

d l<br />

≃ 1. (7.6)<br />

Damköhler does not consi<strong>de</strong>r this case. However Shelkin has proven that the<br />

action of turbulence does not reduce then to an activation of the transport coefficients<br />

but it also influences the combustion time of the mixture, which not only <strong>de</strong>pends on<br />

the reaction velocity of the species but on the rapidity of the turbulent mixing. To date,<br />

no formula is available for this case.<br />

3) Scale of turbulence is large when compared to flame thickness<br />

l<br />

d l<br />

≫ 1. (7.7)<br />

In this case combustion takes place through a laminar front but turbulence wrinkles<br />

the laminar flame front thus increasing its surface. Thereby, the effective combustion<br />

surface of the mixture is increased. Two possibilities should be consi<strong>de</strong>red<br />

<strong>de</strong>pending on the intensity of the turbulence.<br />

3.a) The intensity of turbulence is very large with respect to the laminar propagation<br />

velocity of the flame<br />

v ′<br />

u l<br />

≫ 1. (7.8)<br />

This case was not analyzed by Damköhler. The following consi<strong>de</strong>ration are<br />

owned to Shelkin. When condition (7.8) takes place the structure of the combustion


212 CHAPTER 7. TURBULENT FLAMES<br />

B<br />

A<br />

BURNED<br />

UNBURNED<br />

GASES<br />

GASES<br />

B’<br />

A’<br />

Figure 7.3: Structure of the turbulent flame in the case l ≫ d l and v ′ ≫ u l , according to<br />

Shelkin.<br />

zone shows islands of unburnt gas which have been dragged by the strong turbulent oscillations<br />

towards the region of combustion products without time to burn, see Fig. 7.3.<br />

In this case, the zone between AA ′ and BB ′ can be consi<strong>de</strong>red as a combustion wave<br />

which advances with a propagation velocity u t . The propagation velocity of a combustion<br />

wave is <strong>de</strong>termined by an equilibrium between transport processes, which are<br />

characterized by combustion time τ. In<strong>de</strong>pen<strong>de</strong>ntly from the mechanism which <strong>de</strong>termines<br />

the values for ε and τ, a dimensionless analysis shows that u t must be of the<br />

form<br />

u t ∼<br />

√ ε<br />

τ . (7.9)<br />

For the case shown in Fig. 7.3, ε is clearly the turbulent diffusivity. Combustion<br />

velocity is <strong>de</strong>termined by the mixing rapidity and therefore τ is proportional to the<br />

time of turbulent mixing<br />

τ ∼ l v ′ . (7.10)<br />

When the expressions for ε and τ given by (7.2) and (7.10) are taken into (7.9)<br />

u t ∼ v ′ , (7.11)<br />

Hence, the propagation velocity is proportional to the intensity and in<strong>de</strong>pen<strong>de</strong>nt from<br />

the scale of turbulence. Furthermore, u t results to be in<strong>de</strong>pen<strong>de</strong>nt from laminar velocity<br />

u l . Such conclusion contradicts experimental evi<strong>de</strong>nce.<br />

3.b) The intensity of turbulence is of the same or<strong>de</strong>r of magnitu<strong>de</strong> or smaller<br />

than the propagation velocity of the laminar flame<br />

v ′<br />

u l<br />

1. (7.12)<br />

In this case the islands disappear and the action of turbulence reduces to <strong>de</strong>forming


7.2. TURBULENT COMBUSTION THEORIES 213<br />

BURNED<br />

σ<br />

l<br />

UNBURNED<br />

σ<br />

GASES<br />

GASES<br />

Figure 7.4: Structure of the turbulent flame in the case l ≫ d l and v ′ u l , according to<br />

Shelkin.<br />

the laminar flame front thus increasing its surface as shown in Fig. 7.4. Let σ l be the<br />

effective surface of combustion and σ the surface of access to the flame of the unburnt<br />

gases. Velocity u t is given in this case by expression<br />

u t<br />

u l<br />

= σ l<br />

σ . (7.13)<br />

The problem lies then in estimating the value for σ l , for which several mo<strong>de</strong>ls have<br />

been proposed.<br />

Shelkin assumes that σ l is formed by a system of cones whose base is proportional<br />

to the square of the scale of turbulence l 2 and whose height is proportional to<br />

distance v ′ l/u l travelled by a mass of gas with a diameter l due to turbulent oscillations<br />

during the time l/u l nee<strong>de</strong>d by the laminar flame to cross this mass. Here, σ l /σ<br />

is the ratio of the lateral to the base area of the cones, and from (7.13), it results<br />

√<br />

( )<br />

u t<br />

v<br />

′ 2<br />

= 1 + k , (7.14)<br />

u l u l<br />

where k is a numerical coefficient of the or<strong>de</strong>r of magnitu<strong>de</strong> unity. Therefore, u t is<br />

also in<strong>de</strong>pen<strong>de</strong>nt from the scale of turbulence. When turbulence is weak, Eq. (7.14)<br />

shows that its effect is of the second or<strong>de</strong>r, whilst for very intense turbulence Eq. (7.14)<br />

reduces to (7.11). Through a different analysis M. Tucker [15] obtains the following<br />

expression for the case of weak turbulence<br />

( )<br />

u t<br />

v<br />

′ 2<br />

= 1 + k(θ) , (7.15)<br />

u l u l<br />

which is valid if v ′ /u l ≪ l. In this formula, k(θ) is a coefficient <strong>de</strong>pending on ratio<br />

θ of temperature T f of the burnt gases to temperature T 0 of unburnt gases. It is seen<br />

that when v ′ /u l ≪ l. Eqs. (7.14) and (7.15) agree.<br />

Karlovitz [16] reasons as follows. Let ¯X =<br />

√ ¯X2 be the root mean square<br />

displacement of a particle due to turbulent oscillations. ¯X is a increasing function of


214 CHAPTER 7. TURBULENT FLAMES<br />

time counted from the point at which measurements were initiated. Let t = l/u l be<br />

the time taken by the laminar flame to cross a turbulent vortex. The total path travelled<br />

by the flame in this time is ¯X + l and when this expression is divi<strong>de</strong>d by t one obtains<br />

for u t<br />

That is to say<br />

u t = u l<br />

l<br />

The value for ¯X is given by Taylor’s formula<br />

d ¯X 2<br />

dt<br />

¯X + u l , (7.16)<br />

u t<br />

= 1 + ¯X<br />

u l l . (7.17)<br />

= 2v ′ 2<br />

∫ t<br />

0<br />

R t dt, (7.18)<br />

where R t is the correlation coefficient between the velocities of the same particle at<br />

two different instants. Karlovitz uses from R t the following expression<br />

where<br />

R t = e −tv′ l ′ , (7.19)<br />

∫ ∞<br />

l ′ = v ′ R t dt (7.20)<br />

0<br />

is the Lagrangian scale of turbulence. By substituting (7.19) into (7.18) and with<br />

t = l/v ′ it results for ¯X<br />

√<br />

¯X = l<br />

(<br />

2a v′<br />

1 − au [<br />

l<br />

u l v ′ 1 − exp<br />

(<br />

− 1 )])<br />

v ′<br />

, (7.21)<br />

a u l<br />

Here a is the ratio from the Lagrangian to the Eulerian scales of turbulence. Karlovitz<br />

assigns to a a value 1 while Scurlock and Grover [17] and Wohl [18] choose 1/2.<br />

Taking (7.21) into (7.17) one obtains for u t<br />

√ (<br />

u t<br />

= 1 + 2a v′<br />

1 − au [ (<br />

l<br />

u l u l v ′ 1 − exp − 1 a<br />

)])<br />

v ′<br />

. (7.22)<br />

u l<br />

For very intense turbulence (7.22) reduces to<br />

√<br />

u t<br />

≃ 1 + 2a v′<br />

, (7.23)<br />

u l u l<br />

which is valid provi<strong>de</strong>d that<br />

v ′<br />

≫ 1.<br />

u l<br />

Eq. (7.23) is in contradiction with (7.12) and for weak turbulence is reduces to<br />

u t<br />

u l<br />

≃ 1 + v′<br />

u l<br />

, (7.24)


7.2. TURBULENT COMBUSTION THEORIES 215<br />

valid for<br />

which also contradicts (7.16).<br />

v ′<br />

u l<br />

≪ 1,<br />

Wohl has shown that (7.22) can also be <strong>de</strong>duced from (7.13) through purely<br />

geometric consi<strong>de</strong>rations by assuming that the distortion of the flame front consists in<br />

the formation of a system of prisms instead of the cones proposed by Shelkin.<br />

Scurlock and Graver [17] calculate σ l by assuming with Shelkin that the laminar<br />

surface is formed by a set of cones whose base is proportional to the square l 2<br />

of the scale of turbulence and whose height is proportional to the root mean square<br />

displacement ¯X of the turbulent oscillations suffered by a flame element. Therefore,<br />

as in Eq. (7.14), u t is given by expression<br />

√<br />

u t<br />

= 1 + k ¯X 2<br />

u l l 2 . (7.25)<br />

The problem lies in computing ¯X. Its magnitu<strong>de</strong>, according to Scurlock, is<br />

governed by three different mechanisms:<br />

1) Turbulent diffusivity tends to increase in<strong>de</strong>finitely the value of ¯X as time elapses.<br />

Figure 7.5 shows two consecutive positions of the flame front in Scurlock’s<br />

mo<strong>de</strong>l after the instant at which the flame was plane.<br />

Initial flat flame<br />

Flame after passage<br />

of short time<br />

Flame after passage<br />

of longer time<br />

Figure 7.5: Schematic diagram showing wrinkling with passage of time of an initially flat<br />

flame element exposed to turbulence.<br />

2) Laminar combustion tends to absorb oscillations reducing the value of ¯X. Karlovitz<br />

and his collaborators [19] were the first to acknowledge this fact which is<br />

represented in Fig. 7.6<br />

3) The flame generates turbulence 2 and this tends to increase the value of ¯X.<br />

2 See §4.


216 CHAPTER 7. TURBULENT FLAMES<br />

ARBITRARY WAVY FLAME FRONT<br />

BURNED GAS<br />

FRESH GAS<br />

FLAME FRONT AFTER ∆t TIME<br />

Figure 7.6: The effect of laminar flame propagation on the evolution of a turbulent flame<br />

front.<br />

Since the locus of an element of the flame at two different instants are related<br />

to two different particles of the mixture, when calculating the influence of turbulent<br />

diffusivity on ¯X, a combined time-space coefficient of correlation R tx should be used<br />

substituting coefficient time-correlation R t in Taylor’s formula [17].<br />

The influence of the laminar propagation of the flame in the value of ¯X may<br />

be obtained through purely geometric consi<strong>de</strong>rations. Finally, the influence of the<br />

turbulence generated by the flame is taken into consi<strong>de</strong>ration by substituting into the<br />

expression of ¯X, the intensity v ′ of turbulence of the unburnt gases flow by the resultant<br />

of it, plus the intensity of the turbulence originated by the flame.<br />

The combination of these three effects gives a differential equation which in<br />

turn supplies the law of variation for ¯X as a function of time t during which the<br />

particle has been exposed to turbulent oscillations. This differential equation replaces<br />

Taylor’s equation (7.18) for this case. Through integration of this equation the value<br />

for ¯X is obtained, and when taken into Eq. (7.22) it gives a formula for the propagation<br />

velocity of the turbulent flame.<br />

Now we must <strong>de</strong>termine the time t that each element of the flame has been<br />

exposed to the action of turbulence. Scurlock consi<strong>de</strong>rs only flames inclined respect<br />

to the flow such as the one obtained with a flame-hol<strong>de</strong>r or from the ring of a bunsen<br />

burner. He i<strong>de</strong>ntifies t with the time taken by a gas particle to cross the flame front<br />

from the hol<strong>de</strong>rs section to the point un<strong>de</strong>r consi<strong>de</strong>ration. Let v t by the component<br />

tangential to the flame front of the velocity of the unburnt gases and s the distance to<br />

the flame-hol<strong>de</strong>r measured along the flame front. The following expression is obtained<br />

for t<br />

t =<br />

∫ s<br />

0<br />

ds<br />

v t<br />

. (7.26)


7.2. TURBULENT COMBUSTION THEORIES 217<br />

The selection of t, which is arbitrary to a certain extent, is justified by the<br />

theoretical and experimental studies carried out by Markstein [20] on the propagation<br />

of disturbance along inclined flame fronts. These studies show that such disturbances<br />

propagate with the tangential velocity of the flow as they become amplified.<br />

BURNED<br />

GASES<br />

BURNED<br />

GASES<br />

BURNED<br />

GASES<br />

MEAN POSITION<br />

OF FLAME<br />

MEAN POSITION<br />

OF FLAME<br />

MEAN POSITION<br />

OF FLAME<br />

β<br />

INSTANTANEOUS<br />

FLAME FRONT<br />

INSTANTANEOUS<br />

FLAME FRONT<br />

INSTANTANEOUS<br />

FLAME FRONT<br />

UNBURNED<br />

GASES<br />

UNBURNED<br />

GASES<br />

UNBURNED<br />

GASES<br />

U U U<br />

BURNER<br />

RIM<br />

BURNER<br />

RIM<br />

BURNER<br />

RIM<br />

Figure 7.7: Cross section of stabilized unconfined Bunsen flames constructed on basis of<br />

theory to <strong>de</strong>monstrate un<strong>de</strong>r typical conditions the predicted individual effects<br />

of eddy diffusion, flame propagation and flame generated turbulence.<br />

Figure 7.7 taken from the work by Scurlock and Grover, shows the influence of<br />

the above mentioned facts for the case of an open flame stabilized in a bunsen burner.<br />

The following values were adopted for computations: diameter of the burner = 5 cm,<br />

velocity on the gases in the burner= 200 cm/s, u 1 = 10 cm/s, v ′ = 10 cm/s, intensity<br />

of turbulence 5%, scale of turbulence= 0.25 cm, and width of the flame= 2 ¯X.


218 CHAPTER 7. TURBULENT FLAMES<br />

7.3 Turbulence generated by the flame<br />

Markstein [21] has shown that the disturbances originated at a certain point of an incline<br />

flame propagate and amplify along the same and eventually generate turbulence.<br />

In their studies of turbulent flames stabilized in tubes, Willians, Hottel and Scurlock<br />

[5] recognized the existence of turbulence originated by the flame. They consi<strong>de</strong>red<br />

this turbulence as a consequence of the strong gradients of velocity which originate<br />

through the flame in this case. 3<br />

Scurlock and Grove have given a formula for the<br />

maximum intensity v ′ of the possible turbulence. Such formula was <strong>de</strong>duced from elementary<br />

consi<strong>de</strong>rations on mass and momentum conservation across the flame before<br />

and after the mixing of burnt gases. This formula is<br />

√<br />

( )<br />

v m ′ K m ρ u<br />

= (vu 3 ρ 2 − u 2 l ) ρu<br />

− 1 . (7.27)<br />

b ρ b<br />

Here, K m = ∆p ∆p<br />

−1, where is the relation between pressure jumps across<br />

∆p<br />

′<br />

∆p<br />

′<br />

the flames before and after the turbulent mixing, v u is the velocity of the unburnt gases<br />

normal to the mean flame front and ρ u /ρ b is the ratio between <strong>de</strong>nsities of unburnt and<br />

burnt gases. Formula (7.27) gives a maximum limit to the turbulence that could be<br />

originated in the flame but does not allow the computation of the possible turbulence<br />

for each case. This arises the problem of introducing an additional un<strong>de</strong>termined<br />

element in the theory which adds difficulties to the comparison between theoretical<br />

and experimental results. In practice v ′ m can be consi<strong>de</strong>rably larger than the turbulence<br />

of the inci<strong>de</strong>nt stream.<br />

Karlovitz [16] when comparing the values for u t predicted by his theory with<br />

those obtained from experimenting with flames stabilized in bunsen burner verified<br />

that experimental velocities were much larger than those calculated and he consi<strong>de</strong>red<br />

this discrepancy to be due to the influence of the turbulence originated by the flame.<br />

Karlovitz explains the production of turbulence as follows: through the laminar front<br />

an increase in velocity takes place of the or<strong>de</strong>r of magnitu<strong>de</strong> of ρ u<br />

ρ b<br />

− 1. In a laminar<br />

flame this increase has a constant direction but in a turbulent flame it oscillates since<br />

it must be normal to the temporary position of the flame front at all instants which is<br />

variable. The statistic oscillations of this increase are the source of the turbulence.<br />

Through a not too legitimate calculation Karlovitz <strong>de</strong>duces the following expression<br />

for the maximum intensity of turbulence produced by the flame<br />

v m ′ = √ 1 ( )<br />

ρu<br />

− 1 u l . (7.28)<br />

3 ρ b<br />

3 See chapter 10.


7.4. COMPARISON WITH EXPERIMENTAL RESULTS 219<br />

It is also impossible here to estimate which fraction of the intensity given by (7.28)<br />

will actually turn into turbulence in each case.<br />

7.4 Comparison with experimental results<br />

As aforesaid the available experimental evi<strong>de</strong>nce is not sufficient to establish either<br />

one of the proposed theories in a <strong>de</strong>finite way precluding the others. We owe the most<br />

complete set of experiments to Bollinger and Williams [2] and to Wohl and Shores [7]<br />

referred to herein. Some of the results obtained show laws of variation of u t which<br />

approach those predicted by Karlovitz and by Scurlock and Grover. However, their<br />

effects appear which are either not predicted by theory or in contradiction with it. An<br />

example of effects not predicted is the influence of the mixture’s composition, since<br />

it is verified that the effect of turbulence is not the same for rich mixtures than for<br />

poor ones, even when operating un<strong>de</strong>r i<strong>de</strong>ntical conditions and with the same laminar<br />

propagation velocity. This is especially true for certain combustibles (like butane) and<br />

this fact cannot be explained with any of the proposed theories. Wohl believes this<br />

effect to be due to the different laminar stability in poor and rich mixtures. An example<br />

of effects appearing in contradiction with theory is the influence of the scale of<br />

turbulence, which after Scurlock’s predictions, should be consi<strong>de</strong>rable and which according<br />

to experimental results is very small. Consequently it is necessary to increase<br />

experimental evi<strong>de</strong>nce before final <strong>de</strong>cision may be reached.<br />

References<br />

[1] Mallard, F. and Le Chatelier, H.: Ann. <strong>de</strong> Mines. Vol. 8, Sev. 4, 1884, p. 274.<br />

[2] Damköhler, G.: The Effect of Turbulence on the Flame Velocity in Gas Mixtures.<br />

NACA Tech. Mem. No. 1112, 1947.<br />

[3] Bollinger L. M. and Williams, D. T.: Effect of Reynolds Number in the Turbulent-<br />

Flow Range on Flame Speeds of Bunsen-Burner Flames. NACA Tech. Note<br />

No. 1707, Sept. 1948.<br />

[4] Scurlock, A. C.: Flame Stabilization and Propagation in High Velocity Gas<br />

Streams. Meteor Report No. 19, July 1948.<br />

[5] Williams, G. C., Hottel, H. C. and Scurlock, A. C.: Flame Stabilization and<br />

Propagation in High Velocity Gas Streams. Third Symposium (International)<br />

on Combustion, Williams and Wilkins Co., Baltimore, 1949, pp. 21-40.<br />

[6] Karlovitz, B. Denniston, D. W. and Wells, F. E.: Investigation of Turbulent<br />

Flames. Journal of Chemical Physics, Vol. 19, No. 5, May 1951.


220 CHAPTER 7. TURBULENT FLAMES<br />

[7] Wohl, K., Shore, L., von Rosemberg, H. and Weil, C. M.: The Burning Velocity<br />

of Turbulent Flames. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953.<br />

[8] Leason, D. B.: Turbulence and Flame Propagation in Premixed Gases. Fuel,<br />

Vol. XXX, No. 10, Oct. 1951.<br />

[9] Bowditch, F.: Some Effects of Turbulence on Combustion. Fourth Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore, 1953,<br />

pp. 620-635.<br />

[10] Wohl, K. and Shore, L.: Experiments with Butane-Air and Methane-Air Flames.<br />

Industrial and Engineering Chemistry, April 1955.<br />

[11] Mickelsein, W. R. and Ernstein, N. E.: Propagation of a Free Flame in a Turbulent<br />

Gas Stream. NACA Tech. Note No. 3456, 1955.<br />

[12] von Kármán, Th. and Marble, F.: Combustion in Turbulent Flames. Fourth<br />

Symposium (International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, p. 923.<br />

[13] Sommerfield, M., Reiter, S. H., Kebely, V. and Mascolo, R.W.: The Structure<br />

and Propagation Mechanism of Turbulent Flames in High Speed Flow. Jet<br />

Propulsion, August 1955.<br />

[14] Shelkin, F. I.: On Combustion in a Turbulent Flow. NACA Tech. Mem. No.<br />

1110, Febr. 1947.<br />

[15] Tucker, E.: Interaction of a Free Flame Front with a Turbulence Field. NACA<br />

Tech. Note No. 3407, 1955.<br />

[16] Karlovitz, B.: A Turbulent Flame Theory Derived From Experiments. Selected<br />

Combustion Problems, Vol.I, AGARD, 1954, pp. 248-262.<br />

[17] Scurlock, A. C. and Grover, J. F.: Experimental Studies on Turbulent Flames.<br />

Selected Combustion Problems, Vol.I, AGARD, 1954, pp. 215-247.<br />

[18] Wohl, K.: Burning Velocity of Unconfined Turbulent Flames Industrial Engineering<br />

Chemistry, April 1955, p. 825.<br />

[19] Karlovitz, B.: Open Turbulent Flames. Fourth Symposium (International) on<br />

Combustion, Williams and Wilkins Co., Baltimore, 1953, pp. 60-67.<br />

[20] Markstein, G. H.: Discussion on Turbulent Flames. Selected Combustion Problems,<br />

Vol.I, AGARD, 1954, pp. 263-265.<br />

[21] Markstein, G. H.: Interaction of Flow Propagation and Flame Disturbances.<br />

Third Symposium (International) on Combustion, Williams and Wilkins Co.,<br />

Baltimore, 1949, pp. 162-167.


Chapter 8<br />

Ignition, flammability and<br />

quenching<br />

8.1 Introduction<br />

In the preceding chapter we have analyzed the properties of a plane wave propagating<br />

in a stationary regime, through an in<strong>de</strong>finite combustible mixture whose state and<br />

composition are uniform. Although the problem is far from being completely solved,<br />

the essential characteristic of those waves are well un<strong>de</strong>rstood, and a precise formulation<br />

of the same is available as well as several methods to solve the resulting equations,<br />

including some practical solutions, which may be consi<strong>de</strong>red as fairly acceptable when<br />

compared to experimental results.<br />

However, it must be said that the stage of knowledge is not as favorable for<br />

other fundamental questions relative to flames such as: the problem of ignition of<br />

a combustible mixture; the possibility of a flame to propagate through a mixture of<br />

given state and composition; and the influence of the vicinity of the walls on the<br />

characteristics of the flame. Even when a great amount of attention has been applied<br />

during recent years to the study of these problems, the progress achieved, specially<br />

from the theoretical stand-point, has been consi<strong>de</strong>rably less than for the case of an<br />

in<strong>de</strong>finite plane wave, to the extreme that the causes of some of the phenomena are<br />

still unknown as well as their governing laws.<br />

The present chapter contains a brief <strong>de</strong>scription of each of the above mentioned<br />

problems, accompanied by a selected bibliography which will enable a more <strong>de</strong>tailed<br />

study on each one.<br />

221


222 CHAPTER 8. IGNITION, FLAMMABILITY AND QUENCHING<br />

8.2 Ignition<br />

Consi<strong>de</strong>ring a combustible mixture capable of propagating a flame, in or<strong>de</strong>r for this<br />

to happen it is necessary to supply it with a certain amount of energy, located within a<br />

small volume and reduced interval of time, to initiate the wave. The study of this process,<br />

both theoretically and experimentally, has been the subject of numerous reports.<br />

Among the various sources of energy that may be utilized, the most a<strong>de</strong>quate is the<br />

electrical spark because it is capable of steering great amounts of energy in reduced<br />

spaces and times. For this reason it is the most wi<strong>de</strong>ly used and has been the subject<br />

of more systematic and complete experimental studies, of which a <strong>de</strong>scription may be<br />

found in Refs. [1] and [2].<br />

The fundamental problem of ignition lies on <strong>de</strong>termining the minimum energy<br />

required to ignite a mixture of given composition, at known pressure and temperature.<br />

The results of experiments have pointed out that to each mixture it corresponds a well<br />

<strong>de</strong>fined minimum ignition energy, which increases very rapidly with the <strong>de</strong>crease in<br />

pressure. Such energy results to be in<strong>de</strong>pen<strong>de</strong>nt from the distance between electro<strong>de</strong>s,<br />

provi<strong>de</strong>d this distance exceeds a given minimum value un<strong>de</strong>r which it grows very<br />

rapidly and finally the flame cannot propagate.<br />

Several theories have been attempted for the calculation of this minimum energy<br />

with fair success. The present state of knowledge on this problem may be consi<strong>de</strong>red<br />

as comparable to the situation existing 20 years ago respect to theories on<br />

flames.<br />

All the theories <strong>de</strong>veloped are based on the following mo<strong>de</strong>l. Let us consi<strong>de</strong>r<br />

a small volume V of gas to which energy H is instantaneously communicated, rising<br />

its temperature from T 0 to T f . Starting from this instant, the gas contained in V<br />

tends to cool, by thermal conductivity, heating the gas layers surrounding V . On the<br />

other hand, the chemical reaction which produces in the heated mass releases heat,<br />

which tends to compensate the cooling effect. From the balance between this two<br />

phenomena, it will <strong>de</strong>pend that the flame may progress towards a wave of the type<br />

<strong>de</strong>scribed in Chap. 6, or, to the contrary, that it extinguishes.<br />

The available theories differ in the <strong>de</strong>velopment of this i<strong>de</strong>a and on the conditions<br />

applied to <strong>de</strong>termine the minimum volume V and the necessary energy H. For<br />

instance, Lewis and von Elbe [3] assume that volume V is the one corresponding to<br />

the quenching 1 distance and that the energy is <strong>de</strong>termined by the excess enthalpy of<br />

the flame. The validity of this concept does not appear clearly justified, although the<br />

1 See §4.


8.2. IGNITION 223<br />

comparison between the minimum values of the calculated energy and those measured<br />

experimentally show a surprising agreement [3].<br />

Fenn [4] performs an elementary computation which is actually more of a dimensionless<br />

analysis of the problem. He reasons as follows: be r the radius of volume<br />

V and let us assume that combustion takes place through a second-or<strong>de</strong>r reaction of<br />

the form<br />

w = Aρ 2 Y (1 − Y )e −E/RT . (8.1)<br />

Then, the heat released per unit time in volume V due to the combustion of the mixture<br />

will be<br />

Q = 4 3 πr3 qAρ 2 Y (1 − Y )e −E/RT f , (8.2)<br />

where T f is the temperature of gases, and q the heat of reaction.<br />

In turn, the heat lost by conductivity through the surface limiting V , will be<br />

Q ′ = 4πr2 λ(T f − T 0 )<br />

, (8.3)<br />

cr<br />

where cr is the thickness of the combustion wave, which separates hot from cold<br />

gases.<br />

The condition of minimum ignition energy for propagation, will be<br />

Q = Q ′ , (8.4)<br />

since if it is Q < Q ′ the mass cools and the flame extinguishes.<br />

radius of V<br />

By taking (8.2) and (8.3) into (8.4) we obtain the following expression for the<br />

√<br />

3λ(T f − T 0 )<br />

r =<br />

cqAρ 2 Y (1 − Y )e −E/RT . (8.5)<br />

f<br />

Finally, the minimum energy H is the one nee<strong>de</strong>d to rise the mass contained in V from<br />

temperature T 0 to T f which is<br />

where for r one must use expression (8.5).<br />

H = 4 3 πr3 c p ρ(T f − T 0 ), (8.6)<br />

Recently Swott [5] has exten<strong>de</strong>d this theory to the study of the problem of<br />

the ignition of flowing mixtures including as well the effects of turbulence. Other<br />

studies on the matter, specially directed to the ignition in combustion chambers, with<br />

extensive bibliography, will be found in Ref. [6].<br />

The preceding theories are of the thermal type, since they ignore the influence<br />

of diffusion, which could be important, specially the diffusion of radicals.


224 CHAPTER 8. IGNITION, FLAMMABILITY AND QUENCHING<br />

A correct stating of the problem would require first an a<strong>de</strong>quate formulation of<br />

the same, in an analogous way to the one used in Chap. 6 for the stationary wave. In<br />

the second place it would be necessary to perform an analysis of the type of solutions<br />

for the system of equations obtained un<strong>de</strong>r the set of initial and boundary conditions<br />

corresponding to the mo<strong>de</strong>l previously <strong>de</strong>scribed.<br />

If we consi<strong>de</strong>r the more simple case of an in<strong>de</strong>finite plane wave corresponding<br />

to a one-dimensional problem, in which case the initial energy must be stored between<br />

two parallel layers, it may be easily verified that the system of equations of Chap. 6,<br />

corresponding to a stationary wave, keeping the same notation, must be substituted by<br />

the following:<br />

a) Continuity equation.<br />

b) Energy equation.<br />

c) Diffusion equation.<br />

∂ρ<br />

∂t + ∂(ρv) = 0. (8.7)<br />

∂x<br />

( ∂T<br />

ρc p<br />

∂t + v ∂T )<br />

= ∂ (<br />

λ ∂T )<br />

+ qw. (8.8)<br />

∂x ∂x ∂x<br />

ρ<br />

( ∂Y<br />

∂t + v ∂Y )<br />

= ∂ (<br />

ρD ∂Y )<br />

+ w. (8.9)<br />

∂x ∂x ∂x<br />

Thus we obtain a system of three equations for the three unknowns T , v and Y .<br />

Since the process takes place at constant pressure, ρ is already <strong>de</strong>termined, as function<br />

of T .<br />

With reference to initial and boundary conditions corresponding to the mo<strong>de</strong>l<br />

un<strong>de</strong>r consi<strong>de</strong>ration, if 2d is the width of the heated slab and the origin of coordinates<br />

is fixed at the central point of the slab, we will have:<br />

a) Initial conditions (t = 0).<br />

0 < x < d : T = T f ,<br />

d < x < ∞ : T = T 0 ,<br />

(8.10)<br />

b) Boundary conditions (by symmetry).<br />

0 < x < ∞ : v = Y = 0.<br />

x = 0 :<br />

∂T<br />

∂x = v = ∂Y<br />

∂x = 0.<br />

Furthermore an additional condition must be introduced expressing that w must<br />

be zero for T = T 0 , as it was done for the case of a stationary wave.


8.3. FLAMMABILITY LIMITS 225<br />

The preceding system of equations allows an easy simplification, by using the<br />

stream function ψ, frequently applied to the problem of Fluid Mechanics.<br />

The said function is <strong>de</strong>fined by the following conditions<br />

ρ = ∂ψ<br />

∂x ,<br />

ρv = −∂ψ<br />

∂t , (8.11)<br />

through which Eq. (8.7) is i<strong>de</strong>ntically satisfied. When taking Eq. (8.11) into Eqs. (8.8)<br />

and (8.9), this system reduces to the following<br />

∂T<br />

∂t = ∂ ( ρλ ∂T<br />

∂ψ c p ∂ψ<br />

∂Y<br />

∂t = ∂ ( ρ 2 D<br />

∂ψ<br />

Thus the variable v has been eliminated.<br />

c p<br />

∂Y<br />

∂ψ<br />

)<br />

+ q c p<br />

w<br />

ρ , (8.12)<br />

)<br />

+ w ρ . (8.13)<br />

The integration of the above system has not yet been performed. However,<br />

Spalding [7] has obtained graphical solutions of Eq. (8.12) disregarding diffusion.<br />

These solutions show the existence of a critical value d cr for d un<strong>de</strong>r which the wave<br />

extinguishes, while for d > d cr it propagates in<strong>de</strong>finitely.<br />

There are some experimental observations on the evolution followed by an<br />

ignited mass before the flame establishes [8].<br />

8.3 Flammability limits<br />

The theoretical studies performed so far have shown that the combustible mixture is<br />

always capable of maintaining a flame. Experimentally, however, it has been verified<br />

that it is not so, since there are numerous mixtures which, in practise, are not able to<br />

propagate a flame, even when theory predicts otherwise.<br />

In fact, experiments disclose that when analyzing the behavior of a combustible<br />

mixture un<strong>de</strong>r pressure and temperature constant, but changing its composition, for instance<br />

a mixture of fuel and air, the flame velocity is maximum for a <strong>de</strong>finite value<br />

of the composition which is normally very close to the stoichiometric one, and it <strong>de</strong>creases<br />

for mixtures either rich or loan, up to values known as inflammability limits of<br />

the mixture, beyond which it cannot propagate a flame. Moreover the flame velocity<br />

does not vanish at the inflammability limits but it generally reaches values of a few<br />

centimeters per second. Inflammability limits <strong>de</strong>pend also on pressure and they are<br />

difficult to <strong>de</strong>termine experimentally because it is necessary to operate un<strong>de</strong>r marginal<br />

conditions, on which the experimental techniques might have an influence. Tables


226 CHAPTER 8. IGNITION, FLAMMABILITY AND QUENCHING<br />

and graphics may be found in Ref. [1], giving the inflammability limits of several<br />

mixtures of hydrocarbons with air, un<strong>de</strong>r different pressures. Ref. [3] supplies an<br />

abridged review of the problem. The Proceedings of the Fourth International Symposium<br />

on Combustion inclu<strong>de</strong>, as well, several papers on the subject, Coward and<br />

Jones, [9], reviewed the state of knowledge up to 1952, in a report including an extensive<br />

experimental and bibliographic material. Two recent, very interesting, reviews<br />

are those by Egorton [10] and Linnot-Simpson [11].<br />

At present, the situation of the problem is such, that the causes for the existence<br />

of inflammability limits are still unknown and, even more, it is questioned<br />

whether they actually exit as quantities governed only by the state and composition<br />

of the mixture or, to the contrary, they <strong>de</strong>pend upon the characteristics of the experimental<br />

<strong>de</strong>vice used. Lewis and von Elbe [3] suppose that the existence of such limits<br />

may be due to the internal stability of the combustion wave, which then would be unstable<br />

beyond the limits. Several attempts to study the problem from this stand-point<br />

have been ma<strong>de</strong>, [12] and [13], in addition to the one by Lewis and von Elbe. Of<br />

all, the most satisfactory is own to Werner and Rosen [14], who analyze the internal<br />

stability of the wave with respect to small disturbances of temperature by means of a<br />

linearization of the system of Eqs. (8.7), (8.8) and (8.9) around their stationary solution.<br />

These authors solved the resulting system after introducing several simplifying<br />

assumptions regarding the disturbances of the composition of the mixture and of the<br />

mass flux. They reach the conclusion that the stability of the wave <strong>de</strong>pends on the<br />

behavior of the disturbances of Y with respect to those of T . Although this study<br />

must be consi<strong>de</strong>red as uncompleted it suggests the possibility for some waves to be<br />

internally unstable, which could explain the existence of inflammability limits. An<br />

application of this theory to the ozone-oxygen flame is given in Ref. [15], the result<br />

of which shows that the flame is stable for ozone rich mixtures and unstable for loan<br />

ones. However, the experiments carried out by Streng and Grosse [16] disclose that<br />

the flame is stable even for loan mixtures.<br />

Recently Spalding [17] has analyzed the possible influence of heat losses of<br />

the flame on its behavior. He reaches the conclusion that there exist two different<br />

flame velocities, and that the smaller is unstable. When heat losses augment the velocities<br />

approach one another and finally coinci<strong>de</strong>. Spalding i<strong>de</strong>ntifies the coinci<strong>de</strong>nce<br />

of velocities with the inflammability limits. Although this work does represent an important<br />

contribution, yet an unpublished work, being carried out at the Combustion<br />

Laboratory of the INTA, reveals that his conclusions are not free from criticism.


8.4. QUENCHING 227<br />

8.4 Quenching<br />

So far studies have <strong>de</strong>alt with flames propagating through unlimited mixtures, not<br />

taking into account the influence of the proximity of walls,<br />

A set of phenomena exist, of great importance, relative to the behavior of<br />

flames close to walls which have been the subject of such attention in recent years.<br />

Among those phenomena are quenching, blow-off and flash-back.<br />

Experiments prove that when the diameter of a tube full of a combustible gas<br />

is sufficiently small, it is impossible for a flame to propagate through it. This phenomenon,<br />

which has been recognized for a long time and which has been applied to<br />

the prevention of explosions, is known as “quenching”.<br />

The way in which the walls of the tube act to prevent flame propagation is<br />

not well un<strong>de</strong>rstood at present, The most direct explanation is to attach the effect to<br />

the cooling of the flame due to the proximity of the walls. However, it is though<br />

that chemical action of the wall when acting as chain breaker may also be significant.<br />

Analogously to what happened for some time with flame theory, if one consi<strong>de</strong>rs either<br />

cooling or diffusion as the only acting effect, a “thermal” or “diffusive” theory will<br />

be obtained. Examples of thermal theories are those by Lewis-von Elbe [1] and by<br />

Kármán-Millán [18], and on diffusion theories the one by Simon-Bellos [19].<br />

The difference between this case and the unlimited flame theory lies on the fact<br />

that both the formulation of the equation and boundary conditions and their solution<br />

are far more difficult. Hence, the solutions may only be attempted through drastic<br />

simplifications.<br />

A great number of measurements have been carried out with the purpose of<br />

<strong>de</strong>termining the quenching distance in ducts of different cross-sections and the influence<br />

on it, of the mixture. For an additional study of this question as well as on the<br />

phenomena of blow-off and flash-back we refer the rea<strong>de</strong>r to Refs. [20] through [23].<br />

In particular references [20] and [23] inclu<strong>de</strong> an extensive bibliography.<br />

References<br />

[1] Lewis, B. and von Elbe, G.: Combustion, Flames and Explosions of Gases.<br />

Aca<strong>de</strong>mic Press, New York, 1951.<br />

[2] Lewis, B, Pease, R. N. and Taylor, H,S.: Combustion Processes. Vol. II of High<br />

Speed Aerodynamics and Jet Propulsion, Princeton University Press, 1956.


228 CHAPTER 8. IGNITION, FLAMMABILITY AND QUENCHING<br />

[3] Lewis, B. and von Elbe, G.: Fundamental Principles of Flammability and Ignition.<br />

Selected Combustion Problems, Vol. II, AGARD, 1956, pp. 63-72.<br />

[4] Fenn, J. B.: Lean Inflammability Limit and Minimum Spark Ignition Energy.<br />

Industrial and Engineering Chemistry. Vol. Dec., 1951, pp. 2865-68.<br />

[5] Swott, C. C.: Spark Ignition of Flowing Gases. IV. Theory of Ignition in Nonturbulent<br />

and Turbulent Flow Using Long Duration Discharges. NACA Research<br />

Memorandum No. R M E54F29a, 1954.<br />

[6] Wigs. L. D.: The Ignition of Flowing Gases. Selected Combustion Problems,<br />

Vol. II, AGARD, 1956, pp. 73-82.<br />

[7] Spalding, D. B.: Some Fundamentals of Combustion. Butterworths Scientific<br />

Publications, 1955.<br />

[8] Olsen, E. L., Gayhart, E. L. and Edmoneon, R. B.: Propagation of Incipient<br />

Spark–Ignited Flames in Hydrogen–Air and Propane–Air Mixtures. Fourth<br />

Symposium (International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, pp. 144-148.<br />

[9] Coward, H. F. and Jons, G. W.: Limits of Flammability of Gases and Vapors.<br />

U.S. Bureau of Mines. Bull., 503, 1952.<br />

[10] Egerton, A. C.: Limits of Inflammability. Fourth Symposium (International) on<br />

Combustion, Williams and Wilkins Co., Baltimore, 1953, pp. 4-13.<br />

[11] Linnet, J. W. and Simpson, O. J. S.: Limits of Inflammability. Sixth Symposium<br />

(International) on Combustion, Reinhold Publishing Corp., New York, 1957.<br />

[12] Richardson, J. M.: The Existence and Stability of Simple One–Dimensional,<br />

Steady–State Combustion Waves. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953, pp. 182-189.<br />

[13] Layzer, D.: Journal of Chemical Physics, Vol. 22, 1954, p. 222.<br />

[14] Wehner, J. F. and Rosen, J. B.: Temperature Stability of the Laminar Combustion<br />

Wave. Combustion and Flame, Sept. 1957, pp. 339-345 .<br />

[15] Rosen, J. B.: Stability of the Ozone Flame Propagation. Sixth Symposium<br />

(International) on Combustion, Reinhold Publishing Corp., New York, 1957.<br />

[16] Streng, A. G. and Grosse, A. V.: The Ozone to Oxygen Flame. Sixth Symposium<br />

(International) on Combustion, Reinhold Publishing Corp., New York, 1957.<br />

[17] Spalding, D. B.: A Theory of Inflammability Units and Flame–Quenching .<br />

Proc. Roy, Soc. London, Ser. A, Vol. 240, No. 1220, 1957, pp. 83-100.<br />

[18] von Kármán, Th. and Millán, G.: Thermal Theory of Laminar Flame Front near<br />

a Cold Wall. Fourth Symposium (International) on Combustion, Williams and<br />

Wilkins Co., Baltimore, 1953, pp. 173-177.


8.4. QUENCHING 229<br />

[19] Simon, D. M. and Belles, F. E.: An Active Particle Diffusion Theory of Flame<br />

Quenching for Laminar Flames. NACA Research Memorandum No. RM<br />

E51L18, 1952.<br />

[20] Wohl, K.: Quenching Flash–Back, Blow–Off. Theory and Experiment. Fourth<br />

Symposium (International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, pp. 68-89.<br />

[21] Berlad, A. L. and Potter, A. B.: Effect of Channel Geometry on the Quenching<br />

of Laminar Flames. NACA Research Memorandum RM E54C05, 1954.<br />

[22] Berlad, A. L.: Flame Quenching by a Variable Width Rectangular–Channel<br />

Burner as a Function of Pressure for Various Propane–Oxygen–Nitrogen Mixtures.<br />

Journal of Physical Chemistry, Vol. 58, 1954, pp. 1023-1026.<br />

[23] Massey, B. S. and Lindley, B. O.: Flame Quenching. Journal of the Royal<br />

Aeronautical Society, Jan. 1958, pp. 32-42.


230 CHAPTER 8. IGNITION, FLAMMABILITY AND QUENCHING


Chapter 9<br />

Flows with combustion waves<br />

9.1 Introduction<br />

As previously seen in chapter 6, in a combustible mixture at ambient of higher pressure,<br />

the thickness of the flame is of the or<strong>de</strong>r of a fraction of millimeter. This length<br />

is generally “small” when compared to those of interest in aerothermodynamic problems.<br />

Therefore, when studying these problems it is justified to assume that the flame<br />

thickness is zero. In such a case the flame may be consi<strong>de</strong>red as a surface of discontinuity<br />

of the pressure, temperature and velocity. The same i<strong>de</strong>a is applied, for example,<br />

in the study of gas dynamic problems involving shock waves. The problem has been<br />

studied from this view point mainly by Emmons and his co-operators [1], [2], [3], [4].<br />

In this chapter the general lines of Emmons’s work are followed. First the conditions<br />

that must be satisfied across the flame are established and then some general properties<br />

<strong>de</strong>duced. In the chapter that follows, the method is applied to the study of the<br />

aerothermodynamic field originated by a flame stabilized in a combustion chamber.<br />

This problem has been studied by Scurlock, Tsien, and Fabri-Siestrunck-Fouré.<br />

9.2 Conditions that must be satisfied by the jump across<br />

a flame front.<br />

As aforesaid, the relations <strong>de</strong>duced herein are applicable only if the thickness of the<br />

flame front is small when compared to a characteristic length of the phenomenon un<strong>de</strong>r<br />

study. 1 The flame thickness must also be small compared to the radius of curvature<br />

1 For instance, this does not occur in rarefied mixtures where the thickness of the flame can be large.<br />

231


232 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

of the flame. In fact, if the radius of curvature and the flame thickness are of the<br />

same or<strong>de</strong>r of magnitu<strong>de</strong> the flame speed <strong>de</strong>pends not only on the thermodynamic<br />

conditions of the unburnt gases, but also on the shape of the flame in the neighborhood<br />

of the point un<strong>de</strong>r consi<strong>de</strong>ration. Such is the case in the flame tip of a Bunsen burner.<br />

Here, the flame speed is increased due to the favorable effect of the curvature on heat<br />

transfer and diffusion of the active particles towards the unburnt gases.<br />

V n1<br />

V t1<br />

p , T ,<br />

ρ<br />

1 1 1<br />

V t2<br />

V<br />

n2<br />

p , T ,<br />

2 2<br />

ρ<br />

2<br />

Figure 9.1: Schematic diagram of a surface element of a flame.<br />

Thus, let it be a flame front of small thickness and small curvature as indicated<br />

in Fig. 9.1, and let us consi<strong>de</strong>r a surface element of this front. By applying the continuity,<br />

momentum, and energy equations to the gases at both si<strong>de</strong>s of this element, as<br />

it is done in the study of the invariants across a shock wave 2 , the following conditions<br />

are obtained which relate the state of the unburnt and burnt gases at both si<strong>de</strong>s of the<br />

elements:<br />

1) Continuity equation.<br />

ρ 1 v n1 = ρ 2 v n2 . (9.1)<br />

2) Momentum equation.<br />

a) Normal component.<br />

ρ 1 v 2 n1 + p 1 = ρ 2 v 2 n2 + p 2 . (9.2)<br />

b) Tangential component.<br />

v t1 = v t2 . (9.3)<br />

3) Energy equation.<br />

1<br />

2 v2 1 + h 1 = 1 2 v2 2 + h 2 . (9.4)<br />

Subscripts 1 and 2 refer to conditions before and after the flame, respectively,<br />

v is the velocity of the gases relative to the flame front, assumed stationary, and v n<br />

2 See chapter 6.


9.2. CONDITIONS THAT MUST BE SATISFIED BY THE JUMP ACROSS A FLAME FRONT. 233<br />

and v t are the normal and tangential components of this velocity. Moreover ρ and p<br />

are, as usual, the <strong>de</strong>nsity and pressure of the gases, and h is the total specific enthalpy,<br />

that is, (as seen in chapter 1), the thermal enthalpy h T plus the formation enthalpy h f ,<br />

h = h T + h f , (9.5)<br />

where h T is expressed, in terms of the heat capacity c p at constant pressure and the<br />

absolute temperature T , as follows<br />

h T =<br />

∫ T<br />

0<br />

which, for the special case c p = constant reduces to<br />

c p dT, (9.6)<br />

h T = c p T. (9.7)<br />

At each si<strong>de</strong> of the flame front, the pressure, <strong>de</strong>nsity and temperature are related<br />

by the state equation, which is assumed to be that of the perfect gases, that is<br />

p 1<br />

ρ 1<br />

= R 1 T 1<br />

and<br />

p 2<br />

ρ 2<br />

= R 2 T 2 , (9.8)<br />

where<br />

R 1 = R M u<br />

and R 2 = R M b<br />

(9.9)<br />

are the specific constants of the unburnt and burnt gases, respectively, and M u and M b<br />

their respective average molecular masses.<br />

Since the tangential component of the velocity is continuous throughout the<br />

flame, there results that the incoming and outgoing velocities and the vector normal to<br />

the combustion front are coplanar.<br />

When the state of the unburnt gases and the incline α 1 of the inci<strong>de</strong>nt flow<br />

relative to the flame front, see Fig. 9.2, are known, the system of equations (9.1),<br />

(9.2), (9.3), (9.4) and (9.8) <strong>de</strong>termines the state of the burnt gases.<br />

Let ϕ be the flame propagation velocity corresponding to the composition of<br />

the unburnt gas mixture in the thermodynamic state <strong>de</strong>fined by the values p 1 and ρ 1<br />

of pressure and <strong>de</strong>nsity. Obviously we have<br />

ϕ = v n1 , (9.10)<br />

and for the incline α 1 of the flame front relative to the inci<strong>de</strong>nt flow<br />

sin α 1 = v n1<br />

v 1<br />

= ϕ v 1<br />

. (9.11)


234 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

V t1<br />

V n1<br />

α 2<br />

ϕ = −<br />

V n1<br />

α 1<br />

V 1<br />

α 1<br />

δ<br />

V n2<br />

V 2<br />

V t2 =<br />

Vt1<br />

Figure 9.2: Velocity components at both si<strong>de</strong>s of the front.<br />

Therefore, α 1 is <strong>de</strong>termined by the velocity v 1 of the inci<strong>de</strong>nt flow relative to the<br />

front and by the burning velocity ϕ of the mixture. This velocity <strong>de</strong>pends only on the<br />

composition, pressure and <strong>de</strong>nsity or temperature, and must be consi<strong>de</strong>red as a datum,<br />

<strong>de</strong>termined either theoretically or experimentally (see chapter 6).<br />

In or<strong>de</strong>r to simplify calculations, it is assumed in the following that the thermal<br />

enthalpy can be expressed as indicated in (9.7). Let<br />

q = h f1 − h f2 (9.12)<br />

be the difference between the formation enthalpies of the unburnt and burnt gases. By<br />

substituting equations (9.7) and (9.12) into (9.4), this equation can be written<br />

1<br />

2 v2 1 + c p1 T 1 + q = 1 2 v2 2 + c p2 T 2 . (9.13)<br />

Let T s1 and T s2 be the stagnation temperatures of the unburnt and burnt gases<br />

respectively. By virtue of equation (9.13) they are related by<br />

c p1 T s1 + q = c p2 T s2 , (9.14)<br />

and their ratio n is given by<br />

For the particular case c p1 = c p2 , we have<br />

n = T (<br />

s2<br />

= 1 + q )<br />

cp1<br />

. (9.15)<br />

T s1 c p1 T s1 c p2<br />

n = 1 +<br />

q<br />

c p1 T s1<br />

. (9.16)


9.3. NORMAL FLAME FRONT 235<br />

9.3 Normal flame front<br />

Let us consi<strong>de</strong>r a normal flame front as indicated in Fig. 9.3. In such a case, the<br />

equations that relate the conditions before and after the flame reduce to<br />

ρ 1 v 1 =ρ 2 v 2 , (9.17)<br />

ρ 1 v 2 1 + p 1 =ρ 2 v 2 2 + p 2 , (9.18)<br />

1<br />

2 v2 1 + c p1 T 1 + q = 1 2 v2 2 + c p2 T 2 . (9.19)<br />

V<br />

V<br />

1 2<br />

ϕ<br />

p , T ,<br />

ρ<br />

1 1 1<br />

p , T ,<br />

2 2<br />

ρ<br />

2<br />

Figure 9.3: Schematic diagram of a normal flame front.<br />

Since, v 1 is equal to the flame propagation velocity in the unburnt gases, (which<br />

is normally of the or<strong>de</strong>r of 1 m/s) and v 2 is only a few times larger, both v 1 and v 2<br />

are very “small” when compared to the sound speed in the unburnt and burnt gases<br />

respectively. Therefore, in equation (9.19) the terms containing the kinetic energy of<br />

the unburnt and burnt gases can be neglected. There results<br />

c p1 T 1 + q ≃ c p1 T s1 + q = c p2 T s2 ≃ c p2 T 2 . (9.20)<br />

That is<br />

T 1 ≃ T s1 , T 2 ≃ T s2 . (9.21)<br />

Therefore, by virtue of Eq. (9.15)<br />

T 2<br />

T 1<br />

≃ T s2<br />

T s1<br />

= n. (9.22)<br />

Furthermore, as results from Eq. (9.18) the pressure drop p 2 − p 1<br />

is very small<br />

compared to unity, and, consequently, for the calculation of the thermodynamic state<br />

of the burnt gases it may be assumed to be zero. The <strong>de</strong>nsity ratio is then given by<br />

ρ 2<br />

ρ 1<br />

= T 1<br />

T 2<br />

= 1 n . (9.23)<br />

p 1


236 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

The ratio of velocities is<br />

obtained from Eqs. (9.23) and (9.17).<br />

v 2<br />

v 1<br />

= n, (9.24)<br />

Therefore, once the <strong>de</strong>nsity and temperature of the unburnt gases and the value<br />

of n are known, the temperature of the burnt gases is <strong>de</strong>termined by equation (9.22),<br />

its <strong>de</strong>nsity by equation (9.23), and the pressure drop across the flame by<br />

p 2 − p 1<br />

p 1<br />

= γ 1 M 2 1 (1 − n), (9.25)<br />

where γ 1 is the ratio of the heat capacity at constant pressure to the heat capacity at<br />

constant volume of the unburnt gases, and<br />

M 1 = v 1<br />

a 1<br />

(9.26)<br />

is the flow Mach number for the unburnt gases, where<br />

√<br />

p 1<br />

a 1 = γ 1 (9.27)<br />

ρ 1<br />

is the sound speed in these same gases.<br />

9.4 Inclined flame front<br />

Here, v n1 and v n2 are still small compared to the sound speed of the unburnt and<br />

burnt gases respectively. Therefore, the pressure drop across the flame front is still<br />

small, see Eq. (9.2), and the ratio between <strong>de</strong>nsities on each si<strong>de</strong> of the flame front is<br />

<strong>de</strong>termined by the ratio of temperature T 2 to temperature T 1 . Let λ be this ratio, there<br />

results<br />

Let<br />

λ = T 2<br />

T 1<br />

= ρ 1<br />

ρ 2<br />

= v n2<br />

v n1<br />

. (9.28)<br />

δ = α 2 − α 1 (9.29)<br />

be the velocity <strong>de</strong>viation across the flame front (see Fig. 9.2) . A simple calculation<br />

gives for δ<br />

tan δ = (λ − 1) tan α 1<br />

1 + λ tan 2 α 1<br />

. (9.30)<br />

This ratio has been taken into Fig. 9.4 for different values of λ. As it can easily be<br />

seen, the maximum <strong>de</strong>viation δ max corresponds to<br />

tan α 1 = 1 √<br />

λ<br />

. (9.31)


9.4. INCLINED FLAME FRONT 237<br />

60<br />

50<br />

40<br />

λ = 12<br />

8<br />

6<br />

5<br />

4<br />

δ ( o )<br />

30<br />

3<br />

20<br />

λ = 2<br />

10<br />

δ max<br />

0<br />

10 30 50 70 90<br />

α 1<br />

( o )<br />

Figure 9.4: Velocity <strong>de</strong>viation across the flame front as a function of α 1.<br />

Its value is given by<br />

tan δ max = 1 2<br />

( )<br />

√λ 1 − √λ . (9.32)<br />

Even though the normal velocity is always “small”, the tangential velocity v t , which<br />

by virtue of equation (9.3) is the same at both si<strong>de</strong>s of the front, can be “large”. Therefore,<br />

it must be taken into account when writing the energy equation. As a result, the<br />

temperature at each si<strong>de</strong> of the front can differ, consi<strong>de</strong>rably, from the corresponding<br />

stagnation temperature. This is exactly the opposite to what occurs in the case of a<br />

normal front.<br />

Therefore, two different cases are to be consi<strong>de</strong>red, either<br />

v t ≃ O(v n ), (9.33)<br />

in which case<br />

or<br />

in which case<br />

T 1 ≃ T s1 , T 2 ≃ T s2 , λ ≃ n, (9.34)<br />

v t ≫ v n , (9.35)<br />

v 2 1 ≃ v 2 2 ≃ v 2 t , (9.36)<br />

and equation (9.13) reduces to<br />

c p1 T 1 + q = c p2 T 2 . (9.37)


238 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

By virtue of equation (9.15) and of the relation<br />

the following expression is obtained for λ<br />

T s1<br />

= 1 + γ 1 − 1<br />

M1 2 , (9.38)<br />

T 1 2<br />

λ = n + γ 1 − 1<br />

2<br />

(<br />

n − c )<br />

p1<br />

M1 2 , (9.39)<br />

c p2<br />

which shows that for large values of M 1 , that is, when the flame front is very inclined<br />

to the inci<strong>de</strong>nt flow, λ can appreciably differ from n.<br />

The values of λ/n as a function of M 1 , for c p2 = c p1 and different values of<br />

χ = γ 1 − 1 n − 1<br />

, have been represented in Fig. 9.5.<br />

2 n<br />

λ / n<br />

1.20<br />

1.18<br />

1.16<br />

1.14<br />

1.12<br />

1.10<br />

1.06<br />

1.06<br />

1.04<br />

1.02<br />

χ=(γ 1<br />

−1)(n−1)/2n<br />

χ=0.20<br />

χ=0.18<br />

χ=0.16<br />

χ=0.14<br />

χ=0.12<br />

χ=0.10<br />

1.00<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

M 1<br />

Figure 9.5: Values of the ratio λ/n as a function of inci<strong>de</strong>nt Mach number M 1.<br />

Relation (9.39) is only valid for very inclined flame front, that is, for not too<br />

small values of M 1 . For values of M 1 close to zero, it must be substituted by<br />

λ ≃ n. (9.40)<br />

The first case, that is to say, the case of slow flows (M 1 ≪ 1) with flame<br />

fronts, has been examined in <strong>de</strong>tail by Gross and Esch [4], reaching the following<br />

conclusions:<br />

1) If ϕ, λ and q are constant and the motion is irrotational before the flame, after<br />

the flame it continues being irrotational.


9.5. ENTROPY JUMP ACROSS THE FLAME FRONT 239<br />

PRODUCTS<br />

FLAME FRONT<br />

REACTANTS<br />

Figure 9.6: Straight-line flame flow field according to Gross and Esch for ϕ/v 1 = 0.1 and<br />

λ = 7.<br />

2) Opposite to what occurs with fast flows, if the flow is slow pressure drop across<br />

the flame front cannot be neglected.<br />

In or<strong>de</strong>r to analyze the motion, the flame front can be consi<strong>de</strong>red as a surface with a<br />

distribution of sources. For example, in the case of a plane motion, the strength of the<br />

(λ − 1)ϕ<br />

source per unit length of the front is . Now by expressing the condition that<br />

2π<br />

in the unburnt gases the velocity normal to the front must be ϕ, an integral equation<br />

is obtained which must ϕ satisfied at the front. This integral equation <strong>de</strong>termines the<br />

flame shape which ”a priori” is unknown. Gross and Esch give approximate solutions<br />

for certain cases. For example, for the case in which the flame front reduces to a<br />

straight segment inclined to the inci<strong>de</strong>nt flow. Fig. 9.6, taken from the said work,<br />

shows the shape of the streamlines in this case.<br />

9.5 Entropy jump across the flame front<br />

The entropy of the unburnt gases is given by the expression 3<br />

3 See chapter 1.<br />

S 1 = S 01 + c p1 ln p1/γ 1<br />

1<br />

, (9.41)<br />

ρ 1


240 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

where S 01 <strong>de</strong>pends only on the composition of the mixture. Likewise, the entropy of<br />

the burnt gases is<br />

S 2 = S 02 + c p2 ln p1/γ2 2<br />

, (9.42)<br />

ρ 2<br />

Therefore, the entropy jump ∆S across the flame, is<br />

∆S = S 2 − S 1 = (S 02 − S 01 ) + c p2 ln p1/γ 2<br />

2<br />

− c p1 ln p1/γ 1<br />

1<br />

. (9.43)<br />

ρ 2 ρ 1<br />

As previously seen, the pressure drop across the flame is very small. Therefore, in<br />

(9.43) we can take<br />

p 2 ≃ p 1 = p. (9.44)<br />

Making use of this simplification and of the relation (9.28) between p 1 and p 2 , the<br />

following expression is obtained for ∆S<br />

being<br />

∆S = S 2 − S 1 = (S 02 − S 01 ) + c p2 ln λ + (c p2 − c p1 ) ln p1/γ 12<br />

ρ 1<br />

, (9.45)<br />

γ 12 = c p2 − c p1<br />

c v2 − c v1<br />

, (9.46)<br />

where c v1 and c v2 are the heat capacities at constant volume of the unburnt and burnt<br />

gases respectively.<br />

In the particular case<br />

c p2 = c p1 = c p , (9.47)<br />

the entropy jump across the flame reduces to<br />

∆S = (S 02 − S 01 ) + c p ln λ. (9.48)<br />

Equation (9.39) shows that λ can vary along the flame front when M 1 varies. Therefore,<br />

if the local conditions of the flow vary, the entropy jump across the flame front<br />

can vary consi<strong>de</strong>rably from one point to another. Due to this variation of the entropy<br />

jump along the flame front, the motion after the front can be rotational, even<br />

if the motion before the front is potential. This is similar to what occurs in the case<br />

of supersonic motions with shock waves, when their incline varies from one point to<br />

another. 4<br />

If the motion before the flame is isentropic, the value λ <strong>de</strong>pends only on the<br />

incline α 1 of the flame front. In fact, by using the relation,<br />

57.<br />

tan α 1 = v n1<br />

v t<br />

= ϕ v t<br />

, (9.49)<br />

4 See A. Ferri: Elements of Aerodynamic of Supersonic Flows. Mac Millan Comp., New York, 1949, p.


9.5. ENTROPY JUMP ACROSS THE FLAME FRONT 241<br />

that gives the incline of the flame front, and the relation<br />

M 1 ≃ v t<br />

a , (9.50)<br />

which, as seen in the preceding paragraph, is valid for the very inclined flame fronts,<br />

Bernoulli equation for unburnt gases can be written<br />

1 + γ 1 − 1<br />

M1 2 =<br />

2<br />

(<br />

a01<br />

ϕ<br />

) 2<br />

M 2 1 tan 2 α 1 , (9.51)<br />

where a 01 is the sound speed at the stagnation point of the unburnt gases.<br />

By eliminating M 1 between this equation and equation (9.39), λ can be expressed<br />

as a function of the flame incline, as follows<br />

λ = n +<br />

2<br />

γ 1 − 1<br />

n − c p1<br />

(<br />

a01<br />

ϕ<br />

c p2<br />

) 2<br />

tan 2 α 1 − 1. (9.52)<br />

When the flame incline is increased, α 1 <strong>de</strong>creases, and relation (9.52) show that λ<br />

increases. Therefore, the entropy jump ∆S increases when the flame incline is increased.<br />

Such behavior is the opposite to that of a shock wave where the entropy jump<br />

<strong>de</strong>creases when the wave incline is increased. See Fig. 9.7 [5] . The influence of<br />

this variation of the entropy jump on the flow will be studied in the following paragraph.<br />

The simplified expression (9.48) for the entropy jump will be used, and since<br />

the constant S 02 − S 01 has no influence on the flow, we shall express in short<br />

∆S = c p ln λ. (9.53)<br />

∆ S<br />

ω 2 ∆ S<br />

ω 2<br />

(a) Flame<br />

(b) Shock wave<br />

Figure 9.7: Entropy jump across a flame front and a shock wave.


242 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

9.6 Vorticity across the flame<br />

In the flow of a perfect gas, where viscosity and mass forces are neglected, the variation<br />

of the rotational ¯ω = ∇ × ¯v as a function of the stagnation temperature T s and<br />

specific entropy S can be expressed as follows<br />

∂ ¯ω<br />

∂t + ¯v × ¯ω = c p∇T s − T ∇S. (9.54)<br />

For isoenergetic (T s = const.) and stationary (∂/∂t = 0) flows, only this<br />

case will be consi<strong>de</strong>red in the present study, 5 equation (9.54) reduces to the following<br />

(Crocco’s Theorem)<br />

¯v × ¯ω = −T ∇S. (9.55)<br />

From this equation, the jump of the rotational, across the flame, can be computed<br />

when the values of ¯v, T and S are known. The jump of the rotational is obtained by<br />

(9.55) to both si<strong>de</strong>s of the flame. For the calculation, we shall adopt on each point<br />

of the flame front a cartesian rectangular coordinate system, <strong>de</strong>fined in the following<br />

way:<br />

Axis n: Normal to the flame front.<br />

Axis t: Intersection of the plane tangent to the flame with the plane of the inci<strong>de</strong>nt<br />

and emergent velocities ¯v 1 and ¯v 2 .<br />

Axis τ: Normal to the (n, t) plane forming a positive trihedron.<br />

The velocity components relative to this system, will be v n , v t and 0. The<br />

vorticity components ω n , ω t and ω τ . Therefore, those of the vector product ¯v × ¯ω will<br />

be v t ω τ , −v n ω τ and (v n ω t − v t ω n ). Consequently, equation (9.55) breaks down into<br />

the following three equations 6<br />

v t ω τ = −T ∂S<br />

∂n , (9.56)<br />

v n ω τ =<br />

T ∂S<br />

∂t , (9.57)<br />

v n ω t − v t ω n = −T ∂S<br />

∂τ . (9.58)<br />

The jump of the normal component of the vorticity ω n can be computed directly.<br />

The above equations are not necessary for the computation. In fact, since v t<br />

5 If the flow before the flame is isoenergetic and if the heat q released in the combustion is constant, as<br />

will be assumed hereinafter, the flow after the flame is also isoenergetic, as results from (9.14).<br />

6 The <strong>de</strong>rivative ∂S/∂t in the tangential direction must not be confused with a time <strong>de</strong>rivatives.


9.6. VORTICITY ACROSS THE FLAME 243<br />

is continuous across the flame, by applying Stokes theorem to both faces of a surface<br />

element of the flame, the following is obtained<br />

ω n1 = ω n2 , (9.59)<br />

that is, the normal component of the vorticity is continuous across the flame.<br />

The jump of ω τ can then be obtained from (9.57) by writing this equation for<br />

each si<strong>de</strong> of the flame and forming the difference. Thus, when taking into account that<br />

v n1 = ϕ, v n2 = λϕ, T 2 = λT 1 , S 2 = S 1 + ∆S there results<br />

ω τ2 = ω τ2<br />

λ + T [<br />

1 λ − 1 ∂S 1<br />

ϕ λ ∂t + ∂(∆S) ]<br />

. (9.60)<br />

∂t<br />

Likewise by using (9.58) we obtain for ω t2<br />

ω t2 = ω t1<br />

λ − T [<br />

1 λ − 1<br />

ϕ λ<br />

∂S 1<br />

∂τ + ∂(∆S) ]<br />

. (9.61)<br />

∂τ<br />

On the other hand, equations (9.57) and (9.58) , when written for conditions before<br />

the flame, give<br />

ϕω τ1 = T 1<br />

∂S 1<br />

∂t , (9.62)<br />

ϕω t1 − v t ω n1 = − T 1<br />

∂S 1<br />

∂τ . (9.63)<br />

By eliminating ∂S 1<br />

and ∂S 1<br />

between these two equations and (9.60) and (9.61), we<br />

∂t ∂τ<br />

finally obtain for ω τ2 and ω t2 ,<br />

ω τ2 = ω τ1 + T 1<br />

ϕ<br />

∂(∆S)<br />

, (9.64)<br />

∂t<br />

ω t2 = ω t1 − λ − 1<br />

λ ω n1 cot α 1 − T 1<br />

ϕ<br />

∂(∆S)<br />

. (9.65)<br />

∂τ<br />

In particular, if the motion before the flame is irrotational (ω n1 = ω t1 = ω τ1 =<br />

0), the vorticity produced by the flame results<br />

ω n2 = 0, (9.66)<br />

ω t2 = − T 1<br />

ϕ<br />

ω τ2 = T 1<br />

ϕ<br />

∂(∆S)<br />

,<br />

∂τ<br />

(9.67)<br />

∂(∆S)<br />

.<br />

∂t<br />

(9.68)<br />

Furthermore, if the entropy jump can be expressed as (9.53), we have<br />

ω t2 = − c pT 1<br />

ϕ<br />

ω τ2 = c pT 1<br />

ϕ<br />

1 ∂λ<br />

λ ∂τ , (9.69)<br />

1 ∂λ<br />

λ ∂t , (9.70)


244 CHAPTER 9. FLOWS WITH COMBUSTION WAVES<br />

or else, if the motion is rotational, the equations resulting from (9.64) and (9.65) when<br />

∆S is substituted therein for its value (9.53).<br />

In the special case of a plane motion, the only component different from zero<br />

of the vorticity is ω τ , and its value after the flame is given by (9.64) if the motion<br />

before the flame is rotational, and by (9.68) if it is potential. In Fig. 9.7.a, the rotation<br />

sense of the vorticity generated by the flame has bean indicated, in the case of a plane<br />

motion, assuming that the motion before the flame is potential. In Fig. 9.7.b, the sense<br />

corresponding to a shock wave is shown for the same case.<br />

References<br />

[1] Emmons, H. W., Ball, G. A. and Maier, A. D.: Development of a Combustion<br />

Tunnel. Army Or<strong>de</strong>nance Project Report, Harvard University, Cambridge<br />

Mass., 1954.<br />

[2] Emmons, H. W.: Fundamentals of Gas Dynamics. Sec. E, Vol. III of High<br />

Speed Aerodynamics and Jet Propulsion. Princeton University Press. 1958.<br />

[3] Gross, R. A.: Combustion Tunnel Laboratory Interim Technical Report No. 2.<br />

Harvard University, June 1952.<br />

[4] Gross, R. A. and Esch, R.: Low Speed Combustion Aerodynamics. Jet Propulsion,<br />

March-April 1954, pp. 95-101.<br />

[5] von Kármán, Th.: Aerothermodynamics and Combustion Theory. L’Aerotecnica,<br />

Vol. XXXIII, Fasc. 1st., 1953, pp. 80-86.


Chapter 10<br />

Aerothermodynamic field of a<br />

stabilized flame<br />

10.1 Introduction<br />

As an example of the application of the aerothermodynamic method that was outlined<br />

in the preceding chapter for the study of gas flows, in the present chapter we shall<br />

study the characteristics of the flow in a combustion chamber with a stabilized flame.<br />

y<br />

+h<br />

F<br />

u 0<br />

0<br />

x<br />

−h<br />

F’<br />

Figure 10.1: Schematic diagram of a stabilized flame front in a two-dimensional combustion<br />

chamber.<br />

The problem is outlined in Fig. 10.1. To simplify calculation we shall consi<strong>de</strong>r<br />

the case of a two dimensional chamber of constant width 2h and unit thickness normal<br />

to the plane of the figure. An uniform flow of homogeneous pre-mixed combustible<br />

enters the combustion chamber. The velocity, pressure, <strong>de</strong>nsity and temperature of<br />

the mixture are u 0 , p 0 , ρ 0 and T 0 , respectively. The flame front FOF’ is stabilized at<br />

point O by one of the methods studied in the next chapter. As expressed in the pre-<br />

245


246 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

ceding chapter, this flame front can be consi<strong>de</strong>red as a discontinuity surface between<br />

unburnt and burnt gases.<br />

The problem lies in <strong>de</strong>termining:<br />

a) The width of the flame as a function of the fraction of gas burnt.<br />

b) The pressure drop along the chamber.<br />

c) The velocity distribution in the different sections of the chamber.<br />

d) The shape of the flame.<br />

The following simplifying assumptions are introduced:<br />

1) Combustion efficiency is the same at all points, that is, the heat released in the<br />

combustion per unit mass of fuel is in<strong>de</strong>pen<strong>de</strong>nt from the state of the gas before<br />

the flame.<br />

2) Heat capacity at constant pressure c p is in<strong>de</strong>pen<strong>de</strong>nt from temperature, and has<br />

the same value for the unburnt and burnt gases.<br />

3) Unburnt and burnt gases behave as perfect gases. The constant R g of their state<br />

equation has the same value for both gases<br />

p<br />

ρ = R gT. (10.1)<br />

4) The flame propagation velocity is constant along the front and very small when<br />

compared to the gas velocity.<br />

5) Unburnt and burnt gases are i<strong>de</strong>al fluids, their viscosity and thermal conductivity<br />

are negligible.<br />

Furthermore, in this study only the case of stationary flow will be consi<strong>de</strong>red.<br />

Thus stated, this problem has been studied by A.C. Scurlock [1] and [2], who introduced<br />

further simplification by assuming both unburnt and burnt gases to be incompressible<br />

fluids.<br />

Since the flame speed is small compared to the gas speed, the angle between the<br />

front and the streamlines is very small. Therefore, the streamlines are approximately<br />

parallel to the chamber axis. Furthermore, due to the negligibility of the pressure drop<br />

across the flame front, it can be assumed that pressure is constant at each cross section<br />

of the chamber. Moreover, gas speed can be substituted by its component parallel<br />

to the chamber axis. Therefore, from these assumptions an almost one-dimensional<br />

theory can be worked out.<br />

By using the aforementioned simplifying assumptions Scurlock obtained numerical<br />

solutions for the equations of motion in some typical cases. In or<strong>de</strong>r to per-


10.1. INTRODUCTION 247<br />

form numerical computations, Scurlock substitutes the differential equations for finite<br />

differences.<br />

Some of the results, taken from Ref. [1], are given in Figs. 10.2 to 10.5. These<br />

results correspond to the case where the <strong>de</strong>nsity ratio of the unburnt gases to the burnt<br />

gases is 6, and the ratio of the flame speed to the gas velocity at the chamber’s inlet is<br />

λ = 0.10.<br />

y/h<br />

A B C<br />

1<br />

0<br />

x/h<br />

−1<br />

0<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

∆ P / ∆ P t<br />

1<br />

0 1 2 3 4 5 6 7 8 9 10 11 x/h<br />

Figure 10.2: Streamlines of flow through a flame front in a two-dimensional chamber and<br />

pressure drop along it.<br />

Fig. 10.2 shows the shape of the flame front and of the streamlines, as well<br />

as the pressure drop ∆p along the chamber referred to the total pressure drop ∆p t ,<br />

between the inlet and outlet sections.<br />

Fig. 10.3 shows the velocity distribution for the cross-sections A, B and C<br />

indicated in Fig. 10.2. These sections correspond to the point at which combustion<br />

starts (section A), to the point where the burnt fraction is 0.25 (section B) and to the<br />

point where this fraction is 0.75 (section C). In this figure all velocities are referred to<br />

the velocity u 0 of the gas at the inlet of the chamber. Fig. 10.3 shows that:<br />

1) While the velocity of the unburnt gases is constant at each cross section, the<br />

velocity of the burnt gases increases from the flame front towards the chamber<br />

axis where it is maximum.


248 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

y/h<br />

1<br />

0<br />

A<br />

B<br />

C<br />

U/U 0<br />

−1<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

Figure 10.3: Velocity profiles for three cross-sections of the two-dimensional chamber (see<br />

Fig. 10.2).<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

η / h<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

f<br />

Figure 10.4: Flame width as a function of the burnt fraction for incompressible flow and<br />

λ = 6.<br />

2) Due to the pressure drop produced by combustion, the mass of gas accelerates<br />

downstream of the chamber.<br />

Fig. 10.4 shows the variation law of the flame width, referred to that of the<br />

chamber, as a function of the burnt fraction.<br />

Fig. 10.5 shows the fraction of pressure drop along the chamber, referred to<br />

the total pressure drop between the inlet and outlet sections, as a function of the burnt<br />

fraction.


10.1. INTRODUCTION 249<br />

∆ P /∆ P t<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

f<br />

Figure 10.5: Pressure drop as a function of the burnt fraction for incompressible flow and<br />

λ = 6.<br />

G.A. Ball, [3] and [4], has studied the same problem by applying the relaxation<br />

method.<br />

The assumption that <strong>de</strong>nsity, for both unburnt and burnt gases, is constant,<br />

is permissible only when the variations of the Mach number along the chamber are<br />

small. Due to the acceleration produced by the pressure drop, the local Mach number<br />

close to the walls and near the end of the chamber can be large, even for small entrance<br />

Mach numbers. As a consequence a choking of the chamber can occur preventing the<br />

complete combustion of the gas, as will be shown later. This choking phenomenon<br />

cannot be predicted by Scurlock’s hydrodynamic theory.<br />

H.S. Tsien in the U.S.A. [5], and J. Fabri, R. Siestrunck and C. Fouré in France,<br />

[6], [7] and [8], have studied (simultaneous but in<strong>de</strong>pen<strong>de</strong>ntly) the influence of gas<br />

compressibility on the problem.<br />

By assuming at each section a linear velocity distribution for the burnt gases<br />

(see Fig. 10.3), H. S. Tsien has <strong>de</strong>monstrated that an approximate analytical solution<br />

of Scurlock’s problem can be easily obtained. The results are very close to those<br />

obtained by Scurlock. The linear velocity a distribution is a good approximation to the<br />

profile calculated by Scurlock. Moreover, Tsien has exten<strong>de</strong>d the analysis, taking into<br />

account compressibility effects, by applying the same i<strong>de</strong>a of a quasi one-dimensional<br />

theory. He also assumes a linear velocity distribution for the burnt gases, showing that<br />

in this case choking of the chamber can occur.


250 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

Fabri, Siestrunck and Fouré have <strong>de</strong>veloped a similar theory. They also inclu<strong>de</strong><br />

the effect of <strong>de</strong>nsity variations, but do not postulate a linear velocity profile for burnt<br />

gases. Their stating of the problem leads then to an integral equation for the stream<br />

function. For some typical cases they have integrated this equation numerically. Furthermore,<br />

they exten<strong>de</strong>d the analysis to other types of chambers. For instance, the<br />

cylindrical chamber with circular cross section and different arrangements of the flame<br />

stabilizer.<br />

As a result of their studies Fabri, Siestrunck and Fouré also predict the occurrence<br />

of choking when the velocity at the inlet section is larger than a critical velocity.<br />

In the following sections the approximate method of Tsien will be <strong>de</strong>scribed<br />

first and then the method <strong>de</strong>veloped by Fabri-Siestrunck-Fouré.<br />

10.2 Tsien method<br />

h<br />

p’ p<br />

B<br />

ρ’<br />

1<br />

D<br />

ρ’<br />

2<br />

C<br />

y<br />

u 0<br />

ρ<br />

1<br />

u e<br />

ρ<br />

1<br />

y<br />

u<br />

1<br />

u<br />

0<br />

A<br />

Figure 10.6: Notation for the Tsien method.<br />

Figure 10.6 contains the necessary elements for Tsien’s approximate calculation.<br />

obtained<br />

By applying the continuity equation to section AB, the following equation is<br />

∫ y1<br />

ρ 1 u 1 (h − y 1 ) + ρu dy = ρ 0 u 0 h. (10.2)<br />

0<br />

By introducing η = y 1 /h and U = u 1 /u 0 it can be written in the following dimensionless<br />

form<br />

∫<br />

ρ η<br />

1<br />

ρ u<br />

( y<br />

)<br />

U(1 − η) + d = 1. (10.2.a)<br />

ρ 0 0 ρ 0 u 0 h


10.2. TSIEN METHOD 251<br />

As aforesaid, Tsien assumes that the velocity distribution of the burnt gases<br />

between A and C is linear, that is, u (Fig. 10.6) is given by the expression<br />

u = u 1 + (u e − u 1 )<br />

(1 − y )<br />

, (10.3)<br />

y 1<br />

where u e is the velocity on the axis of the chamber.<br />

Furthermore, Tsien assumes that the ratio λ of the <strong>de</strong>nsity of the unburnt gases<br />

to that of the burnt gases at each point of the flame front is constant. 1 Then, it can be<br />

easily shown that the <strong>de</strong>nsity of the burnt gases is constant between A and C and equal<br />

to ρ 1 /λ<br />

In fact, in D we have<br />

ρ = ρ 1<br />

λ . (10.4)<br />

ρ ′ 1<br />

ρ ′ 2<br />

= λ. (10.5)<br />

But due to the fact that the expansions of both unburnt and burnt gases between<br />

pressure p ′ in D and p in C are isentropic, there result<br />

and<br />

ρ ′ 1<br />

ρ 1<br />

=<br />

( ) 1 p<br />

′<br />

γ<br />

p<br />

(10.6)<br />

The combination of (10.5), (10.6) and (10.7) leads to (10.4).<br />

ρ ′ ( ) 1<br />

2 p<br />

′<br />

ρ = γ . (10.7)<br />

p<br />

Since pressure and <strong>de</strong>nsity of the burnt gases are constant between A and C, the<br />

temperature T of the burnt gases must also be constant between A and C, and equal to<br />

T 2<br />

T = T 2 = λT 1 = T e , (10.8)<br />

where T e is the temperature at A.<br />

The Bernoulli equation, when applied to the unburnt gas between u 0 and u 1 ,<br />

gives<br />

1<br />

2 u2 0 + c p T 0 = 1 2 u2 1 + c p T 1 . (10.9)<br />

If T 1 is eliminated by combining this equation with the equation<br />

( ) 1<br />

ρ 1 T1 γ − 1<br />

=<br />

ρ 0 T 0<br />

(10.10)<br />

1 Relation (9.36) in the preceding chapter proves that this ratio actually varies with the flame front incline.


252 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

of the reversible adiabatics, it results for ρ 1 /ρ 0<br />

(<br />

ρ 1<br />

= 1 − γ − 1 M0 2 (U 2 − 1)<br />

ρ 0 2<br />

) 1<br />

γ − 1 = λ<br />

ρ<br />

ρ 0<br />

, (10.11)<br />

where M 0 = u 0 / √ (γ − 1)c p T 0 is the Mach number at the inlet section of the combustion<br />

chamber.<br />

On the other hand, Bernoulli’s equation applied to the burnt gases between O<br />

and A, taking (10.8) into account, gives<br />

1<br />

2 u2 0 + λc p T 0 = 1 2 u2 e + λc p T 1 . (10.12)<br />

Finally, the elimination of T 0 and T 1 between (10.9) and (10.12) gives for u e<br />

u e<br />

u 0<br />

= √ 1 + λ(U 2 − 1). (10.13)<br />

When (10.3), (10.11) and (10.13) are substituted into (10.2.a) and the integration<br />

performed, the following expression for η as a function of U is obtained<br />

η =<br />

(<br />

U −<br />

1<br />

2λ<br />

1 − γ − 1 M0 2 (U 2 − 1)<br />

2<br />

) 1 −<br />

γ − 1<br />

(<br />

(2λ − 1)U − √ 1 + λ(U 2 − 1)<br />

The fraction f of the burnt gases is given by the expression<br />

f = ρ 0u 0 h − ρ 1 u 1 (h 1 − y 1 )<br />

ρ 0 u 0 h<br />

=1 − ρ 1u 1<br />

ρ 0 u 0<br />

(1 − η)<br />

) . (10.14)<br />

(<br />

=1 − U(1 − η) 1 − γ − 1<br />

) 1<br />

M0 2 (U 2 γ − 1<br />

− 1) . (10.15)<br />

2<br />

System (10.14) and (10.15) allow computation of the fraction f of burnt gas as a<br />

function of the flame width η, through parameter U. Fig. 10.7 shows several of the<br />

results given by Tsien in his work for the case λ = 6.<br />

The main result of Tsien’s work is the following: for each value of λ there is a<br />

critical Mach number M cr for the flow at the inlet section, such that when M 0 < M cr<br />

the combustion is complete and the flame reaches the chamber walls. On the other<br />

hand, if M 0 > M cr the maximum burnt fraction f max is smaller than unity and the<br />

flame cannot reach the walls. This is known as the choking phenomenon. It shows


10.2. TSIEN METHOD 253<br />

1.0<br />

λ=6.0<br />

M 0<br />

=0.2<br />

M 0<br />

=0, 0.1<br />

0.8<br />

η<br />

0.6<br />

M 0<br />

=0.4<br />

0.50<br />

M 0<br />

=0.4<br />

0.4<br />

0.25<br />

M 0<br />

=0.6<br />

0.2<br />

M 0<br />

=0.6<br />

0.00<br />

0.0 0.1 0.2<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

f<br />

Figure 10.7: Flame width as a function of the burnt fraction for <strong>de</strong>nsity ratio λ = 6. For<br />

each value of the initial Mach number M 0, a dot indicates the maximum values<br />

of η and f.<br />

that, asi<strong>de</strong> from the difficulties arising from the problem of flame stabilization at high<br />

speed, 2 other difficulties are to be expected when trying to burn gas at high speed, due<br />

to the interaction between aerodynamic and combustion processes.<br />

Choking occurs due to the fact that the contraction of the unburnt gases originated<br />

by downstream acceleration is not sufficient to compensate the expansion of<br />

the gases as they burn. Compressibility acts because the contraction corresponding<br />

to a given acceleration is smaller in a compressible fluid than in an incompressible<br />

one. This is the reason why the choking phenomenon does not exist in Scurlock’s<br />

hydrodynamic theory.<br />

The critical Mach number that corresponds to each value of λ can be obtained<br />

by expressing the condition that the curve of f as a function of U, obtained by eliminating<br />

η between (10.14) and (10.15), must have a maximum at the point f = 1. That<br />

is, when the conditions<br />

or their equivalents<br />

are satisfied.<br />

df<br />

dU<br />

dη<br />

dU<br />

= 0 and f = 1, (10.16)<br />

= 0 and η = 1, (10.17)<br />

2 The problem of stabilizing a flame at high speed is studied in the next chapter.


254 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

U η=1<br />

10<br />

8<br />

6<br />

4<br />

2<br />

M η=1<br />

M cr<br />

U η=1<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

M cr<br />

, M η=1<br />

0<br />

0.0<br />

0 2 4 6 8 10 12<br />

λ<br />

Figure 10.8: Critical conditions for the complete combustion as a function of the <strong>de</strong>nsity<br />

ratio λ.<br />

Figure 10.8, in which the values of M cr thus calculated have been taken from<br />

[5], shows that the choking effects impose an important limitation to the velocity at<br />

which gases enter the combustion chamber. Choking effects can be avoi<strong>de</strong>d by using<br />

an expanding combustion chamber. It is interesting to point out that when the maximum<br />

Mach number is calculated at critical conditions, that is, at the initiation of the<br />

choking phenomenon, this number is slightly larger than one. This value corresponds<br />

to the unburnt gases at the point where the flame reaches the chamber wall. Therefore,<br />

in the neighbourhood of this point, there exists a small area where the flow is slightly<br />

supersonic. Details of the calculation can be found in Tsien’s work. Furthermore,<br />

Tsien’s analysis shows that Scurlock’s hydrodynamic theory gives satisfactory results<br />

when the entry Mach number is not too close to the critical.<br />

Recently, G. Ernst [9] has applied Tsien’s method to the same problem as well<br />

as to the study of other types of combustion chambers and other arrangements of the<br />

stabilizer. Ernst substitutes Tsien’s linear velocity profile for a parabolic one with<br />

an exponent 1.33, which is the one that best fits the profiles obtained by Scurlock.<br />

Moreover, instead of assuming a constant <strong>de</strong>nsity for the burnt gases at each cross<br />

section of the chamber, he inclu<strong>de</strong>s the effect of the slight <strong>de</strong>nsity variation of the burnt<br />

gases. However, Ernst’s results only differ slightly from those obtained by Tsien.


10.3. METHOD OF FABRI-SIESTRUNCK-FOURÉ 255<br />

10.3 Method of Fabri-Siestrunck-Fouré<br />

The flow in the combustion chamber is two dimensional and stationary. Therefore, a<br />

stream function ψ exists. This function is <strong>de</strong>fined by the following relations<br />

ρu =ρ 0 u 0 h ∂ψ<br />

∂y , (10.18)<br />

ρv =ρ 0 u 0 h ∂ψ<br />

∂x . (10.19)<br />

Thus <strong>de</strong>fined, ψ is non-dimensional.<br />

The x axis is a streamline to which the value ψ = 0 is assigned. The upper<br />

wall of the chamber, y = h, is also a streamline for which ψ = 1, as results from the<br />

integration of (10.18). Therefore, in the upper half of the chamber, ψ varies from the<br />

value zero, at the axis, to the value one, at the wall.<br />

A linearization of the problem, by assuming that all velocities are small compared<br />

to u, leads to the conclusion that pressure is constant at each cross section of<br />

the chamber. Therefore, since the flow of unburnt gases is isentropic, their <strong>de</strong>nsity,<br />

temperature and velocities are equally constant at each cross section. Moreover, the<br />

value of either one of these magnitu<strong>de</strong>s <strong>de</strong>termines the values for the others.<br />

ψ<br />

ψ’<br />

D u’<br />

ρ<br />

1<br />

B<br />

C<br />

ρ’’<br />

E<br />

u<br />

u’’<br />

ψ<br />

ψ’<br />

0<br />

x<br />

A<br />

Figure 10.9: Notation for the method of Fabri-Siestrunck-Fouré.<br />

Let us consi<strong>de</strong>r a cross section AB, at a distance x from the stabilizer, as shown<br />

in Fig. 10.9. Here, evi<strong>de</strong>ntly<br />

∫<br />

1 h<br />

dy = 1, (10.20)<br />

h 0<br />

which, by virtue of (10.18), can be written in the following form<br />

∫ 1<br />

0<br />

ρ 0 u 0<br />

ρu<br />

dψ = 1. (10.21)


256 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

Let us consi<strong>de</strong>r the streamline ψ that crosses the flame front at section AB<br />

(Fig. 10.9) at point C. At this section, the streamline separates the unburnt from the<br />

burnt gases. To the first correspond values ψ ′ of the stream function between ψ and 1.<br />

To the latter correspond values ψ ′ between 0 and ψ. By separating in (10.21) both<br />

intervals, the following is obtained;<br />

∫ ψ<br />

0<br />

∫<br />

ρ 0 u 1<br />

0 ρ 0 u 0<br />

ρ ′′ u ′′ dψ′ +<br />

ψ ρu dψ′ = 1, (10.22)<br />

where ρ ′′ and u ′′ are the corresponding values of ρ and u, on the streamline ψ ′ at point<br />

E on section AB.<br />

As previously said, both <strong>de</strong>nsity and velocity of the unburnt gas are constant<br />

at each cross section of the chamber. If ρ 1 and u are their values at section BC, the<br />

second integral in (10.22) can be integrated, obtaining<br />

∫ 1<br />

This value, when carried into (10.21), gives<br />

∫ ψ<br />

This equation can be written<br />

0<br />

ψ<br />

ρ 0 u 0<br />

ρ 1 u dψ′ = ρ 0u 0<br />

(1 − ψ). (10.23)<br />

ρ 1 u<br />

ρ 0 u 0<br />

ρ ′′ u ′′ dψ′ = 1 − ρ 0u 0<br />

(1 − ψ). (10.24)<br />

ρ 1 u<br />

∫<br />

ρ 1 u<br />

ψ<br />

ρ 1 u<br />

− 1 + ψ =<br />

ρ 0 u 0 0 ρ ′′ u ′′ dψ′ . (10.25)<br />

Hereinafter all velocities will be referred to the maximum velocity u max =<br />

√<br />

2cp T s,1 of the unburnt gases, and the ratio named U. Equation (10.25) is then<br />

written as follows<br />

∫<br />

ρ 1 U<br />

ψ<br />

ρ 1 U<br />

− 1 + ψ =<br />

ρ 0 U 0 0 ρ ′′ U ′′ dψ′ . (10.26)<br />

The problem lies now in expressing ρ 1 U<br />

ρ ′′ U ′′ as a function of U and of the dimensionless<br />

velocity U ′ at the point D where the streamline ψ ′ crosses the flame,<br />

Fig. 10.9. Let us see how it can be done.<br />

Proceeding in the same manner as done for the <strong>de</strong>duction of equations (10.4)<br />

and (10.8) in Tsien’s method, that is, by comparing the expansions along ψ and ψ ′ ,<br />

the following is obtained<br />

λ ′ = ρ ′′<br />

1 T<br />

= . (10.27)<br />

ρ<br />

′′<br />

T 1


10.3. METHOD OF FABRI-SIESTRUNCK-FOURÉ 257<br />

Moreover, it is immediately obtained that<br />

λ ′ = n − U ′ 2<br />

1 − U ′ 2 , (10.28)<br />

from which results<br />

′′<br />

ρ 1 T n − U ′ 2<br />

=<br />

ρ<br />

′′<br />

T 1 1 − U ′ 2 . (10.29)<br />

Bernoulli’s equation when applied to E, gives<br />

U ′′ 2 +<br />

T ′′<br />

T s,1<br />

= n. (10.30)<br />

Likewise, Bernoulli’s equation when applied to C for the unburnt gases, gives<br />

U 2 + T 1<br />

T s,1<br />

= 1. (10.31)<br />

By combining (10.29), (10.30) and (10.31), the following is obtained for U ′′<br />

√<br />

U ′′ nU<br />

=<br />

2 − U ′2 (n − 1 + U 2 )<br />

1 − U ′ 2<br />

. (10.32)<br />

Substituting (10.29) and (10.32) into (10.26), and making use of relation<br />

ρ 1<br />

ρ 0<br />

=<br />

the following equation is finally obtained<br />

( 1 − U<br />

2<br />

1 − U 2 0<br />

) 1<br />

γ − 1 , (10.33)<br />

( ) 1<br />

U 1 − U<br />

2<br />

γ − 1 − 1 + ψ<br />

U 0 1 − U0<br />

2 ∫ ψ<br />

= √ U (n − U ′2 )<br />

√ n<br />

0<br />

(1 − U ′2 )<br />

(U 2 − U ′ 2 n − 1 + U 2 ) dψ′ .<br />

n<br />

(10.34)<br />

This expression is an integral equation which <strong>de</strong>termines ψ as a function of U. Once<br />

the solution is known, the value of ψ gives the burnt fraction as a function of the<br />

velocity U of the gases at the point where the streamline ψ crosses the flame front.<br />

The flame width η = y/h is given by the expression<br />

η = 1 − ρ 0U 0<br />

(1 − ψ). (10.35)<br />

ρ 1 U<br />

To <strong>de</strong>termine the shape of the flame front, the abscissa ξ = x/h corresponding to ψ<br />

must also be known. This abscissa is given by the following expression<br />

ξ =<br />

∫ ψ<br />

0<br />

ρ 0 U 0<br />

ρ 1 ψ ′ dψ′ , (10.36)


258 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

which is <strong>de</strong>duced from (10.19) by making v ≃ −ψ. When computing this expression<br />

the following relation must be used<br />

ρ ′ 1<br />

ρ 0<br />

=<br />

(<br />

) 1<br />

1 − U ′ 2 γ − 1<br />

1 − U0<br />

2<br />

(10.37)<br />

as well as the values ψ ′ = ψ(U ′ ) given by the solution of (10.34).<br />

Equation (10.34) can be integrated numerically by substituting the integral for<br />

a summation with a finite number of terms. Fabri, Siestrunck and Fouré have, thus,<br />

calculated the solutions corresponding to several typical cases. Some of their results<br />

can be found in the references inclu<strong>de</strong>d herein.<br />

When calculating the solution it is found that for each value of n there is a<br />

corresponding value U 0,cr of U 0 such that if U 0 > U 0,cr the curve of ψ = ψ(U)<br />

presents an horizontal tangent for a value of ψ smaller than one. This means that in<br />

such a case the burnt fraction must be smaller than unity. Therefore, the existence<br />

of a critical Mach number for the inlet flow is also obtained here, thereupon choking<br />

occurs. Furthermore, it can be proved that the value of U 0,cr is equal to the value<br />

U 0,cr = (√ n − √ n − 1 ) √ γ − 1<br />

γ + 1 , (10.38)<br />

even by the one-dimensional theory that results from the assumption that at each crosssection<br />

of the chamber the distribution of velocities is uniform. For such a case the<br />

value of the final Mach number of the burnt gases is unity. 3<br />

Fig. 10.10, taken from Ref. [8], shows a solution of (10.34) that corresponds<br />

to the case n = 6 for the three following values of U 0 : subcritical value U 0 = 0.05,<br />

critical value U 0 = U 0,cr = 0.087, and supercritical value U 0 = 0.100. In the latter,<br />

the burnt fraction would be only ψ max = 0.8.<br />

The experimental evi<strong>de</strong>nce available is not enough to judge the good approximation<br />

of these methods.<br />

10.4 Cylindrical chambers<br />

As Fabri, Siestrunck and Fouré have <strong>de</strong>monstrated [6] the case of a cylindrical chamber<br />

of circular cross-section reduces to the two-dimensional problem studied in the<br />

preceding paragraph. In fact, in such case equation (10.18), which <strong>de</strong>fines the stream<br />

3 See chapter 3, §9.


10.5. CHAMBER WITH SLOWLY VARYING CROSS-SECTION 259<br />

0.5<br />

0.4<br />

0.3<br />

U 0<br />

= 0.100<br />

U<br />

0.2<br />

U 0<br />

= 0.087<br />

0.1<br />

U 0<br />

= 0.050<br />

0<br />

0.2 0.4 0.6 0.8 1.0<br />

ψ<br />

Figure 10.10: Velocity profiles as a function of the burnt mass fraction for λ = 6 and three<br />

values of the initial velocity (subcritical, critical and supercritical).<br />

function, must be substituted by<br />

ρu = 1 2 ρ 0u 0<br />

R 2<br />

r<br />

∂ψ<br />

∂r , (10.39)<br />

where R is the radius of the chamber and r is the distance to its axis. Coefficient 1/2<br />

is introduced so that stream function ψ changes from value zero at the axis to value 1<br />

at the chamber wall.<br />

The change of variable<br />

( r<br />

) 2 y =<br />

R h , (10.40)<br />

transforms (10.39) into (10.18) keeping conditions ψ = 0 at y = 0 and ψ = 1 at<br />

y = 1. The remaining calculations are i<strong>de</strong>ntical in both cases. Hence, the problem<br />

reduces to the two-dimensional case. Fabri, Siestrunck and Fouré have also studied the<br />

case where stabilizer is at the chamber wall. The annular stabilizer has been studied<br />

by Ernst [9]. All these cases reduce to the two-dimensional problem by an a<strong>de</strong>quate<br />

change of variables.<br />

10.5 Chamber with slowly varying cross-section<br />

The preceding method can be applied to an approximate study by substituting equations<br />

(10.18) or (10.39) for an equation that inclu<strong>de</strong>s the variation of the cross-section


260 CHAPTER 10. AEROTHERMODYNAMIC FIELD OF A STABILIZED FLAME<br />

along the chamber axis. The same is done for the approximate study of gas motions<br />

within ducts with slowly varying cross-section.<br />

References<br />

[1] Scurlock, A. C.: Flame Stabilization and Propagation in High-Velocity Gas<br />

Streams. Meteor Report No. 19, Fuels Research Laboratory, M.I.T., May 1948.<br />

[2] Williams, G. C., Hottel, H.C. and Scurlock, A. C.: Flame Stabilization and<br />

Propagation in High Velocity Gas Streams. Third Symposium (International)<br />

on Combustion, Williams and Wilkins Co., Baltimore, 1949, pp. 21-40.<br />

[3] Ball, G.: A Study of a Two-Dimensional Flame. Combustion Tunnel Laboratory<br />

Report, Harvard University, Cambridge Mass., 1951.<br />

[4] Emmons, H. W.: Fundamentals of Gas Dynamics. Section F, vol. III of High<br />

Speed Aerodynamics and Jet Propulsion, Princeton University Press, 1956<br />

[5] Tsien, H. S.: Influence of Flame Front on the Flow Field. Journal of Applied<br />

Mechanics, June 1951, pp. 188-194.<br />

[6] Fabri, J., Siestrunck, R. and Fouré, C.: Sur le Calcul du Champ Aérodynamique<br />

<strong>de</strong>s Flammes Stabilisées. Comtes Rendus <strong>de</strong>s Séances <strong>de</strong> l’Académie <strong>de</strong>s Sciences,<br />

Paris, 1951, p. 1263.<br />

[7] Fabri, J., Siestrunk, R. and Fouré, C.: Phénomène d’Obstruction Intervenant<br />

dans la Stabilisation <strong>de</strong>s Flammes. La Recherche Aéronautique, No. 25, 1952,<br />

pp. 21-27.<br />

[8] Fabri, J., Siestrunk, R. and Fouré, C.: On the Aerodynamic Field of Stabilized<br />

Flames. Fourth Symposium (International) on Combustion, Williams and<br />

Wilkins Co., Baltimore, 1953, pp. 443-450.<br />

[9] Ernst, G.: Propagation a Faible Vitesse d’une Flamme dans un Ecoulement<br />

Compressible. Technique et Science Aéronautiques, 1955, pp. 1-12.


Chapter 11<br />

Similarity in combustion.<br />

Applications<br />

11.1 Introduction<br />

The methods of dimensional analysis and physical similarity have been applied with<br />

very good results to the study, both theoretical and experimental, of classical Aerothermodynamics.<br />

Recently, numerous attempts have been ma<strong>de</strong> to extend these methods<br />

to the study of Aerothermochemistry. In some instances, like in the work by Schultz-<br />

Grunow listed in Ref. [1], such an extension was inten<strong>de</strong>d in or<strong>de</strong>r to facilitate the<br />

establishment of a correlation between several experimental results. Other attempts<br />

tried to find a rational basis for the experimentation with mo<strong>de</strong>ls as well as rules to<br />

apply the results obtained through such mo<strong>de</strong>ls to the processes at normal scale.<br />

Before treating a problem of this nature, it is necessary to <strong>de</strong>termine the dimensionless<br />

parameters that characterize the process. In the phenomena whose study<br />

is the purpose of Aerothermochemistry, such parameters are the same studied by classical<br />

Aerothermodynamics, but increased by the chemical parameters introduced by<br />

Damköhler [2] when he applied physical similarity to the study of chemical reactors.<br />

We owe to Penner [3] the first study on the application of the laws of physical similarity<br />

to Aerothermochemistry. A concise work on the subject, analyzing the origin<br />

and significance of these parameters, is own to von Kármán and listed in Ref. [4].<br />

Recently, Weller [5] has performed a review of the practical applications. Several applications<br />

of this theory to the study of technological problem of combustion will be<br />

found in Penner’s work and in the corresponding sections of the Proceedings of Sixth<br />

International Symposium on Combustion and of Second Colloquium on Combustion,<br />

AGARD.<br />

261


262 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

In the following we will first <strong>de</strong>duce the dimensionless parameters of Aerothermochemistry<br />

starting from the fundamental equations governing motion. Thereafter<br />

two practical application will be performed, one to the rule of the scaling of rockets<br />

and another to the problem of flame stabilization.<br />

11.2 Dimensionless parameters of Aerothermochemistry<br />

The physical similarity between two analogous processes consists in the proportionality<br />

between corresponding magnitu<strong>de</strong>s (velocities, temperatures, etc.) through space<br />

and time. In the first place, it implies a geometrical similarity and, moreover, proportionality<br />

of times.<br />

The origin of this similarity and the required conditions for it to occur may be<br />

conceived as follows. Let us consi<strong>de</strong>r the equations of the process and write them in<br />

dimensionless form, referring each one of the variable quantities comprised in them<br />

(length, pressure, velocity, temperature, etc.) to a characteristic value of the same.<br />

Thus, we shall obtain a system of equations between dimensionless variables, with<br />

coefficients that will also be dimensionless combinations of such characteristic values.<br />

All the combinations of these values for which the coefficients are invariable will<br />

correspond to processes represented by i<strong>de</strong>ntical systems of equations. If, furthermore,<br />

we keep invariable the dimensionless expressions of the initial and boundary<br />

conditions, then the solution of the system thus obtained will correspond to a physically<br />

similar set of processes. We will pass from one another of these processes<br />

by introducing changes into the characteristic values which keep invariable the said<br />

coefficients of the equations and their initial and boundary conditions. Hence, these<br />

coefficients are the dimensionless parameters of the physical similarity. Therefore, the<br />

problem reduces to finding, among them, those in<strong>de</strong>pen<strong>de</strong>nt from one another, and to<br />

select the most simple expressions possible for the same.<br />

Let us see now in what way this can be attained, utilizing, for simplicity and<br />

clearness, the equations corresponding to an one-dimensional stationary flow with<br />

only two chemical species, since the generalization to other cases is straightforward<br />

and it only requires multiplying the number of parameters.<br />

The system of equations that we must apply to this study was <strong>de</strong>duced in §2 of<br />

Chap. 5, but it is reproduced here in or<strong>de</strong>r to assist the rea<strong>de</strong>r:<br />

a) Equation of mass conservation.<br />

ρv = m. (11.1)


11.2. DIMENSIONLESS PARAMETERS OF AEROTHERMOCHEMISTRY 263<br />

b) Continuity equation.<br />

c) Momentum equation.<br />

d) Energy equation.<br />

( )<br />

1<br />

ρv<br />

2 v2 + c p T − qε<br />

e) Diffusion equation.<br />

ρv dε = w. (11.2)<br />

dx<br />

p + ρv 2 − 4 3 µ dv = i. (11.3)<br />

dx<br />

D<br />

v<br />

dY<br />

dx<br />

− λ dT<br />

dx − 4 dv<br />

µv = me. (11.4)<br />

3 dx<br />

− Y + ε = 0. (11.5)<br />

Be l 0 , ρ 0 , v 0 , w 0 , p 0 , µ 0 , c p0 , T 0 , λ 0 and D 0 characteristic values of the corresponding<br />

quantities and let us change the variables of the preceding system as follows<br />

x = l 0 x ′ , ρ = ρ 0 ρ ′ , v = v 0 v ′ , w = w 0 w ′ , p = p 0 p ′ ,<br />

µ = µ 0 µ ′ , c p = c p0 c ′ p, T = T 0 T ′ , λ = λ 0 λ ′ , D = D 0 D ′ .<br />

(11.6)<br />

Thus all the “prime” quantities are dimensionless. For ε and Y this change is<br />

not necessary since both are already dimensionless and their variation field has been<br />

selected from zero to one.<br />

When these new variables (11.6) are introduced into the preceding system it<br />

may be written as follows:<br />

a) Equation of mass conservation.<br />

ρ ′ v ′ =<br />

( m<br />

ρ 0 v 0<br />

)<br />

. (11.7)<br />

b) Continuity equation. ( )<br />

ρ0 v 0<br />

ρ ′ w ′ dε<br />

l 0 w 0 dx = w′ . (11.8)<br />

c) Momentum equation.<br />

( )<br />

p0<br />

ρ 0 v0<br />

2 p ′ + ρ ′ v ′ 2 4 −<br />

3<br />

d) Energy equation.<br />

ρ ′ v ′ [ 1<br />

2<br />

( v<br />

2<br />

0<br />

)<br />

v ′ 2 + c<br />

′<br />

c p0 T p T ′ −<br />

0<br />

(<br />

− 4 3<br />

(<br />

µ0<br />

) (<br />

µ ′ dv′<br />

ρ 0 v 0 l 0 dx ′ =<br />

( ) ] (<br />

q<br />

ε −<br />

c p0 T 0<br />

µ 0 v0<br />

2 )<br />

µ ′ v ′ dv′<br />

l 0 ρ 0 v 0 c p0 T 0<br />

i<br />

ρ 0 v 2 0<br />

)<br />

. (11.9)<br />

)<br />

λ 0 T 0<br />

λ ′ dT ′<br />

l 0 ρ 0 v 0 c p0 T 0 dx ′<br />

( )<br />

dx = me<br />

. (11.10)<br />

ρ 0 v 0 c p0 T 0


264 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

e) Diffusion equation.<br />

( )<br />

D0 D<br />

′<br />

v 0 l 0 v ′<br />

dY<br />

dx<br />

− Y + ε = 0. (11.11)<br />

The set of dimensionless coefficients P i characteristic of the system is<br />

P 1 = ρ 0v 0<br />

, P 2 = p 0<br />

l 0 w 0 ρ 0 v0<br />

2<br />

P 5 =<br />

q<br />

c p0 T 0<br />

, P 6 =<br />

, P 3 = µ 0<br />

,<br />

ρ 0 v 0 l 0<br />

P 4 = v2 0<br />

,<br />

c p0 T 0<br />

λ 0<br />

, P 7 = µ 0v 0<br />

,<br />

l 0 ρ 0 v 0 c p0 l 0 ρ 0 c p0 T 0<br />

P 8 = D 0<br />

.<br />

v 0 l 0<br />

(11.12)<br />

The physical similarity requires that the values of all these coefficients be equal for<br />

the two processes compared, when using as characteristic values of the variables for<br />

their calculations those at corresponding points and instants.<br />

Such equality guarentees the i<strong>de</strong>ntity of the left-hand si<strong>de</strong>s of Eqs. (11.7) to<br />

(11.11) in both phenomena. If, furthermore, we guarentee as well the equality of the<br />

right-hand si<strong>de</strong>s, which correspond to the boundary conditions of the problem, we will<br />

have insured the i<strong>de</strong>ntity of the solution, thus reaching the physical similarity of the<br />

phenomena. Let us proceed to discuss the set of parameter of Eq. (11.12).<br />

We readily observe that P 7 is equal to the product of P 3 by P 4<br />

P 7 = P 3 · P 4 , (11.13)<br />

while the remaining parameters are in<strong>de</strong>pen<strong>de</strong>nt from one another. Consequently, the<br />

number of in<strong>de</strong>pen<strong>de</strong>nt dimensionless parameters is seven. Let us see which are these<br />

seven.<br />

by it<br />

It is clear that P 3 is reciprocal to the Reynolds number and may be substituted<br />

1) Re = ρ 0v 0 l 0<br />

µ 0<br />

. (11.14)<br />

The combination of P 2 and P 4 gives the following parameter P ′ 4 which may<br />

substitute P 4<br />

P ′ 4 =<br />

p 0<br />

ρ 0 c p0 T 0<br />

. (11.15)<br />

Yet, the state equation p 0<br />

ρ 0<br />

= R g T 0 allows P ′ 4 to be written<br />

P ′ 4 = R g<br />

c p0<br />

= c p0 − c v0<br />

c p0<br />

= 1 − 1 γ . (11.16)<br />

Consequently, the constancy of P ′ 4 is equivalent to the constantcy of relation (11.16)<br />

between specific heats, which supplies the second characteristic dimensionless parameter


11.2. DIMENSIONLESS PARAMETERS OF AEROTHERMOCHEMISTRY 265<br />

2) γ = c p0<br />

c v0<br />

. (11.17)<br />

Keeping in mind that velocity a 0 of sound in the mixture, at pressure p 0 and<br />

with <strong>de</strong>nsity ρ 0 , is a 0 = √ γp 0 /ρ 0 , we can promptly see that P 2 may be written in the<br />

form<br />

being<br />

P 2 = 1<br />

γM0<br />

2 , (11.18)<br />

3) M 0 = v 0<br />

a 0<br />

, (11.19)<br />

the Mach number of the flow, which will be used as a third characteristic dimensionless<br />

parameter in lieu of P 2 .<br />

The combination of P 3 and P 6 gives the following value<br />

4) Pr = P 3<br />

P 6<br />

= µ 0c p0<br />

λ 0<br />

, (11.20)<br />

which is the Prandtl number, fourth dimensionless parameter the substitutes P 6 .<br />

Likewise, the combination of P 3 and P 8 gives<br />

5) Sc = P 3<br />

P 8<br />

= µ 0<br />

ρ 0 D 0<br />

, (11.21)<br />

which is the Schmidt number, fifth dimensionlees parameter in lieu of P 8 .<br />

The aforegoing parameters are the same used in classical Aerothermodynamics<br />

and the characteristic constants of the chemical reaction have no bearing on them. This<br />

constants act through two other parameter, P 1 and P 5 . The first may be written<br />

where<br />

P 1 = v 0τ ch0<br />

l 0<br />

, (11.22)<br />

τ ch0 = ρ 0<br />

w 0<br />

(11.23)<br />

is a characteristic chemical time. The reciprocal of (11.22) is the first parameter of<br />

Damköhler,<br />

6) Da 1 = l 0<br />

v 0 τ ch0<br />

, (11.24)<br />

and it is the sixth dimensionless parameter of the process.<br />

Finally, P 5 is the second parameter of Damköhler<br />

7) Da 2 = q<br />

c p0 T 0<br />

, (11.25)<br />

this being the seventh dimensionless parameter of the process.


266 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

The interpretation of the meaning of Damköhler two parameters is simple. The<br />

first is a measurement of the relationship between the mean mechanical time required<br />

by a particle to travel the characteristic lenght l 0 at the characteristic velocity v 0 and<br />

the time required by the chemical reaction of the mixture to take place. The second<br />

parameter of Damköhler is a measure of the relationship between the heat produced<br />

by unity of mixture when burning at constant pressure and the heat contained by the<br />

same at characteristic temperature T 0 .<br />

It should be noted that in the non-stationary processes a additional parameter<br />

appears, Strouhal number, which originates from the time <strong>de</strong>rivatives ∂/∂t of the<br />

equations of motions. Likewise, if the mass forces are present as occurs in the phenomena<br />

of free convection, then another dimensionless parameter appears, due to the<br />

gravity terms of the equations, which originates the Frou<strong>de</strong> number.<br />

Finally, when the number of species and chemical reaction is increased, the<br />

number of the Damköhler parameters (of both kinds time and heat) increases as well.<br />

In technical literature we frequently find other parameter which are combinations<br />

of the above mentioned. Por example, occasionally Prandtl number is substituted<br />

by Peclet number, which is a product of the numbers of Reynolds and Prandtl<br />

Pe = Pr · Re = ρ 0v 0 l 0 c p0<br />

λ 0<br />

. (11.26)<br />

Likewise, the Schmidt number is replaced by the Lewis-Semenov number which is<br />

obtained as follows<br />

L = Pr<br />

Sc = ρ 0D 0 c p0<br />

, (11.27)<br />

λ 0<br />

used specially in combustion processes as seen in Chap. 6.<br />

The advantage of utilizing the Peclet and Lewis-Semenov numbers in combustion<br />

phenomena is based on the fact that the conductivity and diffusion coefficients<br />

appear explicity in their expressions. Such coefficients are the ones of interest in these<br />

processes in which the influence of the viscosity coefficient is negligible in many<br />

cases, as said in Chap. 5.<br />

We shall now see some practical applications of this theory.<br />

11.3 Scaling of rockets<br />

The problem of the scaling of rockets has a great practical interest since its solutions<br />

would allow the prediction of the behavior of the rockets after tests ma<strong>de</strong> with reduced<br />

scale mo<strong>de</strong>ls.


11.3. SCALING OF ROCKETS 267<br />

Furthermore, both the scaling un<strong>de</strong>r steady operation as well as un<strong>de</strong>r low and<br />

high frequency oscillations are important in practice. This problem was recently studied<br />

by Penner-Tsien [6], Rose [7], Crocco [8], Barrère [9] and Penner-Fuhs [10].<br />

We have seen in §2 that physical similarity imposses that the seven characteristic<br />

parameters be equal, when passing from the mo<strong>de</strong>l to the rocket un<strong>de</strong>r steady state<br />

conditions and some additional ones un<strong>de</strong>r non-steady conditions.<br />

We shall now study the scaling rules imposed by these equalities. If we assume<br />

that we are working with the same mixture of gases and at the same temperature, the<br />

equality of γ, as well as that of the Prandtl and Schmidt numbers and of Da 2 , is insured.<br />

Consequently we only have to insure the equality of the other three parameters.<br />

Disregarding subscript zero and using subscripts 1 for the mo<strong>de</strong>l and 2 for the rocket,<br />

we will have:<br />

a) Equality of the Reynolds Number. The equality of temperatures makes µ 1 = µ 2 .<br />

Furthermore, ρ may be substituted by p, to which it is proportional. Therefore,<br />

this condition reduces to the following<br />

p 1 v 1 l 1 = p 2 v 2 l 2 . (11.28)<br />

b) Equality of the Mach Number. The equality of temperature insures the equality<br />

of the velocity of sound a 1 = a 2 . Hence, this condition reduces to the following<br />

v 1 = v 2 , (11.29)<br />

this is to say that the velocity must be the same in both the mo<strong>de</strong>l and rocket.<br />

c) Equality of Da 1 . This condition gives<br />

l 1<br />

v 1 τ ch1<br />

= l 2<br />

v 2 τ ch2<br />

. (11.30)<br />

Since l is the length of the combustion chamber and v the gas velocity through it,<br />

relation<br />

τ r = l v<br />

(11.31)<br />

is the resi<strong>de</strong>nce time of the gases in the chamber. τ ch is a physico-chemical time, characteristic<br />

of the process, the so called time lag or <strong>de</strong>lay from the instant at which the<br />

fuel enters the chamber to the instant at which it transforms into products. Therefore,<br />

it accounts for the time nee<strong>de</strong>d by the fuel to form droplets, to evaporate, to mix with<br />

the gases and burn. Hence, Da 1 can be written in the form<br />

Da 1 = τ r<br />

τ ch<br />

. (11.32)


268 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

Let<br />

K = l 1<br />

l 2<br />

(11.33)<br />

be the linear relation of sizes. Then from (11.28), (11.29), (11.30) and (11.33) one<br />

<strong>de</strong>rives<br />

p 1<br />

= 1 p 2 K , (11.34)<br />

τ ch1<br />

= K.<br />

τ ch2<br />

(11.35)<br />

The first condition imposses that the variation of pressure be inversely proportional<br />

to size. The second one cannot be insured since in a rocket, as we have seen,<br />

τ ch is the resultant of a complicated physico-chemical process. Crocco, by means of<br />

a <strong>de</strong>tailed analysis [11], has reached the conclusion that τ ch may be represented, in<br />

many cases, through a <strong>de</strong>creasing function of pressure in the form<br />

where n is an exponent differing slightly from unity.<br />

τ ch ∼ p −n , (11.36)<br />

If n = 1, by taking (11.36) into (11.34) and comparing with (11.35), we see<br />

that the later is satisfied, thus obtaining a complete physical similarity.<br />

If n is different from unity the complete similarity is not possible. In such case,<br />

if we disregard the condition of equality of the Mach numbers, which is actually not<br />

too important un<strong>de</strong>r steady-state conditions, since velocities at the combustion chamber<br />

are very small and compressibility effects can be neglected, the following conditions<br />

are obtained, instead of (11.29). By eliminating v 1 and v 2 between Eqs. (11.28)<br />

and (11.26), which remain valid, and taking into account Eq. (11.33), it results<br />

p 1<br />

p 2<br />

= τ ch1<br />

τ ch2<br />

K −2 . (11.37)<br />

If expression (11.36) is assumed for the time lag, the following condition is obtained<br />

for the ratio of pressures<br />

p 2<br />

1<br />

= K − 1 + n , (11.38)<br />

p 2<br />

whereas, from here and Eqs. (11.28) and (11.33), it results for the ratio at velocities<br />

1 − n<br />

v 1<br />

= K 1 + n . (11.39)<br />

v 2<br />

Penner and Tsien [6] adopt a different view point. They also neglect the equality<br />

of Mach numbers, but assume that both rockets operate at the same pressure<br />

p 1 = p 2 . (11.40)


11.3. SCALING OF ROCKETS 269<br />

From here and Eq. (11.37), one obtains<br />

τ ch1<br />

τ ch2<br />

= K 2 . (11.41)<br />

The authors propose that these condition be satisfied by proper control of the<br />

time lag, which may be reached, for instance, if the mean size of the droplets is conveniently<br />

varied, since the evaporation time of a droplet <strong>de</strong>pends on its size as will be<br />

shown in chapter 13.<br />

Now let us see the scaling rules for the thrust and injector.<br />

Physical similarity implies that the number of injectors be the same in both<br />

rockets and i<strong>de</strong>ntically distributed.<br />

Be e the thrust, g the flow rate of fuel, v 1 its velocity through the injector and<br />

d its diameter. It is promptly verified that the following system of relations is valid<br />

provi<strong>de</strong>d the temperature is preserved.<br />

e ∼ g ∼ v i d 2 ∼ ρvl 2 ∼ pvl 2 , (11.42)<br />

From here, the following system of scaling conditions is <strong>de</strong>rived<br />

e 1<br />

= g 1<br />

= v i1 d 2 1<br />

e 2 g 2 v i2 d 2 = p 1v 1 l1<br />

2<br />

2 p 2 v 2 l2<br />

2 . (11.43)<br />

Moreover, since the velocity fields must be similar, it results<br />

v i1<br />

= v 1<br />

. (11.44)<br />

v i2 v 2<br />

When Eqs. (11.28) and (11.33), which in all cases remain valid, are taken into<br />

Eq. (11.43), one obtains for the ratio of thrusts<br />

e 1<br />

e 2<br />

= K. (11.45)<br />

The combination of this relation with Eqs. (11.43) and (11.44) gives for the ratio of<br />

injector diameters<br />

√<br />

d 1<br />

= K v 2<br />

. (11.46)<br />

d 2 v 1<br />

Hence, the ratio of diameters <strong>de</strong>pends on the scaling law for the velocities at<br />

the chamber, which, as we have seen, <strong>de</strong>pends on the rule applied. For example, for<br />

the Penner-Tsien rule it results, by virtue of Eq. (11.28),<br />

d 1<br />

d 2<br />

= K, (11.47)<br />

whereas for Crocco’s rule, from Eq. (11.39) it is obtained<br />

n<br />

d 1<br />

= K 1 + n . (11.48)<br />

d 2


270 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

11.4 Scaling of rockets for non-steady conditions<br />

The motion on gases through the combustion chamber of a rocket is always strongly<br />

turbulent with oscillations of pressure, velocities, etc., distributed at random in time<br />

and space, which does not cause important disturbances in the normal operation of the<br />

rocket. Un<strong>de</strong>r these circumstances the combustion is called steady.<br />

However, there is also the case, frequently observed, where combustion is unstable<br />

with strong self-excited periodic oscillations of pressure which in some case<br />

<strong>de</strong>stroy the rocket in a few seconds. This phenomenon is at present one of the main<br />

difficulties for the <strong>de</strong>velopment of rockets, specially large ones. It was observed for<br />

the first time in the United States in 1941 by the research staff of the Jet Propulsion<br />

Laboratory of the California Institute of Technology un<strong>de</strong>r the guidance of Professor<br />

von Kármán. Thereafter the great effort applied to the study of this problem has<br />

cons<strong>de</strong>rably increased the technical literature on the subject.<br />

Ross-Datner [11] and Crocco-Gray-Grey [12] have written two excellent reviews<br />

on this problem <strong>de</strong>scribing the kinds of instabilities observed, their causes, experimental<br />

techniques and the results of the measurements performed, in addition to<br />

the fundamentals of the theories available. These reviews inclu<strong>de</strong> as well an extensive<br />

bibliography. The most up to date and complete work on the subject is the one<br />

performed by Crocco and Cheng [13] which was recently published by AGARD, including<br />

an abundant bibliography.<br />

Initially, due to limitations of the instruments available for observation, it was<br />

only possible to isolate one type of low-frequency oscillations of the or<strong>de</strong>r of about<br />

100 cycles per second, which is normally known as chugging. Later on, as the instrumentation<br />

was improved, it became feasible to isolate other high-frequency oscillations,<br />

over 1000 c.p.s., generally called screaming.<br />

Soon, the origine of the low-frequency oscillations was known and it was possible<br />

to <strong>de</strong>rive practical rules to prevent them which resulted efficient when applied to<br />

practice. Summerfield [14] stated the fundamentals of the theory and the preventive<br />

rules born from it. In low-frequency oscillations, the combustion chamber behaves<br />

like a resonating cavity whose oscillations of pressure act on the feeding system,<br />

changing the rate of fuel injected. The influence of this variation reflects on the chamber<br />

with a certain <strong>de</strong>lay which is due partially to the relaxation times of the chamber<br />

and of the feeding system, and partially to the physico-chemical time lag <strong>de</strong>scribed<br />

in the preceding paragraph. If this <strong>de</strong>lay is the a<strong>de</strong>quate one, the oscillations result<br />

self-excited. Consequently, the unit combustion chamber-feeding system behaves as a<br />

dynamic system with a characteristic time lag, able of producing unstable oscillations.


11.4. SCALING OF ROCKETS FOR NON-STEADY CONDITIONS 271<br />

2<br />

0<br />

Chamber pressure<br />

t<br />

−2<br />

−4<br />

−6<br />

Injection pressure drop<br />

Injection rate<br />

τ i<br />

t<br />

t<br />

−8<br />

−10<br />

Burning rate<br />

τ ch<br />

τ c<br />

t<br />

−12<br />

−14<br />

Effect of burning rate<br />

Half period<br />

t<br />

Figure 11.1: Time <strong>de</strong>lays leading to self-exciting oscillationes.<br />

Figure 11.1, taken from Ref. [13], clearly illustrates the mechanisn of the process<br />

capable of producing self-excited oscillations. In this figure, it is assumed that the<br />

feeding system is of the constant pressure type. Therefore, the pressure drop through<br />

the injector, and with it the rate of fuel, <strong>de</strong>creases as the pressure of the chamber is<br />

increased. However, the variation in rate of fuel follows the drop in pressure with a<br />

certain <strong>de</strong>lay τ i , <strong>de</strong>termined by the dynamic characteristics of the feeding system. The<br />

fuel injected turns into combustion products with a <strong>de</strong>lay τ ch , which is the time lag,<br />

and such products act in turn on the pressure of the chamber with a <strong>de</strong>lay τ c , which<br />

<strong>de</strong>pends on the dynamic characteristics of the same. If the sum τ i + τ ch + τ c is equal<br />

to half the period of the pressure oscillations, then there is a phase agreement between<br />

cause an effect and the oscillations amplify with the only limitation of damping effects.<br />

The above analysis shows that in or<strong>de</strong>r to eliminate oscillations it is necessary<br />

to reduce τ i , τ ch and τ c . The measures taken to this effect confirm theoretical prediction.<br />

Two efficient ways, for example, are to reduce the relation between pressure drop<br />

through the injector and the pressure in the chamber, or to reduce τ ch by eliminating<br />

the propellants having an excessive time lag [15].<br />

Asi<strong>de</strong> from the aforegoin causes, Crocco [15] has pointed out an intrinsic cause<br />

for instability, so called because it <strong>de</strong>pends only on the conditions in the combustion<br />

chamber and not on the interaction between it and the feeding system, indispensable,<br />

as it was established, for the previously analyzed case. The intrinsic instability of<br />

Crocco shows the possible existence of low-frequency oscillations in a rocket with a<br />

constant volume feeding system, which is impossible with the preceding mo<strong>de</strong>l.


272 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

We have said in the preceding paragraph of this chapter, that Crocco has proven<br />

that time lag τ ch is not constant but a function of the conditions in the chamber. In<br />

his study he schematizes this by making τ ch <strong>de</strong>pending on the pressure, as shown in<br />

Eq. (11.36). In intrinsic instability the variation of τ ch with p plays the same part than<br />

the consumption of fuel in Figure 11.1, since the element acting on the pressure of<br />

the chamber is not the fuel rate injected into it, but the rate of fuel transformed into<br />

products, which may vary from one instant to another, even if the former is constant,<br />

when τ ch varies with pressure. Un<strong>de</strong>r such conditions, if a coupling of phases is<br />

reached we will have a self-excited oscillation as in the preceding case.<br />

Let us now see the rules obtained for the scaling of rockets, to maintain physical<br />

similarity, so that both rockets be equaly stable with respect to the low-frequency<br />

oscillations.<br />

In such case, asi<strong>de</strong> from the conditions <strong>de</strong>rived in the preceding paragraph, it<br />

is necessary that some additional ones be satisfied relative to feeding system, since its<br />

influence is important.<br />

In particular, we must keep the equality of the ratio of the pressure drop ∆p<br />

through the injector to the pressure in the chamber, that is<br />

the injector<br />

∆p 1<br />

p 1<br />

= ∆p 2<br />

p 2<br />

. (11.49)<br />

From here, one obtains the following scaling law for the pressure drop through<br />

∆p 1<br />

∆p 2<br />

= p 1<br />

p 2<br />

. (11.50)<br />

However, we must keep in mind that the pressure drop through the injector<br />

<strong>de</strong>termines velocity v i of the fuel when passing through the same, and that the ratio<br />

of such velocities has been <strong>de</strong>termined before by the condition of physical similarity<br />

between velocity fields in Eq. (11.44). Consequently we cannot insure the simultaneous<br />

satisfaction of conditions (11.50) and (11.44), except for certain particular cases<br />

or by adopting special precautions. For instance, if the injectors have similar friction<br />

characteristics, then the following ratio would exist between ∆p and v i<br />

and we have<br />

∆p ∼ v 2 i , (11.51)<br />

( ) 2<br />

∆p 1 vi1<br />

= , (11.52)<br />

∆p 2 v i2<br />

which is not compatible with Eq. (11.50), except for n = 2, as it can readily be<br />

verified.


11.4. SCALING OF ROCKETS FOR NON-STEADY CONDITIONS 273<br />

The similarity conditions that still need to be satisfied refer to the dynamic<br />

characteristics of the feeding system which <strong>de</strong>termine the value for τ i in Fig. 11.1.<br />

For their study, we refer the rea<strong>de</strong>r to the papers by Crocco [8] and Penner-Fuhs [10]<br />

previously mentioned.<br />

Low-frequency oscillations may be analyzed by adopting the assumption that<br />

the state of the gases in the chamber is constant, because their period is very long compared<br />

to the propagation time of a wave through the chamber. To the contrary, when<br />

both the period of the oscillations and the propagation time of the wave are of the<br />

same or<strong>de</strong>r of magnitu<strong>de</strong>, it is necessary to consi<strong>de</strong>r the differences in state between<br />

different points of the chamber. Such is the situation in the case of high-frequency<br />

oscillations, where, as shown by experimentation, the chamber oscillates like an organ<br />

pipe with longitudinal and transversal oscillations. The existence of a time lag<br />

makes possible the maintenance of these oscillations in a similar way as it happens<br />

for low-frequency oscillations. High-frequency oscillations are particularly dangerous<br />

because, in addition to pressure oscillations being very large, the transmision of<br />

heat to the walls or injectors increases drastically and they are <strong>de</strong>stroyed within a few<br />

seconds. The study of these oscillations is less advanced than for the low-frequency<br />

ones which are easier to control.<br />

As for the scaling law, it appears evi<strong>de</strong>nt that the resi<strong>de</strong>nce time τ r should be<br />

substituted by propagation time τ p of a pressure wave through the chamber. Therefore,<br />

in or<strong>de</strong>r to mantain the same level of stability for high-frequency oscillation when<br />

passing from the mo<strong>de</strong>l to the rocket, the following condition must be satisfied<br />

τ ch1<br />

τ ch2<br />

= τ p1<br />

τ p2<br />

. (11.53)<br />

However, since the velocity of sound is the same for both rockets, between τ p1 and<br />

τ p2 the following ratio exists<br />

τ p1<br />

τ p2<br />

= K. (11.54)<br />

Consequently we obtain<br />

τ ch1<br />

τ ch2<br />

= K. (11.55)<br />

It happens also as for low-frequency oscillations, that this condition can not always be<br />

satisfied. In particular, it can be verified that for n = 1 and applying Crocco’s rule the<br />

above condition may be attained.


274 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

11.5 Flame stabilization<br />

In the preceding chapter we have studied the aerodynamic field of a flow with a flame<br />

stabilized by a hol<strong>de</strong>r. We verified that the expansion of the burnt gases may originate<br />

a chocking effect which imposses a limitation to the maximum velocity of the flow<br />

entering the chamber.<br />

Another limitation to this velocity, bearing a great practical interest, is that impossed<br />

by the need to fix the flame to the hol<strong>de</strong>r in or<strong>de</strong>r to maintain combustion.<br />

In fact, it has been experimentally observed that when the velocity of the incoming<br />

flow exceeds a certain value, the flame blows out. The blowing velocity of the flame<br />

is <strong>de</strong>termined by the composition and state of the combustible mixture and by the<br />

characteristics of the hol<strong>de</strong>r. The problem arising from this circunstance holds a great<br />

practical interest in the new propulsion systems, specially in ramjets and after-burners,<br />

in which their limits of practical application are often <strong>de</strong>termined by this effect. Consequently,<br />

this phenomenon has been the subject of a particular interest during the last<br />

years and abundant experimental data have been gathered on a great variety of types of<br />

hol<strong>de</strong>r and on the influence of the parameters of the mixture on the blowing velocity<br />

of the flame. An extensive bibliography on the matter can be found in the proceedings<br />

of the congresses and colloquia on combustion celebrated in recent years.<br />

The initial study on this problem was carried out by Scurlock [16], who pointed<br />

out the significance of the recirculation zone that forms in the wake of the hol<strong>de</strong>r. Figure<br />

11.2 taken from Ref. [17] illustrates the phenomenon. Behind the hol<strong>de</strong>r a recirculation<br />

zone forms in which the flame stabilizes. The boundary of this zone consists<br />

in a mixing zone formed by a free boundary layer which separates the unburnt gases<br />

from the combustion products. Within this mixing zone is where the combustible<br />

gases ignite through a process of heat and mass transport, whose characteristics are<br />

not yet well known.<br />

Since the presence of a recirculation zone is essential for the existence of the<br />

flame, the hol<strong>de</strong>rs used for this purpose are given a shape which produces a large<br />

aerodynamic wake.<br />

In many of the tests conducted, attempts were ma<strong>de</strong> to establish a correlation<br />

between the blowing velocity of the flame, in a mixture of given composition and<br />

temperature, and the size of the hol<strong>de</strong>r, characterized by a linear dimension of the<br />

same, for example by its diameter in the case of a hol<strong>de</strong>r of circular section as the one<br />

shown in Fig. 11.2. A summary of the works performed from this stand-point, up to<br />

1953, and of the state or knowledge at the time may be found in Ref. [18].


11.5. FLAME STABILIZATION 275<br />

Flame front<br />

End of recirculation zone<br />

Recirculation zone<br />

Flame hol<strong>de</strong>r<br />

Mixing zone<br />

Propagating zone<br />

Figure 11.2: Zones of a stabilized flame.<br />

The theoretical study of this problem is very difficult. Several theories are<br />

available whose <strong>de</strong>velopment may be found in the works listed un<strong>de</strong>r Refs. [19]<br />

through [21], but none of them has been experimentally confirmed. On the other<br />

hand, the correlation between experimental results as presented by Longwell is very<br />

difficult. At this stage it appears justified to perform the study by means of a dimensionless<br />

analysis, searching for the significant variables and the relation existing<br />

between them.<br />

Inasmuch as we are far from the actual regime in which the compressibility<br />

effect would appear, the dimensionless variable characteristic of the motion is the<br />

Reynolds number referred to the velocity and state of the gases before the hol<strong>de</strong>r and<br />

to a linear dimension of the same. The composition of the mixture is characterized by<br />

the relationship between the fuel fraction and that corresponding to the stoichiometric<br />

one. Finally, the variable we are trying to analyze is the blowing velocity or, if written<br />

in dimensionless form although it is not generally necesary, this velocity referred to<br />

the propagation velocity of the flame.<br />

Zukoski and Marble [22] have performed a very interesting systematic analysis<br />

on the experimental data available. A summary of this analysis appears in Figs. 11.3<br />

and 11.4<br />

Figure 11.3, corresponding to a circular flame hol<strong>de</strong>r as the one shown in<br />

Fig. 11.2, gives the following: a) the influence of the Reynolds number on the maximum<br />

blowing velocity of the flame, and b) the composition of the mixture for which<br />

such velocity is reached, for a given Reynolds number.<br />

Figure 11.4 shows the law of variation of the maximum blowing velocity as a<br />

function of the Reynolds number for a set of experiments performed by other authors<br />

with different types of flame hol<strong>de</strong>rs.


276 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

Maximum blowoff<br />

velocity (ft/s) φ m<br />

2.0<br />

1.5<br />

1.0<br />

600<br />

400<br />

200<br />

100<br />

10<br />

2 4 6 8 10 2 4 6 8 10<br />

Reynolds number<br />

3 4 5<br />

Figure 11.3: Maximum blowoff velocity and corresponding equivalence ratio for a cylindrical<br />

flame hol<strong>de</strong>r.<br />

Maximum blowoff velocity (ft/s)<br />

1000<br />

800<br />

600<br />

400<br />

300<br />

200<br />

150<br />

100<br />

4<br />

2.5 4 6 8 10 1.5 2 3 4 6 8 10 5 1.5<br />

Reynolds number<br />

Figure 11.4: Maximum blowoff velocity for different types of flame hol<strong>de</strong>rs (see [22] for<br />

more <strong>de</strong>tails).<br />

Both figures show clearly the existence of a transition Reynolds number, approximately<br />

equal to 10 4 , which separates two different regions. For Re < 10 4 , the<br />

maximum blowing velocity corresponds to a composition different from the stoichiometric<br />

one, which furthermore varies with the Reynolds number, whilst if Re exceeds<br />

the transition number the composition remains constant. Likewise, for Re > 10 4<br />

the blowing velocity varies with the square root of the Reynolds number, while for<br />

Re < 10 4 it follows a different law.<br />

Zukoski and Marble, Ref. [17], have <strong>de</strong>monstrated that such a change is due to<br />

the fact that the free boundary layer of the mixing zone changes from laminar to turbulent.<br />

This transition is important because the transfer of heat and chemical species between<br />

unburnt gases and combustion products, within the mixing zone, occurs through<br />

laminar diffusion un<strong>de</strong>r subcritical conditions and through turbulence in the opposite<br />

case.


11.5. FLAME STABILIZATION 277<br />

Flame hol<strong>de</strong>r<br />

geometry<br />

D<br />

D<br />

1/8 inch<br />

3/16<br />

1/4<br />

τ i<br />

3.09 × 10 −4 s<br />

2.85<br />

2.80<br />

D<br />

1/4 3.00<br />

mesh screen<br />

D<br />

D<br />

1/4<br />

3/8<br />

1/2<br />

3/4<br />

1/4<br />

3/8<br />

1/2<br />

3/4<br />

2.38<br />

2.70<br />

2.65<br />

2.58<br />

3.46<br />

3.12<br />

3.05<br />

3.03<br />

D<br />

3/4 3.05<br />

D<br />

3/4 2.70<br />

Table 11.1: Ignition time for different types and dimensions of flame hol<strong>de</strong>rs.<br />

Since the existence of the flame is governed by the possibility of igniting the<br />

mixture, while it travels along the mixing zone, the resi<strong>de</strong>nce time τ r of the mixture<br />

within this zone must be an important factor of the phenomenon. The resi<strong>de</strong>nce time<br />

may be expressed by the ratio<br />

τ r = L (11.56)<br />

V<br />

of length L of the mixing zone to the velocity V of the gases when travelling along<br />

it, which is approximately the velocity of the free flow. Consequently, if this interpretation<br />

of the phenomenon is correct, the blowing velocity may be reached when τ r<br />

reaches a limiting value τ i , below which there is no time for ignition. Furthermore,<br />

τ i must <strong>de</strong>pend exclusively on the composition and state of the mixture and on the<br />

conditions at the free boundary layer, but not on the shape of the hol<strong>de</strong>r or any other<br />

mechanical factor.<br />

In or<strong>de</strong>r to check the validity of this assumption, Marble and Zukoski computed<br />

the value of τ i for a set of quite different cases. Results are summarized in Table


278 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

11.1 in which the constancy of the ignition time may be clearly seen. Hence, the<br />

characteristic dimensionless group <strong>de</strong>termining the blowing velocity of the flame is<br />

Damköhler’s first parameter<br />

and blowing velocity is reached for Da 1 = 1.<br />

Da 1 = L/V<br />

τ i<br />

, (11.57)<br />

The advantage of this study lies on the fact that it separates two variables. One,<br />

being mechanical, <strong>de</strong>pends only on the conditions of motion, whereas the other, being<br />

physico-chemical, <strong>de</strong>pends exclusively on the composition and state of the mixture.<br />

Fig. 11.5, from Ref. [17], shows the law of variation for τ i as a function of the mixture<br />

composition for a typical case.<br />

1.8<br />

1.6<br />

Characteristic ignition time (ms)<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0<br />

φ, Equivalence ratio<br />

Figure 11.5: Ignition time as a function of the composition of the mixture (see [17] for more<br />

<strong>de</strong>tails).<br />

With reference to the mechanical time, the problem reduces to a study of the<br />

relationship between the length L of the wake and the characteristics of the hol<strong>de</strong>r.<br />

For a great number of hol<strong>de</strong>rs it has been proved that, in the supercritical regime, this<br />

relation is of the form<br />

L ∼ D 1/2 , (11.58)<br />

which explains the law of variation of V with D observed by experimentation.<br />

There exist, however, other cases for which the following condition is satisfied<br />

L ∼ D, (11.59)


11.5. FLAME STABILIZATION 279<br />

which is the reason for the special behavior of the hol<strong>de</strong>rs observed by Longwell [18].<br />

In this formulae, D is a linear dimension of the hol<strong>de</strong>r.<br />

The problem which remains to be solved is the process in the mixing zone<br />

which <strong>de</strong>termines the value for τ i .<br />

References<br />

[1] Schultz-Grunow, F.: Similarity Laws of Deflagration. Fourth Symposium (International)<br />

on Combustion, Williams and Wilkins Co., Baltimore, 1953, pp. 439-<br />

443.<br />

[2] Damköhler, G.: Einflüsse <strong>de</strong>r Strömung, Diffusion and <strong>de</strong>s Warmeliberganges<br />

auf die Leistung von Reaktionsofen. Z. Elektrochem., Vol. 42, 1936, pp. 846-<br />

862.<br />

[3] Penner, S. S.: Similarity Analysis for Chemical Reactors and the Scaling of<br />

Liquid Fuel Rocket Engines. Combustion Researches and Reviews, AGARD,<br />

1955.<br />

[4] von Kármán, Th.: Dimensionslose Grösen in Grenzgebieten <strong>de</strong>r Aerodynamik.<br />

Z. F. W., Jan-Febr. 1956, pp. 3-5.<br />

[5] Weller, A. E.: Similarities in Combustion, a Review. Selected Combustion Problems,<br />

Vol. II, AGARD, 1956, pp. 371-383.<br />

[6] Penner, S. S.: Mo<strong>de</strong>ls in Aerothermochemistry. International Symposium on<br />

Mo<strong>de</strong>ls in Engineering, Venice, Italy, 1955.<br />

[7] Ross, Ch. C.: Scaling of Liquid Fuel Rocket Combustion Chambers. Selected<br />

Combustion Problems, Vol. II, AGARD, 1956, pp. 444-456.<br />

[8] Crooco, L.: Consi<strong>de</strong>rations on the Problem of Scaling Rocket Motors. Selected<br />

Combustion Problems, Vol. II, AGARD, 1956, pp. 457-468.<br />

[9] Barrère, M.: Similarity of Liquid Fuel Rocket Combustion Chambers. ONERA,<br />

Paris, 1956.<br />

[10] Penner, S. S. and Fuhs, A. E.: On Generalized Scaling Procedures for Liquid-<br />

Fuel Rocket Engines. Combustion and Flame, Vol. I, June 1957, pp. 229-240.<br />

[11] Ross, Ch. C. and Datner, P. P.: Combustion Instability in Liquid Propellant<br />

Rocket Motors. A Survey. Selected Combustion Problems, Vol I, AGARD,<br />

1954, pp. 352-380.


280 CHAPTER 11. SIMILARITY IN COMBUSTION. APPLICATIONS<br />

[12] Crocco, L. and Grey, J.: Combustion Instability in Liquid Propellant Rocket<br />

Motors. Proceedings of the Gas Dynamic Symposium on Aerothermochemistry,<br />

Northwestern Univ., Evanston Illinois, 1956.<br />

[13] Crocco, L. and Cheng, S. I.: Theory of Combustion Instability in Liquid Propellant<br />

Rocket Motors. Butterworths Scientific Publications, 1956.<br />

[14] Summerfield, M.: A Theory of Unstable Combustion in Liquid Propellant Rocket<br />

Systems. Journal of the American Rocket Society, Sept. 1951, pp. 108-114.<br />

[15] Crocco, L.: Aspects of Combustion Stability in Liquid Propellant Rooket Motors.<br />

Journal of the American Rocket Society, Nov. 1951, pp. 163-178.<br />

[16] Scurlock, A. C.: Flame Stabilization and Propagation in High Velocity Gas<br />

Streams. Meteor Report No. 19, Mass. Inst. Tech., May 1948.<br />

[17] Zukoski, E. E. and Marble, F. E.: Experiments Concerning the Mechanism of<br />

Flame Blowoff from Bluff Bodies. Proceedings of the Gas Dynamic Symposium<br />

on Aerothermochemistry, Northwestern Univ., Evanston Illinois, 1956.<br />

[18] Longwell, J. P.: Flame Stabilization by Bluff Bodies and Turbulent Flame in<br />

Ducts. Fourth Symposium (International) on Combustion, Williams and Wilkins<br />

Co., Baltimore, 1953, pp. 90-97.<br />

[19] Longwell, J. P., Frost, E.E. and Weis, M.A.: Flame Stability in Bluff Body Recirculation<br />

Zones. Industrial and Engineering Chemistry, Vol. 45, August 1953,<br />

pp. 1629-1633.<br />

[20] Spalding, D. B.: Theoretical Aspects of Flame Stabilization. Aircarft Engineering,<br />

Sep. 1953, pp. 264-268,276.<br />

[21] Lees, L.: Fluid Mechanical Aspects of Plane Stabilization. Jet Propulsion, July-<br />

August 1954, pp. 234-236.<br />

[22] Zukoski, E. E. and Marble, F. E.: The Role of Wake Transition in the Process<br />

of Flame Stabilization on Bluff Bodies. Combustion Researches and Reviews,<br />

AGARD, 1955, pp. 167-180.


Chapter 12<br />

Diffusion flames<br />

12.1 Introduction<br />

So far we have consi<strong>de</strong>red combustion problems <strong>de</strong>aling with premixed reactant gases.<br />

Yet, it can also happen that species are initially separated, and mixing and combustion<br />

take place simultaneously. Thus, a new type of flame is obtained which Burke and<br />

Schumann [1] called diffusion flame. In these flames the reaction zone separates the<br />

two reacting species which diffuse through inert gases and combustion products from<br />

each si<strong>de</strong> towards the flame. A flame of this type is obtained, for example, when a jet<br />

of combustible gas discharges from a burner into the open atmosphere or into an air<br />

stream parallel to the jet burning at the outlet of the burner, see Fig. 12.1. Such flames,<br />

are largely used in industrial applications. The flame length nee<strong>de</strong>d to burn a given<br />

x<br />

Flame<br />

Air<br />

Air<br />

o<br />

r<br />

Fuel<br />

Figure 12.1: An open diffusion flame.<br />

281


282 CHAPTER 12. DIFFUSION FLAME<br />

quantity of fuel per unit time un<strong>de</strong>r given conditions is highly interesting in these applications,<br />

since the dimensions of burners and furnaces <strong>de</strong>pend on it. The reacting<br />

species burn very rapidly as they reach the reaction zone. Thereby, the combustion<br />

velocity is generally conditioned by the accessibility of the species to the flame. In<br />

other words, by their facility to diffuse across the inert gases and combustion products.<br />

Diffusion can be either laminar or turbulent <strong>de</strong>pending on the conditions of the<br />

phenomenon, originating the flames accordingly.<br />

The actual state of knowledge on diffusion flames is by far less advanced than<br />

on premixed flames The fundamental work by Burke and Schumann [1] has been the<br />

starting point for practically all the later studies. Even though consi<strong>de</strong>rably ahead the<br />

publication of this work a correct qualitative i<strong>de</strong>a on diffusion flames was held, 1 Burke<br />

and Schumann were the first to achieve a quantitative study of the phenomenon. For<br />

the purpose they used a highly simplified mo<strong>de</strong>l that enabled an analytical solution<br />

of the problem. This mo<strong>de</strong>l retains the essential factors of the process and can be<br />

summarized in the following two assumptions:<br />

1) The thickness of the reaction zone is zero.<br />

2) The reacting species burn instantaneously upon reach the flame.<br />

This simple mo<strong>de</strong>l eliminates chemical kinetics from the process thus making<br />

available its computation. This simplification is justified when the reaction rate is very<br />

large compared to the diffusion velocities of the species towards the flame. Thereby it<br />

is not applicable, for example, to the study of flames rarefied or very diluted by inert<br />

gases where the time of reaction can appreciably influence the process. The mo<strong>de</strong>l<br />

disregards the structure of the reaction zone as was done in Chap. 6 for premixed<br />

flames. The difference between both cases lies upon the fact that while in the case<br />

of premixed flames the combustion front is a discontinuity surface for velocity, temperature,<br />

composition, etc., in diffusion flames such magnitu<strong>de</strong>s change continuously<br />

across the front.<br />

The first assumption reduces the reaction zone to a surface called the flame<br />

front, which acts as a sink for reacting species and as a source for combustion products.<br />

The second assumption implies the following consequences:<br />

a) Concentrations of reactants at the flame must be zero.<br />

b) Reactants must diffuse towards the flame in the stoichiometric ratio corresponding<br />

to the produced reactions.<br />

1 See, i.e., Barr’s review paper [2] where data and bibliography on the subject can be found.


12.1. INTRODUCTION 283<br />

In fact, otherwise, some of the species would diffuse across the flame and react with<br />

those on the other si<strong>de</strong>.<br />

In or<strong>de</strong>r to judge the validity of this mo<strong>de</strong>l Hottel and Hawthorne [3] performed<br />

some experiments to <strong>de</strong>termine the composition of the gases in a diffusion flame of<br />

hydrogen burning in the open atmosphere, taking samples of the gas at different points<br />

of the flame by means of a special sound. Details of these experiments can be found<br />

in the said work. Figure 12.2, taken from it, shows the results of the sounding ma<strong>de</strong><br />

in three cross sections of the flame. The results of such measurements prove the<br />

correctness of Burke and Schumann’s mo<strong>de</strong>l.<br />

100<br />

75<br />

50<br />

25<br />

H 2<br />

flame front<br />

distance from port= 30.5 cm<br />

Percentage in dry sample<br />

0<br />

100<br />

75<br />

50<br />

25<br />

0<br />

100<br />

O 2<br />

H 2<br />

N 2<br />

N 2<br />

axis of flame<br />

flame front<br />

O 2<br />

distance from port= 22.85 cm<br />

75<br />

50<br />

25<br />

H 2<br />

N 2<br />

flame front<br />

distance from port= 15.25 cm<br />

0<br />

O 2<br />

0 10 20 30 40<br />

Radial distance (mm)<br />

Figure 12.2: Gas composition in a hydrogen diffusion flame<br />

Burke and Schumann have successfully applied their mo<strong>de</strong>l to the calculation<br />

of the shape and height of laminar diffusion flames for the case where the fuel jet<br />

discharges within a tube where an air stream moves with the same velocity than the<br />

fuel jet, Fig. 12.3. This <strong>de</strong>vice eliminates the difficulties originating from momentum<br />

transfer between the fuel and the surrounding air when their velocities are not equal.<br />

Furthermore, for this reason it is easier to obtain a stable flame. By introducing several<br />

drastic simplifications, to be consi<strong>de</strong>red further on, Burke and Schumann were able<br />

to calculate the shape of the flame. Some of their results are outlined in Fig. 12.3.<br />

Fig. 12.3 (a) corresponds to an overventilated flame, which closes at the axis of the


284 CHAPTER 12. DIFFUSION FLAME<br />

air fuel air air fuel air<br />

(a) overventilated<br />

(b) un<strong>de</strong>rventilated<br />

Figure 12.3: Confined laminar diffusion flames.<br />

tube, and Fig. 12.3 (b) to an un<strong>de</strong>rventilated flame, which ends at the wall of the<br />

exterior tube. The fuel-air ratio can be changed by varying the diameters ratio of both<br />

tubes. Both types of flames were experimentally observed by Burke and Schumann.<br />

The flame length is proportional to the amount of fuel burnt per second and does not<br />

<strong>de</strong>pend on the diameter of the burner.<br />

Burke and Schumann’s analysis on confined flames has been exten<strong>de</strong>d by J.<br />

Barr, [4] and [5], to the case where air and fuel velocities are different. He has experimentally<br />

analyzed the appearance of the flames that form when the streams of air<br />

and fuel change within limits far more exten<strong>de</strong>d than those covered by any previous<br />

investigation. As a result he conclu<strong>de</strong>s that the length of a laminar diffusion flame is<br />

proportional to the fuel consumption not only for the case of short flames, as stated<br />

by Lewis and von Elbe [6], but also for flames of a length as much as a hundred times<br />

the diameter of the burner. The results of his work are summarized in Fig. 12.4, taken<br />

from Ref. [5], which shows the different types of flames observed when the flow rates<br />

of air and fuel are changed. In the same figure, regions 1 and 2 correspond to flames<br />

with excess of air and fuel respectively, of the type studied by Burke and Schumann.<br />

In between these regions a zone of carbon formation exists. Region 3 corresponds to<br />

meniscus flames for which diffusion in the axial direction is important. A meniscus<br />

flame blows-out when the fuel flow rate is reduced. Region 4 corresponds to the socalled<br />

convective or Lambent flames. Within this region, where the fuel and air ratios<br />

are small, buoyancy forces are important and oscillating flames are obtained. Region<br />

5, where fuel flow rate is very large, corresponds to tilted flames preceding blow-off.<br />

Region 6, where the fuel and air streams are very strong, corresponds to lifted flames


12.1. INTRODUCTION 285<br />

which prece<strong>de</strong>d blow-off. Region 7 corresponds to vortex flames. Finally, un<strong>de</strong>r given<br />

conditions, flames similar to the manometric and singing flames were also observed<br />

by Barr. This incomplete enumeration of the types of flames observed gives i<strong>de</strong>a on<br />

the complexity of the phenomenon. Further data can be found in Ref. [4].<br />

100<br />

10<br />

Extinction<br />

2<br />

Smoke<br />

6<br />

Lifted flames<br />

Butane flow (cm /s)<br />

3<br />

1<br />

0.1<br />

5<br />

4<br />

Smoke point<br />

1<br />

Laminar<br />

diffusion<br />

flames<br />

7<br />

Extinction<br />

Meniscus flames<br />

3<br />

Vortex flames<br />

Convective flames<br />

Extinction<br />

0.01<br />

1 10 100<br />

1000 10000<br />

3<br />

Air flow (cm /s)<br />

Figure 12.4: Flow limits for formation of enclosed diffusion flames.<br />

The case of an open diffusion flame where the fuel jet discharges and burns in<br />

an atmosphere at rest has been experimentally analyzed, among others, by Rembert<br />

and Haslam [7] , Cuthbertson [8] , Gaunce [9], Hottel and Hawthorne [3], Wohl,<br />

Gazley and Kapp [10], Barr and Mullins [11], Parker and Wolfhard [12] and Yagi and<br />

Saji [13]. The fundamental results of their experiments are as follows:<br />

1) When the Reynolds number of the jet at the mouth of a burner is small, a laminar<br />

flame establishes.<br />

2) The length of the flame increases when the flow rate of the jet is increased.<br />

3) When Reynolds number is larger than a given value a turbulent region initiates


286 CHAPTER 12. DIFFUSION FLAME<br />

at the flame tip. When Reynolds number of the jet is increased this region propagates<br />

towards the flame base.<br />

4) When the tip of the flame becomes turbulent its length <strong>de</strong>creases as Reynolds<br />

number is increased until turbulence reaches its base. There on it remains constant<br />

in<strong>de</strong>pen<strong>de</strong>ntly from the discharge velocity of the jet.<br />

5) Within the laminar region the length of the flame is in<strong>de</strong>pen<strong>de</strong>nt from the burner’s<br />

diameter and for a given fuel it <strong>de</strong>pends only on its flow rate. Such result agrees<br />

with that obtained by Burke and Schumann for confined flames.<br />

6) Within the turbulent region the length of the flame is proportional to the diameter<br />

of the tube and in<strong>de</strong>pen<strong>de</strong>nt from the fuel velocity.<br />

Diffusion flames<br />

Transition region<br />

Fully <strong>de</strong>veloped turbulent flames<br />

Envelope of flame heights<br />

Height<br />

Increasing nozzle velocity<br />

Envelope of<br />

break points<br />

Figure 12.5: Change in flame type with increasing nozzle velocity.<br />

Fig. 12.5 taken from the work by Hottel and Hawthorne summarizes the experiments<br />

performed by these authors. This figure shows the law of variation of the height of<br />

the flame with the velocity of the jet, and the extension of the turbulent zone. The<br />

following states are observed:<br />

a) Laminar. When the velocity of the jet is small. Here the length of the flame<br />

increases, at first proportionally to the jet velocity and then slows down up to the<br />

maximum height that limits the laminar zone.<br />

b) Transition. The turbulent zone of the flame initiates close to the tip and when<br />

the velocity of the jet increases it propagates towards its base. The total length<br />

of the flame <strong>de</strong>creases when the extension of the turbulent zone increases. This<br />

is such because the diffusivity of the reactant gases is much larger in turbulent<br />

than in laminar state.<br />

c) Turbulent. The whole flame is turbulent throughout its extension except within a<br />

small zone close to the mouth of the burner. In this state the length of the flame


12.1. INTRODUCTION 287<br />

remains appreciably constant as the velocity of the jet increases. Fig. 12.6 gives<br />

some of the results of the measurements done by Gaunce [9] with city gas in a 3<br />

mm diameter burner. The three aforementioned zones are clearly shown in this<br />

figure.<br />

24<br />

20<br />

Distance from nozzle (inch)<br />

16<br />

12<br />

8<br />

4<br />

Total length, on−port flames<br />

Break point length<br />

Flame separation begins here<br />

Upper part of flame blows−off<br />

0<br />

0 50 100 150 200 250 300<br />

Nozzle velocity (ft/s)<br />

Figure 12.6: Effect of nozzle velocity on flame length.<br />

Hottel and Hawthorne [3], Wohl, Gazley and Kapp [10], Yagi and Saji [13]<br />

and Barr [5] have attempted an extension of Burke and Schumann’s method to the<br />

prediction of the length of open flames both laminar and turbulent. Through rudimentary<br />

approximations they obtain an expression for the flame length containing an<br />

unknown function which they <strong>de</strong>termine empirically from the results of their experiments.<br />

Fig. 12.7 taken from Ref. [5] shows the theoretical and experimental lengths<br />

of some laminar open flames as functions of the fuel flow rate. This figure gives an<br />

i<strong>de</strong>a on the agreement to be expected between experimental measurements and those<br />

predicted by semi-empirical formulae. The extrapolation to values not inclu<strong>de</strong>d in<br />

this figure is risky. A summary on the state of knowledge regarding this matter can<br />

be found in the work by Hottel listed in Ref. [14] where additional bibliography is<br />

inclu<strong>de</strong>d.<br />

Lately, J. A. Fay [15] has calculated the shape and characteristics of the laminar<br />

diffusion flame obtained when a fuel jet discharges into the open atmosphere for the<br />

two-dimensional case and for the case with axial symmetry. Fay has taken into account<br />

the influence of the variation of the velocity in cross direction to the jet, computing<br />

the cross distributions of the velocities, concentrations and temperatures. The mo<strong>de</strong>l


288 CHAPTER 12. DIFFUSION FLAME<br />

120<br />

100<br />

Butane tube<br />

A<br />

B<br />

Flame length (cm)<br />

80<br />

60<br />

40<br />

City gas nozzle<br />

C<br />

E<br />

City gas tube<br />

D<br />

B<br />

Theory<br />

Experiments<br />

A & E: Wohl et al. Butane, Wohl et al.<br />

20<br />

B, D & F: Barr City gas, Wohl et al.<br />

C: Hottel et al. City gas, Rembert & Haslem<br />

Gaunce<br />

0<br />

0 50 100 150 200 250<br />

Fuel flow (cm 3 /s)<br />

Figure 12.7: Length of different open flames.<br />

proposed by Burke and Schumann is also used by Fay by reducing the reaction zone<br />

to a surface. This work represents the first serious attempt to analyze the velocity field<br />

induced by a diffusion flame.<br />

So far no attempt has been ma<strong>de</strong> for the study of the influence of free convection<br />

on the phenomenon which can be very important specially when the discharge<br />

velocity of the jet is small.<br />

The structure of the reaction zone of a diffusion flame has experimentally been<br />

analyzed by Wolfhard and Parker [12] by means of a two-dimensional laminar diffusion<br />

flame obtained with two parallel jets, combustible and oxidizer respectively.<br />

This type of flame is specially suitable for spectroscopic studies. Wolfhard and Parker<br />

have mainly experimented with ammonia-oxygen and ethylene-oxygen flames. Their<br />

studies confirm Burke and Schumann’s stand-point. The reaction zone is in thermal<br />

and chemical equilibrium and therein temperature is practically constant. The concentrations<br />

of reacting species are very small within this zone and, except when <strong>de</strong>aling<br />

un<strong>de</strong>r special conditions that reduce the reaction rate, the process is governed by the<br />

diffusivity of the species. The reacting species generally dissociate before reaching<br />

the reaction zone due to temperature. In the case of hydrocarbons such cracking leads<br />

to carbon formation. Temperatures observed agree fairly with those predicted by theory.<br />

2 Zeldovich [16] has taken into consi<strong>de</strong>ration the finite thickness of the reaction<br />

2 See further on §3.


12.2. GENERAL EQUATIONS FOR LAMINAR DIFFUSION FLAMES 289<br />

zone to explain the blowing-off phenomenon. A similar study has been performed<br />

by Spalding to explain the extinction phenomenon in the combustion of fuel droplets<br />

with convection. 3<br />

The following paragraphs are <strong>de</strong>voted to the <strong>de</strong>duction of the general equations<br />

for diffusion flames as well as to the simplified equations on which their study is based.<br />

12.2 General equations for laminar diffusion flames<br />

The general equations given in chapter 3 will be applied here to the study of laminar<br />

diffusion flames by assuming that Burke and Schumann’s conditions stated in §1 are<br />

satisfied. The study limits to the case of steady motions. In the <strong>de</strong>duction that follows<br />

the notation in chapter 3 will apply.<br />

Continuity equations<br />

1) For the mixture.<br />

Since the motion is steady, equation (3.6) reduces to<br />

∇ · (ρ¯v) = 0. (12.1)<br />

2) For the species.<br />

As per Burke and Schumann’s first condition, the reaction rate outsi<strong>de</strong> the flame’s<br />

surface Σ f is zero. Therefore, system (3.7) reduces to<br />

ρ(¯v · ∇)Y i + ∇ · (ρY i¯v di ) = 0, i = 1, 2, . . . , l. (12.2)<br />

Equation of motion<br />

Equation of motion (3.18) takes the form<br />

ρ(¯v · ∇)¯v = −∇p + ∇ · τ ev + ρ ¯F . (12.3)<br />

The following cases are specially interesting:<br />

1) Pressure is constant and the influence of mass forces negligible. Then Eq. (12.3)<br />

takes the form<br />

ρ(¯v · ∇)¯v = ∇ · τ ev . (12.4)<br />

3 See §13 of Chap. 13.


290 CHAPTER 12. DIFFUSION FLAME<br />

Moreover, as done in the theory of free jets, here only some terms of the viscosity<br />

forces need to be preserved. 4<br />

2) The influence of mass forces is important. Hydrostatic equilibrium in the fluid<br />

undisturbed by the flame <strong>de</strong>termines the pressure field. If ρ 0 is the <strong>de</strong>nsity of the<br />

undisturbed fluid<br />

−∇p + ρ 0 ¯F = ¯0, (12.5)<br />

which gives for Eq. (12.3)<br />

ρ(¯v · ∇)¯v = ∇ · τ ev + (ρ − ρ 0 ) ¯F . (12.6)<br />

Furthermore, in each case only the significant terms of ∇ · τ ev will be retained. So<br />

far a solution that inclu<strong>de</strong>s the influence of the last term of Eq. (12.6) has not been<br />

obtained.<br />

Energy equation<br />

Kinetic energy of motion, energy dissipated by viscosity and work done by mass<br />

forces are negligible. Hence, equation (3.38) reduces to<br />

(<br />

ρ(¯v · ∇)h − ∇ · (λ∇T ) + ∇ · ρ ∑ )<br />

Y i h i¯v di = 0. (12.7)<br />

i<br />

Diffusion equations<br />

Such equations are given by system (3.8) which takes the form<br />

∑<br />

(<br />

Y j ¯vdi − ¯v dj<br />

+ ∇Y i<br />

− ∇Y )<br />

j<br />

= ¯0, (i = 1, 2, . . . , l)<br />

M<br />

j j D ij Y i Y j<br />

∑<br />

Y j ¯v dj = ¯0,<br />

j<br />

(12.8)<br />

where pressure and thermal diffusion are not inclu<strong>de</strong>d since they are negligible.<br />

State equation<br />

This is equation (3.39)<br />

p<br />

ρ = R mT. (12.9)<br />

4 See Fay, Ref. [15].


12.3. BOUNDARY CONDITIONS ON THE FLAME 291<br />

The previous system together with the a<strong>de</strong>quate boundary conditions <strong>de</strong>termine<br />

the values for p, ρ, T , ¯v and for the mass fractions Y i at each point as well as the shape<br />

of the flame which so far is unknown.<br />

12.3 Boundary conditions on the flame<br />

Chemical species A i forming the mixture can be classified into three different groups:<br />

a) Reacting species A r i , which diffuses towards the flame where they burn. These<br />

species are unable to cross Σ f .<br />

b) Reaction products A p i , which produce at the flame and diffuse from it towards<br />

both si<strong>de</strong>s.<br />

c) Inert species A d i , which dilute reacting species. These species can cross Σ f .<br />

According to Burke and Schumann’s mo<strong>de</strong>l mass fractions Yi<br />

r<br />

species must be zero on the flame surface, that is<br />

of the reacting<br />

Y r<br />

1 = Y r<br />

2 = · · · = Y r<br />

l r<br />

= 0 (12.10)<br />

on Σ f , where l r is the number of reacting species. 5<br />

position of the flame surface.<br />

These conditions <strong>de</strong>termine the<br />

Let m r i be the mass of species Ar i that reaches Σ f per unit surface and per unit<br />

time. m r i is given by m r i = −ρY i¯v i · ¯n i , (12.11)<br />

where ¯n i is the unit vector normal to Σ f at A r i si<strong>de</strong>. mr i is consumed by the chemical<br />

reactions taking place at Σ f . Let<br />

∑<br />

ν ijA ′ r i → ∑ ν ijA ′′ p i , (j = 1, 2, . . . , r), (12.12)<br />

i<br />

i<br />

be one of these reactions, which transforms reacting species A r i into reaction products<br />

A p i . Since the thickness of the reaction zone is assumed to be zero, it becomes necessary<br />

to <strong>de</strong>fine a surface reaction rate for each reaction. Such rate gives the masses of<br />

the species consumed or produced by the reaction per unit surface and per unit time.<br />

Be r j the value of such surface reaction rate for reaction j. The fraction m r ij of mr i<br />

consumed by this reaction is<br />

where M r i is the molar mass of species Ar i .<br />

m r ij = M r i ν ′ ijr j , (j = 1, 2, . . . , r), (12.13)<br />

5 Actually, these mass fractions are very small but not strictly zero. Their values are practically those of<br />

the equilibrium of species un<strong>de</strong>r the prevailing conditions at the flame.


292 CHAPTER 12. DIFFUSION FLAME<br />

Therefore m r i<br />

is given by<br />

∑<br />

m r i = Mi<br />

r ν ijr ′ j , (i = 1, 2, . . . , l r ). (12.14)<br />

j<br />

Likewise if m p i is the mass of species Ap i produced per unit surface and per<br />

unit time one has<br />

m p i = M p i<br />

∑<br />

j<br />

ν ′′<br />

ijr j , (i = l r + 1, l r + 2, . . . , l a ), (12.15)<br />

where l a = l r + l p is the number of active species (reactants plus products) and l p is<br />

the number of products.<br />

Reaction rates r j are unknown “a priori”. Elimination of the same between<br />

equations (12.14) and (12.15) gives l a −r relations between Yi<br />

r and Y p<br />

i which enables<br />

to express all mass fractions of active species as functions of r selected from them.<br />

Thus the variables of the problem reduce in l a − r.<br />

So far no solutions have been obtained for this general system of equations.<br />

In or<strong>de</strong>r to make computation available, additional assumptions must be introduced<br />

to simplify consi<strong>de</strong>rably the problem. Such simplifications will be performed in the<br />

following paragraph.<br />

12.4 Simplified equations<br />

In this approximate study the number of species is assumed to be three. Namely, fuel<br />

A 1 , oxidizer A 3 and products A 2 . Products inclu<strong>de</strong> not only species resulting from<br />

reactions, but also diluents of fuel and oxidizer initially mixed with them. 6 Flame surface<br />

divi<strong>de</strong>s space into two regions: an interior region at the fuel si<strong>de</strong> and an exterior<br />

one at the oxidizer si<strong>de</strong>. Only A 1 and A 2 species exist within the interior region. In<br />

the exterior one only A 2 and A 3 . Therefore, the composition of the mixture within the<br />

interior region is <strong>de</strong>termined by the value of Y 1 which must satisfy equation (12.2)<br />

ρ(¯v · ∇)Y 1 + ∇ · (ρY 1¯v d1 ) = 0. (12.16)<br />

Similarly in the exterior region one has for Y 3<br />

ρ(¯v · ∇)Y 3 + ∇ · (ρY 3¯v d3 ) = 0. (12.17)<br />

6 A larger number of species could easily be inclu<strong>de</strong>d by distinguishing, for example , between diluents<br />

and products. Such is done, among others, by Zeldovich [17] and Fay [15]. However, this does not represent<br />

any fundamental advantage and computations become more elaborate. A differentiation between products<br />

and diluent can be done once the problem is solved.


12.4. SIMPLIFIED EQUATIONS 293<br />

As for Y 2 , its value is given by expressions<br />

Interior region: Y 2 = 1 − Y 1 . (12.18)<br />

Exterior region: Y 2 = 1 − Y 3 . (12.19)<br />

Diffusion velocities ¯v d1 and ¯v d3 are given by Fick’s law 7<br />

Y 1¯v d1 = −D 12 ∇Y 1 , (12.20)<br />

Y 3¯v d3 = −D 23 ∇Y 3 . (12.21)<br />

On surface Σ f<br />

Y 1 = Y 3 = 0; Y 2 = 1. (12.22)<br />

Furthermore, since Y 1 and Y 3 must diffuse towards the flame in the stoichiometric<br />

ratio from (12.11), (12.20) and (12.21) we have on Σ f<br />

νD 12<br />

∂Y 1<br />

∂¯n i<br />

= D 23<br />

∂Y 3<br />

∂¯n e<br />

. (12.23)<br />

Here ν is the ratio between the masses of oxidizer and fuel nee<strong>de</strong>d for complete combustion.<br />

¯n e and ¯n i are normals to Σ f towards the exterior and interior regions respectively.<br />

System of equations (12.16) and (12.17) can be substituted by a single equation<br />

for a new variable Y <strong>de</strong>fined as follows<br />

⎧<br />

⎪⎨ Y 1 in the interior region,<br />

Y =<br />

⎪⎩ − 1 ν Y 3 in the exterior region.<br />

Y must satisfy the following differential equation<br />

(12.24)<br />

ρ(¯v · ∇)Y − ∇ · (ρD∇Ȳ ) = 0, (12.25)<br />

where D takes the value<br />

⎧<br />

⎨<br />

D =<br />

⎩<br />

D 12<br />

D 23<br />

in the interior region,<br />

in the exterior region.<br />

(12.26)<br />

Furthermore, Y is zero on the flame and takes values of opposite sign at both si<strong>de</strong>s of<br />

it. The <strong>de</strong>rivatives of Y at both si<strong>de</strong>s of the flame in normal direction to it must satisfy<br />

condition<br />

7 See Chap. 2.<br />

D 12<br />

∂Y<br />

∂¯n i<br />

= −D 23<br />

∂Y<br />

∂¯n e<br />

. (12.27)


294 CHAPTER 12. DIFFUSION FLAME<br />

In particular, if condition<br />

D 12 = D 23 = D (12.28)<br />

is satisfied, Y as well as its <strong>de</strong>rivatives are continuous even at the flame. Continuity<br />

Eq. (12.1) for the mixture and Eq. (12.3) for motion still hold unchanged.<br />

As for the energy equation we shall assume that the specific heats at constant<br />

pressure of the three species are constant and equal<br />

c p1 = c p2 = c p3 = c p . (12.29)<br />

In such case, one can immediately check that Eq. (12.7) reduces to the following<br />

which hold throughout space.<br />

ρc p¯v · ∇T − ∇ · (λ∇T ) = 0, (12.30)<br />

Let us assume, that besi<strong>de</strong>s condition (12.28) the following is satisfied<br />

λ<br />

ρDc p<br />

= 1, (12.31)<br />

in accordance with the result of the elementary Kinetic Theory of Gases. Then, comparison<br />

between (12.25) and (12.30) suggests the existence of solutions of the form<br />

T = aY + b, (12.32)<br />

where a and b are constant but can be different for the interior and exterior regions.<br />

These solutions are available for the study of diffusion flames if boundary conditions<br />

can be satisfy through them. In particular, since on the flame Y = 0, temperature T b<br />

of the flame will be constant for the cases where Eq. (12.32) is valid. The study of<br />

diffusion flames reduces, hence, to a computation of the values for ρ , T , ¯v and Y . For<br />

this purpose, system of Eqs. (12.1) and (12.3) is available which must be completed<br />

with the boundary conditions a<strong>de</strong>quate for each case. These refer in general to the<br />

conditions at the discharge sections of fuel and oxidizer and at great distance from the<br />

flame. Let us assume, for example, that in the fuel’s discharge section<br />

and in the oxidizer’s discharge section<br />

Y = Y 10 , T = T 0 , (12.33)<br />

Y = Y 30 , T = T 0 , (12.34)<br />

Then Eq. (12.32) holds and one obtains in the interior region,<br />

T = T b − (T b − T 0 ) Y<br />

Y 10<br />

= T b − (T b − T 0 ) Y 1<br />

Y 10<br />

, (12.35)


12.4. SIMPLIFIED EQUATIONS 295<br />

while in the exterior region<br />

T = T b + ν(T b − T 0 ) Y<br />

Y 30<br />

= T b − (T b − T 0 ) Y 3<br />

Y 30<br />

. (12.36)<br />

In these expressions T b is still un<strong>de</strong>termined. Its value can be obtained by consi<strong>de</strong>ring<br />

an element of the flame as the one in Fig. 12.8 and applying to it the principle of<br />

conservation of energy. One immediately obtains<br />

m 2 h 2 − (m 1 h 1 + m 3 h 3 ) = λ<br />

( ∂T<br />

∂¯n i<br />

+ ∂T<br />

∂¯n e<br />

)<br />

. (12.37)<br />

Here, m 1 and m 3 are the masses of fuel and oxidizer that reach ∑ f<br />

per unit surface<br />

ν 1<br />

m 2,i + m 2,e = =(1+ ) m<br />

m 2<br />

m 2,i<br />

m 1<br />

m 2,e<br />

m 3 = νm 1<br />

n i<br />

λ T n<br />

’<br />

i<br />

n e<br />

λT’ n e<br />

Flame<br />

Figure 12.8: Schematic diagram of an element of a diffusion flame.<br />

and per unit time and m 2 is the mass of products that emerges from it. But<br />

m 3 = νm 1 , m 2 = (1 + ν)m 1 . (12.38)<br />

Furthermore, from Eqs. (12.11), (12.20) and (12.28)<br />

From Eqs. (12.35) and (12.36)<br />

m 1 = ρD ∂Y 1<br />

∂¯n i<br />

. (12.39)<br />

∂T<br />

= −(T b − T 0 ) 1 ∂Y 1<br />

, (12.40)<br />

∂¯n i Y 10 ∂¯n i<br />

∂T<br />

= −(T b − T 0 ) 1 ∂Y 3<br />

. (12.41)<br />

∂¯n e Y 30 ∂¯n e<br />

When expressions for m 1 , m 2 , m 3 , ∂T /∂¯n i and ∂T /∂¯n e are taken into Eq. (12.37)<br />

keeping in mind Eqs. (12.23), (12.29) and (12.31), one obtains for T b<br />

T b − T 0 = q r<br />

c p<br />

Y 10 Y 30<br />

Y 30 + νY 10<br />

, (12.42)<br />

where<br />

q r = h 1 + νh 3 − (1 + ν)h 2 (12.43)<br />

is the heat of reaction per unit mass of fuel.


296 CHAPTER 12. DIFFUSION FLAME<br />

Formula (12.42) shows that T b is the temperature that would correspond to a<br />

premixed flame where fuel and its diluent were mixed with oxidizer and its diluent in<br />

the stoichiometric ratio. In fact, in such mixture, mass fraction of fuel is<br />

Y 10 Y 30<br />

Y 30 + νY 10<br />

and the corresponding temperature of the flame is given by (12.42).<br />

12.5 Solutions of the simplified system<br />

Some approximate solutions have been obtained for the case of two-dimensional and<br />

axis symmetrical flames either confined or open. A <strong>de</strong>tailed study of such flames is<br />

available in the works listed in Refs. [1], [3], [6], [10], [15] and [17]. The following<br />

chapter studies in <strong>de</strong>tail the diffusion flame that forms surrounding a fuel droplet<br />

which evaporates and burns in an oxidizing atmosphere. The problem has great importance<br />

in jet propulsion.<br />

The following simplifying assumptions have generally been adopted:<br />

1) Stream velocity is throughout space parallel to the flame axis and uniform throughout<br />

space or at least at each cross-section of the same.<br />

2) All transport coefficients are in<strong>de</strong>pen<strong>de</strong>nt from the composition and temperature<br />

of the mixture.<br />

3) Diffusion produces only in cross direction to the flame axis.<br />

These assumptions reduce the problem to the integration of the one-dimensional<br />

diffusion equation for the computation of the fuel, oxidizer and temperature distributions.<br />

For example, in the case of an axis symmetrical flame, Fig. 12.1, the preceding<br />

assumptions reduce Eq. (12.25) to<br />

v ∂Y<br />

∂x = D 1 (<br />

∂<br />

r ∂Y )<br />

. (12.44)<br />

r ∂r ∂r<br />

Here, D and v are constant or at least in<strong>de</strong>pen<strong>de</strong>nt from r. In the first case<br />

the integration of Eq. (12.44) is straightforward. This also occurs in the second case<br />

through the previous change of variable<br />

x =<br />

∫ τ<br />

0<br />

v<br />

dτ, (12.45)<br />

D<br />

which to be <strong>de</strong>termined would require knowing v/D as a function of distance x to<br />

the flame base, which has been introduced by Hottel [3], Wohl [10] and Barr [5] as


12.5. SOLUTIONS OF THE SIMPLIFIED SYSTEM 297<br />

an unknown function empirically <strong>de</strong>termined. The aforementioned simplifying assumptions<br />

only represent a poor approximation to reality. In fact, with respect to the<br />

velocity field it can easily be verified that it is only constant in open flames where fuel<br />

and oxidizer move at equal velocity within the discharge section and, furthermore,<br />

where pressure is uniform throughout space. In such case the gas expansion produces<br />

entirely in cross direction to the flame. As for transport coefficients they can change<br />

in the ratio ten to one between flame surface and unburnt gases. While the assumption<br />

that diffusion produces only in cross direction is well justified, except for special cases<br />

like that of meniscus flames which so far have not been theoretically analyzed.<br />

Before ending this brief review on diffusion flames we shall <strong>de</strong>velop a simple<br />

argument by W. Jost [19] which enables an easy establishment of the influence on the<br />

length of a diffusion flame of some of the fundamental variables of the process. In<br />

view of the rudimentary approximations nee<strong>de</strong>d in Burke-Schumann’s calculations in<br />

or<strong>de</strong>r to obtain usable expressions, Jost assumes the same validity for his argument.<br />

R<br />

z<br />

L<br />

vt<br />

z<br />

v<br />

fuel<br />

v<br />

oxigen<br />

burner axis<br />

burner wall<br />

Figure 12.9: Schematic diagram of a overventilated diffusion showing the elements of the<br />

Jost’s criterion.<br />

Let us consi<strong>de</strong>r, for example, a confined flame with an excess of air or an<br />

open flame, Fig. 12.9. According to Jost, the length of the flame is <strong>de</strong>termined by the<br />

condition that the oxygen of the atmosphere surrounding the jet can reach the jet’s axis<br />

through diffusion. Let D be the oxygen diffusion coefficient and z its mean quadratic<br />

displacement due to diffusion at instant t. Diffusion theory gives for z as function of t<br />

z 2 = 2Dt. (12.46)<br />

But, at the point where the flame closes<br />

z = R, (12.47)


298 CHAPTER 12. DIFFUSION FLAME<br />

where R is the burner’s radius, and<br />

t = L v , (12.48)<br />

where v is the mean velocity of the jet between the base and the tip of the flame and L<br />

the length of the same. Through elimination of t and z between (12.46), (12.47) and<br />

(12.48) one obtains<br />

L ∼ R2 v<br />

2D . (12.49)<br />

In laminar flames D <strong>de</strong>pends only on the properties of the gases. Furthermore<br />

R 2 v is proportional to the fuel flow rate G. Consequently, we have<br />

in accordance with the result obtained by Burke-Schumann.<br />

L ∼ G<br />

2D , (12.50)<br />

In a turbulent flame D is the turbulent diffusivity which has the form<br />

When this value is substituted into (12.49) it results<br />

D ∼ vR. (12.51)<br />

L ∼ R, (12.52)<br />

which as previously seen also agrees with experimental results.<br />

References<br />

[1] Burke, S. F. and Schumann, T. E. W.: Diffusion Flames. Industrial and Engineering<br />

Chemistry, Vol. 20, October 1928, pp. 998-1004.<br />

[2] Barr, J.: Diffusion Flames in the Laboratory. AGARD Memorandum No.<br />

AG11/M7, 1954.<br />

[3] Hottel, H. C. and Hawthorne, W.R.: Diffusion in Laminar Flame Jets. Third<br />

Symposium (International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1949, pp. 254-266.<br />

[4] Barr, J.: Diffusion Flames. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953, pp. 765-771.<br />

[5] Barr, J.: Length of Cylindrical Laminar Diffusion Flames. Fuel, Jan. 1954,<br />

pp. 51-59.<br />

[6] Lewis, B. and von Elbe, G.: Combustion, Flames and Explosions of Gases.<br />

Aca<strong>de</strong>mic Press Inc. Publishers., New York, 1951.


12.5. SOLUTIONS OF THE SIMPLIFIED SYSTEM 299<br />

[7] Rembert, E. W. and Haslam, R. T.: Factors Influencing the Length of a Gas<br />

Flame Burning in Secondary Air. Industrial and Engineering Chemistry, Dec.<br />

1925, pp. 1236-1238.<br />

[8] Cuthbertson, J.: Journal Society of Chemical Industries, Transactions, Vol. 50,<br />

1931, pp. 451-457.<br />

[9] Gaunce, H.: Unpublished Research on Flames, M.I.T., 1937.<br />

[10] Whol, K., Gazley, C. and Kapp, N. M.: Diffusion Flames. Third Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore, 1949,<br />

pp. 288-301.<br />

[11] Barr, J. and Mullins, B. P.: Combustion in Vitiated Atmospheres. Fuel, Vol. 28,<br />

1949, pp. 131, 200, 205, 225.<br />

[12] Parker, W. G. and Wolfhard, H. G.: Journal of the Chemical Society, 1950,<br />

p. 2038.<br />

[13] Yagi, S. and Saji, K.: Problems of Turbulent Diffusion and Flame Jets. Fourth<br />

Symposium (International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, pp. 771-781.<br />

[14] Hottel, H. C.: Burning in Laminar and Turbulent Fuel Jets. Fourth Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore, 1953,<br />

pp. 97-113.<br />

[15] Fay, J. A.: The Distributions of Concentration and Temperature in a Laminar Jet<br />

Diffusion Flame. Journal of the Aeronautical Sciences, October 1954, pp. 581-<br />

589.<br />

[16] Gaydon, A. G. and Wolfhard, H. G.: Flames, their Structure, Radiation and<br />

Temperature. Chapman and Hall Ltd., London, 1953, p. 135.<br />

[17] Zeldovich, Y. B.: On the Theory of Combustion of Initially Unmixed Gases.<br />

N.A.C.A. Tech. Memo. No. 1296, June 1951.<br />

[18] Spalding, D. B. : The Combustion of Liquid-Fuels. Fourth Symposium (International)<br />

on Combustion, Williams and Wilkins Co., Baltimore, 1953, pp. 847-<br />

864.<br />

[19] Jost, W.: Explosion and Combustion Processes in Gases. McGraw-Hill Book<br />

Comp., New York, 1946, p. 210.


300 CHAPTER 12. DIFFUSION FLAME


Chapter 13<br />

Combustion of liquid fuels<br />

13.1 Introduction<br />

The present chapter is <strong>de</strong>voted to the study of some of the problems of the combustion<br />

of liquid fuels. In certain cases surface reactions can take place in the liquid phase, for<br />

example, in the combustion of hypergoles. On the contrary, when a fuel burns in an<br />

oxidizing atmosphere the reaction takes place in the gaseous phase. Thereby, the fuel<br />

evaporation must prece<strong>de</strong>d combustion. In such case, if the mixing of the fuel vapours<br />

and the oxidizing gas takes place before the chemical reaction, a flame produces of<br />

the type studied in chapter 6. On the other hand, if mixing and combustion take place<br />

simultaneously, a diffusion flame is obtained, similar to those studied in chapter 12.<br />

This is the type of combustion consi<strong>de</strong>red in the present chapter. Khudiakov [1], [2]<br />

has studied the characteristics of the diffusion flame that forms on the free surface of<br />

a fuel burning in the atmosphere. Spalding [3] has also studied this problem taking<br />

into account the influence of free and forced convection. He ma<strong>de</strong> this study as an<br />

application of his uniform method [4] for the study of all the mass transport processes<br />

(absorption, evaporation, con<strong>de</strong>nsation and combustion).<br />

Of further technical interest is the case where fuel forms droplets of small<br />

diameter. Fuel mists formed by droplets of very small diameter (smaller than a few<br />

hundredths of a millimeter) burn forming a flame similar to the premixed gas flames<br />

[5], [6]. In particular, in the said mists, like in the premixed gas flames, the propagation<br />

velocity of the flame, its flammability limits, etc., are well <strong>de</strong>fined magnitu<strong>de</strong>s. Their<br />

values are close, but not equal, to those corresponding to premixed gas flames [7].<br />

The difference is probably due to the droplets evaporation effect. The diameter of the<br />

droplets is so small that the evaporation produces within the heating zone of the flame,<br />

so that the fuel reaches the reaction zone in gaseous phase.<br />

301


302 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

If the diameter of the droplet is larger than, for example, a few tenths of a<br />

millimeter the phenomenon becomes more complicated. Such size is too large to<br />

allow evaporation across the heating zone of a flame. Therefore, in this case individual<br />

diffusion flames surrounding the droplets are produced.<br />

For technical applications the fuel is normally introduced in the combustion<br />

chamber by means of an atomizer producing droplets of different sizes. These droplets<br />

spray and evaporate in contact with the atmosphere of the chamber, which is strongly<br />

turbulent. Furthermore, in the combustors of jet engines, the gases flow at high velocity<br />

through the combustion chamber. Generally, in such cases not enough space<br />

is available for the complete evaporation of the fuel before it reaches the combustion<br />

zone. Consequently when the fuel reaches the said zone it is only partially evaporated<br />

and the mixing with the oxidizing atmosphere is incomplete. The entire process starts<br />

with the entrance of the fuel in the combustion chamber and ends when the combustion<br />

products are formed. The complete study of this process inclu<strong>de</strong>s the following stages:<br />

atomizing, spraying, mixing, evaporation and combustion of the fuel jet. Due to the<br />

complexity of the process a study in <strong>de</strong>tail can not be attempted with any probabilities<br />

of success. Therefore, we must resort to empiricism and to the study of simplified<br />

physical mo<strong>de</strong>ls which emphasize some of the features of the phenomenon.<br />

Hereinafter we shall first briefly refer to the atomization, spraying and mixing<br />

of fuel jets and then, more closely, to the combustion of isolated droplets. Finally the<br />

evaporation of droplets with no combustion will be consi<strong>de</strong>red.<br />

13.2 Atomization<br />

A great effort has been ma<strong>de</strong> both theoretically and experimentally for the study of<br />

the atomization of fuel jets [8], [9]. The process <strong>de</strong>pends on:<br />

1) The geometrical characteristics of the atomizer.<br />

2) The physical properties of the liquid and the atmosphere in which it discharges.<br />

3) The working conditions.<br />

Depending on the discharge velocity of the liquid and on the conditions of the surrounding<br />

atmosphere, several states of atomization can be observed [9]. In the discharge<br />

at high velocity, normally produced in burners, the atomization starts at the<br />

nozzle of the atomizer. Atomization is produced by the turbulence of the liquid, whose<br />

radial velocities tend to brake the jet, and by the action of the atmosphere at the outlet.<br />

Surface tension and viscosity of the liquid oppose to the jet atomization. The prob-


13.2. ATOMIZATION 303<br />

lem has been reviewed by Heinze [10], who studied in <strong>de</strong>tail the action of its various<br />

factors.<br />

Atomization is characterized by the average size of the droplets and by their<br />

distribution in sizes. The average size of the spray droplets is characterized by the<br />

Sauter diameter d s [11]. This is the diameter that would be obtained in an i<strong>de</strong>al atomization<br />

where all the droplets were of equal size and where the surface and total<br />

volume of the fuel were equal to those of actual atomization. The value d s <strong>de</strong>pends<br />

on the variables that characterize the particular conditions of atomization, according<br />

to rules empirically <strong>de</strong>termined (see Ref. [9], p. 117 and f.).<br />

The size distribution in characterized by the mass fraction R (d/d s ) of the fuel<br />

corresponding to droplets of a diameter larger than d. Therefore, the size distribution<br />

dR<br />

function is<br />

d (d/d s ) . Several empirical formulas have been proposed for R (d/d s) or<br />

dR<br />

, [12]. The two formulas generally used are the Rosin-Rammler formula [13]<br />

d (d/d s )<br />

and the Nukiyama-Tanasawa formula [14]. The results given by these two formulas<br />

are in fair agreement with the distributions experimentally observed [15], [16]. These<br />

formulas are as follows.<br />

Rosin-Rammler:<br />

or else<br />

( ) δ−1<br />

dR d<br />

d (d/d m ) = δ e −(d/d m) δ , (13.1)<br />

d m<br />

R = 1 − e −(d/d m) δ , (13.2)<br />

In these formulas d m is an average diameter, characteristic of the size of atomization,<br />

which is related to the Sauter diameter through formula<br />

d m<br />

d s<br />

(<br />

= Γ 1 − 1 )<br />

, (13.3)<br />

δ<br />

where Γ is the factorial function, 1 and δ is a characteristic parameter of size uniformity.<br />

When δ increases the distribution becomes more uniform.<br />

Nukiyama-Tanasawa:<br />

( ) 5<br />

dR<br />

d (d/d m ) = δ d<br />

e −(d/d m) δ (13.4)<br />

Γ (6/δ) d m<br />

1 Also called gamma function, <strong>de</strong>fined as Γ(a) = R ∞<br />

0 ta−1 e −t dt, Ed.


304 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

or else<br />

R =<br />

γ<br />

(6/δ, (d/d m ) δ)<br />

Γ (6/δ)<br />

(13.5)<br />

Here the parameters have the same meaning that in the Rosin–Rammler formula and γ<br />

is the incomplete gamma function. 2 The Sauter diameter is related with d m by means<br />

of<br />

(<br />

ds<br />

d m<br />

)<br />

= Γ (6/δ)<br />

Γ (5/δ) . (13.6)<br />

13.3 Mixing<br />

Once the fuel is atomized it mixes with the surrounding atmosphere due to the action<br />

of turbulence. Lately, Longwell and Weiss [17] have studied the problem for the case<br />

where the fuel atomizes in a turbulent gas stream flowing with uniform velocity. Let<br />

f be the ratio of the fuel mass to the air mass. Assuming that:<br />

1) Turbulent diffusion produces only transversely to the main flow.<br />

2) Mean motion is stationary.<br />

3) The process has axial symmetry.<br />

The following approximate equation for f is obtained<br />

∂f<br />

∂x = E ( )<br />

1 ∂f<br />

v r ∂r + ∂2 f<br />

∂r 2 . (13.7)<br />

Here v is the motion velocity, E is the coefficient of turbulent diffusion of the<br />

fuel and x and r are the cylindrical coordinates of the system. For the <strong>de</strong>rivation of this<br />

equation E is assumed to be constant. Equation (13.7) is i<strong>de</strong>ntical to the molecular<br />

diffusion equation un<strong>de</strong>r similar conditions.<br />

If the fuel is vaporized, E is the coefficient of turbulent diffusion <strong>de</strong>fined by<br />

Taylor [18]. In this case, E equals the product of intensity by scale of turbulence.<br />

If the fuel is in the liquid state, E is appreciably smaller due to the inertia of the<br />

droplets which are unable to follow the air fluctuations. In a typical case calculated by<br />

Longwell and Weiss, E was only 35% of the value corresponding to the diffusion of<br />

the vapour.<br />

If the boundary conditions in the atomizing section are known, equation (13.7)<br />

can be integrated. For example, if the mixing starts from the origin of coordinates,<br />

which acts as a source of fuel of strength G c , and if the duct radius is large, the<br />

2 Defined as γ(a, x) = R x<br />

0 ta−1 e −t dt, Ed.


13.4. COMBUSTION 305<br />

solution corresponding to Eq. (13.7) is<br />

f = G vr2<br />

c v<br />

e− 4Ex , (13.8)<br />

G a 4πE<br />

where G a is the air mass flux in the duct. This solution corresponds, for example, to<br />

the case of a cylindrical atomizer discharging downstream of the gas flow. Formula<br />

(13.8) can be used to obtained the value for E.<br />

From this fundamental solution, the solutions corresponding to different distributions<br />

of fuel sources in the atomizing section can be found by superposition. Such<br />

as discs, rings, etc. These solutions represent with fair approximation the actual mixing<br />

processes for several practical cases, such as the upstream atomization or the case<br />

of a swirl atomizer, etc. Some of these solutions can be found in the work by Longwell<br />

and Weiss.<br />

13.4 Combustion<br />

The study un<strong>de</strong>r given conditions of the combustion of a fuel spray, for example, un<strong>de</strong>r<br />

the prevailing conditions in the combustion chamber of a jet engine, is very arduous.<br />

At present it can only be attempted empirically. 3 Therefore it is justified as a preliminary<br />

phase to analyze the combustion of isolate droplets un<strong>de</strong>r laboratory conditions.<br />

This problem has lately been studied successfully both theoretically and experimentally<br />

by several investigators. Such problem is the subject of this and the following<br />

paragraphs where it will he consi<strong>de</strong>red in <strong>de</strong>tail. Further on a complementary study<br />

will be ma<strong>de</strong> on the evaporation of the droplet when combustion is absent.<br />

Experimental evi<strong>de</strong>nce shows that combustion takes place within a region of<br />

small thickness surrounding the droplet. From each si<strong>de</strong> towards this region diffuse<br />

fuel vapours and oxygen. Burnt gases diffuse from this region towards the exterior.<br />

Therefore we have a diffusion flame of the type consi<strong>de</strong>red in chapter 12. Thus it<br />

will be assumed that the flame thickness is zero and that the mass fractions of fuel<br />

and oxidizer at the flame are also zero. Moreover, the reacting species must diffuse<br />

towards the flame in the stoichiometric ratio.<br />

The flame must supply the necessary heat for the previous evaporation of the<br />

fuel. This heat is transferred to the droplet surface partially through conduction and<br />

partially through radiation. The energy transmitted through radiation is only a small<br />

fraction of the heat absorbed by the droplet through conduction. Therefore, in the<br />

3 See §14 of the present chapter.


306 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

present approximate study its influence is neglected. Otherwise, the analysis of the<br />

problem becomes far more complicated. By doing so the results obtained do not<br />

substantially varies.<br />

The heat received by the droplet is partially used for its heating and partially for<br />

the evaporation of the fuel. The first fraction increases the temperature of the droplet<br />

up to the boiling point T s of the fuel. Thereon it remains constant and equal to T s . All<br />

the heat received is used up in evaporating the fuel. Hereinafter, the transient initial<br />

period will be neglected by assuming that the droplet temperature is initially T s and<br />

that no thermal gradients exist therein.<br />

Strictly the process is non-stationary since the size of the droplet <strong>de</strong>creases as<br />

combustion progresses. However, the recession velocity of the droplet surface is very<br />

small compared to the velocity of the burnt gases. Thereby, an excellent approximation<br />

is obtained by assuming that the phenomenon is stationary. That is to say by<br />

neglecting the local variations with time of temperature and mass fractions produced<br />

by the droplet reduction in size. This is an important simplification of the problem<br />

since it eliminates an in<strong>de</strong>pen<strong>de</strong>nt variable from the equations.<br />

The existence of free and forced convection introduces a privileged direction<br />

<strong>de</strong>stroying the spherical symmetry of the phenomenon. Even in the case of free convection<br />

with approximately round droplets experimental results show that the shape<br />

of the flame differs appreciably from the spherical within the upper region. A study of<br />

the problem, taking into account these convection effects is very arduous and has not<br />

yet been achieved. Nevertheless, when assuming that both the droplet and the flame<br />

front are spherical, that is when neglecting these effects, the results obtained show a<br />

good agreement with the experimental results as for the overall magnitu<strong>de</strong>s such as the<br />

droplet burning velocity and its law of variation with size. Thereby, neglect of convection<br />

effects is justified. With this assumption and the assumption of quasi-stationary<br />

state previously stated, there is only one in<strong>de</strong>pen<strong>de</strong>nt variable, this is the distance r to<br />

the center of the droplet.<br />

The composition of the fuel vapours, oxidizing atmosphere and burnt gases is<br />

very complex. However, as done in the study of diffusion flames, we shall assume for<br />

simplicity that only three different chemical species exist, namely: fuel, oxidizer and<br />

inert gases, which inclu<strong>de</strong> these in the atmosphere surrounding the droplet as well as<br />

the combustion products.<br />

Summarizing, the following assumptions are adopted:<br />

1) The phenomenon has spherical symmetry.


13.5. NOTATION 307<br />

2) The phenomenon is quasi-stationary.<br />

3) Combustion takes place at constant pressure.<br />

4) The droplet temperature is uniform and equal to the boiling temperature of the<br />

fuel at ambient pressure.<br />

5) The chemical reaction takes place upon a spherical surface, named the flame<br />

front. Fuel vapours and oxygen diffuse in the stoichiometric ratio towards this<br />

flame front from which the burnt gases flow. The gases that reach the flame front<br />

react instantaneously. Thereby, on the flame front both the mass fractions of fuel<br />

vapours and of oxygen are zero.<br />

6) Only three chemical species exist, namely: fuel, oxygen and inert gases.<br />

These assumptions allow a simple analysis of the problem and lead to results<br />

which have been experimentally confirmed. Such a treatment of the problem has been<br />

applied by Godsave [19] and Spalding [3] in England, and by Penner and Goldsmith<br />

[20] in the U.S.A. Asi<strong>de</strong> from the aforementioned assumptions Godsave and Spalding<br />

also assume that transport coefficients are in<strong>de</strong>pen<strong>de</strong>nt from temperature by adopting<br />

a mean value for these coefficients. This rather arbitrary assumption reduces appreciably<br />

the suitability of the method. In fact, for the existing range of temperatures, these<br />

coefficients can vary in a ratio of ten to one. Goldsmith and Penner take into account<br />

this variation as well as that of heat capacities. The following study is mainly based<br />

on the work of these two authors but it assumes that heat capacities are in<strong>de</strong>pen<strong>de</strong>nt<br />

from temperature since their variation has no substantial influence on the results.<br />

13.5 Notation<br />

In the present study the notation used in the preceding chapter will be completed with<br />

the following (see Fig. 13.1):<br />

M = Droplet mass<br />

m = Mass flow across a closed surface surrounding the droplet.<br />

q l = Latent heat of evaporation per unit mass of fuel.<br />

r = Radial distance to the center of the droplet.<br />

v = Radial velocity of the mixture.<br />

v j = Radial velocity of species A j .<br />

v jd = Radial diffusion velocity of species A j .<br />

Y j = Mass fraction of species A j .<br />

ν = Stoichiometric ratio of oxygen mass to fuel mass.


308 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

Subscripts:<br />

1 = Fuel<br />

2 = Inert gases<br />

3 = Oxygen<br />

e = Exterior to the flame front<br />

i = Interior to the flame front<br />

l = On the flame<br />

s = On the droplet surface<br />

∞ = At great distance from the droplet<br />

Y<br />

Y 2<br />

r l<br />

Y 1<br />

Y 3<br />

Y 2 oo<br />

oo<br />

r s<br />

T s<br />

r<br />

l<br />

T<br />

r<br />

Y 3<br />

oo T<br />

T<br />

T l<br />

Figure 13.1: Schematic diagram of the droplet combustion problem.<br />

13.6 Continuity equations<br />

Since the process is stationary, the fluid mass that crosses any spherical surface concentric<br />

to the droplet must be constant. Due to the spherical symmetry of the problem<br />

such condition can be expressed in the form<br />

4πr 2 ρv = const. (13.9)<br />

The constant can be <strong>de</strong>termined by particularizing this equation on the droplet surface<br />

r = r s . This surface acts as a fuel source. The strong of this source is the mass m of<br />

vapour produced per unit time. Therefore, we have<br />

4πr 2 ρv = m. (13.10)<br />

This equation is valid throughout the fluid space surrounding the droplet.


13.6. CONTINUITY EQUATIONS 309<br />

Chemical reactions take place only on the flame surface r = r l , which acts as<br />

a sink for fuel and oxidizer and as a source for the combustion products. Therefore<br />

one obtains for each different chemical species an equation similar to Eq. (13.10). In<br />

this equation ρ must be substituted by the partial <strong>de</strong>nsity ρ i = ρY i corresponding to<br />

species A i , v by velocity v i = v + v di , where v di is the radial diffusion velocity of the<br />

species, and m must be substituted by the partial flow m i of the species. Thus<br />

4πr 2 ρY i v i = m i , (i = 1, 2, 3). (13.11)<br />

Moreover, the following evi<strong>de</strong>nt conditions must be satisfied<br />

∑<br />

Y i v i = v, (13.12)<br />

i<br />

∑<br />

m i = m. (13.13)<br />

i<br />

Equation (13.11) is valid for each species A i , as long as r does not cross the flame.<br />

Let us now apply these equations separately to the interior and exterior regions<br />

of the flame.<br />

Interior region r s ≤ r ≤ r l<br />

In this region only fuel vapours and inert gases exist, since the incoming oxygen is<br />

entirely consumed as it reaches the flame without crossing it.<br />

Equation (13.11) applied to the fuel vapours A 1 gives<br />

4πr 2 ρY 1 v 1 = m 1 = m, (13.14)<br />

since on the droplet surface the mass flow m is the mass flow m 1 of the fuel produced<br />

by evaporation.<br />

By comparing (13.9) and (13.14)<br />

Y 1 v 1 = v. (13.15)<br />

Equation (13.12) gives<br />

Y 1 v 1 + Y 2 v 2 = v. (13.16)<br />

From Eqs. (13.15) and (13.16) results<br />

v 2 = 0, (13.17)<br />

that is to say within the interior region the inert gases are at rest and the fuel vapours<br />

diffuse through them towards the flame.


310 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

Exterior region r ≥ r l<br />

In this region only oxygen and inert gases exist. Equation (13.11) applied to each of<br />

them gives<br />

Inert gases: 4πr 2 ρY 2 v 2 =m 2 . (13.18)<br />

Oxygen: 4πr 2 ρY 3 v 3 =m 3 . (13.19)<br />

Similarly, Eqs. (13.12) and (13.14) give<br />

Y 2 v 2 + Y 3 v 3 = v, (13.20)<br />

m 2 + m 3 = m. (13.21)<br />

Let ν be the stoichiometric ratio of oxygen mass to fuel mass. Since, the fuel<br />

mass reaching the flame per unit time is m, the oxygen mass that reaches the flame<br />

per unit time must be νm. And since the oxygen moves towards the flame, v 3 must be<br />

negative, that is<br />

m 3 = −νm. (13.22)<br />

From here and (13.21) there results for m 2<br />

m 2 = (1 + ν)m. (13.23)<br />

Equations (13.22) and (13.23) <strong>de</strong>termine the constants for the continuity equations<br />

(13.18) and (13.19) as functions of m, thus obtaining<br />

4πr 2 ρY 2 v 2 = (1 + ν)m, (13.24)<br />

4πr 2 ρY 3 v 3 = −νm. (13.25)<br />

Therefore, within the exterior region the flame acts as a sink of strength νm for the<br />

oxygen and as a source of strength (1 + ν)m for the combustion products.<br />

13.7 Energy equation<br />

The process is stationary and it takes place at constant pressure. Moreover the kinetic<br />

energy of the motion and the work of the viscous forces can be neglected. Therefore<br />

the summation of the heat and enthalpy fluxes through any spherical surface concentric<br />

to the droplet must be constant. Due to the spherical symmetry, this condition can be<br />

expressed in the form<br />

∑<br />

i<br />

m i h i − 4πr 2 λ dT<br />

dr<br />

= const. (13.26)


13.7. ENERGY EQUATION 311<br />

The value for the constant can be obtained by applying this equation to the droplet<br />

surface r = r s . Since, here m 1 = m and m 2 = m 3 = 0, we have<br />

(<br />

mh 1s − 4πrs<br />

2 λ dT )<br />

= const., (13.27)<br />

dr<br />

(<br />

where 4πrs<br />

2 λ dT )<br />

is the heat received by the droplet surface per unit time. Since<br />

dr<br />

s<br />

the droplet temperature is assumed to be constant and equal to the boiling temperature<br />

T s of the fuel, this heat must be used in evaporating the combustible. Furthermore<br />

this heat is the only source of energy for the evaporation, since the energy received<br />

from the flame by heat radiation has been neglected. 4<br />

s<br />

Let q l be the latent heat of<br />

evaporation per unit mass at temperature T s . Since the evaporated mass per second is<br />

m, the following condition is obtained<br />

(<br />

4πrs<br />

2 λ dT )<br />

= mq l . (13.28)<br />

dr<br />

s<br />

Consequently, the value for the constant is m(h 1s − q l ), which taken into<br />

Eq. (13.26) gives<br />

∑<br />

i<br />

m i h i − 4πr 2 λ dT<br />

dr = m(h 1s − q l ). (13.29)<br />

The specific enthalpy h i of species A i can be expressed in the form<br />

∫ T<br />

h i = h 0i + c pi dT. (13.30)<br />

T 0<br />

When this expression is substituted into (13.29) we obtain<br />

4πr 2 λ dT ∫ ( )<br />

T ∑<br />

dr − m i c pi dT = m(q l − h 1s ) + ∑<br />

T 0 i<br />

i<br />

m i h 0i . (13.31)<br />

The form taken by this expression within the interior and exterior regions will<br />

be studied separately.<br />

Interior region r s ≤ r ≤ r l<br />

As aforesaid in this region only fuel vapours ant inert gases exist. Furthermore due<br />

to Eq. (13.17) the flow of inert gases throughout the spherical control surface is zero.<br />

4 This matter has been consi<strong>de</strong>red by G.A. Godsave who conclu<strong>de</strong>s that the radiation energy received<br />

by the droplet per gram of evaporated fuel is approximately proportional to the droplet radius. For large<br />

droplets (r s ≃ 1 mm) it can become a 20% of the energy nee<strong>de</strong>d to produce evaporation.


312 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

Therefore, there only remains the enthalpy flux of the fuel vapours and Eq. (13.31)<br />

reduces to<br />

4πr 2 λ dT<br />

dr − m ∫ T<br />

T 0<br />

c p1 dT = m<br />

(<br />

q l −<br />

∫ Ts<br />

T 0<br />

c p1 dT<br />

)<br />

, (13.32)<br />

which <strong>de</strong>termines the distribution of temperature between the droplet surface and the<br />

flame front.<br />

The integration constant of Eq. (13.32) is <strong>de</strong>termined by expressing that the<br />

value of the temperature on the droplet surface must be T s , that is<br />

r = r s : T = T s . (13.33)<br />

Exterior region r ≥ r l<br />

Here, only oxygen and inert gases exist and their flows are given by Eqs. (13.22) and<br />

(13.23) respectively. Therefore, the energy equation (13.31) takes the form<br />

4πr 2 λ dT ∫ T<br />

dr − m ( )<br />

(1 + ν)cp2 − νc p3 dT =<br />

T<br />

(<br />

0<br />

= m q l − ( ∫ )<br />

) Ts<br />

h 01 + νh 03 − (1 + ν)h 02 − c p1 dT<br />

T 0<br />

(13.34)<br />

Let q r be the heat of reaction per gram of fuel in gas state at the temperature<br />

T 0 . This heat is<br />

q r = h 01 + νh 02 − (1 + ν)h 03 . (13.35)<br />

Substituting Eq. (13.35) into (13.34), the latter takes the form<br />

4πr 2 λ dT ∫ (<br />

T<br />

∫ )<br />

Ts<br />

dr − m c p dT = m q l − q r − c p1 dT , (13.36)<br />

T 0<br />

T 0<br />

where, in short<br />

c p = (1 + ν)c p2 − νc p3 . (13.37)<br />

The constant resulting from the integration of this equation is <strong>de</strong>termined by<br />

expressing that the temperature at great distance from the droplet has the value T ∞<br />

r → r ∞ : T → T ∞ . (13.38)<br />

The continuity of the temperature at the flame imposes an additional condition<br />

T (r − l ) = T (r+ l<br />

), (13.39)<br />

which will be used in <strong>de</strong>termining the position of the flame.


13.8. DIFFUSION EQUATIONS 313<br />

13.8 Diffusion equations<br />

To compute the distribution of the mass fractions of fuel vapours and oxygen, the<br />

equations (13.14) and (13.25) must be integrated. For this v 1 and v 3 must be expressed<br />

as functions of the mixture velocity v and of the mass fractions Y 1 and Y 3 by using the<br />

diffusion equations. Both the interior and exterior region will be studied separately.<br />

Interior region r s ≤ r ≤ r l<br />

The continuity equation 13.14) for the fuel vapour can be written in the form<br />

4πr 2 ρY 1 (v + v d1 ) = m, (13.40)<br />

where v d1 is the diffusion velocity of the fuel vapours through the atmosphere of inert<br />

gases. Fick’s law, 5 gives for this velocity<br />

v d1 = − 1 Y 1<br />

D 12<br />

dY 1<br />

dr , (13.41)<br />

where D 12 is the diffusion coefficient for the fuel vapours and the inert gases.<br />

When Eq. (13.41) is substituted into Eq. (13.40) and Eq. (13.10) is taken into<br />

account, one obtains<br />

4πr 2 ρD 12<br />

dY 1<br />

dr = −m(1 − Y 2) (13.42)<br />

for the <strong>de</strong>termination of Y 1 . The integration constant for this equation is <strong>de</strong>termined<br />

by expressing that the mass fraction of the fuel vapours at the flame is zero<br />

r = r l : Y 1 = 0, (13.43)<br />

Exterior region r ≥ r l<br />

This procedure applied to Eq. (13.25) gives the following equation for the distribution<br />

of oxygen in the exterior region of the flame<br />

4πr 2 ρD 23<br />

dY 3<br />

dr = m(ν + Y 3). (13.44)<br />

The solution to this equation must satisfy the condition that on the flame front mass<br />

fraction of oxygen must be zero,<br />

r = r l : Y 3 = 0, (13.45)<br />

which <strong>de</strong>termines the value for the corresponding integration constant.<br />

5 See chapter 2.


314 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

The solution of Eq. (13.44) must satisfy another additional condition. In fact,<br />

the mass fraction of oxygen must tend to the value Y 3∞ corresponding to the composition<br />

of the atmosphere surrounding the droplet at great distance from the same<br />

r → ∞ : Y 3 → Y 3∞ . (13.46)<br />

This equation will be used in <strong>de</strong>termining the burning velocity m of the droplet.<br />

13.9 Combustion velocity of the droplet. Temperature<br />

and position of the flame<br />

From the preceding paragraph it results that the solution of the combustion problem<br />

of a droplet reduces to the following:<br />

1) The integration of the system of differential equations<br />

4πr 2 λ dT ∫ (<br />

T<br />

∫ )<br />

Ts<br />

dr − m c p1 dT =m q l − c p1 dT , (13.32)<br />

T 0<br />

T 0<br />

4πr 2 ρD 12<br />

dY 1<br />

dr = − m(1 − Y 1), (13.42)<br />

for the <strong>de</strong>termination of temperature T and mass fraction Y 1 of the fuel vapours<br />

within the interior region r s ≤ r ≤ r l , with the two boundary conditions<br />

r = r s : T = T s , (13.33)<br />

r = r − l<br />

: Y 1 = 0. (13.43)<br />

2) The integration of the differential equations<br />

4πr 2 λ dT ∫ (<br />

T<br />

dr − m c p dT =m q l − q r −<br />

T 0<br />

with boundary conditions<br />

∫ Ts<br />

T 0<br />

c p1 dT<br />

)<br />

, (13.36)<br />

4πr 2 ρD 23<br />

dY 3<br />

dr =m(ν + Y 3), (13.44)<br />

r →∞ : T →T ∞ , (13.38)<br />

r =r + l<br />

: Y 3 =0, (13.45)<br />

for the <strong>de</strong>termination of temperature and mass fraction Y 3 of oxygen within the<br />

exterior region r ≥ r l .


13.9. COMBUSTION VELOCITY OF THE DROPLET 315<br />

When Y 1 and Y 3 are known, the distribution of inert gases Y 2 is <strong>de</strong>termined by<br />

the following equations<br />

r s ≤ r ≤ r l : Y 2 = 1 − Y 1 ,<br />

r ≥ r 1 : Y 2 = 1 − Y 3 .<br />

(13.47)<br />

The solution of the system must also satisfy conditions<br />

r = r l : T (r − l ) = T (r+ l<br />

), (13.39)<br />

r → ∞ : Y 3 → Y 3∞ . (13.46)<br />

These two additional conditions <strong>de</strong>termine, as aforesaid, the burnt burning velocity m<br />

of the droplet and the position r l of the flame front.<br />

The flame temperature T l is the value for T given by the solution of Eq. (13.32),<br />

or (13.36) since both values are equal for r = r l .<br />

For the integration of these systems the specific enthalpies of the various species<br />

must be known. Therefore, the laws of variation with temperature of the heat capacities<br />

at constant pressure for the said species must be known. Furthermore one must<br />

also know the coefficients of thermal conductivity and diffusion as functions of the<br />

temperature and composition of the mixture. As well as the heat of reaction q r and the<br />

latent heat of evaporation q l of the fuel. Hereinafter it will be assumed that the heat<br />

capacity at constant pressure does not vary with temperature, within the range un<strong>de</strong>r<br />

consi<strong>de</strong>ration. 6 In such case Eq. (13.32) takes the form<br />

and Eqs. (13.32) and (13.36) reduce respectively to<br />

and<br />

h i = h 0i + c pi (T − T 0 ), (13.48)<br />

4πr 2 λ dT<br />

dr − mc p1T = m(q l − c p1 T s ) (13.49)<br />

4πr 2 λ dT<br />

dr − mc pT = m ( q l − q r − c p T 0 − c p1 (T s − T 0 ) ) . (13.50)<br />

Such is the form in which these equations are used in the following calculations.<br />

As for the transport coefficients we shall only consi<strong>de</strong>r the case where λ is<br />

in<strong>de</strong>pen<strong>de</strong>nt from the mixture composition but varies proportionally to the absolute<br />

temperature. That is<br />

6 Goldsmith and Penner in [20] take into consi<strong>de</strong>ration the variation of the heat capacity with temperature.<br />

λ<br />

T<br />

= const., (13.51)


316 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

where, in general, the value for the constant will vary when passing from the interior<br />

to the exterior region.<br />

Moreover, it is assumed that the following condition is satisfied<br />

λ<br />

ρD ij c pi<br />

= δ i = const., (13.52)<br />

in accordance with the results of they elementary theory of diffusion. Experimental<br />

evi<strong>de</strong>nce shows that this condition is approximately satisfied.<br />

13.10 Integration of the equations<br />

Un<strong>de</strong>r the assumptions stated in §9 the system of equations (13.49) and (13.42), corresponding<br />

to the interior region, takes the form<br />

4πr 2 λ s<br />

T<br />

T s<br />

dT<br />

dr − mc p1T = m(q l − c p1 T s ), (13.53)<br />

where the values of λ and ρD 12 are expressed as<br />

4πr 2 (ρD 12 ) s<br />

T<br />

T s<br />

dY l<br />

dr = m(1 − Y 1), (13.54)<br />

λ = λ s<br />

T<br />

T s<br />

, (13.55)<br />

ρD 12 = (ρD 12 ) s<br />

T<br />

T s<br />

, (13.56)<br />

as functions of the corresponding values on the droplet surface and of the ratio of absolute<br />

temperatures T/T s . The integration of Eq. (13.53) is straightforward. Making<br />

use of condition (13.33) one obtains<br />

(<br />

1<br />

− 1 r s r = 4πλ (<br />

s T<br />

− 1 + 1 − q ) [<br />

l<br />

ln 1 + c ( )] )<br />

p1T s T<br />

− 1 . (13.57)<br />

mc p1 T s c p1 T s q l T s<br />

The integration of (13.54) is simplified by previously eliminating r between<br />

(13.53) and (13.54). Thus one obtains for as a function of T<br />

( )δ cp1 (T − T s ) + q 1 l<br />

Y 1 = 1 −<br />

, (13.58)<br />

c p1 (T l − T s ) + q l<br />

where conditions (13.43) has been used and according to (13.44) and δ 1 is given by<br />

δ 1 =<br />

λ<br />

ρD 12 c p1<br />

. (13.59)


13.10. INTEGRATION OF THE EQUATIONS 317<br />

Similarly, the system of Eqs. (13.50) and (13.44) corresponding to the exterior<br />

region takes the form<br />

4πr 2 λ ∞<br />

T<br />

T ∞<br />

dT<br />

dr − mc pT = m ( q l − q r − c p T 0 − c p1 (T s − T 0 ) ) , (13.60)<br />

4πr 2 (ρD 23 ) ∞<br />

T<br />

T ∞<br />

dY 3<br />

dr = m(ν + Y 3), (13.61)<br />

where λ and ρD 23 are expressed as functions of ratio T/T ∞ and of their values at<br />

infinity.<br />

Proceeding as for the interior region and with boundary conditions (13.38) and<br />

(13.45), the following solution is obtained for this system<br />

1<br />

r = 4πλ (<br />

∞<br />

1 − T +<br />

q ln q − c )<br />

pT ∞<br />

, (13.62)<br />

mc p T ∞ c p T ∞ q − c p T<br />

( ( ) ) δ q − cp T<br />

Y 3 = ν<br />

− 1 , (13.63)<br />

q − c p T l<br />

where the following abbreviations<br />

q = q r − q l + (c p − c p1 )T 0 + c p1 T s , (13.64)<br />

δ =<br />

are used for simplicity.<br />

λ<br />

ρD 23 c p<br />

=<br />

λ<br />

ρD 23 c p3<br />

c p3<br />

c p<br />

=<br />

δ 3<br />

(1 + ν) c p2<br />

c p3<br />

ν , (13.65)<br />

Taking Eq. (13.46) into (13.63) one obtains<br />

( (q ) ) δ − cp T ∞<br />

Y 3∞ = ν<br />

− 1 . (13.66)<br />

q − c p T l<br />

This expression gives for T l<br />

(<br />

T l = T ∞ 1 + Y ) (<br />

−1/δ<br />

3∞<br />

+ q (<br />

1 − 1 + Y ) ) −1/δ<br />

3∞<br />

. (13.67)<br />

ν c p ν<br />

Consequently, the temperature of the flame is in<strong>de</strong>pen<strong>de</strong>nt from its position<br />

and from the size of the droplet.<br />

When condition (13.39) is expressed making use of Eqs. (13.57) and (13.62)<br />

one obtains<br />

(<br />

1<br />

= 4πλ s<br />

r s mc p1<br />

T l<br />

T s<br />

− 1 +<br />

+ 4πλ (<br />

∞<br />

1 − T l<br />

+<br />

mc p T ∞<br />

(<br />

1 − q ) [<br />

l<br />

ln 1 + c ( )] )<br />

p1T s Tl<br />

− 1<br />

c p1 T s q l T s<br />

q ln q − c )<br />

pT ∞<br />

.<br />

c p T ∞ q − c p T l<br />

(13.68)


318 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

From which results for m<br />

m<br />

= 4πλ (<br />

∞<br />

1 − T l<br />

+<br />

q ln q − c )<br />

pT ∞<br />

r s c p T ∞ c p T ∞ q − c p T l<br />

(<br />

+ 4πλ (<br />

s T l<br />

− 1 + 1 − q ) [<br />

l<br />

ln 1 + c ( )] )<br />

p1T s Tl<br />

− 1 .<br />

c p1 T s c p1 T s q l T s<br />

(13.69)<br />

Since the right-hand si<strong>de</strong> of this equation <strong>de</strong>pends only on the physicochemical characteristics<br />

of the process, it results that the burning velocity of the droplet is directly<br />

proportional to its radius. There is a simple explanation to this fact. The evaporated<br />

mass is proportional to the droplet surface, Σ s , and to the heat Q received per unit<br />

surface and per unit time, m ∼ Σ s Q. But, Σ s ∼ rs. 2 Furthermore, Q ∼ 1/r s , since<br />

Q is proportional to the temperature gradient and this gradient is inversely proportional<br />

to the distance from the flame to the droplet surface, which is proportional to<br />

r s . Therefore it results m ∼ rs(1/r 2 s ) = r s .<br />

Let ρ c be the <strong>de</strong>nsity of the fuel. The droplet mass M is<br />

M = 4 3 πr3 sρ e (13.70)<br />

and its variation with time is<br />

− dM dt<br />

= −4πr 2 sρ e<br />

dr s<br />

dt . (13.71)<br />

When this expression is combined with Eq. (13.69) the following law is obtained<br />

for the time variation of the droplet radius<br />

2r s<br />

dr s<br />

dt<br />

= −k, (13.72)<br />

where<br />

k =<br />

m = 2λ (<br />

∞<br />

1 − T l<br />

+<br />

2πρ e r s ρ e c p T ∞<br />

(<br />

+ 2λ s T l<br />

− 1 +<br />

ρ e c p1 T s<br />

q ln q − c )<br />

pT ∞<br />

c p T ∞ q − c p T l<br />

(<br />

1 − q l<br />

c p1 T s<br />

)<br />

ln<br />

[<br />

1 + c p1T s<br />

q l<br />

is a constant of the process named evaporation constant. 7<br />

( )] ) (13.73)<br />

Tl<br />

− 1<br />

T s<br />

When (13.72) is integrated the following expression is obtained for the law of<br />

variation of the droplet radius as a function of time<br />

7 The evaporation constant used by Godsave is equal to 4k.<br />

r 2 s = r 2 si − kt, (13.74)


13.11. NUMERICAL APPLICATION 319<br />

where r si is the initial radius. This formula is the fundamental result of the present<br />

theory.<br />

The combustion time t c of a droplet of radius r si is, obviously,<br />

t c = r2 si<br />

k . (13.75)<br />

The position r l /r s of the flame front is obtained by writing T = T l in Eq. (13.62).<br />

Furthermore if m/r s is eliminated from the result by means of Eq. (13.73), one obtains<br />

r l<br />

r s<br />

=<br />

1 − T l<br />

T ∞<br />

+<br />

kρ e c p<br />

2λ ∞<br />

q ln q − c . (13.76)<br />

pT ∞<br />

c p T ∞ q − c p T l<br />

Therefore, the radius of the flame and the droplet are proportional.<br />

Formula (13.73) enables the study of the influence of the physical constants of<br />

the fuel and the state of the surrounding atmosphere on the burning rate of the droplet. 8<br />

In particular, the said formula shows that the heat transmitted from the flame to the<br />

droplet surface is the <strong>de</strong>termining factor for its burning velocity. This is due to the<br />

fact that the burning rate of the droplet is <strong>de</strong>termined by the heat received by the same.<br />

Therefore, the essential factor in the combustion process is not the volatility of the fuel<br />

but its heat of evaporation. Experiments results confirm this conclusion. In fuels at the<br />

high temperature reached when combustion occurs, the evaporation mechanism differ<br />

essentially from the evaporation when the temperature of the surrounding atmosphere<br />

is normal. In the last case the evaporation is maintained by the diffusion of the vapors<br />

from the droplet towards the exterior. The evaporation velocity is <strong>de</strong>termined by the<br />

difference between the vapour pressures on the droplet surface and at infinity. 9<br />

13.11 Numerical application<br />

The above formulas are applied here to the study of the combustion of a gasoline<br />

droplet in the air. The following typical values are adopted<br />

T ∞ = 300 K, T s = 355 K, ρ c = 0.85 gr/cm 3 ,<br />

λ ∞ = λ s = 55 × 10 −6 cal/cm s K, q l = 95 cal/gr, q = 10 000 cal/gr,<br />

c p1 = 0.60 cal/gr K, c p2 = 0.35 cal/gr K, c p3 = 0.26 cal/gr K,<br />

Y 3∞ = 0.23, ν = 3, δ 1 = δ 2 = 0.9.<br />

8 Information on the subject can be found in Godsave’s works [19].<br />

9 See §15, Eq. (13.103).


320 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

obtained<br />

Taking these values into Eq. (13.67) for the temperature of the flame, it is<br />

The evaporation constant is obtained from Eq. (13.73)<br />

The position of the flame is given by (13.76)<br />

T l = 3112 K. (13.77)<br />

k = 2.26 × 10 −3 cm 2 /s. (13.78)<br />

r l<br />

r s<br />

= 9.48. (13.79)<br />

The distributions of temperature and mass fractions are given by Eqs. (13.57) and<br />

(13.58) for the interior region and by Eqs. (13.62) and (13.63) for the exterior region<br />

Interior region<br />

r<br />

r s<br />

=<br />

Y 1 = 1 −<br />

12.42<br />

13.60 − T − 0.66 ln<br />

T ∞<br />

(<br />

1.89 T<br />

T ∞<br />

− 1.24<br />

), (13.80)<br />

( ( ) ) 0.9<br />

1 T<br />

− 0.56 . (13.81)<br />

9.72 T ∞<br />

Exterior region<br />

r<br />

r s<br />

=<br />

Y 3 = 3 ⎣<br />

10.84<br />

1 − T + 53.76 ln<br />

T ∞<br />

⎡(<br />

1<br />

43.38<br />

(<br />

53.76 − T<br />

T ∞<br />

) ) 0.38<br />

52.76<br />

53.76 − T<br />

T ∞<br />

, (13.82)<br />

⎤<br />

− 1⎦ . (13.83)<br />

The distribution of Y 2 is given in the interior region by<br />

Y 2 = 1 − Y 1 , (13.84)<br />

and in the exterior region by<br />

Y 2 = 1 − Y 3 . (13.85)<br />

The distribution of T/T ∞ , Y 1 , Y 2 and Y 3 have been taken into Figs. 13.2 and 13.3.


13.11. NUMERICAL APPLICATION 321<br />

10<br />

8<br />

T/T ∞<br />

6<br />

4<br />

2<br />

T s<br />

/T ∞<br />

T l<br />

/T ∞<br />

r l<br />

/r s<br />

=9.48<br />

0<br />

0 5 10 15 20 25 30<br />

r/r s<br />

Figure 13.2: Temperature profile given by Eqs. (13.80) and (13.82).<br />

1.0<br />

Y 1s<br />

Y 2<br />

Y 2<br />

0.8<br />

Y 2∞<br />

Y 1<br />

, Y 2<br />

, Y 3<br />

0.6<br />

0.4<br />

Y 1<br />

Y 3<br />

0.2<br />

Y 3∞<br />

Y 2s<br />

r l<br />

/r s<br />

=9.48<br />

0.0<br />

0 5 10 15 20 25 30<br />

r/r s<br />

Figure 13.3: Mass fraction profiles given by Eqs. (13.81) and (13.83)-(13.85).<br />

The combustion time for a droplet of a radius r si millimeters is given by<br />

Eq. (13.75). We obtain<br />

t c = 4.42 rsi 2 s. (13.86)<br />

In Fig. 13.4 the variation law of t c as a function of rsi 2 is represented.


322 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

5<br />

4<br />

3<br />

t c<br />

(s)<br />

2<br />

1<br />

0<br />

0 0.125 0.250 0.375 0.500 0.625 0.750 0.875 1.000<br />

r (mm) si<br />

Figure 13.4: Droplet combustion time as a function of the initial size.<br />

13.12 Comparison with experimental results and limitations<br />

of the theory<br />

Godsave [19], [21], Topps [22], Hall and Die<strong>de</strong>richsen [23], Spalding [24], Hottel<br />

et al.<br />

[25], Wise et al [26], Kobayasi [27], Nishiwaki [28] and others have published<br />

experimental results on the combustion of droplets. Such results have been<br />

obtained from different technical procedures and they confirm the main conclusions<br />

of the present theory. For example, Fig. 13.5 obtained in the I.N.T.A Combustion<br />

Laboratory at <strong>Madrid</strong> 10 confirms the linear law (13.74) of variation of the square of<br />

the radius with time. The following table 13.1, taken from the work by Goldsmith and<br />

Penner [20], compares theoretical values of k with these experimentally obtained by<br />

Godsave [21] for different fuels. As seen the agreement is generally excellent.<br />

Fuels k × 10 3 cm 2 /s<br />

Theoretical<br />

Experimental<br />

Benzene 2.5 2.43<br />

Ethyl alcohol 1.98 2.03<br />

Ethyl Benzene 2.13 2.15<br />

N-heptane 2.15 2.43<br />

Table 13.1: Comparison between theoretical and experimental results for k.<br />

10 The results have been obtained by applying Godsave’s technique, that is to say, by suspension of a<br />

droplet from a quartz filament and photographing the variation of radius as a function of time.


13.12. COMPARISON WITH EXPERIMENTAL RESULTS AND LIMITATIONS OF THE THEORY 323<br />

0.5<br />

0.4<br />

r 2 s × 102 (cm 2 )<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6<br />

t (s)<br />

Figure 13.5: Experimental results showing the linear <strong>de</strong>pen<strong>de</strong>nce between r 2 s and time, obtained<br />

for isooctane in air at ambient pressure.<br />

On the other hand a comparison of the theoretical and experimental radius<br />

of the flame [20] shows that the former is two or three times larger than the latter.<br />

Although formula (13.76) shows that ratio r l /r s is practically in<strong>de</strong>pen<strong>de</strong>nt from<br />

pressure, Wise, Lovell and Wood [26] have experimentally observed that this ratio<br />

increases as pressure <strong>de</strong>creases. The authors explain theoretically this effects as a<br />

consequence of free convection. Photographic evi<strong>de</strong>nce shows that combustion is accompanied<br />

by strong free convection flows due to which the spherical flame appears<br />

only in the lower half of the droplet whilst in the upper half it takes a consi<strong>de</strong>rably<br />

elongate shape. The intense formation of carbon observed makes difficult the optical<br />

study of this region. This convection effect may account for the difference between<br />

the theoretical and experimental values of the flame radius making the values of k<br />

predicted by theory agree with the experimental values as previously seen.<br />

Formula (13.73) shows that the burning velocity of a droplet is practically in<strong>de</strong>pen<strong>de</strong>nt<br />

from pressure. In fact, its only influence shows small reduction of latent<br />

heat of evaporation that accompanies and increase in pressure. However, Hall and<br />

Die<strong>de</strong>richsen [23] have experimentally checked that the burning velocity of a droplet<br />

is approximately proportional to the 4th root of pressure. It has been suggested [20]<br />

that the following factors could account for this effect, asi<strong>de</strong> from the aforementioned<br />

<strong>de</strong>creases in evaporation heat:


324 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

1) Augmentation of the energy transmitted by radiation from the flame to the droplet<br />

which increases very rapidly with pressure.<br />

2) Augmentation of the rate of chemical reactions which could be very significant<br />

if the burning velocity of the droplet would <strong>de</strong>pend, even if only slightly, on the<br />

said reaction rate, contrary to what is postulated in the present theory.<br />

3) Augmentation of the free convection effects due to the increase of the Grashof<br />

number with pressure.<br />

The theoretical study of such effects is very arduous as it implies taking into<br />

account the influence of radiation, finite thickness of the flame and free convection.<br />

So far this study has not been satisfactorily accomplished.<br />

13.13 Influence of convection<br />

The present theory exclu<strong>de</strong>s convection effects. Experiments with suspen<strong>de</strong>d droplets<br />

show that the influence of free convection is not significant. Less experimental information<br />

is available for the case of forced convection. Spalding [24] has performed<br />

some measurements for very large Reynolds numbers (from 400 to 4 000). His experiments<br />

show that for this range the burning velocity of droplets with forced convection<br />

can be computed by applying the Frössling formula 11 for the evaporation with no<br />

combustion. Spalding’s experiments bring forth the fact that, at least in the analyzed<br />

range, two different combustion states can exist. For convection velocities lower than<br />

a given critical value, a semi-spherical flame surrounds the front part of the droplet,<br />

whilst in the opposite si<strong>de</strong> a long wake forms with a strong formation of carbon. If the<br />

convection velocity is larger than the critical value the front flame extinguishes and<br />

the wake flame remains. It is questionable whether this second state also produces in<br />

the case of very small droplets.<br />

Spalding explains this extinction as follows. Even in the present theory the<br />

thickness of the flame is assumed to be zero, actually it is finite. Now, the convection<br />

activates evaporation and increases the flame thickness in or<strong>de</strong>r to burnt the largest<br />

quantity of fuel that must be consumed per second. But such an increase in thickness<br />

is accompanied by a <strong>de</strong>crease in the maximum temperature of the flame. Since reaction<br />

rate changes very rapidly with temperature if the said <strong>de</strong>crease is large enough<br />

the reaction is incomplete and combustion extinguishes. The same occurs in diffusion<br />

flames. 12<br />

11 See Eq. (13.106)<br />

12 See chapter 12.<br />

When assuming a reaction rate of the Arrhenius type, Spalding’s calcula-


13.14. COMBUSTION OF FUEL SPRAYS 325<br />

tions <strong>de</strong>monstrate that the extinction velocity is proportional to the droplet diameter.<br />

Experimental results confirm such prediction. Spalding also arrives to the conclusion<br />

that, even when no convection exists, the flame extinguishes if the diameter of the<br />

droplet is very small.<br />

13.14 Combustion of fuel sprays<br />

The present theory enables the calculation with good approximation of the burning<br />

velocity of an isolated droplet un<strong>de</strong>r laboratory conditions. In particular, this theory<br />

allows the study of the influence of the physical characteristics of the fuel and the<br />

state of the surrounding atmosphere on the burning velocity of the droplet. Now these<br />

results must be exten<strong>de</strong>d to the study of the combustion of fuel sprays of the type<br />

existing in the combustion chambers of jet engines [29]. Such an extension has not<br />

yet been achieved. In the actual state of knowledge this seems impossible due to the<br />

interaction of the many factors of the process. In fact, for this extension to become<br />

possible it is necessary to know in advance the distribution of droplets in the spray<br />

and the characteristics of the surrounding atmosphere, in a system in which the fuel<br />

is partially in the liquid phase and partially in the gaseous phase as it enters the combustion<br />

zone. It is also necessary to take into account the interaction of droplets and<br />

the influence of the highly turbulent motion of the gas. All this makes the theoretical<br />

study of the problem extremely difficult. For this reason the studies ma<strong>de</strong> up to date<br />

are of a highly empirical character. These studies limit themselves either to an analysis<br />

of the influence of some fundamental parameters on the combustion efficiency of a<br />

burner [30], or to obtain general conclusions as for the way in which several variables<br />

can influence the process [24].<br />

13.15 Droplet evaporation<br />

When combustion is absent the evaporation process of a fuel spray is also a very<br />

complicated phenomenon. This problem has only been studied empirically for some<br />

typical cases [16]. Bahr [30], for example, has given an empirical formula to express<br />

un<strong>de</strong>r given conditions the evaporated fraction as a function of the distance to the<br />

atomizer. The Bahr formula represents a good approximation to the results obtained<br />

from his own experiments and those performed by Ingebo [16].<br />

As a first step towards the solution of the problem, the evaporation of isolated<br />

droplets can be studied by the same procedure applied in the preceding paragraphs


326 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

to the study of combustion. The extrapolation of the obtained values to the case of a<br />

spray is difficult. In fact, it is necessary to know the size distribution of the droplets<br />

and the composition of the surrounding atmosphere. Furthermore, in or<strong>de</strong>r to account<br />

for the convection effects, which are very important, the motion of the droplets relative<br />

to the atmosphere must also be known. The available information is not sufficient to<br />

show the way in which this extrapolation can be performed [34].<br />

The evaporation velocity of a droplet when convection is absent can be easily<br />

obtained from the formulas <strong>de</strong>duced in the preceding paragraphs. In fact, assuming<br />

that the droplet is isothermal, it is enough to make use of the equations corresponding<br />

to the interior region, enlarged to infinity. Therein the following boundary conditions<br />

must be satisfied which <strong>de</strong>termine the integration constants<br />

r → ∞ : T → T ∞ , Y 1 → Y 1∞ , (13.87)<br />

where Y 1∞ is the mass fraction of the fuel vapours at great distance from the droplet.<br />

Moreover, if one keeps the assumptions previously established with respect to the law<br />

of variation of the values of the transport coefficients as functions of temperature, the<br />

two equations for T and Y 1 are Eqs. (13.53) and (13.54), that is<br />

The last equation when applied to the droplet surface r = r s gives the following<br />

relation<br />

4πr 2 λ ∞<br />

T<br />

T ∞<br />

dT<br />

dr − mc p1T = m(q l − c p1 T s ), (13.88)<br />

4πr 2 (ρD 12 ) ∞<br />

T<br />

T ∞<br />

dY 1<br />

dr = −m(1 − Y 1). (13.89)<br />

The solution of this system that satisfies conditions (13.87) is<br />

1<br />

r = 4πλ (<br />

∞<br />

1 − T + q l − c p1 T s<br />

ln c )<br />

p1(T − T s ) + q l<br />

, (13.90)<br />

mc p1 T ∞ c p1 T ∞ c p1 (T ∞ − T s ) + q l<br />

( ) δ<br />

1 − Y 1 cp1 (T − T s ) + q l<br />

=<br />

. (13.91)<br />

1 − Y 1∞ c p1 (T ∞ − T s ) + q l<br />

(<br />

) δ<br />

1 − Y 1s<br />

q l<br />

=<br />

. (13.92)<br />

1 − Y 1∞ c p1 (T ∞ − T s ) + q l<br />

Let p 1s be the partial pressure of the fuel vapour on the droplet surface. Thermodynamics<br />

shows 13 that p 1s is <strong>de</strong>termined by the temperature T s on the droplet surface.<br />

13 See Prigogine, I. and Defay, R.: Chemical Thermodynamics. Longmans Green & Co., 1954, pp. 332<br />

and f.


13.15. DROPLET EVAPORATION 327<br />

4<br />

0.25<br />

P 1<br />

(kg/cm 2 )<br />

3<br />

2<br />

P 1<br />

(kg/cm 2 )<br />

0.20<br />

n−heptane<br />

0.15<br />

0.10<br />

n−exane<br />

0.05<br />

n−octane<br />

0.00<br />

250 275 300 325 350<br />

T (K)<br />

n−heptane<br />

n−octane<br />

n−exane<br />

1<br />

n−<strong>de</strong>cane<br />

n−do<strong>de</strong>cane<br />

0<br />

250 300 350 400 450 500<br />

T (K)<br />

Figure 13.6: Partial pressure of fuel vapour as a function of temperature.<br />

Therefore a relation exists between p 1s and T s of the form<br />

Figure 13.6 gives Eq. (13.93) for some typical fuels.<br />

f(p 1s , T s ) = 0. (13.93)<br />

Furthermore 14 p 1<br />

p = M a Y<br />

( 1<br />

) , (13.94)<br />

M c Ma<br />

1 + − 1 Y 1<br />

M c<br />

where M c and M a are, respectively, the molar masses of the fuel vapour and of the<br />

gas through which it diffuses.<br />

When (13.94) is particularized on the droplet surface, taking the result into<br />

(13.93), the following relation between Y 1s and T s is obtained<br />

This relation and (13.92) <strong>de</strong>termine Y 1s and T s .<br />

F (Y 1s , T s ) = 0. (13.95)<br />

Once T s is known the evaporation velocity m of the droplet is <strong>de</strong>duced from<br />

(13.90) by making r = r s and T = T s . Thus obtaining<br />

m = 4πλ (<br />

∞r s<br />

1 − T s<br />

+ q )<br />

l − c p1 T s q l<br />

ln<br />

. (13.96)<br />

c p1 T ∞ c p1 T ∞ c p1 (T ∞ − T s ) + q l<br />

14 See chapter 1.


328 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

The parenthesis on the right-hand si<strong>de</strong> of this equation is in<strong>de</strong>pen<strong>de</strong>nt from<br />

r s . Therefore, evaporation velocity, like combustion velocity, is proportional to the<br />

droplet radius.<br />

Making use of Eq. (13.71) a relation similar to (13.74) is obtained for the variation<br />

of r s<br />

r 2 s = r 2 si − k v t, (13.97)<br />

where k v is the evaporation constant, which from (13.96) is given by the expression<br />

k v = 2λ (<br />

∞<br />

1 − T s<br />

+ q )<br />

l − c p1 T s q l<br />

ln<br />

. (13.98)<br />

ρ c c p1 T ∞ c p1 T ∞ c p1 (T ∞ − T s ) + q l<br />

If the following condition is satisfied<br />

a linearization of (13.96) gives for m<br />

A similar linearization of (13.92) gives<br />

c p1 (T s − T ∞ )<br />

q l<br />

≪ 1, (13.99)<br />

m = 4πλ ∞r s<br />

q l<br />

(T ∞ − T s ). (13.100)<br />

c p1 (T ∞ − T s )<br />

q l<br />

= 1 δ<br />

By taking this expression into (13.100), it results for m<br />

or else, as a function of p 1s and p 1∞ by virtue of (13.94),<br />

Y 1s Y 1∞<br />

1 − Y 1∞<br />

. (13.101)<br />

m = 4πr s ρD 12<br />

Y 1s Y 1∞<br />

1 − Y 1∞<br />

, (13.102)<br />

m = 4πr sD 12<br />

R c T ∞<br />

p 1s − p 1∞<br />

p − p 1∞<br />

p, (13.103)<br />

where R c = R/M c is the gas constant for the fuel vapour. Eq. (13.103) is Langmuir’s<br />

evaporation formula valid for small evaporation velocities.<br />

Let<br />

x = c p1(T ∞ − T s )<br />

q l<br />

. (13.104)<br />

If this value is taken into (13.96) and the result divi<strong>de</strong>d by (13.100), the following<br />

relation between m and m 0 is obtained<br />

m<br />

=<br />

q [ (<br />

l<br />

1 − 1 − c ) ]<br />

p1T ∞ ln(1 + x)<br />

+ x<br />

m 0 c p1 T ∞ q l<br />

x<br />

(13.105)<br />

where m is the actual evaporation velocity and m 0 is the velocity that would be obtained<br />

if Langmuir’s formula would apply to all cases.


13.15. DROPLET EVAPORATION 329<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

m/m 0<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

c T /q =2<br />

p1 ∞ l 4 6 8 10<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

x<br />

Figure 13.7: The ratio m/m 0 as a function of x = c p1(T ∞ − T s)/q l for some values of<br />

c p1T ∞/q l .<br />

Figure 13.7 gives m/m 0 as a function of x, for some values of c p1 T ∞ /q 1 .<br />

When the evaporation velocity is high the fast <strong>de</strong>crease of the evaporation constant<br />

can be seen due to the transport of enthalpy done by the vapour motion.<br />

The previous formulae do not take into account the influence of convection.<br />

Such effect has been studied by Frössling [32], who obtained the following relation<br />

between the evaporation velocity m c with convection and the evaporation velocity at<br />

rest<br />

m c<br />

m = (<br />

1 + 0.276 Sc 1/3 Re 1/2) . (13.106)<br />

Here Sc = µ/ρD 12 and Re = ρvd s /µ are, respectively, the Schmidt number<br />

and the Reynolds number of the motion, µ is the gas viscosity coefficient, d s the<br />

droplet diameter and v the motion velocity. Ranz and Marshall [33] have experimentally<br />

verified this formula. Fig. 13.8 gives m c /m as a function of Sc 2/3 Re.<br />

Formula (13.105) was obtained from experiments at ambient temperature. Ingebo<br />

[34] gives the following empirical formula which represents with good approximation<br />

the experimental results obtained by him from 30 ◦ C to 500 ◦ C<br />

m c<br />

m = ( 1 + 0.151(Sc Re) 0.6) . (13.107)


330 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

24<br />

20<br />

Ingebo (S c<br />

=0.65)<br />

m c<br />

/ m<br />

16<br />

12<br />

Frössling<br />

8<br />

4<br />

0<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0<br />

2/3<br />

S Re × 10<br />

−3<br />

c<br />

Figure 13.8: Experimental relations of Frössling and Ingebo, giving m c/m as a function of<br />

Sc 2/3 Re.<br />

This formula is represented by the dash line in Fig. 13.8. It is seen that the Frössling<br />

formula un<strong>de</strong>restimates the evaporation velocity for the large Reynolds numbers.<br />

Miesse [35] has lately calculated the motion of an evaporating droplet by assuming:<br />

1) The coefficient of the aerodynamic drag of the droplet is inversely proportional<br />

to the Reynolds number (laminar flow).<br />

2) The evaporation velocity of the droplet follows the Frössling law.<br />

Miesse has obtained solutions for the case where the air velocity changes linearly with<br />

distance. Such solutions allow the calculation of the distance covered by the droplet<br />

as a function of the reduction of its diameter and the distance nee<strong>de</strong>d for complete<br />

evaporation, etc. His work contains interesting practical applications. However, when<br />

trying to extend his conclusions to the case of sprays the remarks ma<strong>de</strong> at the beginning<br />

of this paragraph should be taken into account. On the other hand the possible<br />

interaction of evaporation and aerodynamic drag of the droplet is neglected.<br />

Penner [36] has calculated the evaporation time of a propellant droplet within<br />

the combustion chamber of a rocket by assuming that the droplet is isothermal but its<br />

temperature changes with time. He has estimated the possible influence of the radiation<br />

energy on the evaporation process reaching the conclusion that such influence is<br />

of no significance.


13.16. APPENDIX: APPLICATION OF PROBERT’S METHOD 331<br />

13.16 Appendix: Application of Probert’s method for<br />

the combustion of fuel sprays<br />

Probert 15 has <strong>de</strong>veloped a theoretical method for the study of evaporation or combustion<br />

of fuel sprays, un<strong>de</strong>r the assumption that evaporation constant is the same for all<br />

droplets. The justification of this assumption and the value that should assigned to the<br />

constant if it is valid, <strong>de</strong>pend on the results the experimental measurements.<br />

After chapter 13 was written, some theoretical works on the application of<br />

Probert’s method have been performed at the I.N.T.A. 16 corresponding to steady burning<br />

as well as to transition from ignition to steady burning and periodic combustion.<br />

In these computations the size distribution function of Mugele-Evans was used with<br />

preference to those of Rosin-Rammler and Nukiyama-Tanasawa, since it allows the<br />

prediction of the sizes with a very good approximation taking into account the maximum<br />

size of the droplets. Let F be the mass fraction of fuel corresponding to droplets<br />

with a diameter smaller than d. Mugele-Evans’ formula gives for F the following<br />

expression<br />

F = 1 2<br />

( (<br />

))<br />

θd<br />

1 + erf ε ln<br />

. (13.108)<br />

d max − d<br />

Here erf is the error integral, d max is the droplet’s maximum diameter and ε and θ<br />

are two parameters characteristic of the distribution. Fig. 13.9 shows some of the<br />

distributions corresponding to typical values for these parameters. It is seen that the<br />

increment of ε increases the uniformity of the spray, whilst when θ increases the mean<br />

diameter of the droplets reduces.<br />

Let G be the volume of fuel, injected to the burner per unit time, and g the<br />

volume of the droplets existing in the burner. It can be verified that g is expressed as a<br />

function of G through formula<br />

where I is given by expression<br />

I =<br />

∫ 1<br />

0<br />

x 4 dx<br />

∫ 1 − x<br />

0<br />

g = εI √ π<br />

G t v , (13.109)<br />

−<br />

e<br />

(<br />

p ) 2<br />

x2 + y<br />

ε ln<br />

1 − p x 2 + y<br />

and t v is the life time of the largest droplets of the spray.<br />

(x 2 + y)<br />

(1 5/2 − √ ) dy, (13.110)<br />

x 2 + y<br />

15 Probert. R.P.: The Influence of Spray Particle Size and Distribution in the Combustion of Oil Droplets.<br />

Philosophical Magazine, February 1946.<br />

16 Millán, G., and Sanz, S.: Analysis of the Combustion Processes in Gas Turbines. Fourth International<br />

Congress of Combustion Engines, Zurich, 1957.


332 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

1.0<br />

F<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

(0.5,1.5)<br />

(1.5,1.0)<br />

(1.0,1.0)<br />

(1.5,1.5)<br />

(1.0,1.5)<br />

(1.0,0.5)<br />

(0.5,0.5)<br />

(0.5,1.0)<br />

(1.5,0.5)<br />

0.0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

d / d max<br />

Figure 13.9: Droplet size distribution without combustion according to Mugele and Evans<br />

(fraction F of droplets with a diameter smaller than d/d max), for several values<br />

of the parameters (ε , θ).<br />

Magnitu<strong>de</strong> (13.109) is interesting since the combustion intensity of the burner<br />

should be consi<strong>de</strong>red inversely proportional to it.<br />

Table 13.2 gives the values for g/G t v for three typical cases corresponding to<br />

θ = 1.<br />

ε 0.5 1 1.5<br />

g<br />

0.129 0.110 0.105<br />

G t v<br />

( )<br />

ds<br />

0.269 0.438 0.895<br />

d max<br />

spray<br />

(<br />

ds<br />

d max<br />

)flame<br />

0.517 0.454 0.401<br />

Table 13.2: Values of g/Gt v for θ = 1 and ε = 0.5, 1, 1.5.<br />

It is seen that when ε increases, that is when the uniformity of the spray increases,<br />

the mass fraction of fuel in the burner <strong>de</strong>creases. Consequently, it is advantageous<br />

to work with spray as uniform as possible in or<strong>de</strong>r to <strong>de</strong>crease the volume of<br />

the primary zone of the burner.


13.16. APPENDIX: APPLICATION OF PROBERT’S METHOD 333<br />

Combustion changes the droplet size distribution, preserving, obviously the<br />

maximum diameter. Figure 13.10 shows the distribution functions at the flame corresponding<br />

to the three cases studied. For comparison this figure inclu<strong>de</strong>s the distributions<br />

corresponding to the spray. Table 13.2 also gives Sauter’s diameters for both the<br />

spray and the flame.<br />

1.0<br />

0.9<br />

0.8<br />

F<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

ε = 1.5<br />

ε = 0.5<br />

ε = 1.0<br />

Spray<br />

Combustion<br />

0.0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

d / d max<br />

Figure 13.10: Effect of combustion on the droplet size distribution for some values of ε.<br />

References<br />

[1] Khudyakov, G. N.: Combustion of Liquids from Free Surfaces. Isv. Acd. Sci.<br />

USER, Div. Chem. Sci., Nos. 10-11, 1945.<br />

[2] Khudyakov, G. N.: Distribution of Temperature within a Liquid Burning from<br />

a Free Surface, and Description of the Flame Formed. R.A.E. Translation no.<br />

422, 1953.<br />

[3] Spalding, B. D.: The Combustion of Liquid Fuels. Fourth Symposium (International)<br />

on Combustion, Williams and Wilkins Co., Baltimore, 1953, pp. 847-<br />

864.<br />

[4] Spalding B. D.: The Calculation of Mass Transfer Rates in Absorption, Vaporization,<br />

Con<strong>de</strong>nsation and Combustion Processes. Proceedings of the Institution<br />

of Mechanical Engineers, Vol. 168, No. 19, 1954.


334 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

[5] Haber, F. and Wolff, H.: Uber neboloxplosionen. Zeitschrift für Angewante<br />

Chemik, July 1923, pp. 373-377.<br />

[6] Jones, G. W. et al.: Research on the Flammability Characteristics of Aircraft<br />

Fuels. WADC Tech Rept., June 1952, pp. 52-35.<br />

[7] Browning, J. A. and Krall, W. G.: Effect of Fuel Droplets on Flame Stability,<br />

Flame Velocity and Inflammability Limits. Fifth Symposium (International) on<br />

Combustion, Reinhold Publishing Corp., New York, 1955, pp.159-163.<br />

[8] The Penn State Bibliography on Sprays. The Texas Company, 1953.<br />

[9] Griffen, E. and Muraszew, A.: The Atomization of Liquid Fuels. Chapman and<br />

Hall Ltd., 1953.<br />

[10] Hinze, J. C.: On the Mechanism of Desintegration of High Speed Liquid Jets.<br />

Proceedings of the Sixth International Congress of Applied Mechanics, 1946.<br />

[11] Sauter, J.: Die Grössenbestimmung <strong>de</strong>r in Gemischnebel von Verbrenungskraitmaschinen<br />

verhan<strong>de</strong>nen Brennstoffteilchon. Forschung Gebiet <strong>de</strong>s Ingenieurswesens,<br />

No. 279, 1926.<br />

[12] Mugele, R. H. and Evans, H. D.: Droplet Size Distribution in Sprays. Industrial<br />

and Engineering Chemistry, June 1951, pp. 1317-1324.<br />

[13] Rosin, P. and Rammler, E.: The Laws Governing the Fineness of Pow<strong>de</strong>red<br />

Coal. Journal of the Institute of Fuel, 1933, pp. 29-36.<br />

[14] Nukiyama, S. and Tanasawa, Y.: An Experiment on the Atomization of Liquid<br />

by Moans of Air Stream. Transactions of the Society of Mechanical Engineers<br />

of Japan, Feb. 1938, p. 86; May 1938, p. 1389; Feb. 1939, pp. 63 and 68.<br />

[15] Bevans, R. S.: Mathematical Expressions for Drop Size Distribution in Sprays.<br />

Conference on Fuel Sprays, Univ. of Michigan, March 1949.<br />

[16] Graves, Ch. C. and Gerstein, M.: Some Aspects of Combustion of Liquid Fuel.<br />

AGARD Memorandum AG 16/M, May 1954.<br />

[17] Longwell, J. P. and Weiss, M. A.: Mixing and Distribution of Liquids in High-<br />

Velocity Air Streams. Industrial and Engineering Chemistry, March 1953, pp. 667-<br />

677.<br />

[18] Taylor, G. I.: Diffusion by Continuous Movements. Proceedings of the London<br />

Mathematical Society, 1921, p. 196.<br />

[19] Godsave, G. L.: Studies of the Combustion of Drops in a Fuel Spray. The Burning<br />

of Single Drops of Fuel. Fourth Symposium (International) on Combustion,<br />

Williams and Wilkins Co., Baltimore, 1953, pp. 818-830.<br />

[20] Goldsmith, M. and Penner, S. S.: On the Burning of Single Drops of fuel in on<br />

Oxidizing Atmosphere. Jet Propulsion, July-August 1954, pp. 245-251.


13.16. APPENDIX: APPLICATION OF PROBERT’S METHOD 335<br />

[21] Godsave, G. A.: The Burning of Single Drops of Fuel. Part. I, Temperature<br />

Distribution and Heat Transfer in the Preflame Region. NGTE Report no. R.66,<br />

1950. Part II, Experimental Results. NGTE Report no. 87, 1951. Part III,<br />

Comparison of Theoretical and Experimental Burning Rates and Discussion of<br />

the Mechanism of the Combustion Process. NGTE Report no. R.38, 1952.<br />

[22] Topps, J. E. C.: An Experimental Study of the Evaporation and Combustion of<br />

Falling Droplets. Journal of the Institute of Petroleum, 1951, pp. 535-537.<br />

[23] Hall, L. R. and Die<strong>de</strong>richsen, J.: An Experimental Study of the Burning of Single<br />

Drops of Fuel in Air at Pressures up to Twenty Atmospheres. Fourth Symposium<br />

(International) on Combustion, Williams and Wilkins Co., Baltimore,<br />

1953, pp. 837-846.<br />

[24] Spalding, D. B. Combustion of Single Droplet and of a Fuel Spray. Selected<br />

Combustion Problems, Vol. I, AGARD, 1954, pp. 340-351.<br />

[25] Hottel, H. C., Williams, G. C. and Simpson, H. C.: The Combustion of Droplets<br />

of Heavy Liquid Fuels. Fifth Symposium (International) on Combustion, Reinhold<br />

Publishing Corp., New York, 1955, pp. 101-129.<br />

[26] Wise, H., Lowell, J, and Wood, B. J.: The Effects of Chemical and Physical<br />

Parameters on the Burning Rate of a Liquid Droplet. Fifth Symposium (International)<br />

on Combustion, Reinhold Publishing Corp., New York, 1955, pp. 132-<br />

141.<br />

[27] Kobayasi, K.: An Experimental Study on the Combustion of a Fuel Droplet.<br />

Fifth Symposium (International) on Combustion, Reinhold Publishing Corp.,<br />

New York, 1955, pp. 141-148.<br />

[28] Nischiwali, N.: Kinetics of Liquid Combustion Processes: I. Evaporation and<br />

Ignition Lag of Fuel Droplets. Fifth Symposium (International) on Combustion,<br />

Reinhold Publishing Corp., New York, 1955, pp. 148-158.<br />

[29] Surugue, J.: Combustion Problems in Turbojets. AGARD AG 5/P2, December<br />

1952, pp. 54-61.<br />

[30] Bahr, D.: Evaporation and Spreading of Isooctane Sprays in High-Velocity Air<br />

Streams. NACA RM E511 E53114, 1953.<br />

[31] Saks, W.: The Rate of Evaporation of a Kerosene Spray. National Aeronautical<br />

Institute of Canada, Note No. 7, 1951.<br />

[32] Frössling, N.: On the Evaporation of Falling Drops. Gerlands Beiträge zur<br />

Geophysik, 1938, pp. 170-216.<br />

[33] Ranz, W. E. and Marshall, W. R.: Evaporation from Drops. Chemical Engineering<br />

Progress, 1952, pp. 141-46 and 173-180.


336 CHAPTER 13. COMBUSTION OF LIQUID FUELS<br />

[34] Ingebo, R. D.: Vaporization Rates and Heat-Transfer Coefficients for Pure Liquid<br />

Drops. NACA Tech. Note No. 2368, 1951.<br />

[35] Miesse C. C.: Ballistics of an Evaporating Droplet. Jet Propulsion, July-August<br />

1954, pp. 237-244.<br />

[36] Penner, S. S.: On Maximum Evaporation Rates of Liquid Droplet in Rocket<br />

Motors. Journal of the American Rocket Society, March-April 1953, pp. 85-88.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!