The oxidation of hydrocarbons: studies of spontaneous ignition I. Ignition limits in small vessels

Studies have been made of the spontaneous ignition of n -heptane+oxygen+inert gas mixtures at temperatures from 440 to 650°C, where ignition takes place by a one-stage process and no cool flames are observed. Detailed measurements have been made of the variation of minimum ignition pressure with such factors as the temperature, the composition of the mixture undergoing ignition and the nature, shape and extent of the surface of the reaction vessel. In particular, experiments in a wide variety of vessels show that the surface parameter which primarily determines the ignition tendency is the average distance of the molecules from the walls, rather than the surface:volume ratio. The quantitative relations observed experimentally are compared with the predictions of two isothermal chain-branching mechanisms involving distinct chemical paths and with the consequences of the theory of thermal ignition. It is shown that the results in small vessels (volume < 500 cm 3 ) are best explained in terms of an isothermal chain mechanism involving hydrogen peroxide as degenerate-branching agent, although in larger vessels thermal factors probably become increasingly important.

Studies have been made of the spontaneous ignition of n -heptane + oxygen + nitrogen mixtures at temperatures in the region of 500 °C in vitreous vessels up to 10 1. in volume. Measurements have been made of the influence of temperature, reactant mixture composition and the size and shape of the reaction vessel on the minimum pressure required for ignition. Experiments have also been carried out to determine how the induction period preceding ignition, the maximum rate of non-isothermal pressure rise and the temperature rise accompanying the non-isothermal pressure increase vary with different parameters. The results show that, even in the large vessels (volume 2 to 101.) used in this work, thermal factors are not of great importance. An isothermal olefin mechanism can still account for the quantitative relations observed experimentally if it is assumed that there is a change in the mode of control of the reaction H 2 O 2 walls → inactive products from efficiency control in small vessels to diffusion control in large vessels.


1932 ◽  
Vol 7 (2) ◽  
pp. 149-161 ◽  
Author(s):  
W. H. Hatcher ◽  
E. W. R. Steacie ◽  
Frances Howland

The kinetics of the oxidation of gaseous acetaldehyde have been investigated from 60° to 120 °C. by observing the rate of pressure decrease in a system at constant volume. A considerable induction period exists, during which the main products of the reaction are carbon dioxide, water, and formic acid. The main reaction in the subsequent stages involves the formation of peroxides and their oxidation products. The heat of activation of the reaction is 8700 calories per gram molecule. The indications are that the reactions occurring during the induction period are heterogeneous. The subsequent reaction occurs by a chain mechanism. The chains are initiated at the walls of the reaction vessel, and are also largely broken at the walls.


Kinetic and analytical studies of the gaseous oxidation of aluminium trimethyl at ambient temperatures and at pressures well below those required for spontaneous ignition have shown that, in contrast to the oxidations of less electron-deficient metal alkyls, no peroxides can be detected and no volatile oxygenated organic compounds are formed. Methane, traces of hydrogen and a solid methoxymethyl compound of aluminium are the only products. The initial rate of reaction is approximately proportional to the first power of the alkyl pressure and to the square of the oxygen pressure; it is little influenced by temperature or by inert gases but is lowered by an increase in surface. The observed kinetic and analytical results can be accounted for in terms of a free radical chain mechanism in which termination takes place predominantly at the walls.


1969 ◽  
Vol 73 (10) ◽  
pp. 3395-3406 ◽  
Author(s):  
Ching-Huan Yang ◽  
Brian F. Gray

Investigation of the kinetics of the oxidation of ethylene and of benzene showed that these reactions are peculiar in the following respects. First, the relation between the rate of reaction and concentration is such that the reactions possess no simple “order,” though the nearest integral value for the order is about the third of fourth. The rate increases very rapidly with increasing hydrocarbon concentration, but is relatively little influenced by oxygen; under some conditions oxygen may have a retarding influence. Secondly, the reactions can be slowed down by increasing the surface exposed to the gases. This indicates that the oxidation occurs by a chain mechanism. Thirdly, the rate of change of pressure accompanying the oxidation only attains its full value after an induction period, during which evidently intermediate products are accumulating. Accepting the fact that the oxidations are probably chain reactions, the relation between rate and concentration shown that the chains are much more easily propagated when the intermediate active molecules encounter more hydrocarbon than when they encounter oxygen. Following the view of Egerton, and consistently with previous work on the combination of hydrogen and oxygen, the working hypothesis adopted is that some intermediate peroxidised substance is responsible for the propagation of the chains. This being so, the question arises whether the peculiarities found in the oxidation of hydrocarbons will also be found with substances already containing oxygen. To investigate, therefore, the influence of chemical configuration on the mechanism of oxidation reactions the following series of compounds has been studied CH 4 CH 3 OH HCHO which represent the stages through which Bone and others have shown the oxidation of methane to occur.


The progressive formation of products in the combustion of benzene and its monoalkyl derivatives has been studied by analytical methods, and the characteristic features of the isothermal reactions at various temperatures have been established. A cool-flame reaction of n -propylbenzene has also been investigated, and by comparison with corresponding isothermal combustions, it is concluded that the propagation of cool-flames is conditioned by the accumulation of a phenylalkyl hydroperoxide. The results are interpreted in the light of the theory of the two-stage process, and a schematic mechanism for the main combustion reaction is outlined. This comprises degradation of the side-chain (if present) and rupture of the benzene nucleus, followed by rapid degradation of the higher aliphatic aldehyde thus formed, yielding finally formaldehyde and the ultimate combustion products CO 2 , CO and H 2 O.


The oxidation of toluene and ethylbenzene has been studied in a static system using a spherical reaction vessel (700 ml.) over the temperature range 300 to 500°C, and at total pressures up to 600 mm. Cool flames were observed in the oxidation of both hydrocarbons, but only the reaction of ethylbenzene gave rise to a ‘blue’ flame at higher temperatures. With neither hydrocarbon did periodicity in light intensity, or pressure pulses, occur. The ignition diagrams for 4 to 1 fuel + oxygen mixtures have been mapped out. With ethyl­benzene, the cool flame was maintained in a flow system, its spectrum was photographed and shown to be similar to that of fluorescent formaldehyde. The products of the reaction con­tained acetophenone, benzaldehyde and benzoic acid, phenol, quinol, hydrogen peroxide and methoxyhydroperoxide. The results have been compared with corresponding data for the oxidation of paraffin hydrocarbons, and it is concluded that, with both aromatic compounds, the processes allowing the possibility of cool-flame formation are themselves secondary in nature.


The investigations described in previous papers on this subject have related mainly to the paraffin hydrocarbons (Townend and Mandlekar 1933 a,b ; Townend, Cohen and Mandlekar 1934; Townend and Chamberlain 1936, 1937). It has been found that mixtures with air of the members containing three or more carbon atoms, while not spontaneously ignitible at low pressures below about 500° C., give rise abruptly to ignition at higher pressures in a temperature range between about 310 and 370° C., where normally only cool flames are initiated; and although neither methane- nor ethane-air mixtures appear to develop cool flames, the latter are ultimately ignitible in a lower temperature system which is less complex than that characteristic of the higher paraffins. Moreover, it is now recognized that “knock” in internal combustion engines arises in circumstances responsible for pronounced chemical reactivity in the unburnt explosive medium characteristic of that occurring in the lower temperature range (cf. Egerton and Ubbelohde 1935; Ubbelohde 1935), and the investigations referred to have indicated that the “knock-ratings” of the paraffins when used as fuels in such engines are related to the pressures requisite for the occurrence of spontaneous ignition in this range within an appropriate short time lag (Townend and Chamberlain 1936, p. 104, cf. Prettre 1936 a and b )


As part of a wider investigation into the influence of pressure on the ontaneous ignition of inflammable gas-air media generally, some time go we studied the behaviour of diethyl ether and found it to simulate at of the higher paraffins, in that at low pressures ignition occurs in a high temperature system and at higher pressures in a low temperature system, which develops in the range where normally only cool flames are propagated. We were also impressed with the analogy between our own observations and those on the limits of inflammability of ether-air mixtures made in 1927 by A. G. White who discovered that at low pressures there are two ranges of explosive mixtures which can propagate ame, one for normal and another for cool flames, separated by a range of mixtures through which no flame can be propagated; with increase of pressure these explosive ranges become superposed. We decided to examine the matter more closely because it seemed likely that an explanation of the analogy referred to would throw light on the whole problem. Moreover, the subject is of practical importance, having in mind the risks inherent in certain circumstances in the use of ether as an anaesthetic, and was hoped that our results might also be of some service in this connexion. A likely interpretation could be based on the thermal theory of flame propagation. This applies to the slow initial stages of gaseous explosions and was developed by Mallard and Le Chatelier who proposed the following well-known equation for the velocity, V, of the “uniform movement”:— V = L/C (T - t )/( t - θ) f (T t ), Where T = the temperature attained in the combustion, t = the ignition temperature of the mixture, θ = the initial temperature, L = the thermal conductivity of the unburnt gas, C = its mean heat capacity between θ and t , and f (T t ) = a function taking into account the change of L and C with temperature. Other modifications have been proposed, but discussion has usually centred round the above equation; and although it is not possible to apply it quantitatively, owing to the insuperable difficulties involved in determining the precise values of the various terms concerned and lack of knowledge concerning the amount of combustion occurring in the flame front and energy losses, etc., it lends itself well to the qualitative interpretation of the effect of the various factors controlling initial slow flame speeds. For example, Mason and Wheeler showed that with mixtures of like thermal conductivity the speed is proportional to (T — t ), and inversely proportional to ( t — θ). Moreover, with combustible-air/oxygen media the mixture giving rise to the maximum flame speed contains an excess of combustible corresponding, owing to the suppression of CO 2 —and H 2 O—dissociation, with that developing the highest temperature; indeed, Bone and Bell have recently shown that with CO-O 2 media the flame speed-composition curve exhibits two maxima corresponding with the suppression of CO 2 dissociation by excess of either CO or O 2 .


Sign in / Sign up

Export Citation Format

Share Document