Sie sind auf Seite 1von 34

Prog.EnergyCombust.ScL 1993,Vol. 19,pp. 71 104 PrintedinGreatBritain.Allrightsreserved.

0360-1285/93$24.00 ~)1993PergamonPressLtd

MODELING A N D SIMULATION OF COMBUSTION PROCESSES OF CHARRING A N D N O N - C H A R R I N G SOLID FUELS


COLOMBA D I BLASI Dipartimento di Ingegneria Chimica, Unioersitd di Napoli, Piazzale V. Tecchio, 80125 Napoli, Italy Received 11 February 1993

Abstract--Some of the progress that, owing to modeling and numerical simulation, has been made to the understanding of chemical and physical processes, which occur during combustion of solid fuels, is presented. The first part of the review deals with thermal degradation processes of charring (wood and, in general, cellulosic materials) and non-charring (poly-methyl-methacrylate) materials. Gas-phase combustion processes (ignition, flame spread and extinction) are discussed in the second part of the review. Solid fuel degradation has been described by kinetic models of different complexity, varying from a simple onestep global reaction, to multi-step reaction mechanisms, accounting only for primary solid fuel degradation, and to semi-global reaction mechanisms, accounting for both primary solid degradation and secondary degradation of evolved primary pyrolysis products. Semi-global kinetic models have been coupled to models of transport phenomena to simulate thermal degradation of charring fuels under ablation regime conditions. The effects of bubble formation on the transport of volatiles during thermal degradation of non-charring fuels, described through a one-step global reaction, have also been modeled. On the contrary, very simplified treatments of solid phase processes have been used when gas phase combustion processes are also simulated. On the other hand, the latter have also always been described through one-step global reactions. Numerical modeling has allowed controlling mechanisms of ignition and flame spread to be determined and the understanding of the interaction between chemistry and physics during thermal degradation of solid fuels to be improved. However, the chemical processes are not well understood, the few kinetic data are in most cases empirical and variations of solid properties during degradation are very poorly known, so that even the most advanced models do not in general give quantitative predictions. CONTENTS 1. Introduction 2. Thermal Degradation Processes of Solid Fuels 2.1. Charring materials 2.2. Kinetic modeling 2.2.1. One-step global models 2.2.2. One-stage, multi-reaction models 2.2.3. Two-stage, semi-global models 2.3. Energetics of the pyrolysis reactions 2.4. Modeling of chemical and physical processes 2.4.1. The equations of wood pyrolysis 2.4.2. Numerical simulation of wood pyrolysis 2.5. Non-charring materials 2.6. Kinetic modeling 2.7. Modeling of chemical and physical processes 3. Gas Phase Combustion Processes 3.1 Modeling and simulation of ignition processes 3.2. Flame spread modeling 3.2.1. Gas phase models 3.2.2. Solid phase models 3.2.3. A mathematical model of flame spread 3.3. Opposed flow flame spread 3.4. Flow assisted flame spread 3.5. Microgravity flame spread 4. Conclusions and Further Developments Acknowledgements References 71 73 73 74 75 75 76 79 80 82 84 85 86 87 87 89 91 91 93 93 94 97 99 100 102 102

l. INTRODUCTION T h e r m a l d e g r a d a t i o n o f solid fuels is a process pertinent to energy recovery t h r o u g h biomass conver71

sion a n d to solid municipal waste disposal. It is also o f interest to safety problems since solid thermal d e g r a d a t i o n is the first step o f u n w a n t e d fire events,

72
external flaming combustion ~

C. D! BLASl
1

heating

air +l

SOLID ~

thermal degradation - - ~

volatiles

|
l

l 1~
. . . . .

ugh ~,~,~ + chars / to~he,t~ ~ + air smolderingcombustion / ~


energy feed back

~
o" ~

~-

l
.,,._.

Fro. 1. Schematic diagram of combustion processes of charring materials. such as smoldering combustion, ignition, flame spread and burning. Thus it is very useful to understand the mechanisms controlling the interaction betwen chemical and physical processes and also the composition of evolved gases. This latter point is important both in the context of maximizing the yields of chemicals from material conversion and minimizing air pollution from material burning (incineration and stoves), When subjected to external heating, solid fuels start to decompose, giving a mixture of volatile species and, sometimes, a solid carbonaceous residual (char) as products. Combustion processes of charring materials, exposed in an oxidizing flow environment, could proceed by two alternative pathways involving flaming combustion and smoldering (or glowing) combustion, as shown, from a qualitative point of view, by the schematic diagram of Fig. I. Conditions of flaming combustion are achieved when the heat released by gas phase combustion of volatile products provides the heat flux needed for solid fuel degradation and flame spread. When the temperature or the intensity of the heat flux are below certain levels, oxidation of the char could result in smoldering combustion. Indeed, for porous or fibrous materials, such as wood, air may diffuse inside the solid matrix and cause slow oxidation: the even, low heat release rate, in the absence of significant radiation heat losses, convection and conduction, provides the heat flux needed for further charring and propagation of the smoldering combustion. For charcoal production or biomass gasification, oxygen flow and other heat and mass transfer conditions are often adjusted in order to obtain partial combustion and reaction heat, together with substantial pyrolysis. 1'2 Most of the biomass conversion processes are aimed at obtaining solid fuel pyrolysis to volatile species and char in a non-oxidizing environment, Analyses of solid pyrolysis in a non-oxidizing environment are also of interest to combustion processes related to fire safety issues. At intermediate temperatures, gas evolution from cellulosic fuel pyrolysis is sufficiently rapid to prevent significant air diffusion within the solid matrix, yet not rapid enough to

produce ignition. 3 Under such conditions, degradation processes in an inert atmosphere and in air are not significantly different. A rapid heating of the solid fuel, to high temperatures, leads, in air, to flaming combustion. 4 Even in this case, degradation of cellulosic materials in air would not be very different from that in an inert atmosphere since, if the flame envelops the degrading solid, the diffusion of oxygen into the subsurface layer can be neglected'5 For subliming or melting solid fuels, such as polymethyl-methacrylathe (PMMA), processes occurring inside the solid are somewhat simplified since vaporization (and pyrolysis) are confined to a thin layer of the fuel sample, at the condensed phase/gas phase interface. In the melt layer, bubbles grow and transport volatiles towards the surface. Thermal degradation processes in an oxidizing environment may be different from those occurring in an inert atmosphere. Indeed, oxygen may diffuse through the melt layer, favored by the large holes left at the surface by burst bubbles. 6 Flaming combustion processes are the result of complex interactions of transport phenomena in the gas phase (momentum, mass and heat transfer) and in the solid phase, thermal degradation of the solid and chemistry of the oxidation of the fuel vapors. Processes which can lead to gas phase ignition include the vaporization of the solid, the formation of a flammable mixture adjacent to the solid surface and the initiation and the sustainment of oxidation reactions. The characteristics of these processes determine whether ignition will occur and the ignition delay time. In general, to get flaming ignition, three conditions must be met: 7 (1) fuel and oxidizer must be available at a proper level of concentration to give a mixture within the flammability limits, (2) the gas phase temperature must attain values sufficiently high to initiate and accelerate the combustion reaction, and (3) the extent of the heated zone must be sufficiently large to overcome heat losses. The temperature of the mixture above the solid surface plays a key role. Its increase above certain levels can occur by heat transfer from the hot degrading surface and/or by devices capable of creating a region of very high temperature in the gas phase, such as pilot flames, sparks and hot wires (piloted ignition). Ignition can also be caused by hot air stream, hot surface and thermal radiation (auto-ignition). In the last case, the absorption of radiation by the fuel vapors is another very important mechanism for increasing the gas phase temperature to levels high enough to produce ignition. After ignition, proper conditions may allow flame spread and solid burning. In general, the flame spread rate is determined by the energy feed-back from the burning region to the unburned solid ahead of the flame while the burning rate is determined by the

Combustion processes of solid fuels rate of energy transfer from the flame to the degrading solid beneath the flame, a Combustion kinetics may also become very important for flame spread processes. It is often very difficult to determine which one among the many types of processes is the controlling one. Most analyses have been devoted to the determination of the mechanism of energy feed back from the burning region to the solid ahead of the flame. This energy transfer occurs by radiation, convection and conduction, the importance of each mechanism being related to geometrical factors, such as the thickness of the solid specimen or the scale of the fire, or environmental conditions, such as free or forced flow convection, oxygen level etc. For instance, radiative heat transfer is controlling for large scale turbulent flames whereas gas phase heat conduction is the major heat transfer mechanism for flames spreading over very thin fuel beds. This review article on modeling and simulation is divided into two parts, the first dealing with solid phase processes and the second with gas phase processes. For the solid phase part, thermal degradation of cellulosic materials, which give rise to char formation, and PMMA, a non-charring material will be treated. The analysis is focused on these two types of materials because most of the experiments and numerical simulations of flame spread over solid fuels have been carried out with paper, a charring material, and PMMA. Modeling and simulation of gas phase combustion processes (ignition, flame spread and extinction) will be presented in the second part of the review,

73

2. THERMALDEGRADATIONPROCESSES OF SOLID rUEtS In this section the analysis of pyrolysis processes in a non-oxidizing environment will be presented since modeling of smoldering combustion has already been treated in an excellent review by Ohlemiller. 9 Modeling of chemical kinetics and both physical and chemical processes of solid thermal degradation will be discussed,

2.1. Charring Materials


Wood, or more generally cellulosic fuels, are to be considered as the most representative of charring materials. Wood has an anisotropic structure with different property (thermal diffusivity, permeability) values along and cross grain. It is basically composed of cellulose (ca. 50~o), hemicellulose (ca. 25~) and lignin (ca. 25~), the proportion of these constituents varying to some extent among species.l'11 Smaller amounts of extractive and inorganic compounds are also present, Thermo-gravimetric analyses (TGA) indicate that hemicellulose is the least thermally stable component

of wood. The decompositon temperature increases from 117C to 1700C in the order hemicellulose< lignin < cellulose. 12 Furthermore the lignin component has a higher tendency for charring, whereas cellulose and hemicellulose readily decompose to volatile products at temperatures above 300C. Pyrolysis products are often grouped into a few main components. Each component represents a sum of numerous species which are lumped together to simplify the analysis. Generally, the product groups considered are: char, gas and tar. Chars are the carbon-rich non-volatile residues. Tars are any of several high molecular weight products (rich in 1,6anhydro compounds) that are volatile at the pyrolysis temperature but condense near room temperature. 13 Gases include all lower molecular weight products (mainly CO and CO2 and also water), which have a measurable vapor pressure at room temperature. Conventional pyrolysis, as usually carried out for conversion purposes, produces gases, tars and chars in approximately equal proportions. Also, flash (or fast) pyrolysis,14 based on high heating rates, leading to moderate temperatures (from 400C to 600C), and short residence times of volatiles, is used to produce high yields of tars. For both conventional and flash pyrolysis, heat transfer occurs from the gas film surrounding the solid surface. A variation of fast pyrolysis is represented by ablative pyrolysis, 15.16 based on direct contact of solid fuel with a high pressure, moving, hot surface. Heat transfer occurs through a thin film of liquid oil between the hot surface and the solid fuel. This results in much higher heat transfer coefficients, depending on the pressure and relative velocity of the hot reactor surface. The ablation rate and the thickness of the reacting zone are the two basic parameters of the problem. 16 According to the analyses conducted at the University of Aston, 17 such a conversion process seems to be more promising for obtaining high yields of liquid product than conventional and flash pyrolysis. In fact, the higher heat transfer coefficients allow higher reactor specific capacities (smaller equipment sizes) and avoid the need for a hot carrier gas as used in most flash pyrolysis units. In general, two different regimes during cellulosic material pyrolysis can be observed, as a result of the relative importance between heat transfer through the degrading medium and the overall chemical reaction rate. The two regimes which can be established are: (1) the chemical regime, when the rate of heat transfer is larger than the overall reaction rate and the thermal conversion process is controlled by chemistry and (2) the heat transfer controlled or ablation16 regime, when the rate of heat transfer is slow with respect to the overall reaction rate and strong temperature gradients are established along the fuel thickness, as the narrow reaction front propagates. Time evolution of the thermal degradation is the result of a strong coupling between transport phenomena and chemical reactions.

74

C. DI BLASI
SOLID PHASE chars primary degradation ~ (cracking - polymerization)

CHEMICAL PROCESSES heat flux ~

m~

of the solid

volatiles solid pre-heating (heatconduction) ~heat transfer by secondaryreactions I diffnsion and convection ~throughthe hot charlayeru ~

PHYSICAL PROCESSES m ~

Iconductlon,convectin Idiffnsion convection and


( and radiation I through the virgin solid I pressure gradients / interior to the degrading solid I condensation I surface regression crack formation shrinkage and/or swelling

FIG. 2. Schematic diagram of (solid phase) degradation processes of charring materials. Qualitative descriptions of charring material degradation, caused by a radiative heat flux applied on one side of the fuel sample, under ablation regime conditions, are given by Kanury s't and Roberts, H among others, and are summarized in the schematic diagram of Fig. 2. Initially the solid is essentially interested by transient heat conduction. Then a region, in the neighborhood of the heated side, undergoes thermal degradation. When all the volatiles are removed from the solid, a char layer is formed. Two further spatial zones can be seen: the region where pyrolysis reactions are active and the virgin wood region. Volatile species, generated in the pyrolysis region, may, because of pressure gradients 18-2, be forced to flow towards both the unreactcd solid and the already charred region. However, because of the much higher char permeability (lower mass flow resistance), the flow of products occurs mainly towards the heated surface. As volatile species flow through this high temperature region, secondary reactions can occur both homogeneously in the gas phase or heterogeneously on the surface of the char.t'2 In particular, heterogeneous reactions may give rise to exothermic char gasification by reducing the oxygenrich volatile products of primary pyrolysis, Volatile products may also migrate through the unreacted virgin solid where, because of tbe low temperature, they may condense and, subsequently, as the pyrolysis front progresses, evaporate. The soaking of some of the pyrolysis products and the subsequent redrying cause physical and chemical changes of material characteristics, even before the pyrolysis process starts. 5'1'~1 Apart from heat, momentum and mass transfer, changes in the physical structure of the reacting solid are observed with the development of a network of cracks in the already pyrolyzed region) 1 Surface regression and internal shrinkage and/or swelling are also possible. 11.1s.20 As the pyrolysis front progresses into the solid, the extent of the char layer, the residence time of the volatiles inside the porous char matrix and the size and the number of surface fissures increase. Thc increase in the char layer thickness causes a larger mass flow resistance, while the increase of volatile residence time enhances the possibility for further reactions to occur. The formation of cracks in the partially reacted wood naturally reduces the effects due to volatile reactions because of the reduced residence times (lower resistance to mass flow). It also alters the mechanisms of heat transfer, since radiation and convective flow of hot gases inside the solid may be enhanced.

2.2. Kinetic Modeling Detailed numerical simulations of solid fuel combustion are made complex not only by computational requirements, but also by the formulation of mathcmatical models from the complex physical and chemical processes and by the acquisition of reliable data.18 The determination of kinetic mechanisms and kinetic constants, for cellulosic material degradation, has been pursued mainly under chemical regime conditions (intrinsic kinetics) even though some efforts have also been devoted to the modeling of chemical processes under ablation regime conditions (apparent kinetics). Several studies have been conducted on the pyrolysis of wood and, in general, on biomass materials and their main components, cellulose, hemicellulose and lignin. Chemical processes of such materials can roughly be described as two stages, related to primary reactions of virgin solid degradation and secondary reactions of evolved degradation products.l'2 The description of both primary and secondary reactions has been afforded mainly by lumping the pyrolysis products into a few main groups (tar, gas and char) and by means of semi-global kinetic mechanisms. Considerable scatter can be observed on the reported kinetic data because of ~1 (a) the great vail-

Combustion processes of solid fuels ety of experimental techniques which give rise to different types of pyrolysis on dependence of solid and gas residence times, (b) the type of experiment (isothermal, dynamic), (c) the experimental conditions (temperature, pressure, heating rate), (d) the physical properties of the solid (mainly moisture content and particle size) and (e) the chemical composition of the solid (contents of cellulose, hemicellulose, lignin and inorganic components), In general, kinetic studies can be classified into three maingroups: (i) one step globalmodels, when a one-step reaction is used to describe degradation of the solid fuel by means of the experimentally measured rates of weight loss; (2) one-stage, multi-reaction models, used to correlate reaction product distributions. These are onestage simplified kinetic models, made of several reactions, describing the degradation of the solid to char and several gaseous species; and (3) two-stage, semi-global models, when kinetic mechanisms of solid degradation include both primary and secondary reactions,

75

2.2.1. One-stepglobalmodels Studies of group (1) propose a very simple kinetic scheme to model wood and cellulosic material thermal degradation, that is a global, one-step reaction: SOLID k--~VOLATILES + CHAR (AI)

whose rate is proportional either to weight residue or to weight loss after infinite time and presents an Arrhenius law dependence on temperature. Such models have been used to describe the chemistry of solid degradation for conditions of the chemically controlled regime as well as the ablation regime, that is for conditions where secondary reactions have a significant role. Thus, in the first case, the kinetic model (AI) describes the primary stage of the degradation process (gas residence times are very short) while in the second case, volatiles and char include products proceeding from the initial degradation (primary gases and chars) and those proceeding from the cracking and the repolymerization of primary volatiles. Early studies were reviewed by Roberts.11 Subsequently, other investigations have been conducted, most of them using TGA, 12"22-26 other devices such as fluidized bed reactors, 27 a tube furnace 2s'29 and other approaches based on in situ measurement techniques, s.3 Experimental results of solid fuel degradation can also be interpreted in terms of a single reaction accounting for the different fractions of gases; tars and chars formed: 3t SOLID k a GASES + b TARS + c CHAR (AI')

where a, b and c are the yield coefficients, expressed as grams of gas, condensable species and solid per grams of reacted solid. Table 1, where some data are reported, shows large differences in the estimated values of the kinetic constants. As anticipated, this is due to the complex chemistry of the thermal degradation, the complicated heat, mass and momentum transport phenomena occurring within the reacting medium and the effects of the size of the particle and environment (heating rate) conditions. In particular, studies have been conducted to analyze the effects of migration, condensation and regasification of volatile products inside the virgin solid on the apparent kinetics of degradation. Kanury and Blackshear 32 studied the pyrolysis of large radius, highly permeable cellulosic cylinders, uniformly heated along the surface, and observed that the reaction rate increases, at a particular temperature, passing from the surface to the center of the sample. Regasification of condensed pyrolysis products, which migrated in the virgin wood region, was conjectured to be the cause of this behavior. A further experimental study of the effects of migration, condensation and regasification of volatile species in the virgin solid fuel on the apparent kinetics of pyrolysis was conducted by Min. 29 He studied the pyrolysis of filter paper and filter paper smeared with condensable pyrolysis products, heated on one side, the other being pressed against a permeable chalk (where pyrolysis product could migrate) or an aluminium foil (non-permeable). Condensable species deposition on the virgin solid was seen to reduce the rate of gaseous species generation and to shift the Arrhenius curve towards lower temperatures. The trend is similar to that observed by Kanury and Blackshear, 32 however changes in the apparent kinetics were found to be negligible. The different conclusions reached by the two analyses can be attributed again to the different experimental conditions and thickness of the degrading solid, changing from very thick (cellulosic cylinders) to very thin (paper sheets)with trasnport phenomena becoming less important. 2.2.2. One-stage, multi-reaction models The studies of group (2) are related to the systematic analyses of the effects of temperature on the thermal degradation of small particles of wood, 33'34 celulose 14'3s'36 and lignin, 37~* analyzed through product yields and volatile composition. Under high heating rates, even though the onset of degradation and conversion of wood, cellulose and lignin occurs at different temperatures, the same qualitative behaviour of products, lumped as tar, char and gas, is observed. Tars and gases initially evolve at the same rate, but tar production becomes much larger as the temperature is increased. For even higher temperatures, tar yield reaches a maximum, then decreases

76

C. DI BLASI TABLE1. Kinetic data for one-step global models of charring fuel thermal degradation (X is solid conversion) Ref. [5] [23] [24] [25] [27] [27] [27] 127] [28] [29] [29] [30] [31] Sample ~t-cellulose Cellulose Fir wood Cellulose Beech sawdust Beech sawdust Cellulose Cellulose Cellulose Lignin Hemicellulose Wood Almond shell T(K) 550-1000 600-850 300-1100 580-1070 450-700 450-700 450-700 520-1270 520-1270 520-1270 520-1270 321-720 730-880 E(kJ mol-l) 79.4 100.5 101.7 + 142.7X 8.8-33.4 18 (T < 600) 84 (T > 600) 71 139.6 166.4 141.3 123.7 125.4 95-121 A(s- t ) 1.7 x 10* 1.2 x 106 2.1 x 108 0.019-0.14 0.0053 2.3 x 104 6.79 x 103 6.79 x 109 3.9 x 1011 1.2 x 10s 1.45 x 109 1.0 x l0 s 1.8 x 106

and finally attains an asymptotic value. The decrease of the tar yield at high temperatures is due to secondary cracking to light volatiles. The total gas yield also increases with temperature reaching a maximum in the rate of increase at the temperature of maximum tar yield. This is a consequence of secondary reactions, For temperatures above 800-850 K, tar is the major pyrolysis product while carbon monoxide dominates gas yield. Lower amounts of carbon dioxide, methane, ethylene and aldehydes are also observed, Product distribution is also dependent on the heating rate. Among others, such effects are presented for cellulose in Ref. [35]. Instead of the maximum in the curve of the tar yield as function of temperature observed for high heating rates, a constant value is obtained for low heating rates ( < 100C s-l). This behavior is believed to occur because volatile residence times are shorter than the time required for the attainment of temperatures sufficiently high for secondary reactions to occur. Furthermore, for a given temperature, the total volatile yield (tars +gases) increases as the heating rate decreases, as a consequence of more time available for thermal conversion, Most of the studies of group (2) were mainly related to the determination of product distribution and only very simplified kinetic modeling was proposed. Even though secondary reactions were believed to contribute extensively to the overall production of light gases, such processes as well as transport phenomena were neglected. The proposed models 3335.37 assume that the virgin solid fuel (wood or its components) decomposes directly to each reaction product i, except tar, by a single independent reaction according to k~ V I R G I N F U E L -----,PRODUCT i. (A2)

are the pre-exponential factor and apparent activation energy, respectively. The quantity Vi* is the ultimate attainable yield of the species i, that is, the yield at high temperatures for long residence times. Theoretical curves, obtained by best fit values of kinetic parameters, correlate well with experimental measurements. However, since at high temperatures and long residence times secondary reactions effects are not negligible, a rigorous kinetic model should include multi-step reactions for both the primary and the secondary stage of the degradation, as well as transport processes. Therefore the authors 33'35'37 pointed out that the values of the kinetic parameters are valid only for correlating experimental data under the operative conditions from which they were derived and not representative of the true physicochemical processes governing the degradation of solid fuels. 2.2.3. Two-stage, semi-global models Studies of group (3) are related to the determination of semi-global reaction mechanisms including the primary and, sometimes, the secondary stage of the solid degradation process together with the estimation of kinetic data. Among wood components, cellulose is the one most studied. One of the first semi-global models proposed for primary degradation of such a material is due to Kilzer and Broido: 41 k, CELLULOSE ~ A N H Y D R O C E L L U L O S E (A3) CELLULOSE - - ~ TAR (A4)

The kinetics are modeled through a unimolecular first-order reaction rate which can be expressed as: d Vi = Ai e x p ( - E i / R T ) ( V ~ * - VO (1) dt where Vi is the yield of the product i, and Ai and Ei
-

k3 ANHYDROCELLULOSE----o GAS + CHAR. (A5) According to the analyses conducted by Kilzer and Broido, 41 thermal degradation of cellulose starts at about 220"C with endothermic intermolecular water elimination to produce anhydrocellulose, whereas the higher temperature (about 280C), 'unzipping'

Combustion processes of solid fuels (propagation of the reaction chain), more endothermic step results in the formation of volatile ievoglucosan, the major constituent of tar. The anhydrocellulose undergoes further exothermic pyrolysis to form char and gases. Dehydration reactions are predominant at low temperatures and lead ultimately to char formation and, in air, to glowing combustion, Depolymerization reactions, predominant at higher temperatures, lead to the formation of volatile tars and, in air, to flaming combustion.a A subsequent study by Arseneau, 42 based on the analysis of thermograms of cellulose and levoglucosan, confirmed only partially the interesting study by Kilzer and Broido. 4~ For the case of thin cellulose samples, only endothermic process were observed, whereas for thick cellulose sample and levoglucosan a strong exotherm was observed for the range of temperature 300-350C. This behavior led to the conclusion that the anhydroceIlulose is highly reactive and rapidly undergoes further reactions with gas and char formation. However, such processes are not exothermic, while the levoglucosan decomposes exothermally. Thus the exotherm observed at high temperatures during the pyrolysis of thick cellulose samples is the result of secondary reactions and not representative of primary pyrolysis processes, The degradation of cellulose was also extensively investigated by Shafizadeh and coworkers L2'43 in the temperature range 260-340C. Similarly to the kinetic scheme (A3-A5), two main pathways were individuated. The first pathway, which dominates at low temperatures, involves reduction in the degree of polymerization by bond scission; appearance of free radicals; elimination of water; formation of low molecular weight gases and a char residue. At temperatures of about 300C, the second pathway starts to become competitive and rapidly dominates. The primary reaction in this pathway involves depolymerization by transglycosylation, followed by dehydration and formation of char and gas. As the temperature is increased, the tar forming reactions accelerate rapidly and overshadow the formation of char and gases. Interesting details on the chemistry of the process are given in Refs [1,2] and have also been critically analyzed in the review papers by Antal. 44'4s In Refs [1,2,43] the chemical kinetics of cellulose pyrolysis are represented by the following scheme: CELLULOSE - - ~ ACTIVE CELLULOSE (A6) ACTIVE CELLULOSE ~ks TARS (A7)

77

formation of an 'active cellulose', which is essentially the result of a strong reduction in the degree of polymerization (El = 242.4 kJ moi -1, A ~ = 2.8 l019 s-l). The degradation of active cellulose (E2 = 196.5 kJ tool -1, Az = 3.28 x l014 s-1, Ea = 150.5 kJ mol -~, Aa = 1.3 x l0 ~ s-~) leads to volatile, gas and char formations. Both mechanisms (A3-A5) and (A6--AS) account for the decrease of the char yield as the reaction temperature increases indicating, in agreement with experiments, 14'aa'aS'37 that the weight loss reactions (tar and gas formations) are competitive with the char formation reactions. However, it has been suggested 13 that the ratio of gas to char yield may not be constant, because low pressure and high temperature favor cracking reactions of the active cellulose to gases, while low temperature favors crosslinking and aromatization of the active cellulose to char. In order to examine this point, Agrawal~3 plotted some data about the ratio of the char yield to the total yield of char and gas as functions of temperature, showing values from 1 (300C) to 4 (400C). The conclusion was that if char and gas reactions were linked this ratio would be independent of temperature. Consequently, a three-reaction scheme, leading separately to primary tar, char and gas formation was proposed: C E L L U L O S E - - ~ GAS k2 CELLULOSE - - * CHAR k~ CELLULOSE - - * TAR. (A9)

(Al0)

(Al l)

The study, ta based on a re-examination of the weight loss measurements by Lipska and Parker, 46 shows how to evaluate the kinetic parameters for the reaction scheme (A9-A1 l). Also a comparison is made between the data estimated for the cases of: (linked char and gas formation) Yc~o - 0.35 Yc~0 + YGoo (not linked char and gas formation) Yc -- f ( T ) }'coo + YGoo
(3)

(2)

k3 ACTIVE CELLULOSE----. CHAR + GASES. (A8)

The primary degradation of cellulose is described by a first step reaction mechanism accounting for the

where f(T) is determined through experimental observation. The difference between the unmodified (Eq. (2)) and the modified (Eq. (3)) values results in small changes only in the amounts of tars and gases. The unmodified assumption overestimates the tar yields at high temperatures and underestimates the tar yields atlowtemperatures. Given the narrow tempera-

78

C. DI BLASI The second approach considers the fuel as a single homogeneous species which undergoes thermal degradation according to semi-global kinetic schemes. The primary pyrolysis rate of small biomass particles, in the temperature range 200--700C, was considered to be the sum of the rates of main biomass components by Koufopanos e t al. 48 Each component contributes to the formation of the pyrolysis rate to an extent proportional to its contribution to the composition of the virgin biomass. The interactions among the components as well as the possible bonds among them were assumed negligible. The kinetic model 48 for the description of the kinetic rate of each component is schematized as: VIRGIN M A T E R I A L - ~ INTERMEDIATE (AI5) I N T E R M E D I A T E - - ~ GASES + TARS (A16) INTERMEDIATE --~ CHAR. (A17)

ture range tested (250-360C), differences between the two approaches are not very large. However, they may be greater at higher reaction temperatures, The interesting feature of the kinetic model (A9Al 1) is that formation of tars, chars and gases may not be entirely linked so that the variation in the percentage of volatile products and chars with the operating conditions can be predicted. The model should not be viewed as suggesting that certain experimental conditions permit the entire conversion of cellulose to an individual product species at the expense of the other two species. 13 This is highly unlikely because the rates and the activation energies for the formation of various product species are comparable (the values of the activation energies of primary reactions (A9-AI1) are reported t3 to be 191,211 and 171 kJ mol 1, respectively). Agrawal 4~ also determined the kinetic data for the original Kilzer-Broido model (A3-A5) and a modifled Kilzer-Broido model with competitive reactions to form char and gas. It is shown 47 that the modified Kilzer-Broido model predicts well the experimental weight loss data, confirming that at low temperatures cellulose decomposition is dominated by reactions of formation of anhydrocellulose and tar. A tentative semi-global model for primary lignin pyrolysis has been proposed by Anta145 as a result of his review of literature data:

L I G N I N - - ~ CHAR + GAS LIGNIN ~-~ TAR

(A12) (A13)

The first reaction is similar to that proposed for cellulose 43 and describes the changes in the chemical structure of the solid fuel observed at low temperatures. However, the assumption of linked gas and tar formation does not come from considerations based on the chemistry of the process but simply from limitations of the experimental technique (TGA) allowing only weight loss measurements. Then, the pyrolysis of different biomass fuels is described according to the following rule: BIOMASS = ~ CELLULOSE + fl L I G N I N + ~,HEMICELLULOSE (A 18) where ~, fl and ~, represent the contribution of each component in the biomass composition. The determination of kinetic parameters was achieved by fitting experimental TGA curves with numerical solutions of the mass balance equations for chemical species. Six independent, first-order reactions for the pyrolysis of small dry pine wood sawdust, under negligible temperature gradients, were presented in Ref. [49]. Each reaction corresponds to each main wood component, that is hemicellulose (one reaction), cellulose (one reaction) and four species describing parts of the lignin macromolecule (or stages in its degradation). In Ref. [50"1, two sets of wood thermal degradation data, at subatmospheric pressure, obtained at constant temperature and at constant heating rate, were used to develop a model of weight loss rate, based on the contributions of the three main wood components. Experimental studies have also been conducted to determine the reaction rate of wood thermal degradation, summing reaction rates for cellulose and a 'second constituent' (hemicellulose and lignin), sl

LIGNIN --~ GAS + REACTIVE VAPORS. (A 14) Low temperature processes (AI2) are essentially represented by dehydration reactions. At higher temperatures, the formation of a variety of lignin monomers is described (AI3), which may undergo secondary degradation and condensation reactions for temperatures above 500C. At very high heating rates, fragmentation reactions are described by a further pathway (A14) which does not lead to char formation. Secondary char formation is, however, possible by the condensation of reactive vapors, Hemicellulose is the less studied wood component. It is generally believed that its pyrolysis mechanism is similar to that of cellulose 17 with the levoglucosan replaced by a furan derivative. However, experimental studies have not yet confirmed this supposition, Two different approaches have been employed in the modeling of the thermal degradation of complex solid fuels such as wood and biomass in general. The first approach considers the fuel composed of three chemical components, cellulose, hemicellulose and lignin, each of them present in different amounts,

Combustion processes of solid fuels An alternative description of the thermal degradation of wood and biomass materials considers the solid as a single homogeneous species. In some cases, only primary 3~'37's2 or secondary s3 reactions have been studied, in other cases, both primary and secondaryreactions ~5'54-57 have beenexamined. One of the most used primary wood degradation mechanisms, originally proposed by Shafizadeh and Chin, s8 is based on the following reactions: WOOD k , TAR (A19)

79

WOOD k2 GAS

(A20)

WOOD - - ~ CHAR.

(A21)

for in Refs [54,57]. Diebold Is modeled tar cracking as two competitive reactions to form gases and secondary tars, while Koufopanos et al. 56 describe secondary gas, tar and char formations as a result of interactions among primary pyrolysis products. Models [53-57] assume first order reactions and analyses indicate that tar degrades essentially to gas. Depending on reaction conditions, intra- and/extraparticle secondary reactions have a different influence on the product yields and distributions from wood pyrolysis. In particular, Boroson et al. 53 observed that tar conversion is strongly dependent on reaction temperature. For residence times of 1 s, homogeneous conversion is 30 wt% at 600C and increases to 88 wt% at 740C. They also analyzed the composition of evolved gases from tar cracking. It was found that carbon monoxide is the major product at all temperatures, accounting for 50-70 wt% of the tar converted, with low amounts of carbon dioxide, ethylene, acetylene, ethane and hydrogen. An approach, similar to that proposed by Diebold, ~5 has been used by Anta159 to model the dependence on temperature and residence times of gas yields from cellulose- and lignin-derived volatile matter: ~, VOLATILE ~ P E R M A N E N T GASES (A24)

Experimental verification of the model and determination of kinetic parameters require simultaneous collection of tar and gas and measurements of wood weight loss rate as a function of time. In Ref. [52] experimental measurements of tar, gas and residue mass fractions were made for temperatures varying in the range 300C-400C, while the range of evolution time, in the estimation of kinetic parameters, was chosen to avoid secondary reactions. Activation energies (A~ = 1.43 x 104 s ~, A2 = 4.1 x 106 s-1, A3 = 7.4 x l0 s s-1, E1 = 88.6 kJ mo1-1, E2 = 112.7 kJ mo1-1, E3 = 106.5 kJ mol 1) are comparable, indicating only a weak dependence of the distribution of pyrolysis products upon the pyrolysis temperature, for the small range of values considered, Two reactions, the first for tar formation and the second for a linked gas and char formation, :5'54's5 on the analogy of the active cellulose degradation mechanism of Ref. [43], or a single reaction, 57 as the kinetic scheme (AI'), have also been used to model primary wood degradation. Finally, Koufopanos et al. s6 proposed a kinetic scheme accounting for primary and secondary reactions, specifically to describe thermal degradation of large biomass particles under ablation regime conditions. The kinetic model ineludes two primary reactions for char formation and linked gas and tar formations, respectively, with rates showing a power law dependence on virgin solid concentration and a modified, three parameter Arrhenius law dependence on temperature. In general, secondary reactions describe tar cracking to lighter gases and tar repolymerization to char: TAR-~ CHAR (A22)

V O L A T I L E - ~ R E F R A C T O R Y CONDENSABLE MATERIAL. (A25)

The first reaction produces more permanent gases by cracking the reactive material to lighter, less reactive species. The second step produces refractory condensable materials, which may be a tar or some combination of water-soluble organic compounds. First order rates were employed and kinetic parameters were estimated for cellulose. As for lignin, only the difference (El - E2)and the ratio In(A1/A2)were determined.

2.3. Energetics o f the Pyrolysis Reactions Differential thermal analysis (DTA) has been widely used to investigate the energetics of the thermal degradation of charring cellulosic materials. Some of the early results have been summarized by Roberts. 6 In general, it has been found that pyrolysis of hemicelluiose and lignin is an exothermic process, while cellulose pyrolysis is an endothermic process 'at low temperatures and becomes an exothermic process at high temperatures.12'42"49'6 Endothermic as well as exothermic processes of wood pyrolysis reactions at different temperatures have been found (see, for example, Refs [10,32,60]). Also very large differences are observed in the measured values for the same pyrolysis temperature. The effects of reaction temperature on the ener-

T A R - ~ GAS.

(A23)

Studies s3'55 only consider tar cracking to light gases, Tar polymerization to secondary char is accounted
JPC$ 19:I-F

80

C. DI BLASI

getics of the pyrolysis of large wood samples have been investigated in Ref. [20]. At low radiative heat fluxes, the process is endothermic ( - 6 1 0 kJ kg -t is the computed value). At high radiative heat fluxes (high temperatures), even though no specific value of the heat of reaction is given, it appears that the process is globally exothermic. Also, a dependence of the energetics of cellulose pyrolysis on the sample thickness was found by Arseneau. 42 Thin cellulose samples exhibit endothermic degradation while thick samples exhibit endothermicity followed by a significant exothermicity, The changes in the energetics of cellulosic material degradation, as temperature and size of the sample vary, are believed to be due to the different role played by primary and secondary reactions. At low temperatures and short residence times of volatiles, only primary (endothermic) reactions are active, while the high temperatures cause secondary (exothermic) reactions. Also, the occurrence of secondary charring reactions and the lower medium permeability to gas flow (that is, the increased residence times) may cause changes in the global energetics of the process. These findings are also confirmed by a recent study s6 where, through measurements of temperature-time history inside the degrading solid, it was shown that the process is initially endothermic and then weakly exothermic. Estimation of the heat of reaction leads to a value of - 2 5 5 kJ k g ~ for primary reactions and to a value of 20 kJ kg -1 for secondary reactions,

2.4. Modeling of ChemicalandPhysical Processes Detailed mechanisms controlling transport phenomena (momentum, heat and mass transfer) as well as chemical processes occurring during thermal conversion of wood and related substances, under ablation regime conditions, are not yet available to improve fire prevention and control and to support chemical reactor design and scale up, where the use of large particle feed is preferred due to the high cost of size reduction. 6~ From a practical point of view, processes affecting thermal conversion of particles in biomass reactors are related to temperature and velocity fields established around the particle. The analysis in such cases is made complex by the non-uniformity of temperature, species concentrations, velocity and pressure profiles and the variation of reaction rate along the particle thickness. In order to study the thermal conversion from a quantitative point of view and more controllable conditions, heating fluxes have been varied in the range of values 30-250 kJ m -2 s-~ which are of interest both for simulating fire conditions and for thermal conversion and combustion of biomass. Measurements of the effects of ambient oxygen level and heat flux on wood gasification,6~-6a of the application of the heat flux with respect to wood grain direction 2'6t gave a relevant contribu-

tion to improve the understanding of physical and chemical processes. Several models of chemical and physical processes occurring during cellulosic material thermal degradation have also been published either related to fire safety issues and to biomass conversion for energy. Based on the properties of the Oregon Iodgepole pine and on the two length scales of interest, that is the pore diameter (10 /lm), through which volatile products escape, and the characteristic size of typical feed particles (1 cm), Chan et al. 64 gave an estimation of the order of magnitude of main process characteristic times, which can be used as a guideline in the formulation of a mathematical model. The explicit effects of temperature on transport characteristic times are small while, given the typical values of the activation energy (125-170 kJ mol-~), they are strong on the chemical reaction rate. More complex is the implicit dependence of transport properties on the pore distribution which is, in turn, dependent on reaction conditions. Following the changes in the medium density, a global increase in the porosity of a factor two is expected. Permeability and effective diffusivity, on the other hand, depend on the third 65 and second order power of porosity. The increase in the medium porosity lowers the transport times and increases heat transfer times, unless radiation is important. Also, the volumetric thermal capacity of the solid decreases with temperature because of the decrease in the solid density during thermal degradation. In conclusion, at high temperatures, the heat transfer rate is several orders of magnitude slower than the chemical reaction rate while, at low temperatures, the degradation reaction rate is about four times slower than that of heat transfer. For large particles a zone, where the characteristic times of chemical reactions and heat transfer are comparable, propagates through the particle. Thus in the development of a mathematical model both effects should be accounted for. Another important aspect of the wood degradation is moisture release which is a low temperature phenomenon and affects local thermal properties. From the analysis of characteristic times it is seen that, for dry particles, convective mass transport is faster than diffusive transport, indicating that this last effect can be disregarded. For wet particles, since moisture evaporation occurs at temperature values much lower than those allowing pyrolysis, volatile diffusion may compete with convective flow. Indeed, at this stage of the process, the porosity and permeability have not yet changed. In general, even for flow parallel to the wood grain direction where permeabilities are very large, gas velocities are rather low, resulting in gas residence times sufficiently long for secondary reactions to occur. The characteristic time of convective heat transfer is also very short at the microlevel and, given that the solid volumetric heat capacity is about 600 times larger than that of the gas phase, it may be assumed that volatile products, as they flow, are

Combustion processes of solid fuels rapidly heated to the temperature of the char (local thermal equilibrium). Consequently a heat flow from the solid phase char to the volatiles occurs and the heat transfer towards the virgin solid region is lowered. Mathematical models, available to date, use simplifying assumptions for the description of both chemical processes and transport phenomena. As for chemical reactions, in most cases, studies of thermal degradation of wood and its main components have been grouped as studies on cellulosic materials and chemical processes have been modeled according to a one-step, first order, Arrhenius reaction (see, for instance, more recent publications66-7), In a few cases, the description of physical processes is combined with multistep reaction schemes for chemical processes. Cellulose decomposition has been modeled in Ref. [71] according to a scheme where the virgin solid, considered as a single species, decomposes by a first reaction into a second solid and a gas and, by a second reaction, into another gas. The solid product of the first reaction may further react to give a final (inert) solid and another volatile species. Even though the proposed scheme was not based on experimental evidence, the analysis is interesting because it presents the first attempt to account for a multi-step reaction scheme in the modeling of both chemical and physical processes occurring during cellulosic material decomposition, The assumption that the virgin solid can be described as a single homogeneous species has been made in Refs [64,72-74]. All these analyses include primary and secondary reactions for solid fuel pyrolysis, and use the assumption of lumping the products of degradation only into tar, char and gas. Even though this is still a rough approximation, the upproach allows main experimental observations to be accounted for. The contribution of each wood component is considered in Ref. [75] where kinetics of wood pyrolysis, described according to Ref. [49], are coupled to enthalpy and moisture/vapor balances, A different degree of approximation has also been employed in the modeling of transport phenomena, The problem of the pyrolysis of a wooden sample has always been simulated as one-dimensional, since this condition is often met in the experiments. 2'62'63 Most of the analyses take the viewpoint that the porosity is fine and uniformly distributed. The material is considered as a homogeneous medium, with specified porosity and permeability, where gas and solid are in good thermal contact (solid and gas are at the same temperature). The simplest approach consists of a heat conduction equation with a source term accounting for heat release due to chemical processes, written for a non-porous, constant property solid. 69 71 An improvement in the description of transport phenomena is presented by Kung 66 and Villermaux et. al.,ra who also include convective heat transfer due to the outward flow of volatiles generated during pyrolysis. The assumption of quasi-steady gas

81

phase processes is still made, that is the accumulation of mass and energy of the gaseous species within the solid is neglected, the gas density being three orders of magnitude lower than that of the solid. This assumption is removed in the model by Kansa eta/. 67 which also accounts for pressure variations inside the porous solid according to the Darcy law. Therefore this model presents a rather complete description of transport phenomena, even though a one-step global reaction is still used for the chemistry. Models accounting for multi-step reaction schemes of wood pyrolysis still make some assumptions in the description of physics. In Ref. (64) the gas is assumed to flow only towards the heated side of the particle and there is no description of mass transfer resistance. No convective transport of condensable species, products of the pyrolysis process, is included, and the volatile release is modeled as instantaneous (dpg/3t = 0). Apart from the interesting analysis of the characteristic time scales of the different phenomena, the model is not used to make extensive simulations of the thermal degradation process. A comparison between the predicted and the measured temperature history, at two locations along the particle thickness, and pyrolysis product distribution for a couple of heat flux values and particle sizes are given. The model by Hastouglu and Berruti 73 is quite complete, that is, all of the main chemical and physical processes are included. In particular, it accounts for the non-isobaric mass transport through the porous medium by means of the 'dusty gas' flux equation. The model also tries to account for the fibrous structure of wood by assuming that the reacting particle is made of hollow cylindrical fibres which, after pyrolysis, again leave a homogeneous charred solid. However the analysis is limited to chemically controlled small particle pyrolysis and only the conversion time history for several bulk stream temperatures is given. More interesting is the analysis presented by Curtis and Miller 72 which, even if assumes quasi-steady gas phase equations and constant pressure and is still focused on very small biomass particles (from 5 x l0 -s to 4 x 10-2 cm), presents parametric results of the effects of particle thickness, total pressure and sample heating rate. It is shown that significant gradients are established within the sample, which affect product yields. Secondary reactions are negligible only in the limit of very low total pressure, which reduces gas densities and residence times in the region of high temperature. Also the role played by radiative heat transfer and pyrolysis was examined by the use of two different models, the first using an effective thermal conductivity and the second the method of zones. 76 This effect, for the temperatures considered in the analysis, was found to be negligible. The model by Di Blasi et al.74 includes convective and diffusive heat transfer, unsteady terms in the gas phase equations, pressure variations and variable properties. The most restrictive assumption is that of

82

C. DI BLASI (4) no diffusive transport of volatile species, generated during the reaction process, occurs; (5) condensation of volatile species in the virgin solid region is neglected; (6) kinetic and potential energy are neglected in the energy balance equation and internal energy is replaced by enthalpy; and (7) gases behave according to the ideal gas law. The energy, chemical species mass, momentum and ideal gas law equations and relations giving solid volume, porosity and medium property variations constitute the mathematical description of the problem: - m a s s balance for wood species dpw -- (Kx + K2 + K3)pw (4) dt - m a s s balance for char species

condensed phase tar species. Thermal decomposition of wood indeed starts at rather high temperatures (above 550 K), thus tar is present as gas in the pyrolysis region and in the char layer where even higher temperatures are reached. Tar might be present as a condensed phase species only in the virgin wood region where, after migration, volatile species may condense because of the low temperatures. The study was developed with the aim of clarifying the role played by the different assumptions generally made in the formulation of wood pyrolysis mathematical models, that is (1) one-step kinetics, (2) quasi-steady gas phase with constant porosity and (3) constant properties. All these assumptions were found to affect even the qualitative behavior of the predictions. The assumption of condensed phase tar has been removed in Refs [77,78]. Apart from the prediction of the wood thermal degradation, the model has been used to investigate the coupling of heat transfer and secondary reaction processes to the flow field by varying wood and char permeabilities. Internal flow convection and volatile residence times mainly depend on char permeability. As char permeability is decreased, for a fixed wood permeability, larger pressure peaks, ahead of the pyrolysis front, and lower velocities, in the char region, are observed. Residence times of tar gases inside the char region become longer and secondary reactions are favored. The effects of the application (parallel or perpendicular to wood grain direction) of the radiative heat flux, used to cause thermal degradation, have also been investigated.78 The effects of moisture vaporization have been accounted for only by Chan et a1.,64 who describe the process as an additional chemical reaction and by Alvesand Figueiredo75 whoassumethatwatervapori _ zation is controlled by heat equilibrium relation, by a local moisture-vapor supply and can be described

Opc - 0t = K3pw + eKspr - m a s s balancefortarspecies 0(epT) C3(pTU) O-----~ + 3x - tOT - t o t a l continuity

(5)

(6)

O(eps)
dt - energy conservation

(pgu) + -= cog 0x

(7)

OT (pwCw + pccc + e(cGp~ + CvPr)) tgt

+ (T-

I" OOw Op ecOpC Opt T o ) ~ C w ~ + Cc-- + -- + ecr-dt dt dt 0t de) (

+ (pGc6 + pT)--- + (T-- To) cr

+ c6 Opru_ Ox /
k= *.3

2.4.1. Theequationsofwoodpyrolysis A one-dimensional model of the thermal degradation of a dry wooden slab in inert atmosphere subjected on one side to a radiative heat flux has been recently developed. 77 78 Effects of variable properties, unsteady gas phase processes, pressure and velocity variations and convective transport of tar species have been accounted for. Chemical processes have been modeled according to primary reactions (A19A21) and secondary reactions (A22-A23). The model is based on the following assumptions: (1) the volume occupied by the cell wall sample does not change as the solid undergoes pyrolysis (thermal swelling and/or shrinkage and surface regression are neglected, that is V = Vs(t) + V~(t) = const); (2) the gas and the solid matrix are in local thermal equilibrium; (3) inertial terms in the momentum balance equations are negligible;

/ .OT\ + u(prcT + P~C~)OT=ox Ox~k ~ x ) + ~ rkAhk + ~, erkAhk


k=4.5 - Darcy law u = - - tCap g 0x -ideal gas law p = p,RT/W s - V s variation Vs (Mw + Me) -- = Vso Mso where e = Vd V (11) (10) (9) (8)

Combustion processes of solid fuels

83

K = rIKw + (1 - q)Kc
1110. d = r/dw + ( I r/)dc

8I~.- 7o0. I-

~"~'~-~'-'C
, 2~

wooo

5,'JF/A ~///-1

lO2O.
930.

,1 = Mw/Mwo

600.~ , , , 5oo. 2, ', ,,

U/__///// ..] 840. /// -t 7 o.

'~
-

Kk=Akexp(--Ek/RT) k= Akexp(--Ek/RT)pw k rk = Ak exp(-- Ek/R T)px

1,5,
=

1,3

8_ 400. ~ " ', "l.,~.~"v/~

/'( /

-~ 660. ~

k = 4,5

100. 0.
.000

390. ~ '
.005

tog = (K~ + K2)pw - eKspT pw = Mw/V,

'
.010 .015

'
.020 .025

300.

pc = Mc/V

X [m] FIG. 3. Temperature and wood mass concentration as functions of the sample length from t = 3 rain and with step 3 rain. 2.00 [
1.80

Pc = M~/VG = M~/(eV),

pr = Mr~ Vg = MT/(e V), pg = p c +Pr.

400.
360. 320. 280. ~
-

, = ~ [.,i ,, /lC~A~' !l i'~ l .... ~, , ,~n,~r~, , , a 1.40 t t i j i I 1;, , , , , ,


~E 120

[-,7 - 7 - ~ ~

"~"

x .~ I"

1.00

, , ~ u.~.-.-~ .8o ~ .60 ~ ~ .40

}r- J ~ I 1 I

~ r ~ i'~ ~ ~ . . \ ~ N ~ x ~

I ~

I-

240.
200.

s ~

~2~

16o.
120.

ff
~

80.

.20 ~ / 7 , / i
.00 ~ - " .000

.~/'~ ~ X k , k ~
~"" "~ ~ ~ " ~ " ' ~ ' ~ : .O05 .010 .015 .02O .025

40.
0.

x [m] FIG. 4. Tar and char mass concentrations as functions of the sample length from t = 3 rain and with step 3 rain. .024 "~=.. p -,".,- * ~.. . t=3mln/ r. ,, .016 - . . ,, ~ ~ . ",, ,, .I, , , j // -~ "" ' ~*'*,A/~" .012 ~ ~, ,, '**,, ~ -, ,'. , , ; ~ / ~ . - , "-, ," ,~ z .ooa i ' , "" "~
.004

30.0 25.0 20.0


15.0

c~ z 10.0 o.
5.0

In the equations given above, pw a n d pc are the a p p a r e n t wood a n d char densities, px, p c a n d p, are the mass c o n c e n t r a t i o n s of volatile species, u the gas velocity, e the porosity, e the heat capacity, T the temperature, k the thermal conductivity, K the permeability,/z the viscosity, R the universal gas constant, W s the m e a n molecular weight, V the volume, M the mass, d the pore diameter, ~ the S t e f a n - B o l t z m a n n c o n s t a n t a n d to the emissivity. The subscript 0 refers to initial conditions, s to solid phase, g to gas phase, W to wood, T to tar a n d C to char. In the energy balance Eq. (8), the first two terms account for the accumulation o f the enthalpy o f condensed phase a n d gas phase species, the third a n d fourth terms for the convective t r a n s p o r t o f the gas phase species, the fifth for the conductive t r a n s p o r t o f heat a n d the last two for the heat release associated with chemical reactions. Radiative heat transfer, inside the solid, is described t h r o u g h a n effective radiative c o n t r i b u t i o n to the thermal conductivity. 64 A linear variation o f the conductive c o n t r i b u t i o n to effective thermal conductivity, permeability a n d pore diameter with the m e d i u m composition, between the virgin wood value a n d the char value, is assumed. Also, the decrease of the solid volume is p r o p o r t i o n a l to the decrease in the mass o f the solid, due to thermal degradation. In order to define the problem, initial a n d b o u n d ary conditions should also be assigned. Initially (t = 0) the solid is in a quiescent e n v i r o n m e n t at a m b i e n t conditions: T= To, pw=pwo,

i" - " - /
-"

--

/
~ I .010 I .015

.~
I 020

~ / .025

p=po,

u=0.

(12)

.000 ~ -.002 .000

I 005

.0

F o r t > 0, the fuel slab is subjected on one side (x = L) to an assigned, constant, radiative heat flux, q. A t this side, radiant a n d convective heat transfer from the surface a n d c o n s t a n t a m b i e n t pressure conditions are used: c3T k * - - = q - tr(T 4 - To ) - h .... ( T - To); p = p o ~ Ox (13) where h .... is the convective heat transfer coefficient a n d tr the S t e f a n - B o l t z m a n n constant.

X [m] F~G. 5. Gas velocity and overpressure as functions of the sample length from t = 3 min and with step 3 min.

k*

= k~o. + k,=d

ko, = r/kw + (1 - rl)kc + ek, kraal---- o'T 3 d/to

84

C. DI BLASl

.0010 [ .o008 .00o6 o 0 ,~


x

i"''i /
I

P " ~

30. 25. 20. \


15. ,, 10.

.ooo4 .o0o2 -

,,

,, x-

,-/ .o0o0 ~ -.0002 - NXk, juJ -.0004 j 2.00


1.60

~'E ~ 5. ~ o.

~ 13oo.

T
t I

/
TAR

110o. 9o0. _
~"

1.2o x w .80 .40 .o0 .010 x ~ xx -

7o0. ~ 500.

,I I 300. .015 .020 .025 X [m] FIG. 6. Gas velocity, overpressure, temperature and tar concentration as functions of the sample length for t = 3 min. 1.00 800.
.90 580.

primary reactions and by Liden e t al. ss for secondary reactions. The time and space evolution of the pyrolysis process is shown through main variable distributions (temperature, chemical species concentrations, gas overpressure and velocity) in Figs 3-5, for certain times, as a function of the distance from the heated surface. Initially the time evolution of the phenomenon is controlled by the increase of the temperature at the surface as a result of the applied radiative heat flux, while chemical species concentrations are constant and equal to the initial values, indicating that no reaction takes place. As the surface temperature reaches a value of about 550 K, primary wood decomposition rates start to increase, initiating the pyrolysis process. Then, for times shorter than 15 min, three main regions are present in the computational domain: a virgin fuel region, a primary pyrolysis region and a char layer. In the virgin wood region, due to the low temperature values (T < 400 K), reaction rates are negligible. The boundary between this region and the primary pyrolysis region is characterized by a maximum in the gas overpressure. The location of the gas overpressure peak also separates two velocity distributions, one directed towards the cold side of the sample (virgin wood region), the other towards the irradiated surface (pyrolysis and char regions). The maximum
in the pressure distribution is caused by gas produc-

.80 Lno,.~ .70 x .80 E .50 ~_..40 ~ .30 .20 .10


.00

T ,-'" - ~" "~

'2 560. L=:3min- 540. 520. 5O0. ~ 480. 480. 440. 420.
I 400.

.000

.005

. 0 1 0 .015

.020

.025

X [m] FIG. 7. Reaction rate of tar formation and temperature simulated at the maximum reaction rate position as functions of the sample length from t = 3 rain and with step 3 min. On the cold side of the sample (x = 0), convective and radiative heat transfer and zero velocity conditions are considered: k* d T - - = a ( T 4 - T ~ ) + h .... (T - To); u = 0. 8x (14) 2.4.2. N u m e r i c a l s i m u l a t i o n o f w o o d p y r o l y s i s In this section some results of a wood pyrolysis simulation, 7a are presented. The radiative heat flux is 84 kW m 2 and the fuel sample is 0.025 m thick, Property values are taken mainly from Lee e t al., 2 kinetic data are those by Thurner and Mann 5z for

tion in the very low permeability neighborhood of the boundary between reacting and non-reacting regions. Here, both the gas volume and the convective transport are small. This is shown by the magnified view of the gas overpressure, velocity, temperature and tar concentration distributions for t = 3 min, reported in Fig. 6. Indeed, even though in this zone the temperature is still low, the concentration of gaseous species is found to be about five times larger than those in the virgin wood region, immediately ahead of the maximum pressure. Notwithstanding the significant pressure gradients in the virgin wood region, only very low gas velocities are computed because of the very low values of permeability. The maximum pressure increases slightly with time because of the increased resistance to mass flow. The virgin wood region is followed by the region of primary pyrolysis characterized by temperatures in the range 400 K < T < 700 K. At these temperatures, only primary reactions, accounting for wood degradation to tar, char and gas, are active. From the solid phase species distribution, it appears that a (primary) pyrolysis wave, about 0.25 x 10 2 m thick, propagates through the virgin solid with a decreasing spread rate (on the average, during pyrolysis of the whole sample, the propagation rate of the reaction front is about 2 x 10-5 m s-~). The propagation of the (primary) pyrolysis front through the virgin wood region is also shown in Fig. 7, where the reaction rate of tar formation and the temperature value, simulated at the maximum tar reaction rate position,

Combustion processes of solid fuels .40 .38 .32


~o .28 .24

85

t=Bmin~J1050.
/t
r4 10oo. 950.

1100.

having gone to completion and secondary reactions not active, only heat transfer occurs. Given the rather high temperature, enhanced by exothermic secondary reactions, heat transfer in the char region occurs not
only by convection and conduction, but by radiative

mE .20
.16

900. - b-850.

.12 .08 .o4

800. 750. .00 I 700. .000 . 0 0 5 .010 .015 .020 .025

X [m] FIG. 8. Reaction rate of tar cracking and temperature simulated at the maximum reaction rate position as functions of the sample length from t = 3 min and with step 3 rain. as a function of the fuel sample length, for certain times, are plotted. As time increases, primary reactions occur at lower temperatures and a decreasing maximum is shown. Indeed, temperature variations of only 30 K cause strong reductions in the reaction rate because of the exponential dependence, Due to the increase in the medium permeability along the primary pyrolysis region, the gas velocity starts to increase from zero to positive values, allowing convective transport towards the irradiated surface. The energetics of the conversion process, heat convection and conduction determine the temperature distribution in this region. Tar species reaches a maximum, as does the maximum wood decomposition rate, and goes to zero as the temperature decreases, towards the virgin solid region. A smoother decrease in the tar distribution is observed at the boundary between the pyrolysis region and the third region of the computational domain, the char layer, characterized by temperatures larger than 700 K. Tar concentration values along the char layer are the result of convective transport, which, given the larger permeabilities, is much more active, and of secondary reactions. The occurrence of secondary reactions, as volatiles are convected through the already charred zone, is confirmed by the distribution of reaction rates of tar cracking to lighter gases, reported in Fig. 8 for certain times as a function of the distance from the irradiated surface. The figure also shows a variation of 150 K in the temperature values simulated at the location of the peak in the reaction rate. As time increases, the secondary reaction front slowly enlarges through the char layer with a decreasing maximum in the reaction rate and then propagates behind the primary reaction front. Thus two subregions can be seen in the char layer, the first characterized by very high temperatures (800 K < T < 1100 K) where both secondary reactions and heat transfer take place, and the second characterized by lower temperatures (700 K < T < 800 K), where the primary reactions

From Fig. 4 it appears that primary char formation does not lead to the attainment of a constant value but a slight increase is observed as the pyrolysis front extends along the sample length. This is due to primary reactions occurring at lower temperatures which favor char formation. Secondary char formation, shown in Fig. 4, as well as secondary gas formation, are strongly dependent on the residence times of the gas phase tar species inside the porous solid matrix. Thus, the longer the residence times, the greater the extent of secondary reactions. Formation of secondary char also alters the structure of the solid matrix. Indeed porosity first decreases, due to the primary wood pyrolysis, and then increases. For heating times longer than 15 min, the primary reaction front enlarges to the whole sample and the decrease of permeability leads to a continuous decrease in the gas overpressure until the total conversion of wood to char and gaseous species (t = 27 min). Quantitative comparison of numerical predictions and experimental measurements is difficult because of the variation of physical properties and kinetic data among cellulosic materials and the dependence of kinetic data on experimental conditions. A comparison between the predictions of the model presented here and experiments by Lee et al. 2 is, however, given by Di Blasi and Russo. 78 Good quantitative agreement is obtained for temperature and pressure distributions for short times as long as variations in the physical structure of the solid are not significant. Both theory and experiments predict a maximum overpressure but, whereas the model simulates a propagating pressure front with a slightly increasing maximum, in the experiments the propagation of the pressure front occurs with a decreasing maximum. The char permeability is not constant (as in the model) but increases with time, due to crack formation, and the char layer thickness, due to surface regression and internal shrinkage, is narrower. The decrease of char permeability leads to a decreasing maximum overpressure while the reduced char layer thickness leads, on average, to slightly larger temperature values. Finally, predictions and measurements of the time history of total gas production show fairly good agreement.

2.5. Non-Charring Materials Complete degradation of some thermoplastic polymers occurs with breaking of the main chain and no char formation, so that heat transfer conditions at the condensed phase/gas phase interface do not

86

C. DI BLASI

change, allowing degradation and pyrolysis processes to be studied under quasi-stationary conditions, Quasi-stationary conditions, of course, cannot be established for char-forming polymers because heat and mass transfer processes are strongly affected by the formation of the char layer. In this review, P M M A degradation will be briefly analyzed because this material has been widely used to investigate flammability limits and flame spread characteristics both experimentally and theoretically. A more extensive review of thermal pyrolysis of thermoplastic polymers was presented by Khalturiskii and Berlin 79 with reference to experimental techniques, surface temperature measurements and modeling of apparent kinetics, On dependence of heating exposure, two types of thermal degradation of P M M A can be observed: 79 isothermal bulk degradation and surface degradation or linear pyrolysis, which, respectively, correspond to the chemical regime and the ablation regime, previously introduced for charring materials, When P M M A is exposed to a radiant flux, under linear pyrolysis conditions, a narrow layer, near the heated surface, liquifies. Small clusters of degradation products must escape from the surface. Since the boiling point of the degradation products is much less than the polymer degradation temperature and since a significant quantity of only partially degraded polymer is still available at the surface, the degradation products are superheated. Consequently they nucleate and form bubbles which grow due to mass diffusion through the molten layer, mass vaporization and thermal expansion. 6 Furthermore, because of the higher temperatures, the degree of superheating and bubble nucleation rate increase while surface tension decreases, as the surface is approached. Transportation of bubbles from the interior of the polymer towards the surface is also affected by the existence of viscosity gradients, a Bubbles very close to the surface are able to burst directly through it while those well below burst through small holes. The size and rate of formation of bubbles depend on the polymer characteristics and environmental conditions, in particular the oxygen level.6 The presence of oxygen in the environment causes a lower viscosity of the molten layer, with higher bubble frequency, Formation of large holes, caused by burst bubbles, extends the effects of the oxygen level to the interior of the sample, Even under conditions of bulk degradation, the gaseous degradation products should escape from the molten polymer, but experiments are conducted with thin samples in order to avoid making transport phenomena the rate-controlling step for weight loss.

2.6. Kinetic Modeling In general, the degradation of P M M A occurs according to the following main stages: depolymeriza-

tion initiation, propagation of reaction chain and termination. Depolymerization is believed to initiate at chain ends, at random points along the chain, or at isolated 'weak links'. 81 Free radicals, formed during initiation, give rise to unzipping, that is, the propagation of the reaction chain. Chemical proesses at this stage may be characterized by the length of the reaction chain, that is the number of monomer units produced on average for one initiation. Termination reactions account for the stabilization of free radicals which occurs through the combination with an H atom. The abstraction of an H atom can occur either from an inactive chain, thus causing another initiation, or at a random point along the chain, or from the same chain of the free radical, leading to products larger than the monomer unit. Kinetic modeling studies of P M M A degradation can be divided into two main categories: (1) One-step global models which employ a onestep, global, Arrhenius rate reaction to account for all chemical processes. Such an approximation has been used both with the description of physical proesses, to model solid phase processes only, and also in all computer models which couple solid phase processes (degradation and usually heat transfer) to gas phase combustion and transport phenomena. Studies on P M M A degradation which model chemical processes by means of a one-step global reaction have been employed both for isothermal bulk degradation and linear pyrolysis. Explicit relations were given between surface regression rate and surface temperature, in the limit of large activation energies, al This treatment is based on the assumption that the linear pyrolysis mechanism is the same as that of homogeneous bulk degradation. The study by Krishnamurty and Williams 82 makes the same assumption since it uses a pre-exponential factor and an activation energy, developed for bulk degradation, for correlating linear pyrolysis data. The extrapolation of the isothermal bulk degradation kinetics data to the range of regression rates observed in the linear pyrolysis is, however, considered invalid. 79 This is confirmed by a comparison of the measured time evolution of surface temperature and mass loss rate. a3 In fact, mass loss rate, for linear pyrolysis, still increases while the surface temperature reaches a constant value. This indicates that the mass loss rate cannot be related to the surface temperature through an Arrhenius expression. Therefore the sub-surface of the sample contributes to the rate of gasification. (2) Detailed degradation models where kinetic schemes, accounting for chain initiation reactions, depropagation reactions and termination reactions, have also been proposed. Such models have never been coupled to the description of physical processes. Studies belonging to this second group are those performed by Kashiwagi et al. s4 as P M M A depolymerization, under vacuum conditions, may initiate at chain ends or at random points, the former mechanism being favored at low temperatures, the latter at

Combustion processes of solid fuels high temperatures. Possible sources of random scission were analyzed in Ref. 1-89]. In Ref. [84] the mechanisms of thermal and oxidative degradation were investigated by measuring the molecular weight, For both thermal and oxidative degradation, random scission is the initiation step. However, the oxidative environment causes a much faster reduction in the degree of polymerization, and the activation energy, for random scission of polymer chains, was about four times lower than that for degradation in nitrogen. A further study on the effects of oxygen on P M M A degradation was presented in Refs 1,85,86] through the measurements of weight loss and molecular weight. Oxygen has a different effect on the temperature dependence for polymer degradation: at low temperatures the stability to loss of weight is increased whereas at high temperatures random scission is enhanced with a decrease in the polymer stability. A theoretical model of thermal degradation of P M M A 87 accounting for three main chain reactions: (1) random initiation, (2) depropagation of free radicals and (3) termination of free radicals, has also been proposed. Numerical solutions of a large system of ordinary differential equations allow the changes in the molecular weight distribution and in the sample volume to be predicted as a function of the initial molecular weight, average zip length, etc. The assumption of a steady state radical concentration made in Ref. 1,87] was removed in Ref. 1-88].

87

tions in the limit of low Reynolds numbers and under the assumption that surface tension forces dominate buoyancy forces. After the bubble has nucleated, the expression for the growth rate is obtained by an approximate balance between the rate of heat flow into the bubble from the liquid and the latent heat required to supply bubble vapor. By means of the translational velocity and the bubble growth rate expressions, the bubble distribution function is determined. For this derivation it is assumed that bubbles are spherical and that the rate of binary bubble collisions is negligible. The expression, giving the bubble distribution function, is coupled to the balance equations for the melt layer through sink terms which describe, on average, the effects of bubble distribution on mass, energy and momentum transfer from the liquid to the bubbles. Expressions for the regression rate, the non-dimensional melt velocity and the liquid volume fraction were obtained by means of the method of matched asymptotic expansions. The regression rate was the result of a balance between the surface heating rate, the rate of heat removal into the condensed phase and the rate at which the monomer is lost from the polymer with bubble formation.

3. GAS PHASE COMBUSTION PROCESSES A mathematical model of ignition and flame spread over solid fuels should include the description of transport phenomena and chemical processes occurring in the solid and gas phase and should account for their interaction. Fuel vapors, produced by solid fueldegradation, mix with air and, under appropriate conditions, form a flammable mixture above the surface. At the same time, the gas layer, adjacent to the surface, is also heated by both heat conduction from the solid and, sometimes, the same external source used to heat the solid. As a consequence of the heating and the existence of a flammable layer in the gas phase, the rate of exothermic gas phase reaction rapidly increases together with the heat release rate. A runaway condition can be achieved and ignition occurs (auto-ignitionT), or ignition can be caused by an external device such as a pilot flame. In this case, the flame may be quenched by heat losses at the surface once fuel vapors have been depleted (flash) or, if enough heat is supplied to the solid, it may become a sustained diffusion flame (piloted ignition91). Sometimes ignition is only the first step of further gas phase combustion processes, such as flame spread. The exothermicity of gas phase reactions and of possible oxidation reactions inside the solid fuel cause a heat flux that, through solid and gas phase processes, heats the unburned fuel. To have flame spread, the burning region must supply enough heat to the unburned solid to cause degradation. At the same time, proper conditions in the gas phase should be met. In fact, flame spread characteristics are af-

2.7. Modeling of Chemical and Physical Processes Most of the models of thermoplastic polymer degradation, available to date, are based on a simple energy balance equation for the solid with degradation localized at the surface, and have been coupled to gas phase equations for simulating flame spread, The very few models, specifically formulated for improving the knowledge of solid phase chemical and physical processes, will be reviewed here. Vovelle et al.S3 proposed a model of the mass loss rate of PMMA, subjected to a radiant heat flux, based on a constant property one-dimensional energy equation accounting for in-depth degradation of the solid. Surface regression was modeled by introducing the Landau transformation. A comparison of simulated and measured temperature profiles shows good agreement. Surface temperature is predicted to undergo a rapid increase followed by the attainment of a constant value which depends on the incident heat flux. The effects of bubbles inside the molten layer on the steady-state transport ofvolatiles, during degradation of thermoplastic polymers, have been predicted by means of a one-dimensional model in Ref. (90). The model includes the description of individual bubble characteristics in terms of translational velocity and growth rate. The translational velocity is determined from the steady-state Navier-Stokes equa-

88

C. DI BLASI Another important factor for flame spread processes is fuel thickness. In general the definitions of 'thermally thin' and 'thermally thick' fuel are used. 92 The term thermally thin indicates that the fuel thickness is small compared to the characteristic thickness of the thermal diffusion layer in the solid, along which no significant variation of solid properties is observed. Depending on the thickness of the solid fuel, the role played by the solid fuel thickness in the path for heat transfer, through the solid, to the unburned fuel changes from being of negligible importance for thin fuels to becoming of primary importance for thick fuels. If the combustible is thin, flame spread is further characterized by the consumption of the solid fuel in the burning region. This produces a propagating burn-out front which affects the flame and pyrolysis front propagation. Material properties influence the behavior of the sample under fire conditions. Most experiments on flame spread over thick solid fuels have been carried out with PMMA, a non-charring material, while experiments related to thin fuels have mainly used paper, a charring material. Gas phase combustion processes are very complicated because of the very high number of chemical species evolved from degrading solid both for cellulosic materials and thermoplastic polymers. For example, in the flame above P M M A surfaces, 95 apart from the fuel monomer, these include hydrogen, methane, ethane, ethylene, acetylene, propene, propylene, allene, propyne, methanol, formaldehyde, carbon monoxide and possibly some fuel-produced carbon dioxide. Few attempts have been made to investigate the detailed structure of the flame 95'96 and, for PMMA, to propose a detailed mechanism of gas phase thermal decomposition. 95 On the other hand, numerical modeling of hydrocarbon fuel oxidation is becoming possible for many types of fuels. Reaction mechanisms are extended over wide ranges of experimental conditions and are being used for problems of much greater complexity than those addressed previously97 but numerical models of combustion processes of solid phase fuels have always described gas phase oxidation chemistry by means of a one-step global reaction. There are two reasons for this simplification. The first is that it is generally acknowledged that the fundamental process of solid fuel combustion is controlled more by physical mechanisms (heat transfer phenomena) than chemical mechanisms. 98 In the second place, in order to apply detailed kinetic models for gas phase combustion, a very accurate description of the chemistry of solid degradation is needed, that is accurate predictions of evolved combustible volatiles have to become available. In some way, solid phase models should give the 'input', through the boundary conditions at the solid/gas interface, for detailed modeling of gas phase combustion processes. However, as stressed in the first part of this review, only global or semi-global kinetic models and composition of evolved gases under speci-

fected not only by the mechanisms of solid degradation but also by other factors, such as the flow configuration, oxygen level and orientation of the solid fuel. Flame spread over solid fuels can be classified into two main categories according to flow conditions. 92 One mode occurs when flame spread is in the same direction as the oxidizing flow (flow assisted flame spread). The second mode occurs when the flame spreads against the oxidizing gas flow (opposed flow flame spread), In the flow-assisted mode of flame spread, the concurrent flow pushes the flame ahead of the vaporizing fuel surface. The heat transfer from the hot mixture of reacting gases and the combustion products above the vaporized region to the unburned fuel surface favors the propagation of the flame. The resulting flame spread process is very rapid and consequently of great importance to fire safety science. For a very large range of flow velocities and oxygen concentrations, this phenomenon appears to be controlled simply by heat transfer mechanisms. 92 However, the flow remains laminar only in the initial stage when heat transfer from the flame to the fuel is due mainly to convection. When the dimensions of the flame increase, the flow becomes turbulent and flame radiation appears to be the dominant mode of heat transfer, In the case of opposed-flow flame spread, the heat transfer to the unburned fuel is more difficult since the flame and pyrolysis fronts are in the same location. The opposed flow pushes the flame into the burning region, and heat transfer to the unburned fuel, which occurs through solid or gas phase conduction, is very slow. The flame front is well defined and generally the size of the fire is easily controlled. The phenomenon shows an interesting dependence on environmental conditions: 92 it is dominated by heattransfer mechanisms at relatively low opposed-flow velocities and high oxygen concentrations and by chemical kinetics at high flow velocities or low oxygen concentrations, Ignition and flame spread processes strongly depend on the presence of gravity and, in a normal gravity environment, on the orientation of the solid. 92 Upward flame spread, that is, in the direction opposed to the gravity vector, is faster than downward flame spread, that is, in the same direction as the gravity vector. In the case of upward flame spread the heat transfer from the hot combustible gases to the fuel is enhanced by natural convection while, in the opposite case, natural convection slows the spread process, taking away the hot combustion products from the unburned fuel. Upward and downward modes of flame spread show the same characteristics of flow-assisted and opposed-flow flame spread. Different characteristics of spread are observed 93'94 in microgravity environments when no natural convection is present; radiative heat losses are believed to play an important role.

Combustion processes of solid fuels 2800.0


solid f ~ (-'~ gas

89

23oo.o 1800.0 T[K] 1300.0


ooo

12.6 12.4 12 ..

~il!il
"

~i~ ~ ~

300.0

-.7

0.0

x[cm]

20.0

40.0

FIG. 9. Solid and gas phase temperature as functions of the space coordinate for several times (s).

tic conditions are available. Recently, an attempt has been made to estimate the kinetic constants for pyrolysis, thermal oxidative degradation and char oxidation for cellulosic materials. 99 Yields of CO, total hydrocarbons, char and products (CO2 and H20) from each reaction have also been evaluated. The study is interesting in that it represents the first step towards the introduction of two gas phase oxidative reactions, one for total hydrocarbons and the other for CO, instead of the one-step reaction usually employed. The guidelines of this part of the review are essentially those followed in Ref. [100]. Analyses of recent simulation results of auto-ignition as well as piloted ignition will be presented. Then, the characteristics of computer models of flame spread available to date and some results of numerical simulations will be analyzed. 3.1. Modeling and Simulation of lgnition Processes

yl.O ~ 8[

" i I ~ ~ ~ .2 l ~ ~ ' - ) S f / l ~ / ' 0 0 [6t( ~2"~k-~x'~k'l"~'~'~/~'~ 15 20.0 40.0 x[em] FIG. 10. Fuel (solid lines) and oxygen (dashed lines) mass fractions as functions of the space coordinate for several times (s). 0.0

9o.o 75.0 12.61 6o.o u [cm/s] I 45.0 ~ 3o.o 15.0 0.0 0.0 5 . 20.0 x[cm] 40.0 \ 3

15

FIG. 11. Velocity as a function of the space coordinate for several times(s),

The simplest mathematical models of ignition only consider one-dimensional solid energy balances and use ignition criteria based on the solid surface reaching a 'critical' temperature. Asymptotic techniques, valid in the limit of large activation energies, have been applied for the solution of more advanced onedimensional models, including finite-rate gas phase kinetics. 1L12 A numerical solution of the problem is presented in Ref. 1-103]. Although in Refs [101-103] equations both for the solid phase (energy balance) and gas phase (chemical species mass and energy) are solved to model noncharring solid radiative ignition, the gas phase absorption of radiation is not accounted for and the only mechanism that can lead to ignition is heat conduction. Consequently, such models predict ignition only if very high surface temperatures are reached or if high reaction rates are simulated. This last condition can be achieved only by the use of values for kinetic parameters which are not realistic and thus not useful for comparisons with experimental data. The model proposed by Fernandez-Peilo and coworkers t4 is interesting in that, for the first time, they tried to account for gas phase absorption of radiation in an oxidizer stagnation point flow. However, solid phase processes are not described at all and the temperature at the solid/gas interface is assumed equal to an estimated vaporization temperature. Solid phase processes are described in Ref. [105]. Thermal radiation is absorbed in-depth by the solid fuel according to the Beers's law. Solid degradation is also an in-depth process, and is modeled with a zeroorder Arrhenius rate reaction. The density and the solid properties are constant while surface regression is described, in the solid energy balance equation, by a convective term formed by the product of the regression rate and the temperature gradient. The radiation available for heating the solid fuel decreases with time because, as fuel vapors are

90

C. Dl BLASI front, giving a premixed zone from t = 12.61 s (Fig. 10). As long as a premixed zone exists ahead of the flame front, the flame shows a quite high propagation rate, decreasing with time according to the decrease in the temporal gradient of density, velocity, and amount of fuel ahead of the flame front. For t > 15 s, the premixed zone ahead of the flame no longer exists and the typical structure of a diffusion flame is observed. As ignition occurs in the gas phase, the heated layer in the solid increases. The predicted ignition delay time as a function of the intensity of the radiative heat flux shows rather good agreement with experiments, and gas phase absorption of radiation is found to play a controlling role for ignition. The dependence of the ignition phenomenon on kinetic parameters 15 has also been simulated. A one-dimensional ignition model of PMMA, exposed to a radiatively active high temperature source, has been proposed by Back and Kim. 16 Even though in-depth degradation and surface regression of the solid are neglected, the vaporized fuel absorbs, emits and scatters radiation. Ignition delay times are found to be dependent on temperature and emissivity of the hot surface but not on the distance between the fuel surface and the hot source. Gas phase ignition of cellulosic materials has received less attention than PMMA. A very simplified description of radiative ignition for such materials is proposed by Ghandi and Kanury. 17 Finite rate pyrolysis (first order) and combustion (second order) reactions are considered. The assumptions of the Boussinesq approximation (constant gas phase density except for variations due to buoyancy) and of a well-stirred boundary layer in the longitudinal direction reduce the mathematical model to the constant property and constant pressure one-dimensional flow equations. Solid phase processes are described by the transient one-dimensional heat conduction and mass balance equations. A mixed analytical/numerical upproach is used for the computation of the solution. The analysis of the simulations has allowed the validity of the ignition criterion based on a reverse in the sign of the gas phase temperature gradient at the solid/gas interface to be assessed. Piloted ignition has been simulated in Ref. [91]. Solid phase processes are modeled through the formulation of boundary conditions on temperature and chemical species to describe fuel injection through a porous plate at a known temperature. The problem is thus described through one-dimensional gas phase equations accounting for combustion processes by means of a second order, one-step irreversible Arrhenius reaction. Simulations of piloted ignition show a time evolution of main variables in some way similar to that observed for auto-ignition with a premixed flame, propagating both towards and away from the porous plate. Furthermore, the effects of the ignition source location, the fuel flow rate and the temperature at the surface have been investigated.

produced, they absorb some of the incident radiation, A treatment, valid in the limit of an optically thin medium, has been applied to describe such a process, that is the absorbed radiation is proportional to the fuel vapor concentration and the radiation intensity:

dI~ -- flpflg = fllopfexp f f ' -flpfdx t~x

(15)

where pf is the partial pressure of fuel vapors and f l a constant, global absorption coefficient. The gas phase mathematical model includes onedimensional continuity, mass balances for chemical species, an energy balance and the ideal gas law, written under the assumptions of unit Lewis number, constant specific heats, thermal conductivity and pressure. A second-order finite rate Arrhenius reaction is used to model combustion. Numerical solution of the problem is simplified through the introduction of a Lagrangian coordinate which uncouples velocity from the other unknown variables, An example of the simulated ignition phenomena 15 for PMMA, radiatively heated with a heat flux of 8.3 W cm -2, is shown through Figs 9-11, where solid and gas temperature, fuel and oxygen mass fractions and velocity are reported as a function of x for selected times. Until t = 5 s, the solid fuel absorbs all the external radiation and displays ternperatures higher than those of the gas phase. When solid phase temperatures larger than 600 K are reached, the degradation process starts and, as the fuel vapors diffuse, they also absorb some of the external radiation in the gas, which shows temperatures higher than those of the solid. Because the amount of absorbed radiation depends on the thickness of the fuel layer and the fuel concentration, the temperature in the gas phase shows a maximum at 2.5 cm from the surface for t = l0 s. Until this time, the reaction rate is negligible and no disappearance of fuel and oxygen is observed. Both of them, indeed, give rise to a flammable mixture layer 7.5 cm thick, Ignition occurs for t = 12.61 s and is characterized by a large instantaneous increase in the gas temperature (the temporal gradient of gas temperature is about 70,000 K s-~) and by a noticeable consumption of fuel and oxygen. Because ignition occurs in the gas phase (about 3 cm from the surface) and a flammable mixture layer is present, the propagation of two flame fronts, towards the surface and towards the unreacted gases, is observed. Oxygen, close to the surface and in the whole mixture layer, is burned very fast. Then the former flame front approaches the surface and the latter continues to propagate. The thickness of the fuel layer under the propagating flame front increases and absorption of external radiation makes the temperature rise. Ignition is also characterized by strong temporal gradients of density that cause high gas phase velocities, as shown in Fig. ! 1. High gas velocity and spatial fuel concentration gradients, push fuel ahead of the propagating flame

Combustion processes of solid fuels 3.2. Flame Spread Modeling In order to simplify the mathematical treatment of the problem, early studies of flame spread employed several assumptions. The most commonly used approximations are constant surface temperature of the solid during thermal degradation, a flame sheet and a boundary layer. The assumption of constant vaporization temperature has been criticized since, although solid fuel reaches constant surface temperature in the pyrolysis zone during the vaporization process, this value depends on material properties, ambient pressure and temperature, solid fuel thickness and heating conditions. The flame sheet assumption implies the existence of an infinite gas phase reaction rate, whereas boundary layer theory assumes that diffusive processes along the streamwise direction are less important than those occurring in the cross direction, However, at the flame leading edge, both for opposed flow and flow assisted flame spread, and at the flame tip, for flow assisted flame spread, the assumptions of infinitely fast kinetics and boundary layer are not justifiable, a Extensive literature on simplified models, reviewed in Refs [8,108,109], is available. Sometimes explicit expressions for spread rates have been obtained. Generally, these formulae are able to describe qualitatively the dependence on environmental conditions and property values of the solid fuel, when phenomena are controlled by heat transfer mechanisms, Simplified models of flow assisted flame spread, still based on the boundary layer and infinite kinetics assumptions, have also been recently proposed. 1lO.111 More comprehensive mathematical models include balance equations for both gas and solid phases and remove some approximations. These more complete models are not amenable to analytical solutions and it is necessary to use numerical approaches. Detailed mathematical models, solved numerically, do not give explicit expressions for global parameters. On the contrary, these can be derived from the predicted time and space evolution of the phenomenon, often by means of the same approaches used in experiments, 3.2.1. Gas phase models Models of flame spread are two-dimensional balance equations for the gas phase coupled through the boundary conditions at the interface to solid phase balance equations. As for the gas phase, most advanced models published to date, include momenturn, energy and chemical species mass balance equations. All analyses are for laminar flow, and finite rate combustion kinetics are described through an overall, second order reaction: F + voO ~ vpP. Viscous dissipation and compressive work are neglected, Furthermore, the coupling between the momentum equations and the state equation due to pressure

91

terms, when momentum balance equations are included in the mathematical formulation of the problem, is neglected. Pressure variations in space are very small and, since in general the system is open, the mean pressure reduced to the specified ambient pressure. As in Ref. I-112], the pressure excess with respect to the ambient value is neglected in the state equation, while it is retained in the momentum equations. The decoupling of the momentum equations from the state equations cuts off the acoustic waves and the determination of the pressure field becomes an elliptic problem. Apart from Ref. [113-1, no model of flame spread accounts for gas phase heat radiation losses. To date, flame spread models numerically solved and treated the velocity field differently. The simplest models consider the solution of species and energy equations assuming that the gas density and pressure are constant and the velocity field is known.l ~4-12~ The streamwise component of velocity is assigned according to a chosen profile, while the effects of the mass outflow from the vaporizing surface on the velocity field are neglected. Consequently, these models account essentially for thermo-diffusive aspects of flame spread phenomena. The computational costs are very low and have therefore been widely used. Predicted values of the opposed-flow spread rate depend strongly on the velocity profile used in the computations. However, the use of flow fields assigned according to the Oseen hypothesis (uniform velocity profile) and to the Hagen-Poiseuille profile (parabolic velocity profile) do predict the same functional dependence of the spread rate on the maximum oxidizing flow velocity 122 as observed in the experiments. A second type of model, 123 which has been used to simulate flame spread, assumes the Boussinesq approximation for the formulation of the momentum equations. With this approximation, the pressure can be decoupled from the velocity field if the unsteady Navier-Stokes equations are expressed in vorticity and stream function transport form and the continuity equation is identically satisfied. The most complete models are those which consider the solution of the full Navier-Stokes equations with density and pressure variations. 113.! 24-131 As for the numerical techniques, all differential models of flame spread have been approximated to systems of algebraic equations by means of finite differences. Details on the numerical schemes and solution proceduresaregiven in Ref. [100-1. Models of flame spread essentially differ in that they use either steady l13-115A24-129'lal or unsteady~ tr-tz3,130 formulations of the gas phase equations. They can be roughly classified as: (a) quasi-steady, constant spread rate models, (b) quasisteady, variable spread rate models and (c) unsteady models. Quasi-steady, constant spread rate models. From a rigorous point of view, flame spread is an unsteady

92

C. DI BLASI

phenomenon. However, a quasi-steady treatment of ment of the coupling between pressure and velocity, the gas phase equations can be used, based on the which is the most critical aspect to be faced in the fact that the characteristic evolution times of the computation of the numerical solution, can be solid are much longer than those of the gas phase, avoided. Compressible Navier-Stokes equations may given that the flame spread rate is much smaller than be again formulated in terms of two equations for the gas velocity. In this approach, gas phase condi- the vorticity ~ and the stream function ~,: tions can be considered as a series of steady states resulting from small changes in the solid.~ 14 Besides ~ux = pu ~ur = - p v ~ = uy - ux. the assumption of a steady gas phase, if a reference In this way the continuity equation is identically frame attached to the flame front is employed, the satisfied even though the gas density varies with flame becomes stationary, while the oxidizing environ- temperature. The calculation of the pressure field is ment and the solid fuel move towards the flame with not required for the computation of the velocity field: velocity equal to the spread rate. It should be ob- if desired it can be readily computed from a Poisson served that such a formulation is possible only under equation, obtained by taking the divergence of the the assumption of constant spread rate. The flame momentum equations. spread rate becomes an eigenvalue of the problem Unfortunately, apart from the model in Ref. [131], and it is not known a priori. From a computational all models of flame spread, based on the assumption point of view, a reference velocity is introduced as an of quasi-steady gas phase processes, do not use the initial guess, then its value is continuously adjusted compressible vorticity-stream function formulation in order to meet the true value of the spread rate. of the Navier-Stokes equations. Numerical methods Most of the numerical models of flame spread belong used to solve the momentum equations account for to this category. ~13'~24 129 Also, the models of Refs the pressure-velocity coupling by iterating the solu[132-135], used to predict the thermal and convective tion for each time step, according to the SIMPLE structure of a diffusion flame in a counterflow environ- and SIMPLER approaches.137 Iteration procedures ment, can be included in this category. In these are not only very expensive in terms of computer models, only the gas phase equations are solved and time but convergence is not always guaranteed, unless solid phase processes are described by boundary a d h o c u n d e r r e l a x i n g f a c t o r s a r e u s e d . 138'139 conditions at the solid/gas interface. Using this apUnsteady spread rate models. The problem becomes proach, the most important feature in the modeling more complex from a numerical point of view when of flame spread, that is, the coupling between the the full unsteady Navier-Stokes equations with dendegradation of the solid fuel and the gas, is not sity and pressure variations are to be considered. described. However, interesting information has been Such a formulation is the proper one to be used for obtained on the extinction and blowoff of diffusion highly unsteady gas phase combustion processes such flames, as ignition and, sometimes, extinction. The coupling Quasi-steady, variable spreadratemodels. The mathbetween pressure and velocity cannot be avoided, ematical formulation of model equations adopted in however an effective technique which avoids iteraRefs [113,124-129"1 allows only predictions of con- tions by splitting computational operations into few stant spread rate problems. However the spread rate steps, was proposed by lssa for non-reacting is not always constant. For example, in concurrent flOWS. 138'139 This technique, named PISO (Pressureflame spread over thin solid fuels, an asymptotic Implicit with Splitting Operators), has been applied value of the spread rate is achieved only after a for simulating opposed flow flame spread. ~3 relatively long accelerative stage. 136 Of course, The PISO technique is an extension of classical models [113,124-129] cannot predict such behavior, splitting procedures used to solve discretized equaTo retain the unsteady character of the spread rate, tions, such as ADI, to the treatment of the coupling however, it is sufficient to use a fixed coordinate between pressure and velocity in the momentum system, that is to write unsteady balance equations equations by predictor-corrector steps. One predictor for the solid phase only, by making use of the quasi- and two corrector steps are considered. Starting from steady approximation for the gas phase. Thus, as the a known pressure field, the velocity distribution is solution in the solid is advanced of a time step, a new computed by an implicit solution of momentum equadistribution of gas phase variables is determined. The tions (predictor step). In general, this velocity field spread rate does not appear any longer in the balance does not satisfy the continuity equation. Hence a equations and can be determined from the distribu- corrector step is performed where a new pressure tion of simulated time evolution of the main variables distribution is computed by means of the latest avail(for instance, from the different positions of the flame able velocities. Then the flow velocities are updated leading edge). This approach has been used in early and temperature and species mass fractions are also thermo-diffusive models by T'ien ~~4.115 and, more computed. At this point, the continuity equation is recently, in the formulation of a more advanced satisfied but to improve the accuracy of the solution, model, including the Navier-Stokes equations. T M An- a new corrector step for pressure and velocities is other very important point to be observed is that, if performed. It is possible to demonstrate that soluthe steady gas phase equations are written, the treat- tions obtained with one predictor and two corrector

Combustion processes of solid fuels steps are second-order accurate in time. ~ss'139 A higher formal order of accuracy can be reached by increasing the number ofcorrector steps, 3.2.2.

93

Solidphase models

From a mathematical point of view, models of thermally thin fuels can take advantage of the uniformity along the fuel thickness, observed in the experiments, and use one-dimensional balance equations. However fuel consumption (burn-out) cannot be neglected. Two-dimensional heat transfer must be taken into account for thermally thick fuels but, in this case, fuel consumption is generally neglected, Modeling fuels of intermediate thickness is very difficult and both two-dimensional and solid consumption effects are important. The complex models of chemical and physical pr0cesses presented for thermal degradation of charring materials have never been used together with gas phase models to simulate flame spread. Also, the models of flame spread over P M M A do not account for the effects of bubble formation in the degrading solid. Very simple descriptions for the solid phase processes have been used. Thermally thick materials were modeled by neglecting in-depth mass transfer and the regression rate, by describing the chemical kinetics with a one-step Arrhenius reaction, and including heat transfer. 116 12o,122,123,13o,131 In the modeling of cellulosic materials, char formation was described in a very simple fashion and only for thin fuels.113 115,121,124 131 Thermally thick fuel models assume that the heated polymeric fuel remains solid until it finally gasifies at the surface according to a zero-order Arrhenius pyrolysis reaction giving the corresponding monomer. In such a way, the use of a vaporization temperature is not needed and the temperature at the interface, which depends on fuel properties and environmental conditions, is found as a result of the solution. The external thermal radiation, which is used to cause ignition, is absorbed by the solid fuel according to Beer's law. Heat capacity, thermal conductivity and density are assumed constant. Processes related to surface consumption of the solid fuel are neglected. By these hypotheses, and bearing in mind that pyrolysis is a surface process and must be taken into account through the boundary conditions, the mathematical model for the solid phase can be expressed by a simple twodimensional heat conduction equation. Properties of the solid and kinetic data have been chosen to describe PMMAXlS and cellulosic materials.116-12o,131 The models for thermally thin fuels assume that variable distributions across the solid thickness are uniform. For chemical processes, only pyrolysis is described: S---, vGG + vcC with kinetics expressed according to a first-order Arrhenius law. Most of the models [113-115,124-129] only consider two ordinary differential equations describing solid density and temperature changes associated with the chemi-

cal reaction. On the contrary, the model proposed by Kung 66 for wood pyrolysis, was adapted to describe thin paper degradation in Refs [121,131]. As the solid pyrolyzes, the density, specific heat and thermal diffusivity change from their initial values to their final values for the char. The fuel vapors, flowing out of the solid immediately after vaporization, are in thermal equilibrium with the char. Also, the char surface is at the same location where the virgin solid surface was, but the burn-out front propagates as burning occurs. The energy balance accounts for the variation of the solid fuel enthalpy, energy convection due to the flow of volatile gases through the porous solid, heat diffusion along the direction of flame propagation, and energy release due to the pyrolysis reaction. 3.2.3.

A mathematical model offlame spread

An example of a mathematical model accounting for main chemical and physicalprocessesofflame spread over thermally thick solid fuels, based on unsteady gas phase equations, has been proposed in Ref. [130]. Apart from basic assumptions, already outlined in Section 3.2.1, constant thermal properties and viscosity and validity of the ideal gas law are assumed. Gas phase equations are written for: -continuity

Op O(pu) O(pv)
-- + 0t
-

+ 0x dy

= 0

(16)

momentum along the x direction

(On

uOu your= p 3 (On


+ 3 ~x \~xx +

~p
-

fOZu ~2u~
(17)

P ~ t + --t3x + 0y/'

---63x+ /.t k~xx2 + ~y2]

O~y) g(p - po) sin 0

- momentum along the y direction

(Or

uOv vOv']

Op

fO2v

P ~t + --Ox + Oy / = ---Oy q- II ~ X 2 +

02v"] 0y2,]
(18)

p~(Ou O~y) ++ - g ( P - po)cos0 3 ~y \Ox


- chemical species

( t3Yi + - - Y, + vOY,~ uO P\ Ot Ox t3y /


[- 0 / t3Yi'X Of

= wi
6 Yi'~7 ~

+ L ~ x ?D~ x ) + ~y~pD~y)J
i=O,F -energy

(19)

(OT uOT vOT') (O2T 02T ") PCv Ot + - - + =q+k + Ox Oy f \OX 2 Oy2 f
(20)

94 -- state equation
p T = cos t.

C. DI BLAS! (26) The cross component of velocity is given by a balance of mass fluxes:
pv = m

(21)

The solid phase model, for thermally thick fuels, includes a balance equation for: - energy
c~Ts f~ZTs ~Ts~

(27)

(22)

-Ctsk~Sx2 + ~y2 j " Chemical production terms are expressed as:


Wi=
~ A

while for the streamwise component a no-slip condition is usually assigned. Downstream of the leading edge of the flame, the boundary layer approximation is adequate and consequently the first derivatives along the streamwise direction are zero. The remaining boundary conditions are based on the physical problem to be modeled (open boundaries or walls). The treatment of the far field boundary conditions is particularly complicated for the velocity components. In some cases these conditions are derived from the solution of simplified mathematical models; that is mixed humerical/analytical solutions of simplified theories are used as far field boundary conditions for the solution of more complete mathematical models. Examples of such an approach can be found in the numerical prediction of the burning of vertical cellulosic cylinders 14 or in the modeling of free convection along a heated vertical plate (see for example Ref. [141]) where the Ostrach's 142 laminar flow solution is used. In most cases, zero normal gradients for the velocity components and other main variables are specified at the far field boundary, lt3'lls'119'132 134 On the other hand, for the prediction of laminar free convection in a heated cavity, it has been shown 143 that, while for a realistic simulation of the far field flow pattern, proper specification of the far field boundary condition is required, the characteristics of the flow inside and immediately in front of a cavity do not differ from those simulated through the specification of zero normal gradient along the far field boundary.

exp ( - E ' ~ YoYvp 2 ---viMi i -- O, F. MF \RTJ

q = -wFAH.

In Eqs (16-21), u and v are the longitudinal and normal velocity components, # the viscosity, Y the species mass fraction, p the density, D the diffusion coefficient, cp the specific heat at constant pressure, T the temperature, k the thermal diffusivity, g the acceleration due to gravity, 0 the angle of solid fuel orientation with respect to gravity, A and E the preexponential factor and the activation energy i~, the gas-phase reaction, AH the heat of combustion, R the universal gas constant, v the stoichiometric coefficient and M the mean molecular weight. The subscript 0 refers to initial conditions and i to chemical species: F (fuel), O (oxygen), I (inert), P (product). In the solid phase energy balance (22), ~s is the thermal diffusivity and Ts the temperature, The initial conditions correspond to ambient conditions, with ignition caused by an external radiative heat source. 13 In order to describe ignition and flame spread over solid fuels, Eqs (16)-(21) must be coupled to the solid phase balance Eq. (22) through boundary conditions at the solid-gas interface. Usually, the changes in the shape of the degraded solid behind the flame are neglected and the boundary conditions are obtained by writing species mass flux and heat flux balances. For the chemical species, the mass fluxes of species i convected out and diffused out must be equal to the mass flux of species i generated on the surface:
pD ~ YF = m(Yv -- 1)

3.3. Opposed Flow Flame Spread An example of the predicted flame structure, for opposed flow flame spread over PMMA (property values and kinetic data are those reported in Ref. [144]), as computed with the model presented above is given in Fig. 12, showing the solid and gas phase temperatures, vector velocity field and the chemical species mass fractions. Opposed flow flame spread is dominated by the interaction of fluid-dynamics, chemical reactions and heat transfer at the flame leading edge (position of the maximum heat flux from the flame to the fuel), where a combustible mixture is established. Downstream of this position, the flame structure is that typical of a diffusion flame in a counterflow environment: isolines of reaction rate correspond to the flame position characterized by the maximum temperature and the minimum fuel and oxygen mass fractions. From the vector velocity field, two main flows appear, one of vapor fuels from the pyrolyzing surface and the other of air at ambient conditions from the inlet. From Fig. 13, where the heat flux from the flame to the solid and the surface

(23)

?~y
pD t3Y = m Yo.

(24)

~y For thermally thick fuels, the heat flux balance is expressed as:
_ k o T = _ks &Ts + m Q

(25)

~y

Oy

where Q is the heat of pyrolysis and ks the solid thermal conductivity, At each instant at the solid-gas interface, the solid temperature must be equal to the gas temperature: T = Ts.

Combustion processes of solid fuels

95

f .730 I~m .610 "~ ~ .490 .370 250

' t__= _~ _ _-- ~ .~ _ z z - - .850 . . . . . . _ _" _" ." . "[3 .730 -- -- .61~1 ~. ~- _ - _ ~ --- : . 4 ~ ~ ' ~ ~
-

.37 "a2~5 L=n,i


125

" - -- ~ / /

z ~ !

" -

.000 .000

' 200 .400 .f:~0

I , .800 1.000

x 1~] Fro. 12. Opposed flow flame spread structure and vector velocity field: (A) fuel mass fraction (solid lines): 0.0002, 0.075 and then with step 0.075, and oxygen mass fraction (dashed lines) from 0 with step 0.033 and (B) gas phase temperature (K) from 300 with step 250 and solid phase temperature (K) from 300 with step 50; maximum inlet velocity 35 cm s-1.

and the flame. The opposed velocity goes to zero on the surface, facilitating this fuel transport upstream. In addition, this process is favored by the very steep gradients of fuel mass fraction in this region, while it does not appear to depend on the surface temperature gradient. This mechanism of flame spread has also been predicted by early numerical analyses. 114,115 An important effect is the existence of a region of elevated pressure upstream of the flame leading edge, caused by gas expansion. It causes the outward deflectionoftheflowneartheflamefrontandareduction of the gas velocity encountered by the flame at its leading edge. Studies of flame stabilization at the leading edge of a fuel plate have contributed to the understanding of chemical and physical mechanisms controlling flame spread. The paper by Mao et al. 132 presents the first detailed description of the convective and thermal structure of the flame leading edge over a solid combustible surrounded by inert material. The elimination of inert material allowed blowoff phenomena to be predicted by Chen and T'ien 133 as the velocity of the oxidizing flow opposing the flame is increased. The transition from envelope diffusion flames (high Damk6hler numbers) to open-tip flames (critical Damk6hler number) up to blowoff is also predicted. Comparison between theoretical and experimental extinction limits is presented by Kodama et al., 134 as opposed velocity and oxygen concentrations are varied. Flame stabilization mechanisms at the leading edge o f a fuel plate, under free convection, as gravity level and oxygen concentration are varied, are presented in Ref. [135]. At very low gravity, the flame extinguishes while a blowoff limit is predicted at very high gravity levels with the flame changing from envelope to open-tip configuration. A blowofflimit is also determined at very low oxygen concentrations. Further investigation on the effects of gravity level on flame spread over thin cellulosic materials is presented in Ref. [128]. The first computer model of flame spread is a quasi-steady, variable spread rate model, based on the thermo-diffusive equations. It4 It has been used to analyze the dependence of spread on gas phase parameters 1~4 and solid temperature for thermally thin fuels. Unsteady thermo-diffusive models have also been used to analyze flame spread characteristics over thermally thick fuels. ~t6-tts Particularly interesting is the prediction of the dependence of the flame spread rate on the velocity and oxygen concentration of the opposed gas flow. The spread rate for thick PMMA, as measured x45 and predicted by a thermo-diffusive model, lxa are reported in Fig. 14 as a function of the opposed-flow velocity and for several values of the oxygen mass fractions. The analysis, which qualitatively describes the experimental results, predicts a spread rate that varies with the flow velocity for fixed oxygen concentration. This spread rate first increases and then decreases as the flow velocity increases. For a given flow velocity, the

2.2 2.0 ~ 1.8 ~" 1.6 ~ 1.4 1.2 o


1.0
.8

~ l

'~t qs
:'

.0050 .0045
.0040

, ~' /
,

.0035 ~ t'M .oo3o 'E


.0025 ~
o
.0020 u')

'

'

.0015

cr

.6 ,E .4 . . . . - " ~ , .0010 .2 , .oo05 .0 I I I I .0000 .000 200 .400 .600 .800 1.000 x [era]

FIG. 13. Heat flux and surface pyrolysis mass flux as functions of the x coordinate. pyrolysis mass flux as functions of the specimen length are reported, it is worth observing that the pyrolysis mass flux, proportional to the normal component of velocity at the interface, reaches the highest values behind the flame leading edge. Consequently, the formation of the upstream combustible mixture comes from fuel diffusion from the flame zone through a quenching layer between the solid surface
d1~$ 19:I'~

96

C. DI BLAS1 the momentum equations, formulated according to the Boussinesq approximation, have been made 123 in order to analyze extinction caused by high flow velocities or low oxygen concentrations. Although the mechanisms which lead to extinction are very different for the two cases, extinction can always be related to a continuous decrease of the ratio of the flow time O(/U2 to the chemical time pg/wr, that is, the Damkrhler number. In both extinction cases simulated, the flame leading edge recedes with respect to the heated solid fuel, thus the solid heat conduction is not controlling at near-extinction conditions. The interactions between fluid dynamics and chemical processes in the gas phase, which are much faster than in the solid phase, are responsible for extinction. In the former case, when the flow velocity is increased, the slight decrease in the chemical time, due to the increase in the fuel concentration, is much lower than that in the flow time. Thus the Damkrhler number at the flame leading edge continuously decreases, leading to extinction. In the latter case, the flow time is unchanged, but the strong increase in the chemical time, due to the decrease of the reaction rate, again gives decreasing values of the Damk6hler number and extinction occurs. In conclusion, if the chemical time is much smaller than the flow time, the spread of the flame will be controlled by the rate of fuel gasification, that is by the heat transfer from the flame to the fuel. On the other hand, if the chemical time is of the same order as the flow time, the spread of the flame will be controlled by the rate at which the fuel can be consumed, that is by chemical kinetics. More recent simulations of flame spread, including momentum transfer, are based on the assumption of quasi-steady, constant spread-rate equations. The effects of the gas velocity on the opposed flow flame spread over thin paper have been presented in Ref. 1-126].In qualitative agreement with experiments, the flame spread rate decreases astheopposed ftowvelocity is increased. The equations do not account for buoyancy terms, thus the region of constant spread rates, observed in the experiments ~45 at low opposed velocities, is not predicted. The extinction limit is associated with small Damkrhler numbers. The more interesting result of the analysis is that solid phase heat conduction, which does not play any role at high Damk6hler numbers, becomes dominant near the blowoff limit. Flow recirculation is predicted ahead of the flame leading edge, caused by thermal expansion of hot product gases, producing high pressure inside the flame. The model is not able to identify the quenching limit in the slow opposed flow regime, observed under microgravity conditions. The downstream spread over thin paper as the gravity level is varied is presented in Ref. [127]. The same qualitative behavior as for the case of forced flow conditions is predicted. Even in this case the model is not capable of predicting an extinction limit at low gravity level. The only significant difference

F~ 0.3 \ z U
> 0.e4

oo O n A
, ~33

. 329 .23o

~ . aao A~~. E6 ( f f w O O

0.18

~ ~ 1 1 0~2

0.0E

O j Mn l l ~
i i itll

O O ,,L ~0 3
i

k I I

~0

,,.~ ~ , ,,~11~ to1 t?

I II~l

UMAX [CM/S]

FIG. 14. Predicted I ,s (full symbols) and measured 145 (empty symbols) spread rates for opposed flow flame spread over PMMA. spread rate increases with the oxygen concentration and the location of the maximum is also displaced towards the higher-velocity region as the oxygen concentration increases. The major quantitative discrepancies between theory and experiment occur in the very low and very large velocity regions where the approximations of the thermo-diffusive model are not valid. Indeed, at low flow velocities, the phenomenon is controlled by buoyancy, not accounted for in the model, while at high flow velocities, the flame structure is very sensitive to the gas velocity profile, not well described in the model. For intermediate gas flow velocities, the flame spread rate is controlled by the transfer of heat from the flame to the solid, which is less sensitive to the relatively small variations of the velocity profile. Thus, in this range of velocities, the analysis of the spread of the flame is quantitatively better, The results of numerical analyses 116-x18 have also been employed to determine the controlling mechanisms of flame-spread over thick solids and thus to verify the phenomenological arguments used to explain the experimental results in terms of thermally and chemically controlled mechanisms. Both the maximum surface heat flux and maximum surface temperature increase with the oxygen concentration and the flow velocity. The increase with the oxygen concentration is due primarily to the increase in the flame temperature. The increase with the flow velocity is due to the flame moving closer to the surface, Both of these effects increase the heat transfer from the flame to the fuel and the gasification rate of the solid combustible. Consequently they would tend to increase the spread rate for all values of flow velocity. Detailed simulations of the flame-spread process over thick solids by a more advanced model including

Combustion processes of solid fuels


.750

97

t=0.05s

.25O oOOO .750 . .400 .800 1.200 1.(500 2.000

.25O
.000 .750 = o .400 .aO0 1.200 :t..00O 2.000

.000 .750

.400

1.200

],.(~30

2.000

.OOO

.400

.OO0

=,.200

1.000

2.000

I~1

FIG. 15. Dynamics of the gas phase temperature (K) fields for flow assisted flame spread, from 300 with step 250 (maximum concurrent velocity 35 cm s-1). with respect to the case of forced flow is that the flame, in the natural-convective environment, is weaker (lower spread rates) and no recirculation ahead of the flame leading edge is predicted. Inclusion of surface radiation in the mathematical model ~2s allows an extinction limit to be predicted at low gravity levels, while radiative processes are unimportant at high opposed flow velocities, 3.4. Flow Assisted Flame Spread Only thermo-diffusive models have been used to simulate flow-assisted flame spread over thick ~~9.~2o and thint21 fuels. The dynamics of the early stage of concurrent flame spread, over thick solids, are shown in Figs 15 and 16 through the gas phase temperature and fuel and oxygen mass fraction distributions. Initially, the solid is at ambient conditions and ignition is caused by an external radiative heat source, As time increases, a rapid propagation of the flame front followed by a slower advancement of the pyrolysis front is observed. The gas phase isotherms and contours of equal species concentrations show that the flame, except at its upstream leading edge and tip, has a diffusional character with a relatively large reaction rate. Also, the characteristic shape associated with a diffusion flame over a pyrolyzing surface in a concurrent environment is shown: the hot gases (reacting species and combustion products) are pushed downstream in the direction of the propagating pyrolysis front. Three regions exist in the integration domain: the vaporization (or pyrolysis) region, the 'combusting plume', and the 'thermal plume'. The pyrolysis region is characterized by surface temperatures larger than 600 K. In this region, an almost constant value of the surface temperature is reached and maxima of gas temperature and reaction rate are computed. At the flame leading edge there is a

98
.7~O
t=O.OSS

C. DI BLASI

.OOO .750 . . .

.400 .

.,

.800

"i200

1.600

2.000

t=O.O75s

.soo

-.

_------~

-'>."-

, ,,,,
.000 .750
t--O.ls

.400

.O00

1.200

:1.,600

2.000

7
\ ( . ~ _ ------. _-~ ---<.....,. /

-.---..,,,,
~50 .750.000 , I .400 ,% % 1 % % % , % % % ~ I ~, -800 1,200 I, [ fd'/~f_aL,(~00 . 2.000

[o.]

FIG. 16. Dynamics of fuel (solid lines) and oxygen (dashed lines) mass fraction distributions for flow assisted flame spread, from 0 with step 8.9 x 10-2 (fuel) and from 0 with step 2.3 x 10 2 (oxygen); maximum concurrent velocity 35 cm s-L premixed zone similar to that shown in opposed-flow flame spread, where the flame is thick. In the combusting plume region, the surface pyrolysis rate is negligible. The fuel vapors, however, are pushed beyond the vaporization region by the external flow rate, and burn, giving high gas temperatures. The consequent high values of the heat flux at the surface heat the solid fuel which also reaches rather high temperatures. In the final part of the combusting plume region, due to the proximity of the flame to the surface and to the relatively low surface temperatures, the reaction rate assumes low values, indicating that the flame sheet approximation is not valid here. The last zone at the left of the flame tip is the so-called thermal plume. Because of the absence of fuel vapors in this zone, the reaction rate is zero. The species present are essentially the oxidizer and the combustion products at a fairly high temperature, The predicted flame and pyrolysis spread rates, 119'12 in general agreement with experiments,146 increase with oxygen concentration and flow velocity and vary over a very large range of values. Both the increase of oxygen concentration and concurrent flow velocity increase the heat flux from the flame to the fuel and the spread rate is enhanced. This indicates that, unlike the opposedflow case, flow-assisted flame spread is controlled essentially by heat transfer. The effects due to finiterate kinetics are of increasing importance as extinction is approached, as a consequence of low oxygen concentrations or very large flow velocities. These effects appear mainly at the flame leading edge. Here the extinction length, that is, the distance of the flame leading edge from the edge of the fuel slab, increases as the flow velocity is increased or the oxygen concentration is lowered. Again, the extinc-

Combustion processes of solid fuels tion process can be explained in terms of a decrease, below a critical value, of the Damk6hler number at the flame leading edge. The flame counteracts the decrease by moving into a zone of high fuel concentration. The extinction process begins in this way, but the downstream spread process goes on increasing flame and pyrolysis lengths as long as the pyrolysis spread rate is greater than the upstream extinction rate. Complete extinction occurs when the extinction distance extends to the position of the pyrolysis front. 12 The characteristics of flame structure predicted for thin fuels T M are qualitatively similar to those computed for the case of thermally thick fuels, except for the presence of a burn-out region. The predicted T M and measured 136 variation with time of the flame, pyrolysis, and burn-out distances to the location of the flame spread initiation under several flow conditions are in good agreement. The spreading process initially accelerates and then tends towards a steady state as the flame spreads over the solid surface,

99

equations, are all based on the formulation of quasisteady, constant spread rate equations and also consider thermally thin fuels. 113.124-126 Also, simulations of flame stabilization at the leading edge of a fuel plate 13s present conditions of interest to microgravity studies. By lowering the gravity level, the flame size increases with a decrease of the heat flux to the solid. Different possible explanations of the weakness of the flame at low gravity level were given. The first was a reduction in the fuel supply rate below a limit value, so that the flame cannot be sustained. The second suggests a reduction in the oxidizer supply rate to the flame, due to the absence of an external driving force: the rate of oxidizer depletion is larger than the feeding rate and the flame extinguishes. Also, heat radiation losses are supposed to plan an important role. In Ref. [124], Bhattacharjee et al. used both finite and infinite rate kinetics to model flame spread over a thermally thin fuel in a quiescent environment at zero gravity. The flame structure is similar to that predicted for opposed flow flame spread. For the case of infinite kinetics, the premixed zone disappears 3.5. Microgravity Flame Spread and higher temperatures are observed at the flame leading edge. The increased gas temperature leads to Studies of flame spread under microgravity condi- higher heat fluxes and consequently to larger spread tions have been greatly increased in the last few years rates. A comparison between predicted and measured because of the revived interest in fire-safety concerns spread rates is also presented. Even for the case of in spacecraft and for the space shuttle. In a micrograv- finite gas phase kinetics, spread rates two or three ity environment, in the absence of forced flow, the times larger than the measured values are predicted. velocity distribution is mainly determined by the fuel Furthermore, in disagreement with experiments, only vaporization at the surface, possible pressure gradi- a weak increase of the spread rate with pressure is ents and flow induced by the flame motion with observed. Simulations of flamespread over thin paper respect to the solid and the quiescent environment, for very low opposed velocities126 and in a quiescent Among the experimental studies on microgravity environment, 127 in the absence of or at very low flame spread, those conducted by Olsen et al.,93'94 gravity, show the common feature of being incapable where the flame propagation over a thin paper sheet of predicting the quenching limit observed in the under quiescent and slow forced flow conditions is experiments. The reason for such disagreement, as investigated, are very interesting. As a result of the well as for those in Ref. [124], was attributed to the analyses, besides flame structure, an extinction limit neglect of radiation in the gas phase and the solid at very low opposed flow velocities was determined, phase equations. Very low propagation rates and flame shapes, differThe model in Ref. [124] was modified to include the ent from those at normal gravity, with larger stand- effects of radiative heat transfer from the fuel suroff distances, were observed, face. 125 Spread rates are predicted as a function of the Early thermo-diffusive models of flame spread do solid fuel emittance and oxygen level. For any oxygen not include momentum equations and buoyancy level, the spread rate is found to decrease as the solid terms, consequently, in principle, they can be consid- emittance is increased, with a higher rate of decrease ered as microgravity models. Almost all simulations at lower oxygen levels and higher values of the emithave been performed for conditions dominated by tance. Changes, as the surface emittance increases, are the assigned forced flow, however, simulations of also shown in the flame structure. Indeed, the gas flame spread over thermally thick P M M A at low phase temperature decreases and the flame shrinks in opposed flow velocities, as the oxygen level is varied, size and is located closer to the fuel surface. Near-limit have been presented in Ref. [l 18]. The unsteady flame spread is believed to be controlled by the ratio mathematical model is able to predict an extinction of radiation loss to conductive heat flux to the surface. limit at very low opposed velocities and oxygen The extinction is still due to chemical kinetics but concentrations above ambient conditions (see Fig. reaction rate slows down through temperature reduc14). Also, a reverse U-shape curve of the spread rate tion as a result of radiative heat losses. In other words, as a function of the opposed flow velocity is pre- the extinction is caused by a strong decrease in the dicted, surface temperature with the consequent reduction of More recent studies, including momentum balance the fuel supply rate which attains very low values.

100

C. D! BLASl is made more complex by gas phase turbulence. Detailed modeling of combustion processes of charring and non-charring solid fuels is a formidable task and models available to date only partially describe the problem. A one-step global reaction is the most widely suggested kinetic model for the thermal degradation of both charring and non-charring solid fuels. It is the simplest but offers only a very rough description of the problem with reaction products (usually volatiles and chars) accounting both for primary and secondary chemical processes. This approach might suffice for fire safety issues, however it is certainly not appropriate when the study of solid degradation is aimed at obtaining information on the yields of cornbustible components evolved from reaction products. Multi-step, one-stage models have made a significant contribution to the understanding of the elemental composition of the reaction products from cellulosic material pyrolysis. However, their validity is limited to the range of experimental conditions under which they have been determined, and usually their role has been that of correlating the experimental data. The chemistry of the depolymerization process has also been studied for the case of non-charring materials. Notwithstanding the noticeable effort undertaken in this direction, due to complexity of the chemical mechanisms, it is unlikely that in a short time such a type of models can be coupled to solid and/or gas phase transport phenomena to have great insight into solid fuelcombustion. Two-stage, semi-global models for thermal degradation of charring fuels are based on lumping the reaction products into three main components: gases, tars and chars. As both primary and secondary reactions are modeled, they are, in some way, the most useful and advanced models to be used together with the description of physical processes for predicting product yields from solid fuel thermal degradation. However, despite the limited number of chemical species used, there is large uncertainty in the kinetic schemes, mainly in the kinetic data. For the primary stage of wood (or biomass) degradation, a different evolution of tar, char and gas yields, as functions of temperature, has been observed. Thus, on a very general basis, in the group of two-stage, semi-global models, the three-reaction model accounting separately for gas, tar and char formation from the complex chemistry of wood could be the most general. However, few estimations of kinetic data are available and their validity is strongly impugned by the very narrow range of temperature variation considered in the experiments and by the assumption of isothermal conditions, no longer verified during the heating period, and by a final char yield not depending on temperature. Another drawback in the estimation of kinetic parameters is that the data are determined and are valid only for the specific wood considered. The alternative approach, based on the contribution of the reaction rate of cellulose, hemicellulose and lignin

A further improvement is presented in Ref. [113] where gas phase radiation effects have been included in the modeling of flame spread over thermally thin fuels at low gravity. Radiation is coupled to hydrodynamics using the emission approximation through three simple parameters: a Planck mean absorption coefficient, the fraction of total emission directed towards the surface and a shape function that represents the distribution of gas-to-surface radiation. In order to study the effects of radiation on the predictions of flame spread, a set of simulations, made in a 50Y/ooxidizer ambient, with different models: (a) no heat radiation, (b) surface heat radiation losses only, (c) gas phase heat radiation losses only and (d) gas phase and solid phase heat radiation losses and gas to surface radiative heat flux, has been performed. At high opposed velocities, a decrease of the spread rate is observed for the cases (a), (b) and (c), indicating that gas phase kinetics are the controlling factor, At low opposing velocities, the flame becomes controlled by radiation. It appears that only models including radiation effects 113"128"131 can predict the increase of the spread rate as the opposed velocity is increased from zero to a limit value with the inverted 'U' shape of the spread rate curve as a function of the opposed velocity observed in the experiments. A parametric study of the coupled effects of surface emission and absorption of radiation is also available, t13 Finally, the model,~ 13 with only gas phase radiation heat losses, has been used 129 to predict the structure of the flame leading edge which compares favorably with that experimentally observed,

4. CONCLUSIONSAND FURTHER DEVELOPMENTS Combustion of solid fuels is the result of complex interactions among many chemical and physical processes. For the solid phase, these include heat and mass transfer, chemistry of homogeneous and heterogeneous reactions interior to the solid (thermal degradation and oxidation) and surface regression, Furthermore, for PMMA, internal absorption of radiation and formation of bubbles in the melt layer are also to be accounted for, while, for celluloic charring materials, evaporation of absorbed water, pressure variations, cracks caused by thermal stresses in the char layer, internal shrinkage and/or swelling, and migration of volatiles in the virgin solid region are other important processes to be considered. For the gas phase, processes include convection and diffusion of mass, momentum and energy, combustion reactions, ignition and flame propagation, radiation absorption by the degradation products, and heat transfer by radiation. Processes occurring at the solid-gas interface include convection and diffusion of mass, momentum and energy, reradiation from the surface, and heterogeneous reactions between the gas phase and the solid phase. In some situations, especially when the scale length is large, the problem

Combustion processes of solid fuels to the primary reaction rate of the wood, seems to be more general. Once the amounts of components participating in the composition of the specific wood are determined, the degradation rate of every type of wood may, in principle, be determined. The role played by possible interactions among the wood cornponents as well as possible bonds among them, however, should be investigated, Extensive studies on cellulose have been conducted and there is some agreement about a semi-global primary pyrolysis mechanism which considers two competitive pathways leading, respectively, to tar formation and to a linked char and gas formation, Kinetic data for such a mechanism have also been estimated. A tentative, two-stage, semiglobal model has been proposed for lignin pyrolysis, but kinetic constants are not available. Less information is available on hemicellulose. Therefore the approach based on the contribution of the single components to model wood or biomass degradation rate also requires further study on the kinetics of each component. Further investigation is needed in the analysis of secondary reactions. Almost all authors agree that tar degradation to gases and chars occurs. Several estimations of kinetic data for tar cracking to light gases, considered only as preliminary, have been made, while no quantitative measurement of the secondary reaction rate of tar deposition to char has ever been attempted. The role of heterogeneous reactions of char gasification, due to volatile products of primary pyrolysis, should also be studied, Models describing both chemical and physical processes of charring material degradation and accounting for transport phenomena generally use a very simplified description of chemistry (one-step global pyrolysis reaction), whereas models accounting for pyrolysis through two-stage, semi-global models, often do not describe other important aspects of the phenomenon such as property variations, convective transport of gaseous species or momentum transfer, In the last few years, models accounting for both the main chemical and physical processes have been proposed. They have proved capable of predicting the proper trend of product yields to the changes in the reaction conditions (heating, solid and gas residence times), Quantitative agreement between predictions and experiments is believed to be very difficult to achieve due to the lack of reliable data and the dependence of the data on operating conditions and type of wood. Under ablation regime conditions, for very large samples subjected to high radiative heat fluxes, structural changes may be important. In this case, even the qualitative agreement with experiments may be lost because no model accounts for volume variation effects. Mathematical models have not yet been widely used to perform systematic studies of the influences of the heating level, particle size, moisture content, etc., which are useful to evaluate the effects

101

of the feed heterogeneity on the pyrolysis of biomass fuels. Such results could be useful to suggest possible pretreatments to improve process control and product quality. The analysis of the external conditions, in particular of the heating level, could also be of great importance to fire safety science. As for non-charring materials, only one-step kinetic models have been coupled with physical modeling, which in most cases is only a simple heat conduction equation. Indeed, apart from the analytical work in Ref. [90], no effort has been made to bring together chemical processes, heat transfer and bubble formation effects. Ignition simulations have been made mainly for non-charring fuels by means of one-dimensional models. Flame spread has been simulated through two-dimensional models accounting for momentum, mass and heat transfer. In all cases, gas phase combustion and solid phase pyrolysis processes have been described according to one-step global reactions. The predictions of ignition delay times and spread rates are in qualitative agreement with experiments under almost all conditions, and, in some cases, in quantitative agreement, indicating that mathematical models could be a powerful tool for the analysis of controlling mechanisms, provided all main chemical and physical processes are taken into account even if with some simplifications. An improved description of the chemistry can of course give more information on the details of ignition and extinction and the emission of gaseous species (mainly CO and CO2) useful for toxicity evaluations. Among the physical processes, deserving an improved treatment, there is radiation transport which is believed to play a controlling role for microgravity flame spread. Solid phase degradation is very important not only for conversion of biomass fuels to energy, but it is also the first step in the complex phenomena of gas phase combustion. However, while it is well known that under ablation regime conditions physical processes are controlling and that the details of the chemistry of solid phase degradation are very important for wood and biomass conversion processes, it is not yet sufficiently clear how these aspects of solid phase process can affect gas phase flaming combustion. The coupling of solid and gas phase processes occurs through the boundary conditions at the interface, that is through the solid surface temperature, the blowing velocity due to the devolatilization process and the chemical composition of the mixture adjacent to the solid surface. Chemical and physical processes occurring inside the solid are, to a significant extent, responsible for the surface conditions and composition of the evolved gases. Particularly interesting is a pioneering work by Martin t47 where the implications of the details of the temperature evolution along cellulose samples radiatively heated and the transients and the chemical composition of

102

C. DI BLASI solid degradation, were available. Once the submodels were validated, they could be included in a general, comprehensive model. The complexity of the complete problem and the more detailed descriptions of physical processes require the use of advanced numerical techniques and high-speed computers, lmprovement in quantitative predictions, however, do not depend only on the availability of comprehensive models, efficient numerical solution techniques and high-performance machines, but also on the determination of reliable data from experimental investigations.
Acknowledgements--The work was supported by the Euro-

primary and secondary pyrolysis products on the gas phase ignition processes are presented. For a wide range of radiative heat fluxes, the surface is always well-charred before ignition. It was observed to be a site for secondary reactions (tar degradation and char gasification) which generate a low, but abruptly increasing, concentration of reactive substances which lead to ignition. This seems to indicate that secondary reactions play an important role not only for the yields of products from biomass thermal degradation but also for the flammability characteristics of charring materials, Formation of char make the solid phase processes highly transient and affects heat transfer processes with changes in the amount of combustible volatile species generated. The assumption that the pyrolysis is a surface process, made in the models of flame spread over thick charring and non charring materials, is not confirmed by experimental measurements which show that the mass loss rate still increases even though a constant surface temperature is reached. For thin charring fuels, exothermic char oxidation reactions are supposed to affect flame spread at microgravity conditions. 94'99A29 Therefore, among the improvements to be made in solid phase models used in conjunction with gas phase models to predict flame spread, the effects of surface oxidative reactions, for thin fuels under microgravity conditions, and in-depth pyrolysis and char formation, for thermally thick fuels, should be included. In this case, bubble formation might be another important process because it plays a role on the transport of degradation products towards the surface and finally on the amount and composition of combustible gases. Finally, more accurate models of evolved combustible products from solid degradation have to be used in order that the detailed kinetic models, available for hydrocarbon fuel oxidation, can be used in the modeling of flames over solid surfaces. As a concluding remark, it can be observed that, on the basis of the present information, the details of solid phase processes are very important for gas phase flaming combustion processes mainly at nearlimit conditions, but further studies are required to assess more clearly their role. Despite the noticeable efforts undertaken, the chemistry of solid degradation and gas phase combustion is still not yet well understood. The large uncertainty existing on the kinetic data of global and semi-global models does not in general allow quantitative predictions to be made. However, detailed numerical modeling has proved a useful tool for determining the controlling mechanisms of combustion processes of charring and noncharring solid fuels. In the long term, there would be great progress in modeling of solid phase thermal degradation and gas phase ignition and flame spread phenomena if a more detailed description of the different processes, especially of chemical kinetics for combustion and

pean Economic Community under Contract Joule No. 0035 and by the National Research Council of Italy under Contfibutions N.90/4100 and N.91/68.

REFERENCES 1. SHAFIZADEH, In: The Chemistry of Solid Wood, R. F., Rowell (Ed.), Advances in Chemistry Series 207, American Chemical Society, Washington, D.C. (1984). 2. SHAFIZADEH, In: Wood andAgricultural Residues, J. F., Soltes (Ed.), Academic Press, 415, 1983. 3. BROIDO,A. and NELSON,M. A., Combust. Flame 24, 263, (1975). 4. SHIVADEV,U. K. and EMMONS,H. W., Combust. Flame 22, 223 (1974). 5. KANURY,M. A., Combust. Flame 18, 75 (1972). 6. KASHIWAGI,T. and OHLEMILLER,T. J., Nineteenth Symposium (International)on Combustion, p. 815, The Combustion Institute, Pittsburgh (1982). 7. KASHIWAGI, Fire Saf. J. 3, 185 (1981). T., 8. WILLIAMS,F. A., Sixteenth Symposium (International) on Combustion, p. 1281, The Combustion Institute, Pittsburgh(1976). 9. OHLEMILLER,T. J., Prog. Energy Combust. Sci. 1!, 277 (1985). 10. KANURY,M. A. and BLACKSHEAR,P. J., Combust. Sci. TechnoL 1,339 (1970). 11. ROBERTS,A. F., Combust. Flame 14, 261, (1971). 12. RAMIAH,M. V., J. App. Polym. Sci. 14, 1323 (1970). 13. AGRAWAL, K., Can. J. Chem. Eng. 66, 403 (1988). R. 14. SCOTT,D. S., PISKORZ,J., BERGOUGNOU, A., GRAHAM, M. R. and OVEREND, R. P., Ind. Eng. Chem. Res. 27, 8 (1988). 15. DIEBOLD,J. P., M.S. Thesis, Colorado School of Mines, Golden, CO (1985). 16. LEDE,J., PANAGOPOULOS, LI, H. Z. and VILLERMAUX, J., J., Fuel64, 1514 0985). 17. BRIDGE,S. A., M.S. Thesis, The University of Aston in Birmingham (1990). 18. ROBERTS, F., Thirteenth Symp. (International) on ComA. bustion, p. 893, The Combustion Institute, Pittsburgh (1971). 19. TINNEY,E. R., Tenth Symposium (International)on Cornbustion, p. 925, The Combustion Institute, Pittsburgh (1965). 20. LEE,C. K., CHAIKEN,R. F. and SINGER,J. M., Sixteenth Symposium (International) on Combustion, p. 1459, The Combustion Institute, Pittsburgh (1976). 21. BILBAO,R., ARAUZO,J. and MILLERA,A., Thermochim. Acta 120, 121 (1987). 22. EAIRBRIDGE,C., ROSS, R. A. and SOOD, S. P., J. AppL Polym. Sci., 22,497 (1978). 23. TABATABAIE-RAISSI, MOK, W. S. L. and ANTAL,M. A., J., Ind. Eng. Chem. Res. 28, 856 (1989). 24. TRAN,D. Q. and RAI, C., FuelS'l, 293 (1978).

Combustion processes of solid fuels 25. GULLET,B. K. and SMITH,P., Combust. Flame 67, 143 (1987). 26. ANTAL, M. J., FRIEDMAN, H. L. and ROGERS, F . E . , Combust. Sci. Technol. 21,141 (1980). 27. BAROOAn,J. N. and LONG,V. D., Fuel 55, I 16 (1976). 28. LEWELLEN,P. C., PETERS,W. A. and HOWARD, J.B., 16th Symposium (International) on Combustion, p. 1 4 7 1 , The Combustion Institute, Pittsburgh (1976). 29. MIN, K., Combust. Flame 30, 285 (1977). 30. NOLAN,P. F., BROWN,D. J. and ROTHWELL,E., Fourteenth Symposium (International) on Combustion, p. 1143, The Combustion Institute, Pittsburgh (1973). 31. FONT,R., MARClLLA,A., VERDU,E. and DEVESA,J., Ind. Eng. Chem. Res. 29, 1846 (1990). 32. KANURY,A. M. and BLACKSHEAR,P. L., Pyrodynamics 4, 285 (1966). 33. NUNN, T. R., HOWARD, J. B., LONGWELL,J. P. and PETERS, W. A., Ind. Eng. Chem. Process Des. Dev. 24, 836 0985). 34. SAMOLADA,M. C. and VASALOS, I. A., Fuel 70, 883 (1991 ). 35. HA.IALIGOL, R., HOWARD,J. B., LONGWELL,J. P. and M. PETERS, W. A., Ind. Eng. Chem. Process Des. Dev. 21, 457 (1982). 36. FUNAZUKURL,T., HUDGINS,R. R. and SILVERSTON,P. L., Ind. Eng. Chem. Process Des. Dev. 25, 172 (1986). 37. NUNN, T. R., HOWARD, J. B., LONGWELL,J. P. and PETERS, W. A., Ind. Eng. Chem. Process Des. Dev. 24, 844 (1985). 38. AVNI, E., COUGLIN,R. W., SOLOMON,P. R. and KING, H. H., Fuel64, 1495 (1985). 39. JEGERS, H. E. and KLEIN, M., Ind. Eng. Chem. Proc. Dev. 24, 173 (1985). 40. IATRIDIS,B. and GAVALAS,G. R., Ind. Eng. Chem. Prod. Res. Dev. 18, 127 (1979). 41. KILZER, F. J. and BROIDO, A., Pyrodynamics 2, 151 (1965). 42. ARSENEAU, F., Can. J. Chem. 49, 632 (1971). D. 43. BRADBURY, G. W., SAKAI,V. and SHAEIZADEH, J. A. F., AppL Polym. Sci. 23, 3271 (1979). 44. ANTAL, m. J., In: Advances in Solar Energy, Boer, K. W. and Duffle, J. A. (Eds), p. 61, American Solar Energy Society, Boulder, CO (1982). 45. ANTAL,M. J., In: Advances in Solar Energy, Boer, K.W. and Duffle, J. A. (Eds), p. 175, American Solar Energy Society, NY (1985). 46. LIPSKA,A. E. and PARKER,W. J., 3'. AppL Polym. ScL 10, 1439 (1966). 47. AGRAWAL,R. K., Can. J. Chem. Eng. 66, 413 (1988). 48. KOUFOPANOS,C. A., MASCHIO, G. and LUCCHESl,A., Can. J. Chem. Eng. 67, 75 (1989). 49. ALVES, S. S. and FIGUEIREDO,J. L., J. Analyt. AppL Pyrolysis 13, 123 (1988). 50. WARD,S. W. and BRASLAW,J., Combust. Flame 61,261 (1985). 51. VOVELLE,C., MELLOTTE,H. and DELBURGO,R., Nineteenth Symposium (International)on Combustion, p. 7 9 7 , The Combustion Institute, Pittsburgh (I 982). 52. THURNER,F. and MANN, U., Ind. Eng. Chem. Process Des. Dev. 20, 482, (1981). 53. BOROSON,M. L., HOWARD,J. B., LONGWELL,J. P. and PETERS,A. W., AICHEJ. 35, 120 (1989). 54. KOSSlTRIN,H., Proc. Spec. Workshop on Fast Pyrolysis ofBiomass, p. 105, Copper Mountain, CO (1980). 55. LIDEN,A. G., BERRUTI,F. and ScoTr, D. S., Chem. Eng. Commun. 65, 207 (1988). 56. KOUFOPANOS, A., PAPAYANNAKOS,N., MASCHIO,G. C. and LUCCHESI,A., Can. J. Chem. Eng. 69, 907 (1991). 57. GORTON, C. W. and KNIGHT, J. A., Biotech. Bioeng. Syrup. 14, 14 (1984). 58. SHAF1ZADEH, and CHIN,P. P. S., ACS Syrup. Set. 43, F. 570977). 59. ANTAL,M. J., Ind. Eng. Prod. Res. Dev. 22, 366 (1983).

103

60. ROBERTS,A. F., Combust. Flame 17, 79 (1971). 61. CHAN,W. R., KELBON,M. and KRIEGER-BROCKETT, B., Ind. Eng. Chem. Res. 27, 2261 (1988). 62. OHLEMILLER,T. J., KASHIWAGI,T. and WERNER, K., Combust. Flame 69, 155 (1987). 63. KASHIWA61,T., OHLEM1LLER,T. J. and WERNER, K., Combust. Flame 69, 331 (1987). 64. CHAN,W. R., KELBON,M. and KRIEGER,B. B., Fuel64, 1505 (1985). 65. COLLINS,R. E., Reinhold Publ. Corp., New York (1961). 66. KFNG, H., Combust. Flame 18, 185(1972). 67. KANSA,E. J., PERLEE,H. E. and CHAIKEN,R. F., Combust. Flame 29, 311 (1977). 68. VILLERMAUX, ANTOINE,B., LEDE,J. and SOULIGNAC, J., F., Chem. Eng. Sci. 41,151 (1986). 69. MILLER, C. A. and RAMOHALLI,N. R., Combust. ScL Technol. 46, 249 0986). 70. WICHMAN, I. S. and ATREYA,A., Combust. Flame 68, 231 (1987). 71. PANTON,R. L. and RITTMANN,J. G., Thirteenth Symposlum (In ternational) on Combustion, p. 88 l, The Combustion Institute, Pittsburgh 0971). 72. CURTIS,L. J. and MILLER, D. J., Ind. Eng. Chem. Res. 27, 1775 (1988). 73. HASTAOGLU,M. A. and BERRUTI, F., Fuel 68, 1408 (1989). 74. Dt BLASI,C., CRESCITELLI,S., RUSSO,G. and MAGLIONE, A., In: Phase Change and Combustion Simulation, Advanced Computational Methods in Heat Transfer, VoL 3, Wrobel, L. C., Brebbia, C. A. and Nowak, A. J. (Eds), Comp. Mech. Publ. and Springer, 209 (1990). 75. ALVES,S. S. and FIGUEIREDO,J. L., Chem. Eng. Sci. 44, 2861 (1989). 76. HOTTEL,H. C. and SAROEIM,A. F., Radiative Transfer, McGraw-Hill, New York (1967). 77. D1 BLASt,C., Combust. ScL TechnoL, 90, 315 (1993). 78. D! BLASt, C. and Russo, G., Proc. of the International Conference on Advances in Thermochemical Biomass Conversion, Bridgwater, A. V. (Ed.), Elsevier, Amsterdam, in press (1992). 79. KHALTURINSKII, A. and BERLIN,AI. AI., In: DegradaN. tion and Stabilization of polymers, A Series of Comprehensive Reviews, 1, Jellinek, H. H. G. (Ed.), Elsevier, Amsterdam 0983). 80. KASHIWAGI,T., OMORI, A. and NANBU, H., Combust. Flame 81,188 (1990). 81. LENGELLE,G., AIAA J. g, 1989 (1970). 82. KRISnNAMURTHY,L. and WILLIAMS, F. W., Combust. Flame 20, 163 (1973). 83. VOVELLE,C., DELFAU,J. C., REUILLON,M., BRANSlER,J. and LARAQUI,N., Combust. Sci. TechnoL 53, 187 (1987). 84. KASHIWAGI,T., HmATA, T. and BROWN, J. E., Macromolecules 18, 131 (1985). 85. HIRATA,T., KASHIWAGI,T. and BROWN, J. E., Macromolecules 18, 1410 (1985). 86. KASHIWAGI,T., INABA,A., BROWN,J. E., HATADA,K., KITAYAMA, and MASUDA, E., Macromolecules 19, T. 2160 (1986). 87. INABA,A. and KASHIWAGI,T., Macromolecules 19, 2412 (1986). 88. INABA,A. and KASH1WAGI,T., Eur. Polym. d. 23, 871 (1987). 89. MARTIN,J. W., DICKENS,B., WAKSMAN,D. BENTZ,D. P., BYRD, W. E., EMBREE,E. and ROBERTS,W. E., J. Appl. Polym. Sci. 34, 377 (1987). 90. WICHMAN,I. S., Combust. Flame 63, 217 (1986). 91. TZENG,L. S., ATREYA,A. and WICHMAN,I. S., Combust. Flame 80, 94 (1990). 92. FERNANDEZ-PELLO, C. and HIRANO,T., Combust. ScL A. TechnoL 32, 1 (1983). 93. OLSON, S. L., FERKUL, P. V. and T'IEN, J. S., TwentySecond Symposium (International) on Combustion, p. 1213, The Combustion Institute, Pittsburgh (1988).

104

C. DI BLASI

94. OLSON, S.L.,Combust. Sci. Technol. 76, 233 (1991).

95. SESHADR1,K. and WILLIAMS,F. A., J. Polym. Sci. 16,


1755 (1978). 96. P1TZ, W. J., BROWN,N. J. and SAWYER,R. F., Eighteenth Symposium (International) on Combustion, p. 1871, The Combustion Institute, Pittsburgh(1981). 97. WESTBROOK,C. K. and PITZ, W. J., In: AIAA progress in aeronautics and astronautics N. 135, Numerical Approaches to Combustion, Oran, E. S. and Boris, J.P. (Eds), p. 57 (1991). 98. AKITA,K., In: Aspects of Degradation and Stabilization o f Polymers, Jellinek, K. (Ed.), p. 500, Elsevier, New York, (1978). 99. KASHIWAGI,T. and NAMBU, H., Combust. Flame 88, 345 (1992). I00. DI BLASI, C., In: AIAA progress in aeronautics and astronautics N. 135, Numerical Approaches to Combustion, Oran, E. S. and Boris, J. P. (Eds), p. 643, (1991). 101. KINDELAN, M. and WILLIAMS, F. A., Combust. Sci. Technol. 10, 1(1975). 102. KINDELAN, M. and WILLIAMS, F. A., Combust. Sci. Technol. 16, 47 (1977). 103. KASHIWAGI, Combust. Sci. Technol. 8, 225 (1974). T., 104. AMOS,B. and FERNANDEZ-PELLO, C., Combust. Sci. A. Technol. 62, 331 (1988). 105. DI BLASI,C., CRESCITELLI,S., RUSSO,G. and CINQUE, G., Combust. Flame 83, 333 (1991). 106. BALK,S. W. and KIM, J. S., Combust. Sci. Technol. 75, 89 (1991). 107. GANDHI, P. D. and KANURY, M. A., Combust. Sci. Technol. 50, 233 0986). 108. SIRIGNANO,W. A., Combust. Sci. Technol. 6, 95 (1972). 109. FERNANDEZ-PELLO,A. C., Combust. Sci. Technol. 39, l l9 (1984). 110. WICHMAN,I. S. and AGRAWAL,S., Combust. Flame 83, 127 0991). 111. AGRAWAL,S. and WICHMAN,I. S., Combust. Sci. Technol. 81, 25 (1992). ll2. RAMSHAW,J. D., O'ROURKE,P. J. and STEIN,L. R., J. Comp. Phys. 58, 361 (1985). I 13. BHATTACHARJEE, S., ALTENKIRCH, R. A., Twenty-third Symposium (International) on Combustion, p. 1627, The Combustion Institute, Pittsburgh (1990). 114. FREY, A. E. and T'IEN, J. S., Combust. Flame 36, 263 (1979). I 15. BORGESON, A. and T'IEN,J. S., Combust. Sci. Technol. R. 32, 125 (1983). l l6. DI BLASI,C. and CONTINILLO,G., The Use of Computers in Chemical Engineering, Proc. of Chemical Engineering Fundamentals XVIII Congress, GiardiniNaxos, 261 (I 987). ll7. Dl BLASI, C., CONTINILLO, G., CRESCITELLI,S. and Russo, G., Combust. Sci. Technol. 54, 25 (1987). 118. DI BLASI, C., CRESCITELLI, S., Russo, G. and FERNANDEZ-PELLO,A. C., Proc. of the Second International Symposium on Fire Safety Science, Hemisphere, 119 (1988). 119. Dl BLASI,C., CRESCITELLI,S. and Russo, G., Comput. Meth. Appl. Mechan. Eng. 5, 481 (1989). 120. DI BLASI,C., CRESCITELLI,S. and Russo, G., Combust. Flame 72, 205 (1988). 121. DI BLASl, C., CRESCITELLI, S., RUSSO, G. and FERNANDEZ-PELLO,A. C., Twenty-Second Symposium (International) on Combustion, p. 1205, The Combustion Institute, Pittsburgh (1988).

122. DI BLASI, C., CRESCITELLI, S., Russo, G. and FERNANDEZ-PELLO, C., Combust. Sci. Technol. 64, A. 289 (1989). 123. DI BLASI,C., CRESCITELLI, and Russo, G., In: NumeriS. cal Combustion, Lecture Notes in Physics, Springer, 151, Dervieux, A., and Larroutorou, B. (Eds), 233 (1989). 124. BHATTACHARJEE, ALTENKIRCH,R. A., SRIKANTAIAH, S., N. and VEDHA-NAYAGAM, M., Combust. Sci. Technol. 69, I (1990). 125. BI4ATTACHARJEE, and ALTENKIRCH,R. A., Combust. S. Flame&l, 160 (1991). 126. CrlEN, C. H., Combust. Sci. Technol. 69, 63 (1990). 127. Dun, F. C. and CHEN, C. H., Combust. Sci. Technol. 77, 291 (1991). 128. WEST,J., BHATTACHARJEE, and ALTENK1RCH,R. A., S. Combust. Sci. Technol. 83, 233 (I 992). 129. BHATTACHARJEE, and ALTENKIRCH,R. A., TwentyS. Fourth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, in press (1992). 130. DI BLASl,C. CRESCITELLI,S. and Russo, G., Proceedings of the European Conference on Computer Application in the Chemical Industry, Dechema Monographs, Frankfurt, 357 (1989). 131. DI BLASI,C., GALOTrA,N. and Russo, G., Proc. of the 6th International Conference on "Computational Methods and Experimental Measurements', CMP and Elsevier, in press (1993). 132. MAO, C. P., KODAMA,H. and FERNANDEZ-PELLO,A. C., Combust. Flame 57, 209 (1984). 133. CHEN, C. H. and T'IEN, J. S., Combust. Sci. Technol. 283 (1986). 134. KODAMA,H., MIYASAKA,K. and FERNANDEZ-PELLO, A.C., Combust. Sci. Technol. 54, 37 (1987). 135. CrmN, C. H. and Hou, W. H., Combust. Flame 83, 309 (199 l). 136. LOll, H. T. and FERNANDEz-PELLO,A. C., Proc. of the First International Symposium on Fire Safety Science, Hemisphere, 65 (1986). 137. PATANKAR,S. V., Numerical Heat Transfer and Fluid Flow, Hemisphere, London (1980). 138. ISSA,R. I., J. Comp. Phys. 62, 40 (1985). 139. ISSA, R. I., GOSMAN, A. D. and WATKINS, A. P., 3.. Comp. Phys. 62, 66 (1986). 140. KOSDON,F. J., WILLIAMS,F. A. and BUMAN,C., Twelfth Symposium (International) on Combustion, p. 253, The Combustion Institute, Pittsburgh (1969). 141. To, W. M. and HUMPHREY,J. A., Int. J. Heat Mass Trans. 29, 573 (1986). 142. OSTRACH,S., NA CA TN 2635 (1952). 143. HUMPHREY, J. A. and To, W. M., Int. Z Heat Mass Trans. 29, 593 (1986). 144. FERNANDEZ-PELLO,A. C. and WILLIAMS,F. A., Combust. Flame 28, 251 (1977). 145. FERNANDEZ-PELLO, C., RAY, S. R. and GLASSMAN, A. I., Eighteenth Symposium (International) on Combustion, p. 579, The Combustion Institute, Pittsburgh (1981). 146. Loll, H. T. and FERNANDEZ-PELLO,A. C., Twentieth Symposium (International) on Combustion, p. 1575, The Combustion Institute, Pittsburgh(1984). 147. MARTIN,S., Tenth Symposium (International) on Cornbustion, p. 877, The Combustion Institute, Pittsburgh (1965).

Das könnte Ihnen auch gefallen