PERGAMON Progress in Energy and Combustion Science 27 (2001) 1–74 www.elsevier.com/locate/pecs Fundamental aspects of the heterogeneous flame in the self-propagating high-temperature synthesis (SHS) process A. Makino* Department of Mechanical Engineering, Faculty of Engineering, Shizuoka University, Hamamatsu 432-8561, Japan Received 30 April 1999; received in revised form 7 February 2000; accepted 7 February 2000 Abstract Recent progress on understanding fundamental mechanisms governing the Self-propagating High-temperature Synthesis (SHS) process, which is characterized by the flame propagation through a matrix of compacted reactive particles and is recognized to hold the practical significance in producing novel solid materials, is reviewed. Here the focus is not only on the theoretical description of the heterogeneous nature in the combustion wave, which has not been captured by the conventional premixed-flame theory for a homogeneous medium, but also on the extensive comparisons between the predicted and experimental results in the literature. Topics included are the statistical counting procedure used for deriving governing equations of the heterogeneous theory, flame propagation in the adiabatic condition, flame propagation and extinction under heat loss conditions, effects of bimodal particle dispersion on the combustion behavior, those of external heating by electric current, the transition boundary from steady to pulsating combustion, and the initiation of the combustion wave by use of the external heating source. The importance of heterogeneity in the combustion wave, that is the particle size of the nonmetal or the higher melting-point metal, has been emphasized for fundamental understanding of such combustion behavior as flame propagation, extinction, and initiation. Potentially promising research topics are also suggested. 䉷 2000 Elsevier Science Ltd. All rights reserved. Keywords: Combustion synthesis; Self-propagating high-temperature synthesis; Flame propagation; Burning velocity; Range of flammability; Extinction; Ignition; Ignition delay time Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. The SHS process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. General characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Scope of the present survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Phenomenological description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Formulation of the heterogeneous theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Surface regression rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Statistical counting procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1. Change of distribution function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2. Conservation equations in general form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.1. Overall continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.2. Species conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.3. Momentum conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2.4. Energy conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . * Tel.: ⫹ 81-53-478-1050. E-mail address: [email protected] (A. Makino). 0360-1285/00/$ - see front matter 䉷 2000 Elsevier Science Ltd. All rights reserved. PII: S0360-128 5(00)00004-6 3 6 6 9 10 11 11 12 12 12 12 12 12 12 2 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 4. 5. 6. 7. 3.3. Heterogeneous theory for the SHS process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1. Size distribution function in quasi-one-dimensional form . . . . . . . . . . . . . . . . . . . . . . . 3.3.2. Conservation equations for steady, one-dimensional form . . . . . . . . . . . . . . . . . . . . . . . 3.3.2.1. Overall continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2.2. Species conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2.3. Momentum conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2.4. Energy conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.3. Further simplification of the governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Flame propagation in the adiabatic condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Dominant parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Burning velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Range of flammability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4. Theoretical and experimental results for Ti–C system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5. Applicability of the heterogeneous theory for other systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.1. Available experimental data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.2. Effects of mixture ratio, m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.3. Effects of degree of dilution, k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.4. Effects of initial temperature, T0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.5. Effects of particle radius, R0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6. Approximate expression for the burning velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6.1. Derivation of the approximate expression for m ⱕ 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6.2. General behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6.3. Comparisons with numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6.4. Approximate expression for m ⱖ 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.7. Some remarks on the flame propagation in the adiabatic condition . . . . . . . . . . . . . . . . . . . . . . Flame propagation in the nonadiabatic condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Governing equations and boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Characteristics in the nonadiabatic condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Turning-point determined by the thermal theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Experimental comparisons for Ti–C system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1. Range of flammabilty and extinction limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2. Burning velocity under heat loss condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5. Experimental comparisons for other systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.1. Boride synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.2. Synthesis of intermetallic compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6. Approximate expression of the heat loss parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6.1. Asymptotic expansion analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6.2. Correction term for the heat loss parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6.3. General behavior of the heat loss parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6.4. Applicability of the analytical expression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7. Some remarks on the flame propagation under heat loss condition . . . . . . . . . . . . . . . . . . . . . . . Other aspects of the nonadiabatic flame propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1. Bimodal particle dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1.1. Equivalent particle radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1.2. Experimental comparisons for spherical carbon particles . . . . . . . . . . . . . . . . . . . . . . . . 6.1.3. Experimental comparisons for diamond particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Representative length of the cross-sectional area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Correspondence between the heterogeneous theory and the homogeneous theory . . . . . . . . . . . . 6.4. Field activated SHS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1. Heat-input parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.2. Temperature profiles outside the combustion wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.3. Effective range of electric field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.4. Experimental comparisons for Si–C system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5. Some remarks on the several, other, important factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Boundary between steady and pulsating combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1. Criterion for the appearance of cracks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 13 13 13 14 14 14 14 15 16 16 17 18 21 21 22 23 23 23 26 26 28 28 29 30 30 30 31 32 33 33 36 36 36 37 38 38 39 40 40 41 41 42 43 43 44 46 47 48 49 49 50 50 51 51 52 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 7.2. Linear stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1. Basic solution for planar propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.2. Asymptotic solution and linear stability boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.3. Experimental comparisons for the boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.4. Boundary at the adiabatic condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3. Transverse effects on the range for steady combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4. Some remarks on the boundary of steady combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. Initiation of the combustion wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1. Flame initiation induced by igniter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.1. Inert stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.2. Transition stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.3. Limiting solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.4. Experimental comparisons for the ignition delay time . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.5. Expression for the ignition energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1.6. Results for the ignition energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2. Flame initiation by use of radiative heat flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.1. Model definition and inert stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.2. Transition stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.3. Results and experimental comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.3.1. Situation without heat loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.3.2. Situation with heat loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3. Some remarks on the flame initiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.1. Summary of the present survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2. Area for further research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. Introduction There has been increasing interest in the combustion synthesis of new materials which would play important roles in future technologies in various fields. Many of the materials available today, such as ceramics (inorganic compounds) and intermetallic compounds, fall into a category of chemical compounds, so that we can expect that combustion would make a great contribution in producing new materials and/or enhancing productivity of these materials, because of the existence of strong chemical reactions accompanied by the combustion. The combustion synthesis, defined as “materials synthesis” by combustion, is therefore an interdisciplinary research subject in the fields of combustion and material science, for producing ceramics and/or intermetallic compounds in lumps, thin films of oxides and/or diamond, ultra-fine powder of inorganic compounds, etc. as combustion products. Research for the combustion synthesis has been very active in the field of material science. On the other hand, research from the viewpoint of combustion has not been so active, in spite of the important fact that synthesis of materials by combustion can only be accomplished by virtue of combustion. Then, it remains unsolved whether the knowledge of combustion hitherto obtained is applicable to the 3 52 52 53 55 56 56 58 58 59 59 60 61 61 63 63 64 65 66 67 67 68 69 70 70 71 72 72 combustion synthesis in exploring and/or explaining its combustion phenomena. This may be attributed to the fact that the main concern of the combustion has been on producing heat, for the sake of heating and/or heat-supply to heat engines. As for the combustion synthesis, its main concern is not the heat, but the materials produced as combustion products. However, it should be recognized that combustion synthesis is one of the research subjects of combustion science and technology because of the existence of combustion phenomena, although objects to be produced by combustion might be very different from each other. Furthermore, since it is a new research field in combustion, combustion synthesis should be studied systematically, from the viewpoint of combustion science and technology. This trend appeared in the late 1980s and has been reflected by the establishment of a session on Materials Synthesis in the International Symposium on Combustion, since 1990. Generally speaking, methods of combustion synthesis can be divided into two main classes: one is burner flames which produce thin films of oxides and/or diamond, and ultra-fine powder of inorganic compounds; the other method is to produce ceramics (inorganic compounds) and/or intermetallic compounds in lumps, by use of a kind of powder metallurgy called the Self-Propagating High-Temperature 4 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Nomenclature A a a0 B B~ b C c D Da E F f fst G H h hrad I Jn K k k~ L Le M m n p P Q q0 R R_ r S s s~ T T0 Ta Td t U u v W X x Reduced surface Damko¨hler number Da exp ⫺Ta =T Parameter a3 f ⫺ 1 in Section 4; length of the rectangular cross section in Section 6; thermometric conductivity in Section 8.1; radius of heating area by radiative heat flux in Section 8.2 Leading term of the normalized mass burning rate m=ma in Sections 5 and 7 Frequency factor for the surface reaction in Section 3; constant in Section 7 Frequency factor for the “reaction” term in the conventional homogeneous theory Parameter b3 f ⫹ 2g ⫹ 1 in Section 4; width of the rectangular cross section in Section 6 Constant; maximum temperature in Section 7 Specific heat Mass diffusivity in Sections 3–7; constant in Section 8. Surface Damko¨hler number; BR=D Electric field in Section 6.4; energy for establishing combustion wave in Section 8 Thermal runaway criterion in Section 8.2 Distribution function Stoichiometric mass ratio Size distribution function Heat-input parameter in Section 6.4 Enthalpy Radiative heat-transfer coefficient Integrated quantity Bessel function of the first kind of order n in Section 8.2.2 Constant in Section 7; derivative of temperature in nondimensional form in Section 8.1.2 Inverse of the arithmetic mean of melting point and adiabatic combustion temperature in nondimensional form Over-all transverse wave-number in Section 7.3; nondimensional parameter related to normalized mass burning rate in Section 8.1 Volumetric heat loss Lewis number Melting parameter Mass burning rate for deflagration Number of N particles Static pressure in Section 3; perimeter of rectangular cross section in Section 6.3 Size ratio defined in Eq. (143) in Section 6.1.1 Heat flux Heat of combustion per unit mass of N species Average radius of N particles Rate of change of particle size R dR=dt Radius of compacted specimen Cross-sectional area of compacted specimen Variable which represents a change of particle size in Section 3; stretched variable of time in reactive/diffusive region in Section 8.2 Stretched variable of time in transient/diffusive region in Section 8.2 Temperature Standard temperature Activation temperature for the reaction Activation temperature for the condensed phase mass diffusivity Stretched variable of temperature in Section 5.6.1; time in Sections 7 and 8. Diffusion velocity of species Heterogeneous burning velocity Velocity of a particle Molecular weight Normalized heat loss, defined as X exp ⫺2bC=a 02 exp ⫺C1 in Sections 5 and 7; stretched variable of location in reactive/diffusive region in Section 8.2 Physical spatial coordinate A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Y y Z Mass fraction Stretched variable of mass ratio of fluid in Section 5.6.1 Mass ratio of fluid Vectors and tensors E F f P q T U u v Unit tensor Force vector Body force vector Total pressure tensor Heat flux vector Viscous stress tensor Diffusion velocity vector Veclocity vector for fluid Veclocity vector for a particle Greek symbols a b G g D d( ) d e z h u k L0 L l l u, l d m n j 1⫺j r s sc s ST s t t xx f w w~ x C c V v Normalized temperature rise Spalding transfer number in Section 3; Zeldovich number in Sections 4–8 Mass ratio of small particles to large particles in Section 6 Mass ratio of fluid to solid Damko¨hler number for flame initiation in Section 8 Delta function Thickness of combustion wave Emissivity Normalized mass fraction Variable which represents a change of particle size in Section 3; inner variable of length in Section 5.6.1 Normalized temperature Degree of dilution defined as the initial mass fraction of the diluent Mass burning rate eigenvalue Heat penetration coefficient in Section 8 Thermal conductivity Exponents which represent temperature distribution in Section 6.4 and Section 7 Mixture ratio defined as the initial molar ratio of nonmetal, N, to metal, M, divided by the corresponding stoichiometric molar ratio Stoichiometric coefficient Normalized mass ratio of fluid Normalized mass ratio of solid Density Nondimensional coordinate Electric conductivity in Section 6.4 Stefan–Boltzmann constant Stretched coordinate in the downstream in Section 5.6.1. Nondimensional time xx component of the viscous stress tensor in Section 3. Inverse of the mixture ratio m in Section 4.6 and 5.6; heat loss in Section 8.2. Function which represents dependence of the “reaction” term on reactants Nondimensional parameter related to heat loss in Section 8.1 Surface regression rate Heat-loss parameter Variable for particle size distribution in Section 3; stretched variable for temperature in Section 8 Constant Mass rate of production by homogeneous chemical reaction in Section 3; stretched variable of location in transient/diffusive region in Section 8.2 5 6 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Subscripts a Adiabatic condition c Initiation of the combustion wave cr State of extinction d Downstream E Equivalent e Edge of boundary f Fluid I Inert heating ig Igniter in Material flow to the fluid from particles L Large particles M Metal or lower melting-point metal m Melting point max Maximum N Non metal or higher melting-point metal P Combustion product p Perturbed term r Representative value in Section 6.4; radial direction in Section 8.2 rel Relative S Small particles s Solid or surface sb Neutral stability boundary sp Specimen, that is, medium to be ignited TMD Theoretical maximum density t Differentiation with respect to time u Upstream z Differentiation with respect to location in the axial direction j With respect to mass fraction of fluid Z u With respect to temperature t Ignition delay 0 Initial or unburned state ∞ Final or burned state in the adiabatic condition Superscripts t Nondimensional — Average over all velocities in Section 3. Synthesis (SHS) process. In the present review, emphasis is put on the latter, because of its various advantages, to be examined. 2. The SHS process 2.1. General characteristics Among various types of combustion synthesis, wide attention has been given to the SHS process, proposed by Merzhanov and co-workers [1–3] in Russia in the late 1960s, because not only bulk processing of materials, but also formation of some elemental combinations not previously synthesized, can be performed [4–9] by initiating and passing a combustion wave through a matrix of compacted reactive particles (cf. Figs. 1 and 2). From the viewpoint of combustion science and/or technology, as shown in Fig. 2 photographically, the SHS process falls into a category of the flame propagation, because a reaction initiated at one end of a compacted medium self-propagates through an unburned medium, in the form of a combustion wave. The velocity of the combustion wave, depending on systems, varies from 1 to 250 mm/s. The high temperature needed for synthesis (perhaps more than 2000 K for inorganic compounds and more than 1000 K for intermetallic compounds) can be supplied by the self-sustained exothermic chemical reactions. The potential advantages of this process are as follows [4–9]: rapid synthesis; self-heating; energy savings; self-purification due to enhanced impurity outgassing; near-net-shape fabrication, etc. In addition, more than 500 kinds of materials, including carbides, A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 1. Schematic drawing of combustion synthesis by the SHS process [5]. 7 borides, silicides, nitrides, sulfides, hydrides, intermetallics, and complex composites, are reported [10] to be synthesized by applying the SHS process not only for solid–solid systems but also for solid–gas and/or solid–liquid systems. Table 1 shows only a few of examples of inorganic compounds synthesized by the SHS process. Compounds synthesized are being considered for use as electronic materials, materials resistant to wear, corrosion, and heat. Furthermore, it can even be applied to the synthesis of shape-memory alloys, hydrogen-storage alloys, and hightemperature superconductors [7–9]. Thus, it is now well recognized that the SHS process can be of practical significance in producing novel solid materials. However, if restricted to synthesizing homogeneous materials in composition and/or texture, as is the case for conventional materials, the SHS process would not be attractive to researchers in the field of combustion. It was in the late 1980s when various attempts began to discover possibilities for improving the specific nature of the conventional, homogeneous materials for the purpose of producing advanced materials with useful, multifunctional characteristics generated by heterogeneity in composition and/or texture. Among these advanced materials, Functionally Graded Materials (FGMs), which are composed of different material components such as ceramics and metals with continuous profiles in composition, structure, texture, material strength, and thermophysical properties, have attracted special interest as advanced heat-shielding/structural materials in future space applications. At that time, Fig. 2. A sequence of the advance of a combustion wave in the SHS process for Ti–C system; stoichiometric mixture without dilution; diameter is 18 mm and length is about 50 mm; the initial temperature T0 is 300 K and radius R0 of the carbon particle is 5 mm; the maximum temperature is about 3200 K and the burning velocity u is 17 mm/s. (a) Just after the ignition; (b) 1 s; (c) 2 s; and (d) 3 s. TiB2 –Al2O3, etc. MoB–Al2O3, NiS, Ni3S2, CoS, CoS2 MnS WC–Al2O3, MoS2, WS2, MoSe2, WSe2 SiC–Al2O3, SiC–MgO, TiC–ZrO2, CsH2 TiC–TiB2, Hydride Compound TiC–SiC, TiC–Al2O3, PrH2, NdH2 TiH2, ZrH2 B4C–Al2O3, TaS2, NbSe2, TaSe2 TiS2, ZrS2, PbS, TiSe2 ZnS, CdS, ZnSe, CdTe Cu2S, CuInSe2 Chalcogenide CeS FeSi2, CoSi2 MnSi CrSi2, MoSi2, Mo3Si, WSi2 VSi2, V3Si, NbSi2, TaSi2 TiSi, TiSi2, ZrSi, ZrSi2 Mg2Si, CaSi2 Cu2Si Silicide YSi2, LaSi2, CeSi2, DySi2 CrN VN, NbN, TaN, Ta2N Si3N4, TiN, ZrN, HfN Mg3N2, SrN BaN CuN Nitride BN, AlN, NdN CaC2 Carbide B4C, Al4C3 SiC, TiC, ZrC, HfC VC, Nb2C, NbC, TaC Cr3C2, MoC, Mo2C, WC Mn7C3 FeB, NiB MnB, MnB2 CrB2, MoB, Mo2B5, WB VB, NbB, NbB2, TaB TiB, TiB2, ZrB2, HfB2 AlB2, LaB6, CeB2 MgB2, MgB4, MgB6, CaB6 Boride VII VI V IV III II I Group in periodic table Complex Table 1 Examples of inorganic compounds synthesized by the SHS process Fe4N, Fe8N A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 VIII 8 production of FGMs by use of the SHS process (especially for producing ceramics in FGMs with gradual profiles in composition) was considered because of its rapidity in production. Although it has proven effective, it was also recognized that in order to control the manufacturing process, dependence of flame propagation speed on various dominant parameters is indispensable. In addition, range of flammability with respect to various parameters is urgently required because the combustion wave cannot be maintained outside the range of flammability. Of course, dependence on various parameters is also required in preventing flame extinction. From an academic point of view, these areas are within the parameters combustion researches have pursued. That is, researchers in the field of combustion are strongly encouraged to clarify these problems through knowledge of combustion. In addition to its practical utility, the SHS process commands fundamental interest because of its interesting and diverse phenomena, such as steady planar propagation, pulsation, spinning, and repeated combustion, which can be observed during the flame propagation in the solid compacted medium [1–3,11,12]. In pulsating combustion which is also called self-oscillation combustion, the combustion wave travels in a planar, but pulsating manner, which frequently results in materials with a laminated structure (cf. Fig. 3). In spinning combustion the combustion wave is nonplanar, and one or more hot spots are observed to move along a spiral path over the surface of the sample specimen, inside of which reactants experience only heating, without reaction. In repeated combustion, after the passage of the combustion wave, another combustion wave is initiated and propagates through the burned medium, yielding the final combustion products from the intermediates. Fundamental understanding of flame propagation, extinction, and/or initiation in the SHS process is not only of interest for its own sake, but also of great use for potential improvement in the manufacturing process. Various experimental works have been conducted, and many of the accomplishment have been summarized in several good review papers [5–9]. In order to elucidate effects of various system parameters on the flame behavior, theoretical studies have also been conducted keeping up with experiments. The basic theory, originally derived by Novozhilov [13] for the combustion of solid propellant, and extended by Merzhanov [14] for the SHS process, has been applied to examining stability [15] and unsteady behaviors, such as pulsating [16] and spinning [17] combustion. Effects of competing reactions [18], particle size distribution [19], heat loss [20,21], Arrhenius mass diffusion [22], etc. have also been studied. Subsequent accomplishments by theoretical works are well summarized in recent review papers [23,24]. Note here, however, that theoretical description in those works has primarily been based on the premise that synthesis is accomplished through the passage of a premixed flame in a homogeneous medium in which reactants are well mixed at the molecular level before arrival of the flame front of the combustion wave. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 9 Fig. 3. Laminated structure in the combustion products; Ti–C system for stoichiometric mixture without dilution; the initial temperature T0 is 300 K. (a) Radius of the carbon particle R0 is 0.5 mm and the relative density r rel is 0.56. (b) R0 0:5 mm and rrel 0:63: (c) R0 0:5 mm and rrel 0:72: (d) R0 2:5 mm and rrel 0:47: (e) R0 2:5 mm and rrel 0:64: 2.2. Scope of the present survey Although much work has been conducted both experimentally and theoretically, a general lack of consistency has been noted [25] in the combustion data and no firm understanding of the combustion mechanism has been achieved. This may be attributed to the fact that comparisons between experimental and theoretical results have been rare, regardless of their importance for comprehensive understanding of various combustion characteristics. Through agreement between theoretical and experimental results over wide range of parametric quantities for various combinations of metal and nonmetal (or another metal of a higher melting point) systems, appropriateness of theories should therefore be examined. When such comparisons are conducted with conventional models in the homogeneous premixed-flame theory, there exists an inherent deficiency that cannot account for size effect of particles, while experiments have firmly established that the flame propagation speed increases with decreasing particle size in a sensitive manner [26–28]. This inherent deficiency comes from insufficient treatment, based on the homogeneous premixed-flame theory. Even if fine powder is used, for example 0.1 mm in particle diameter, each particle contains more than 10 8 atoms. This situation is far from that to which the homogeneous premixed-flame theory can be applied. That is, we cannot regard the SHS process as a homogeneous phenomenon. Therefore, we are urgently required to re-examine the appropriateness of describing the SHS process as the premixed flame propagation in the homogeneous medium. In another mathematical model for the SHS process, it is proposed to assume the compacted medium as laminae of reactants [29,30]. The combustion situation for the usual SHS process, however, is quite different from this model because the combustion wave propagates through the medium of compacted reactive particles. Again we are required to construct an appropriate model which can properly represent the combustion situation in the SHS process, by taking its heterogeneity into account in an essential manner. The present paper, restricted to theoretical accomplishment, aims to complement the extensive reviews by Frankhouser et al. [5], Munir and Anselmi-Tamburini [7], Merzhanov [8] and Varma and Lebrat [9] regarding the experimental accomplishment, and those by Merzhanov [4], Merzhanov and Khaikin [23] and Margolis [24] regarding the theoretical accomplishment. Indeed, because of the broad coverage of some of these reviews, a narrow range of topics will be discussed here, focused on the heterogeneity in the SHS process, to which little attention has been paid, although it is indispensable for realistic descriptions of combustion behavior. The present survey also emphasizes understanding of the fundamental mechanisms governing the flame propagation, extinction, and/or initiation in the SHS process, by being restricted to the solid–solid systems. It is also suggested that the reader refers to Ref. [31] for topics about the SHS process for the gas–solid systems, which are not covered in the present review. In the next section, a phenomenological description of the flame propagation in the SHS process is presented. In Section 3, the statistical counting procedure is reviewed and its usefulness is assessed. In Section 4, theoretical approaches are primarily aimed at studying the SHS flame propagation in the adiabatic condition. Following this, several important effects on the SHS flame propagation (such as the heat loss, size distribution of particles, and external heating by electric current) are discussed in relation to flame extinction. In Section 7, the transition boundary between steady and pulsating combustion, which is indispensable in bulk processing, is discussed. Up to this point, 10 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 4. Schematic flame structure in the heterogeneous flame (crystallization not described) [32]. discussion is confined to the flame propagation and extinction. In the next section, initiation of combustion wave is discussed for two different ignition methods: igniter and radiative heat flux. 2.3. Phenomenological description In order to recognize differences between the heterogeneous flame in the SHS process and the homogeneous premixed flame, let us first make a detailed phenomenological consideration for the SHS flame propagation. It can be considered that the SHS process typically involves reaction between particles of a metal and a nonmetal (or another metal of a higher melting point). Generally speaking, before the arrival of the combustion wave, the effects of the reaction cannot be remarkable due to the low temperature and small area of contact between particles. However, as this particle matrix is heated by the approaching combustion wave, the lower melting point metal melts first, resulting in slurry consisting of the higher melting-point nonmetal particles suspended in the molten metal. In addition, due to an increase in the area of contact, as well as that in temperature, the reaction between nonmetal particles and molten metal is greatly enhanced. Then, it is anticipated that the subsequent reaction between them will take place mainly over the surface of the nonmetal particles; this situation is schematically shown in Fig. 4 [32]. Provided that the nonmetal particles do not dissolve in the molten liquid, and that they are not too small, compared to the thickness of the combustion wave, it is easily seen that the flame propagation through the compacted medium is basically heterogeneous in nature, involving a premixed-mode of propagation for the bulk flame supported by the nonpremixed reaction of the dispersed nonmetal particles in the liquid metal. Therefore, it is reasonable to anticipate that the resulting flame structure and response should be quite different from those obtained by assuming a premixed-mode of reaction at the molecular level. That is, the present phenomenological consideration strongly suggests that a satisfactory formulation must take into account the nonpremixed mode of particle reaction. Here, it may be of great importance to recognize that the flame propagation in the SHS process is actually quite similar to that in fuel spray combustion, if we identify the fuel droplets as the nonmetal (or higher melting point metal) particles and the oxygen in air as the molten metal, restricting ourselves to the flame zone in the combustion wave. It is also appropriate to recognize that across the combustion wave, dozens of nonmetal particles [33] can exist in the “reaction” zone in which the temperature is higher than the melting point of metal. This can easily be confirmed by comparing a representative particle size, for instance 10 mm, to the approximate “flame thickness” d of about 0.3 mm, evaluated from a relation d ⬇ l= rcu derived by the phenomenological analysis for gaseous premixed flames [34], with representative values of thermometric conductivity l /(r c) of 3 × 10⫺6 m2 =s and the burning velocity u of 0.01 m/s for Ti–C system. This recognition further suggests that a statistical counting procedure for spray combustion, explained in Ref. [35], is indispensable and of great use in describing the combustion behavior of the heterogeneous flame in the SHS process. The degree of similarity is achievable if we restrict ourselves to fundamental principles without probing too far into specific details of the physical processes. Another important factor that can influence the SHS flame propagation is the mass diffusion in liquid phase, which is closely related to the nonpremixed reaction of nonmetal particles in the combustion wave. Since the mass diffusivity in the SHS process is anticipated to increase markedly over a relatively thin, high temperature combustion wave, use of the temperature-dependent liquid-phase mass-diffusivity in Arrhenius fashion [22] is appropriate and indispensable. As for the surface reaction on each nonmetal particle, related to the nonpremixed reaction in the combustion wave, it is recognized to be quite similar to the combustion A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Ta;s ; Ts of a single solid particle, having been examined in the field of solid combustion, if the oxygen in air is identified as the molten metal. This recognition further facilitates inclusion of the finite-rate reaction at the particle surface when formulation is conducted. Finally, in closing this section, it may be informative to describe a simplified model [32] which has been adopted for investigating combustion behavior of the heterogeneous flame in the SHS process. As shown in Fig. 4, the combustion situation modeled is the planar, heterogeneous flame in an infinite domain of a compacted medium, originally consisting of a mixture of particles of nonmetal N, metal M, and an inert which can be the product P of the reaction according to nM M ⫹ nN N ! nP P; where n i is the stoichiometric coefficient. For simplicity, it is assumed that there is no reaction until the mixture has been heated to the melting point Tm of the metal, at which all the metal particles melt instantly. The reaction that follows is assumed to take place at the particle surface at a finite rate and proceed until all of the nonmetal (or higher melting-point metal) particles, or metal species, are consumed. A Das exp ⫺ 3. Formulation of the heterogeneous theory r~ r vR Y n W ; vr r ; Y~ M M ; fst M M ; R De fst nN WN 3.1. Surface regression rate r~ rM D ; D~ : De rM;e Before describing the statistical counting procedure used in spray combustion, let us first confirm the description of a single solid particle undergoing combustion. Since it is unnecessary to consider the velocity difference between particles and molten metal in the SHS combustion wave, only combustion behavior in a quiescent fluid will be examined. For a single, spherical, nonpermiable, solid particle of radius R in a quiescent fluid, the rate of change of its mass caused by the surface reaction nM M ⫹ nN N ! nP P is expressed as _ ⫺ d rN 4 pR3 ⫺rN 4pR2 R; _ M 1 dt 3 which yields ⫺RR_ rM D x~ ; rN x~ ⬅ _ M : 4prM DR 2 Here, x~ is the surface regression rate, which coincides with the definition of the nondimensional combustion rate, commonly used in the conventional solid combustion. Then, we have x R_ ⫺ ; R x rM D x~ : rN 3 The specific form of x~ can be obtained as [32,36] x~ A Y~ M;e ⫺ b ; 1⫹b 11 b exp x~ ⫺ 1; 5 by solving the quasi-steady liquid-phase conservation equation in spherical coordinates ! ~ ~ ~ d r~2 dY M ⫹ x~ dY M 0; ⫺ r~ D 6 dr~ dr~ d~r under the isothermal condition, with the boundary condition at the particle surface ~r 1 ! ~ ~ s dY M ⫹x~ Y~ M;s ⫺x~ ; ⫺ r~ D x~ ⬅ r~ vr r~2 AY~ M;s ; dr~ s 7 and that at the outer edge ~r ! ∞ of the boundary layer around the particle Y~ M Y~ M;e : 8 In the above, A is the reduced surface Damko¨hler number, Das the surface Damko¨hler number ( B·R/D), b the Spalding transfer number, and other variables and parameters are defined as where r is the radial coordinate, vr the radial velocity, Y~ M the stoichiometrically weighted mass fraction, fst the stoichiometric mass ratio, W the molecular weight, D the mass diffusivity, and the subscript s and e designate, respectively, the particle surface and the outer edge of the boundary layer. An introduction of the isothermal condition, which enables us to put r~ 1 and D~ 1; can be justified when the thickness of the bulk flame is much larger than the particle size because we can anticipate that both a nonmetal solid particle and molten metal around the particle are heated equally as the combustion wave arrives. The use of the conventional constant property assumption can also be justified if attention is confined to a restricted region around the single particle. Note that changes in thermophysical properties in the bulk during the flame propagation can be incorporated. The analytical expression in Eq. (4), however, is an implicit expression with respect to the surface regression rate, which is difficult to have a clear image of its dependence on other parameters. Therefore, an attempt has been made to obtain an explicit, approximate expression [32,36] as A ~ Y M;e ; 9 x~ ⬅ ln 1 ⫹ b ⬇ ln 1 ⫹ 1⫹A by use of the approximate relation 4 b 1 ⫺ exp ⫺x~ ⬇ x~ 1⫹b 10 12 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 for small x~ ; as is the case for most solid combustion. Note that the expression in Eq. (9) for the surface regression rate has the same form as that commonly used in droplet combustion. The error between Eqs. (4) and (9) is within 2% for Ti–C system. It may be informative to note that Y~ M;e in Eq. (9) can change as the combustion wave propagates, being governed by the species conservation equations, to be mentioned in the next section. 3.2. Statistical counting procedure 3.2.1. Change of distribution function In spray combustion [35] a statistical description is given by the distribution function f R; x; v; t dR dx dv which is the probable number of particles in the radius range dR about R, located in the spatial range dx about x with velocities in the range dv about v at time t. Here, dx and dv are the three-dimensional elements of physical space and velocity space, respectively. The variables R, x, and v must appear in the distribution function because conditions are not known well enough to permit specification of the exact size, position, or velocity of each particle. An equation governing the time rate of change of the distribution function f, which is called the spray equation in spray combustion, is given as 2f 2 _ ⫺ Rf ⫺ 7 x · vf ⫺ 7 v · Ff ; 2t 2R 11 when there are no particle formation/destruction by processes nor collisions with other particles. Here, R_ ( dR/dt) is the rate of change of the particle size R at (R, x, v, t) and F dv=dt the force per unit mass on this particle. The terms in the RHS respectively represent the changes in f resulting from the change of particle size, the motion of particles into and out of the spatial element dx by virtue of their velocity v, and the acceleration of particles in the velocity element dv caused by the force F. Note that both R_ and F are allowed to depend on R, x, v, t, and the local properties of the fluid. 3.2.2. Conservation equations in general form In analyzing behavior of the combustion wave, hydrodynamic equations for the fluid are necessary. If we restrict ourselves to the situation in which the statistical fluctuations in the fluid, induced by the random motion of individual particles, may be neglected, these governing equations for the local average properties in the fluid are equivalent to the ordinary equations of fluid dynamics, with suitably added source terms accounting for the average effect of the dispersed particles. 3.2.2.1. Overall continuity By adding the mass of material per unit volume per unit time from the particles to the fluid, the overall continuity equation for the fluid is given as ZZ 2rf _ dR dv; rN 4pR 2 Rf ⫹ 7 x · rf u ⫺ 12 2t where r f is the fluid density defined as the mass of fluid per unit volume of physical space. Note that the fluid and liquid densities are related by the expression ZZ 4 pR3 f dR dv : rf rM 1 ⫺ 3 13 3.2.2.2. Species conservation When chemical reactions occur at the particle surface and species M is consumed at the surface reaction according to n MM ⫹ n NN ! n PP, the species conservation equation for component M is given as 2 r Y ⫹ 7 x ·rf u ⫹ UM YM 2t f M ZZ n W _ dR dv; v⫹ M M rN 4pR 2 Rf nN WN 14 where YM is the mass fraction of species M in the fluid, UM the diffusion velocity, and v the mass rate of production of species M by homogeneous chemical reactions in the fluid. Note that for steady flow, by use of Fick’s law, UM is expressed as UM YM ⫺D 7 x YM : 15 3.2.2.3. Momentum conservation conservation equation is given as rf The momentum N X 2u ⫹ rf u·7 x u ⫺7 x ·P ⫹ rf Yk f k 2t k1 ⫺ ⫺ ZZ ZZ 4 rN pR 3 Ff dR dv 3 _ ⫺ uf dR dv; rN 4pR2 R v 16 in which the third term in the RHS represents the average force per unit volume exerted on the particles by the surrounding fluid, and the last term accounts for the momentum carried to the fluid by the material from the particles. In the above, P is the total pressure tensor ( pE ⫹ T), related to the hydrostatic pressure p and the viscous stress tensor T, where E is the unit tensor. In addition, fk is the external body force per unit mass acting on species k in the fluid. 3.2.2.4. Energy conservation The energy conservation A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 equation is given as " !# " !# 2 u2 u2 r h ⫹ ⫹ 7 x · rf u hf ⫹ 2t f f 2 2 and the bar denotes an average over all velocities, that is 1 Z_ 1 Z Rf dv; v vf dv: 20 R_ G G Eq. (18) can be further simplified as N X 2p ⫹ rf Yk u ⫹ Uk ·f k ⫺7 x ·q ⫺ 7 x · T·u ⫹ 2t k1 2 _ 1 2 0; RG ⫹ SvG 2R S 2x ZZ 4 rN pR 3 F·vf dR dv 3 ! ZZ v2 2_ ⫺ rN 4pR R hin ⫹ f dR dv; 2 ⫺ 17 where hf is the total enthalpy per unit mass of the fluid, q the heat flux vector, and hin the total enthalpy per unit mass of material flowing from the vicinity of a particle to the fluid. In the above, the last two terms accounts for the work done on the fluid by particle and the energy added to the fluid by the material from the surface. 3.3. Heterogeneous theory for the SHS process In the SHS process, it is usual to consider a situation in which a reaction, initiated at one end of a matrix of compacted reactive particles, moves through the adjacent unburned matrix in the form of a self-sustained combustion wave. In this situation, even a quasi-one-dimensional treatment has practical importance, and its simplicity is favorable for clarifying dominant parameters that influence the flame propagation. 3.3.1. Size distribution function in quasi-one-dimensional form The velocity dependence of the distribution function f, which is not of primary interest, can be eliminated from the governing equations by integrating Eq. (11) over all velocity space. Since f ! 0 as 兩v兩 ! ∞ for all physically reasonable flows, the integral of the last term in Eq. (11) is zero. As for the unsteady change of the distribution function f, it is of considerable significance in spray combustion, in relation to such phenomena as diesel-engine combustion or combustion instability in liquid-propellant rocket motors. However, in the SHS process it may be excluded from our consideration, because a matrix of compacted reactive particles through which a combustion wave propagates is expected to have a certain distribution of particles when it is produced by compaction, as far as the distribution of particle sizes is concerned. Then, we have 2 _ RG ⫹ 7 x · vG 0; 2R 13 18 where the number of particles per unit volume per unit range of radius is Z G f dv; 19 21 when the local cross-sectional area S(x) only depends on the variable x. Here, v is the x component of v ; and the quantities _ v; and G are averages over the cross section, independent R; of the spatial coordinates normal to x. By using a relation that R_ ⫺x=R in Eq. (3), Eq. (21) becomes ⫺x 2c 2c ⫹ vR 0; 2R 2x 22 where c⬅ SvG : R 23 Upon transformation to the new independent variables Zx x Zx x s 2 R2 ⫺ 2 dx; dx; h2 R2 ⫹ 2 v v ⫺∞ ⫺∞ 24 Eq. (22) becomes 2c=2s 0; which means that c only depends on h . Letting the subscript 0 identify conditions at the unburned state, we have c c 0(h ). Since h R0 as x ! ⫺ ∞, we see from Eq. (23) that the expression S 0 v0 R G G h 25 Sv h 0 determines the size distribution G(R, x) at any position x in terms of the distribution G0(h ) at the unburned state x ! ⫺∞: 3.3.2. Conservation equations for steady, one-dimensional form When we restrict ourselves to steady, one-dimensional, constant-area flows in which all particles travel at the same velocity, the distribution function f is expressed as f R; x; v G R; xd v ⫺ v; 26 where x is the flow direction, v the velocity of all particles, and d (·) the delta function. 3.3.2.1. Overall continuity The integral over v in Eq. (12) yields Z∞ d _ dR; rf u ⫺ 27 rN 4pR 2 RG dx 0 where u is the x component of u. The RHS in Eq. (27) also appears in Z∞ d _ dR; r v rN 4pR2 RG 28 dx s 0 14 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 which is derived by multiplying Eq. (11) by r N(4/3)pR 3 and integrating over v and R. Note that r s is the mass of solid per unit spatial volume defined as Z∞ 4 pR3 G dR: rs rN 0 3 rf u hf ⫹ u2 2 ! ⫹ rs v hN ⫹ v2 2 ! ⫹ qx ⫹ txx u constant: 37 29 Eqs. (27) and (28) show that the overall continuity equation is written in the form rf u ⫹ rs v m constant; which yields 30 where m is the total mass flow rate (i.e. mass burning rate) per unit area. Note that in deriving Eq. (36), use has been made of the relation Z∞ _ in GdR d rs vh N ; rN 4pR2 Rh 38 dx 0 obtained by equating Z∞ _ N G dR N 7 x · rs vh rN 4pR2 Rh 0 3.3.2.2. Species conservation When there exists no homogeneous chemical reactions in the fluid, the species conservation Eq. (14) is simplified as d nM WM d rf u ⫹ UM YM r v; 31 dx nN WN dx s which yields rf D dYM ⫺ rf uYM ⫺ rs v fst constant; dx 32 3.3.2.3. Momentum conservation The momentum conservation Eq. (16), by neglecting influences of body forces fk, becomes as 33 which yields rf u2 ⫹ rs v2 ⫹ p ⫹ txx constant; 34 where t xx is the xx component of the viscous stress tensor T. Note that in deriving Eq. (33), use has been made of the relation Z∞ 0 4 dv rN pR3 FG dR rs v ; 3 dx 35 which is the momentum equation of the dispersed particles, derived by multiplying Eq. (11) by rN 4=3pR3 v and integrating over v and R, as well as the relations in Eqs. (27) and (28). 3.3.2.4. Energy conservation When radiative heat transfer is neglected, the energy conservation Eq. (17) becomes as " ! !# d u2 v2 ⫹ rs v hN ⫹ rf u hf ⫹ dx 2 2 dq d ⫺ x ⫺ t u; dx dx xx and Z∞ 0 where UM is the x component of UM and can be expressed by Fick’s law. d dp dt r u2 ⫹ rs v2 ⫺ ⫺ xx ; dx f dx dx ⫹ 36 ZZ 4 2 hN rN pR3 R_ ⫹ v·7 x hN ⫹ F·7 v hN f dR dv; 3 2R 39 _ in G dR rN 4pR2 Rh ⫹ ZZ Z∞ 0 _ N G dR rN 4pR2 Rh 4 2h rN pR3 R_ N ⫹ v·7 x hN ⫹ F·7 v hN f dR dv: 2R 3 40 Here, Eq (39) is obtained by multiplying Eq. (11) by rN 4=3pR3 hN and integrating over v and R, while Eq. (41) is obtained by multiplying d 4 r pR3 hN dt N 3 _ N ⫹ rN 4 pR3 R_ 2hN ⫹ v·7 x hN ⫹ F·7 v hN ; rN 4pR2 Rh 3 2R 41 _ in rN 4pR2 Rh by f and integrating over v and R. Note that Eq. (41) gives a relation between hin and hN. 3.3.3. Further simplification of the governing equations In the SHS process, all particles in the combustion wave are reasonably assumed to travel at the same velocity as the fluid; that is, v u: Then, the overall mass conservation in Eq. (30) becomes rt u m constant; 42 where rt ⬅ rf ⫹ rs is the total density. As for the mass consumption of solid N particles, by introducing the mass fraction of fluid as rf Z ; 43 rf ⫹ rs Eq. (28) reduces to d 1 ⫺ Z 1 Z∞ _ dR: ⫺ r 4pR2 RG dx m 0 N 44 Substituting Eqs. (3) and (25) into Eq. (44) and by A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 integrating, we have 2! d 1 ⫺ Z 4p Z∞ u0 R ⫺ G0 h dR: rN x dx m u h 0 species conservation equations, such as 45 When the particles are monodispersed with number density n0, that is G0 R n0 d R ⫺ R0 ; Eq. (45) reduces to d 1 ⫺ Z 4prN n0 R0 u0 1 ⫺ Z 1=3 ⫺ x ; 46 m u dx 1 ⫺ Z0 where the following relation has been made use of ! Zx x 1=2 Z∞ R2 dx G0 h dR n0 R20 ⫺ 2 v h 0 ⫺∞ ! Zx x 3=2 1⫺Z dx ; R30 R20 ⫺ 2 1 ⫺ Z0 v ⫺∞ 48 50 when u2 =2 p hf and/or hN. The enthalpy of the fluid is given as hf N X Yk hk ; hk h0k ⫹ c T ⫺ T 0 ; k M; P 51 k1 and that of the particles is hN h0N ⫹ cN T ⫺ T 0 : {YP ⫺ 1 ⫹ fst }Z ⫹ as well as the relation between the heat of combustion and enthalpies as ⫺fst hM ⫹ 1 ⫹ fst hp ⫺ hN nP WP hP ⫺ nM WM hM ⫺ nN WN hN nN WN 56 l=c dT~ T~ ⫺ T~ 0 ⫺ Z ⫺ Z0 ; 57 m dx where T~ is the nondimensional temperature cT=q0 and q 0 the heat of combustion per unit mass of N species. Here, it is assumed that enthalpy of phase change is negligible because it is usually much smaller than the heat of combustion, that thermal conductivity across the combustion wave is constant, and that the specific heats are the same, that is c cN : Note that in the nonadiabatic condition, it is necessary to take account of heat loss L in the lateral direction, so that the energy conservation equation is expressed as " # d l=c dT~ L ⫺ T~ ⫺ T~ 0 ⫹ Z ⫺ Z0 : 58 dx m dx mq0 In closing this section, it should be mentioned that these governing equations can simply be derived if it is a priori assumed that N particles are the same size in counting for the heat and mass balance around them [38,39]. Regardless, here statistical counting procedure has been used, to show the mathematical description in a general form, which is applicable to systems in which particle size distribution is not monodispersed. 4. Flame propagation in the adiabatic condition Mathematical description [32,37] of the flame propagation in the SHS process has been made by use of Eq. (46) for Nconsumption, Eq. (49) for M-conservation, and Eq. (57) 1 for energy conservation, as well as the relation of constant mass 1 N X dT ⫹ hk rf Yk Uk ; dx k1 rf YP UP {YP;0 ⫺ 1 ⫹ fst }Z0 ; 55 m 52 As for the x component of heat flux vector, it is expressed as qx ⫺l 54 Eq. (50) reduces to where Y~ M is the stoichiometrically weighted mass fraction YM =fst and the constant of integration is evaluated at the unburned state x ! ⫺∞: Note that the use of the Arrhenius mass-diffusivity defined as D D0 exp ⫺Td =T is indispensable for the appropriate description of the SHS process. The momentum conservation equation has already been expressed in Eq. (34), which states that constant pressure is a good assumption when the velocity is low and the velocity gradient is small. The energy conservation equation (37) in the adiabatic condition becomes qx Z0 hf;0 ⫹ 1 ⫺ Z0 hN;0 ; m rf YM UM YM;0 ⫹ fst Z0 ; m ⫺q0 ⫹ c ⫺ cN T ⫺ T 0 ; derived by the integration of Eq. (29). The species conservation equation (32) for M species is given as rf D dY~ M 49 Y~ M ⫹ 1Z ⫺ Y~ M;0 ⫹ 1Z0 ; m dx Zhf ⫹ 1 ⫺ ZhN ⫹ YM ⫹ fst Z ⫹ 47 and rt u r t;0 u0 15 53 when there is no radiation heat transfer. Then, by the use of It should be noted that the energy conservation equation first reported [32] was incorrect, due to inappropriate treatment in accounting for a term representing the sensible heat of N particles. This was pointed out and corrected in the subsequent paper. [37] Figs. to be presented in this review have already been corrected by use of the energy conservation Eq. (57) although quantitative trends are the same. 16 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 burning rate expressed in Eq. (42). The combustion situation modeled, as shown in Fig. 4, is that of the steady, adiabatic, one-dimensional, planar, heterogeneous flame propagation in an infinite domain of a compacted medium, originally consisting of a mixture of particles of nonmetal N, metal M, and an inert which can be the product P of the reaction according to vM M ⫹ vN N !vp P: Then the flame propagation can be examined by solving these governing equations, subject to boundary conditions at the unburned and burned states. 4.1. Dominant parameters Before describing numerical results, let us first reconfirm dominant parameters, because it is important in analyses to use parameters which can fairly represent system conditions for the initial and final states of the SHS process, as well as their clear definitions from the physical point of view. Furthermore, in order to make comparisons between theoretical and experimental results, definite relations between parameters used in analyses and those in experiments are strongly required. At the burned state, we have T~ ∞ Z∞ ⫺ Z0 ⫹ T~ 0 ; 59 Z Y~ M;∞ Y~ M;0 ⫹ 1 0 ⫺ 1; Z∞ 60 which are obtained, respectively, from the energy conservation equation (57) and the species conservation equation (49) at a completely reacted state. We see that both T~ ∞ and Y~ M∞ are expressed in terms of not only their initial values, but also the mass fractions of fluid, Z, at the burned and unburned states. As for Z∞, by counting for the mass fraction of the the remaining N particles after the completion of the reaction, we have 8 m ⱕ 1 > <1 Z∞ ; 61 1 ⫺ Z0 > m ⱖ 1 : Z0 ⫹ m where m is the mixture ratio defined as the initial molar ratio of nonmetal, N, to metal, M, divided by the corresponding stoichiometric molar ratio. Note that it is common in experiments to specify the nature of the initial (unburned) compact by this mixture ratio m and the degree of dilution k , defined as the initial mass fraction of the diluent. These experimental parameters are related to the initial mass fraction of N particles, 1 ⫺ Z0, used in analyses, as [37] 1 ⫺ Z0 m 1 ⫺ k : m ⫹ fst 4 rN pR30 n0 1 3 m 4 3 nN WN = vM WM r YM;0 1 ⫺ pR0 n0 3 64 4 r YP;0 1 ⫺ pR30 n0 3 k 1 ⫺ YM;0 Z0 ; 65 4 3 4 r 1 ⫺ pR0 n0 ⫹ rN pR30 n0 3 3 where r is the fluid density that consists of metal M and inert P. It is seen that by specifying m and k , both the mass fraction of N particles, 1 ⫺ Z0 ; and the mass fraction YM,0 of M species in fluid at the unburned state can be specified. It may be noted that the present consideration of dilution, through which the mass fraction of molten M is reduced due to addition of combustion product, implicitly assumes that the product is soluble in the liquid metal, and that the dissolution process occurs much faster than the reaction of N particles. The first requirement depends on the specific metal/nonmetal (or metal/metal) system, and can be assessed from its phase diagram, while the second requirement is favored for fast rates of interfacial dissolution and dispersion through mass diffusion, especially for small product particles. For example, in the Ti–C system, TiC melts in the range 1918–3340 K, depending on composition. Since Ti melts at about 1950 K, and the mixture temperature increases due to reaction, TiC in the combustion wave is expected to exist in the fluid phase. As another dominant parameter, we can point out the initial temperature T0. The effect of T0, which determines T∞ through Eq. (59), may be significant because T∞ can exponentially influence the reaction rate and mass diffusivity in the condensed phase. It should be noted that T0 is the temperature of the unburned medium, just before the arrival of the flame front of the combustion wave. 4.2. Burning velocity Because the spatial coordinate x appears in none of these equations, except as d/dx, the problem can further be reduced to having Z instead of x as the independent coordinate, for adiabatic flame propagation. Dividing Eqs. (49) and (57) by Eq. (46), and defining the nondimentional variables as 62 This relation can be derived from the following expressions for their definitions [32]: 4 r 1 ⫺ pR30 n0 3 Z0 ; 63 4 3 4 r 1 ⫺ pR0 n0 ⫹ rN pR30 n0 3 3 fst 1 ⫺ Z0 ; YM;0 Z0 u T~ ⫺ T~ 0 ; T~ ∞ ⫺ T~ 0 z Y~ M ⫺ Y~ M;0 ; Y~ M;∞ ⫺ Y~ M;0 j Z ⫺ Z0 ; Z∞ ⫺ Z0 66 we have two first-order differential equations as [32] ! L r =r du T~ d u ⫺ j 0 t;0 1=3t exp ; 67 dj 1 ⫺ j x~ T~ 1 ⫺ Z 0 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 dz Le 0 L0 rt;0 =rt 2 2T~ d exp 1=3 dj {Z0 ⫹ 1 ⫺ Z0 j} 1 ⫺ j x~ T~ Z∞ Z0z z⫺ ; j⫹ Z∞ ⫺ Z0 1 ⫺ Z0 ! 68 when the mixture ratio is smaller than or equal to unity. When the mixture ratio is larger than unity, we have [40] ! L 0 rt;0 =rt du u ⫺ j T~ d ; 1=3 exp dj Z∞ ⫺ Z0 T~ Z ∞ ⫺ Z0 1⫺ j x~ 1 ⫺ Z0 69 dz dj Le 0 L0 rt;0 =rt 2 1=3 Z ∞ ⫺ Z0 {Z0 ⫹ Z∞ ⫺ Z0 j} 1 ⫺ j x~ 1 ⫺ Z0 ! 2T~ d Z∞ Z0z : exp z⫺ j⫹ Z∞ ⫺ Z0 Z∞ ⫺ Z0 T~ 70 In the above, L 0 is the mass burning rate eigenvalue defined as L0 Z∞ ⫺ Z0 2 m2a ; 4p rM D0 l=cn0 R0 71 Le0 the Lewis number as Le0 l=c ; rt D0 72 and x~ the surface regression rate as expressed in Eq. (9). In Eq. (71), ma is the total mass burning rate in the adiabatic condition. The boundary conditions are j 0; u um; j 1; u 1; z 0; z 1: 73 74 Note that the cold boundary difficulty [34] can be eliminated by assuming that the reaction is initiated at the melting point Tm of the metal [32]. Therefore, it is seen that the problem is reduced to solving the mass burning rate ma from the two first-order equations of Eq. (67) and (68) for under-stoichiometric mixture ratios, or Eqs. (69) and (70) for over-stoichiometric mixture ratios, satisfying the four boundary conditions in Eqs. (73) and (74). For spray combustion, the surface regression rate x~ is independent of the droplet size and L 0 becomes the eigenvalue. Since n 0 R03 is fixed for a given stoichiometry, L 0 implies that ma ⬇ R⫺1 0 : For the present problem, the surface Damko¨hler number A, and the surface regression rate x~ ; depend on the particle radius R because of the existence of the surface reaction; the particle radius varies as R R0 1 ⫺ j 1=3 : Then the dependence of ma on R0 is more complex. In the limit of small R0, x~ ⬃ A ⬃ R 0 ; that is L 0/ R0 becomes the eigenvalue such that ma ⬃ R⫺1=2 : In the 0 17 diffusion-controlled limit, x~ becomes independent of R0 and we have ma ⬃ R⫺1 0 : Once the specific value of L 0 is obtained by a numerical calculation, the burning velocity is then obtained as [40] p s 1 ⫺ Z0 rM =rN ; 75 u0 ·R0 D0 3L0 Le0 Z∞ ⫺ Z0 1 ⫺ Z0 which is simply expressed as [32,37] s p rM =rN u0 ·R0 D0 3L0 Le0 ; 1 ⫺ Z0 76 when the mixture ratio is under-stoichiometric. We see that the burning velocity is inversely proportional to the particle size. The parameter u0·R0 is sometimes called the “SHS rateconstant”. Note that in the experiments the burning velocity u0 should be determined as the velocity at which the dazzling combustion wave moves normal to its surface through the adjacent unburned compacted medium. As for the effect of the relative density r rel, which is also one of the important system parameters in experiments and is defined as the ratio of the apparent density of a compacted medium to the theoretical maximum density r TMD, its influence on the burning velocity is anticipated to appear through the thermophysical properties in the Lewis number Le0 l=c= rt D0 which appears in Eqs. (75) or (76). Since the total density r t is expressed as rt rrel ·rTMD and the specific heat c is proportional to r rel, while the thermal conductivity is a weak function of r rel, the burning velocity is considered to be inversely proportional to the relative density r rel, as the first approximation for the usual range of r rel in the SHS process. 4.3. Range of flammability The regime in which a solution of the present problem can exist is anticipated to coincide with the range of flammability. If we consider the general restriction that the temperature of the completely reacted state should be higher than the melting point T∞ ⱖ Tm [32,37], we have fst ! ⱕ m ⱕ 1; 1⫺k ⫺1 T~ m ⫺ T~ 0 f 0 ⱕ k ⱕ 1 ⫺ 1 ⫹ st T~ m ⫺ T~ 0 ; m 77 78 for the under-stoichiometric mixture ratios; 1ⱕmⱕ 1⫺k ⫺ fst ; T~ m ⫺ T~ 0 0 ⱕ k ⱕ 1 ⫺ m ⫹ fst T~ m ⫺ T~ 0 ; for the over-stoichiometric mixture ratios. 79 80 18 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 5. Representative flame structure for the temperature u , the carbon concentration 1 ⫺ j; and the reaction rate dj /ds , along the nondimensional coordinate s , for the stoichiometric mixture m 1:0 without dilution k 0 when the initial temperature T 0 450 K; the Lewis number Le0 100; the initial Damko¨hler number Da0 10 9 ; and the adiabatic mass burning rate eigenvalue L0 6:109 × 10⫺5 : Dashed curves are the results in the adiabatic condition [32] and solid curves are those in the nonadiabatic condition [64]. 4.4. Theoretical and experimental results for Ti–C system Eqs. (67) and (68) have been numerical integrated to determine the mass burning rate eigenvalue for Ti–C system [32]. Values of the physicochemical parameters are: q 15 MJ=kg; c 1 kJ= kg:K; rM 4:50 × 103 kg=m3 ; rN 2:25 × 103 kg=m3 ; WN 47:9 × 10⫺3 kg=mol; WM 12:0 × 10⫺3 kg=mol; Td 1973 K; Ta 3 × 104 K: For the mass diffusivity D D0 exp ⫺1:66 × 104 =T m2 =s is used, following Hardt and Phung [29]. Since apparent thermometric conductivity l /(r tc) is of the order of 10 ⫺5 m 2/s at 3000 K [41], the Lewis number Le0 is estimated to be 50 or more. Total densities before and after combustion are assumed to be equal. Note that the thermophysical properties used here implicitly account for effects of compact density, gases in void spaces, gas evolution, etc. Fig. 5 shows a representative flame structure with profiles of temperature u , carbon mass fraction, 1 ⫺ j; and the particle consumption rate dj /ds , along the nondimensional distance s m a 1 ⫺ Z0 = l=cx: The dashed curves represent the results in the adiabatic condition, while solid curves are those of the nonadiabatic condition, to be discussed in Section 5. Position at which u u m is set at s 0: With the monotonic increase in the temperature u from the melting point u m to the adiabatic combustion temperature u 1; the mass fraction of carbon 1 ⫺ j rapidly decreases from 1 to 0. The particle consumption rate dj /ds exhibits the initial increase due to carbon depletion and consequent decrease due to reactants depletion, as R ! 0: Its profile spans approximately the same range over which the temperature increases. Compared to the typical premixed flame structure, characterized by a narrow, highly peaked reaction rate profile located close to the maximum temperature, the present result demonstrates the diffusive nature of the heterogeneous flame structure supported by particle reaction, so that the region of the particle consumption in the combustion wave can be called the “consumption zone.” For the prescribed conditions with R 0 10 mm; the flame thickness and burning velocity are predicted to be about 1 mm and 10 mm/s, respectively. Fig. 6 shows mass burning rate eigenvalue L 0 as a function of the initial surface Damko¨hler number Da0 B·R0 =D0 for an initial temperature T~ 0 0:03 T0 450 K; with the Lewis number Le0 and the mixture ratio m taken as parameters. With increasing Da0, L 0 initially increases and attains a constant value for Da0 ⬎ 10 5. This implies that for higher values of Da0, the reaction, and hence the mass burning rate, becomes diffusion limited, herein A= 1 ⫹ A ! 1 in x~ which appears in Eqs. (67) and (68). Since Da0 is closely related to B·R0, its value is also shown in the abscissa. Fig. 6 further shows that due to the very large value of Le0, L 0 is insensitive to Le0. It is, however, significantly increased with increasing m because of the increased amount of volumetric heat generation. Since L 0 is basically independent of Le0, in order to determine appropriate values of B and Le0, numerical results are compared with the experimental results [33] for various radii (from 0.5 to 12.5 mm) of specimen and it is confirmed [32] that close correlation can be obtained for Le0 100 and B ⱖ 10 3 m=s when the relative density r rel is about 0.6. Note that the value of B ⱖ 103 m=s implies that the combustion is diffusion-controlled. Fig. 7 shows the effect of the initial radius R0 of carbon particle on the burning velocity u0 of the stoichiometric mixture m 1 for various B. The comparison between the predicted and experimental results [26–28,33,42] demonstrates the importance of heterogeneity in the flame propagation process, and shows the fact A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 6. Mass burning rate eigenvalue L 0 for Ti–C system as a function of the initial surface Damko¨hler number Da0 for the nonmetal particle, with the Lewis number Le0 and the mixture ratio m taken as parameters [32]; the initial temperature T~ 0 0:03 (T0 450 K). Fig. 7. Burning velocity u0 for Ti–C system as a function of the initial radius R0 of carbon particle, with the frequency factor B taken as a parameter [32]; the Lewis number Le0 100; the initial temperature T~ 0 0:03 (T0 450 K), and the mixture ratio m 1:0: Data points are experimental [26–28,33,42]; solid curves are calculated from theory. 19 that u0 is inversely proportional to R0 as shown in Eqs. (75) or (76). The theory breaks down as R0 ! 0; and ma ! ∞; which corresponds to the theoretical fact that the flame propagation mode should become homogeneous and the combustion be reaction-controlled (cf. Fig. 6). Experimental observation [33], however, shows that there occurs a transition from the steady, one-dimensional flame propagation to the so-called pulsating combustion [1–3] when the particle radius is less than about 2.5 mm. Fig. 8 shows u0 as a function of the mixture ratio m , with R0 taken as a parameter. Comparison between the calculations and experimental results [33] is again satisfactory. The lower limit of flammability given by Eq. (77) is 0.448, which is close to the lowest m of about 0.47 obtained experimentally [33]. Calculations have also been performed with a constant diffusivity of D 10⫺9 m2 =s; which is representative of its value in the consumption zone. Results for the constant diffusivity are more “curved” than the somewhat linear variations shown in both the experimental data and calculated results with Arrhenius mass-diffusivity. This therefore demonstrates the need [22] to realistically include the temperature-dependent mass diffusivity, without which u0 becomes large because of the overestimated mass diffusivity at relatively low temperatures. As for the dependence of u0 on m in its wide range, Fig. 9 shows the SHS rate-constant u0,a·R0 as a function of 2m= 1 ⫹ m for T0 450 K; with the degree of dilution k taken as a Fig. 8. Burning velocity u0 for Ti–C system as a function of the mixture ratio m , with the initial particle radius R0 and the frequency factor B taken as parameters [32]; the Lewis number Le0 100 and the initial temperature T~ 0 0:03 (T0 450 K). Data points are experimental [33]; solid curves are calculated with Arrhenius diffusivity; dashed curves are calculated with representative diffusivity. 20 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 9. Burning velocity multiplied by particle radius (SHS rate-constant) u0·R0 for Ti–C system as a function of 2m= 1 ⫹ m; with the degree of dilution k taken as a parameter [40]; the initial temperature T0 450 K: Data points are experimental [33,43] for different particle radii of R0 5 mm and 12.5 mm. parameter. Introduction of 2m= 1 ⫹ m; which is a harmonic mean of m and 1=m; enables us to discuss the effect of m , which can change from zero to infinity, in a finite domain. With increasing 2m= 1 ⫹ m; the SHS rate-constant u0,a·R0 first increases, reaches the maximum, and then decreases. The increase in u0,a·R0 is attributed to an increase in the volumetric heat generation, as well as an increase in the heat generation rate in the combustion wave, related to the total surface area of N particles. The peak of u0,a·R0 occurs on the rich side m ⬎ 1 of N particles, because of the increase in the total surface area of N particles, in spite of the decrease in volumetric heat generation. Further increase in 2m= 1 ⫹ m results in a decrease in u0,a·R0, because of the reduced volumetric heat generation. It is also observed that u0,a·R0 decrease with increasing k , and that dilution moves the position of the maximum u0,a·R0 towards the stoichiometric mixture. Experimental data [33,43] for two different particle sizes are also shown in Fig. 9 and fair agreement is observed as far as the general trend and the approximate magnitude are concerned. It is pointed out, however, that discrepancies near the maximum u0,a·R0 may be attributed to the assumption of constant thermophysical properties throughout the combustion wave. Overestimation of the range of flammability by Eqs. (77) and (79) is attributed to the heat loss to be discussed in Section 5. Fig. 10 shows u0 as a function of the initial temperature T0, for m 0:6 and 1.0. The trend of increasing u0 with increasing T0 is as expected, as the preheating effect. The comparison with experimental data [33,44] is again satisfactory. Dilution by the final product (TiC), to control the burning velocity and/or combustion temperature has also been investigated; in the field of material processing, this corresponds to controlling grain size in materials through control of the maximum temperature. Fig. 11 shows u0 as a function of k for m 1:0; with T0 taken as a parameter. It is seen that u0 gradually decreases until k reaches the dilution limit given by Eq. (78), due to reduced volumetric heat generation. Experimental results shown in Fig. 11 are those by Bloshenko et al. [44] at T0 973 K and those by Makino et al. [33] at T0 450 K; the experimental results [26,45,46] with carbon particles of 0.1 mm in diameter are excluded because steady propagation cannot be expected. Fig. 12 shows u0 as a function of the relative density r rel for m 1:0; with 2R0 taken as a parameter; u0 gradually decreases when 2R0 is larger than 10 mm and the steady Fig. 10. Burning velocity u0 for Ti–C system as a function of the initial temperature T0, with the mixture ratio m and the frequency factor B taken as parameters [32]; the Lewis number Le0 100 and the initial radius R0 10 mm: Data points are experimental [33,44]; solid curves are calculated from theory. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 21 densities it is caused by reduced temperature in the combustion wave due to enhanced heat transfer ahead of the combustion wave. 4.5. Applicability of the heterogeneous theory for other systems Agreement with a single system is not adequate to establish the validity of the heterogeneous theory, especially since this theory represents a fundamentally different description of the SHS process. To this aim, the heterogeneous theory has been applied to compare with experimental results obtained from the syntheses of several borides and intermetallic compounds for which reliable experimental data are available. It is attributed that the viability of the theory will be demonstrated through extensive, quantitative comparisons in the range of flammability with respect to effects of mixture ratio, degree of dilution, initial temperature, and/or particle size on the burning velocity. Fig. 11. Burning velocity u0 for Ti–C system as a function of the degree of dilution k , with the initial temperature T0 and the frequency factor B taken as parameters [32]; the Lewis number Le0 100 and the mixture ratio m 1:0: Data points are experimental [33,44]; solid curves are calculated from theory. propagation is maintained. When the particle size is small, pulsating combustion occurs and the trend is quite different from that of the steady propagation. It is reported [26] that the reduction of the burning velocity is caused by the reduced contact area between the molten metal and solid particles for low relative densities, as well as the reduced volumetric heat generation, whereas for high relative 4.5.1. Available experimental data Borides and intermetallic compounds have been chosen [47] because of the practical interest for advanced applications. For example, the boride ceramics of titanium diboride (TiB2), zirconium diboride (ZrB2), and hafnium diboride (HfB2), due to their refractory nature and hardness, can be used as heating elements for high-temperature electricresistance furnace, as components and reinforcements for furnaces in metal industry, and as cutting tools and abrasives. Further, TiB2 is considered an important component in the fabrication of Functionally Graded Materials (FGMs), for use as heat-shielding/structural materials in future space applications. Alloys based on the intermetallic compound of nickel aluminide (NiAl) possess an attractive combination of Fig. 12. Burning velocity u0 for Ti–C system as a function of the relative density r rel, with particle diameter 2R0 taken as parameters; the Lewis number Le0 100 and the mixture ratio m 1:0: Data points are experimental [26,33]; open symbol designates the steady propagation and half-open symbol the pulsating combustion. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 3:00 × 104 3:83 × 104 3:73 × 104 4:79 × 104 1:68 × 104 4:28 × 104 5:61 × 104 Azatyan et al. [46] Borovinskaya et al. [51] Borovinskaya et al. [51] Borovinskaya et al. [51] Naiborodenko and Itin [58] obtained with data from Itin et al. [53] obtained with data from Itin et al. [53] mechanical and physical properties, such as high strength, low density, and high thermal conductivity, that make them ideal candidates for a number of high-temperature applications in aviation, automobile, and nuclear power engineering. The intermetallic compounds such as titanium cobaltide (TiCo) and titanium nickelide (TiNi) are of interest because of their shape-memory properties. It is even suggested [5] that formation of these compounds by the SHS process has distinct advantages over other conventional methods in terms of stronger shape-restoring force and enhanced homogeneity. The values of physicochemical parameters for these systems are listed in Table 2. Some of the transport parameters listed in Table 3 are those in the solid phase of the metal near the melting point, because of very limited published data for the mass diffusivity. For the activation energies, it has been found that they have nearly the same values as those obtained from the Arrhenius plot of (u/T) vs. (1/T) at high temperatures in which mass diffusivity can be expected to influence the burning velocity, where u is the burning velocity and T the maximum temperature. The thermometric conductivity l= rt c are those at the melting point of the lower melting-point metal [48]. Representative values of the Lewis number Le0 are also listed. In order to apply the heterogeneous theory [32], which allows for the presence of finite-rate surface reactions, kinetic parameters for the individual surface reactions are also required. However, because of the high temperature in the consumption zone in the SHS process, for instance, higher than about 2000 K for boride synthesis and 1000 K for the intermetallic compounds, the reaction is likely to be fully activated, such that the combustion behavior is expected to proceed in the diffusion-controlled regime. Satisfactory agreement in the subsequent comparisons substantiates this assumption. a b M-species represents metal or lower melting-point metal. N-species represents nonmetal or higher melting-point metal. 1973 1973 2125 2500 934 1767 1728 12:0 × 10⫺3 10:8 × 10⫺3 10:8 × 10⫺3 10:8 × 10⫺3 58:7 × 10⫺3 47:9 × 10⫺3 47:9 × 10⫺3 47:9 × 10⫺3 47:9 × 10⫺3 91:2 × 10⫺3 178:5 × 10⫺3 27:0 × 10⫺3 58:9 × 10⫺3 58:7 × 10⫺3 2:25 × 103 2:35 × 103 2:35 × 103 2:35 × 103 8:85 × 103 4:50 × 103 4:50 × 103 15.0 12.9 14.9 16.6 2.0 2.6 2.9 Ti ⫹ C ! TiC Ti ⫹ 2B ! TiB2 Zr ⫹ 2B ! ZrB2 Hf ⫹ 2B ! HfB2 Al ⫹ Ni ! NiAl Co ⫹ Ti ! TiCo Ni ⫹ Ti ! TiNi 1.00 0.99 0.46 0.45 1.18 0.55 0.52 4:50 × 103 4:50 × 103 6:53 × 103 13:09 × 103 2:69 × 103 8:90 × 103 8:85 × 103 Mol. wt. of N-species WN (kg/mol) Mol. wt. of M-species WM (kg/mol) Density of N-species r N (kg/m 3) b Density of M-species r M (kg/m 3) a Specific heat c (kJ/kg ·K) Heat of combustion q 0 (MJ/kg) Reaction nM M ⫹ nN N ! nP P Table 2 List of physicochemical parameters [47] Melting point of M-species Tm (K) Activation temperature of reaction Ta (K) Remarks with respect to Ta 22 4.5.2. Effects of mixture ratio, m Fig. 13(a) shows the burning velocity u0 for Ti–B system as a function of stoichiometry, with initial temperature T0 taken as a parameter. It has been found that flame propagation becomes diffusion-controlled when the frequency factor B exceeds 10 4 m/s. The particle radius is set to R0 0:5 mm in accordance with experimental data; some of the experiments [45,49,50] were conducted with 1 mm boron particle in diameter, others [51,52] with 0.1 mm boron particle. The comparison between predicted and experimental results is considered to be satisfactory, hence demonstrating the importance of stoichiometry in the flame propagation process. The trend of decreasing u0 with decreasing m is as expected, because of reduced volumetric heat generation, and the lower limits of flammability as given by Eq. (77) fairly agree with those of the experiments. As for the preheating effect, it is found to be relatively small although u0 increases with increasing T0 at a certain stoichiometry. Further, since there is no discernible difference in the A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 burning velocities for different sizes of the boron particles, it is pointed out that there can be agglomeration for small particles in the experiments. Fig. 13(a)–(d) shows similar comparisons for Zr–B, [3,45,50–52] Hf–B, [51,52] and Co–Ti [53,54] systems. The values of R0 used are indicated in the respective figures. The comparison between predicted and experimental results, including the limits of flammability, can be considered to be satisfactory. Hardt and Phung [29] Hardt and Phung [29] Hardt and Phung [29] Samsonov and Vinitskii [62] Naiborodenko and Itin [58] Graham [63] Graham [63] Touloukian et al. [41] and Touloukian et al. [48]; values at the melting point of the M-species. Touloukian et al. [41] and Touloukian et al. [69]. a 1:0 × 10⫺5 1:6 × 10⫺5 1:8 × 10⫺5 1:8 × 10⫺5 7:0 × 10⫺5 8:0 × 10⫺6 1:6 × 10⫺5 1:66 × 104 1:54 × 104 1:74 × 104 2:40 × 104 9:01 × 103 1:87 × 104 1:92 × 104 2:0 × 10⫺7 8:9 × 10⫺9 1:3 × 10⫺8 5:0 × 10⫺7 4:8 × 10⫺6 1:1 × 10⫺5 2:9 × 10⫺5 Ti–C Ti–B Zr–B Hf–B Al–Ni Co–Ti Ni–Ti Activation temperature Td (K) Pre-exponential factor D0 (m 2/s) 4.5.3. Effects of degree of dilution, k Fig. 14(a) shows the burning velocity u0 for Ti–B system as a function of k for m 1:0; with T0 taken as a parameter. The experimental results are from Refs. [45,51,52,55,56]. It is seen that the burning velocity gradually decreases until k reaches the dilution limit given by Eq. (78), due to reduced volumetric heat generation, as well as the reduced mass diffusivity caused by the reduced temperature in the consumption zone. Fig. 14(b)–(e) shows similar comparisons for Zr–B [3,45,50–52], Hf–B [50–52], Al–Ni [57], and Co–Ti [53] systems. Fair agreement is demonstrated as far as the trend and approximate magnitude are concerned. b 0.4 0.6 0.7 0.7 0.8 0.8 0.8 100 1000 500 4 10 10 4 Remarks with respect to D0 and Td Emissivity, e b Representative Lewis number Le0 Thermometric conductivity l /(r t c) (m 2/s) a Diffusivity System Table 3 List of transport parameters [47,66] 23 4.5.4. Effects of initial temperature, T0 Initial temperature T0 is also one of the controllable parameters in experimentation, and its effects on the burning velocity have been examined extensively. Fig. 15(a) shows u0 for Al–Ni system as a function of T0, for stoichiometric mixture m 1:0: Values of B in the diffusion-controlled regime is 10 5 m/s and more. The particle radius is set to R0 10 mm in accordance with the experimental data; Hardt and Holsinger [45] used the Ni particle of 6 mm in diameter, Naiborodenko and Itin [58,59] used 20 mm Ni particle, and Maslov et al. [57] used particle less than 45 mm in diameter; the relative density was about 0.4. The trend of increasing u0 with increasing T0 is due to the preheating effect, as expected. Fair agreement is demonstrated when T0 is lower than about 600 K. The deviation at high temperatures can be attributed to the experimental determination of T0; that is, the initial temperature T0, which should be that before arrival of the combustion wave, could be higher than the ambient temperature, which might be the reported T0, because of the heat associated with the ignition source. Fig. 15(b) shows similar comparison for Co–Ti system. The particle radius is chosen to be R0 30 mm: Experimental results are from Ref. [53]. The trend is the same as that of Al–Ni system. 4.5.5. Effects of particle radius, R0 Particle radius R0 is the important parameter which specifies the heterogeneity in the combustion wave, such that the burning velocity can be controlled by varying the particle size. Fig. 16(a) shows the effect of particle radius R0 on u0 for Co–Ti system when m 1:0 and T0 573 K: The comparison between predicted and experimental [53] results 24 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 13. Burning velocity u0 as a function of the mixture ratio m , with the initial temperature T0 taken as parameters [47]; solid curves are calculated from theory. (a) Ti–B system for Le0 103 and R0 0:5 mm; (W) data from Hardt and Holsinger [45]; (A) from Kirdyashkin et al. [49]; ( ) from Zenin et al. [50]; (K) from Borovinskaya et al. [51]; (S) from Novikov et al. [52]. (b) Zr–B system for Le0 500 and R0 0:5 mm; (A) data from Merzhanov and Borovinskaya [3]; (W) from Hardt and Holsinger [45]; ( ) from Zenin et al. [50]; (K) from Borovinskaya et al. [51]; (S) from Novikov et al. [52], (c) Hf–B system for Le0 4 and R0 0:5 mm; (K) data from Borovinskaya et al. [51]; (S) from Novikov et al. [52]. (d) Co–Ti system for Le0 10; R0 30 mm and 50 mm; ( ) data from Itin et al. [53]; (A, W) from Itin et al. [54]. Fig. 14. Burning velocity u0 as as a function of the degree of dilution k , with the initial temperature T0 taken as a parameter [47]; stoichiometric mixture m 1:0; solid curves are calculated from theory. (a) Ti–B system for Le0 103 and R0 0:5 mm; (W) data from Hardt and Holsinger [45]; (K) from Borovinskaya et al. [51]; (S) from Novikov et al. [52]; ( ) from Shcherbakov and Pityulin [55]; (L) Maksimov et al. [56]. (b) Zr–B system for Le0 500 and R0 0:5 mm; (A) data from Merzhanov and Borovinskaya [3]; (W) from Hardt and Holsinger [45]; ( ) from Zenin et al. [50]; (K) from Borovinskaya et al. [51]; (S) from Novikov et al. [52]. (c) Hf–B system for Le0 4 and R0 0:5 mm; ( ) from Zenin et al. [50]; (K) from Borovinskaya et al. [51]; (S) from Novikov et al. [52]. (d) Al–Ni system for Le0 10 and R0 10 mm; (A, S) data from Maslov et al. [57] for T0 298 K and 740 K. (e) Co–Ti system for Le0 10 and R0 30 mm; (W) data from Itin et al. [53] for T0 873 K: A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 25 26 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 15. Burning velocity u0 as a function of the initial temperature T0 [47]; stoichiometric mixture m 1:0; solid curves are calculated from theory. (a) Al–Ni system for Le0 10 and R0 10 mm; (W) data from Hardt and Holsinger [45]; (A) from Maslov et al. [57]; ( ) from Naiborodenko and Itin. [58,59]. (b) Co–Ti system for Le0 10 and R0 30 mm; (W) data from Itin et al. [53] for k 0:0; ( ) from Itin et al. [53] for k 0:2: can be considered to be satisfactory, hence demonstrating the essentially heterogeneous nature of the flame propagation process. Specifically, u0 is inversely proportional to R0 when combustion is diffusion-controlled (say, B ⱖ 107 m=s: Fig. 16(b) shows similar comparison for Ni–Ti system [60,61] when m 1:0 and T0 750 K: The inverse dependence of u0 on R0 is again observed. 4.6. Approximate expression for the burning velocity Mass burning rate eigenvalue L 0, the square root of which is the nondimensional burning velocity, has been obtained numerically [32,47] and hence presented Fig. 16. Burning velocity u0 as a function of the initial particle radius R0 [47]; stoichiometric mixture m 1:0; solid curves are calculated from theory. (a) Co–Ti system for Le0 10 and T0 573 K; data from Itin et al. [53]; open symbol designates the steady propagation and solid symbol the unsteady combustion. (b) Ni–Ti system for Le0 4 and T0 750 K; (K) data from Bratchikov et al. [60]; (W) from Itin et al. [61]. parametrically, by solving the two first-order equations of (67) and (68) under the four boundary conditions in Eqs. (73) and (74). An approximate analytical expression for the burning velocity is also obtainable, which can explicitly display the functional dependence of the burning velocity on the various system parameters. 4.6.1. Derivation of the approximate expression for m ⱕ 1 To this aim, we should first note that for the usual SHS process the diffusion-controlled reaction prevails, as mentioned in the previous sections, and the Lewis number is very large, say, Le ⬎ 104 in the consumption zone. By A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 assuming diffusion-controlled limit and infinite Lewis number, we have x~ ⬇ ln 1 ⫹ Y~ M z⬇ 81 j Z 0 ⫹ 1 ⫺ Z0 j 82 respectively. The former is derived from Eq. (9) by setting the reduced Damko¨hler number A to be infinity. The latter is derived by setting the last part of the RHS in Eq. (68) to be zero, because of the excessively slow rate of diffusion as compared to thermometric conductivity and reaction rate. Numerical results obtained by using Eqs. (81) and (82), instead of Eqs. (9) and (68), are even found to be practically indistinguishable. By introducing these relations, Eq. (67) is expressed in integral form as ! Z1 T~ d exp ⫺ du T~ um 1 ⫹ gL0 rt;0 Z1 rt 0 t ⫺ j dj f⫺j 1 ⫺ j 1=3 ln 1 ⫹ j⫹g (83) where 1 f ; m Z0 mk ⫹ f g 1 ⫺ Z0 m 1 ⫺ k 84 27 and the complementary error function 2 Z∞ exp ⫺t2 dt erfc z p p z 89 In integrating the RHS of Eq. (83), an attempt [37] has been made to express u with 1 ⫺ j and 1 ⫺ j 1=3 ; and it is reported that by comparing the numerical results, the relation 1 ⫺ u 1 ⫺ um 1 ⫺ j 2 90 fairly represents the temperature profile at high temperatures (cf. Fig. 17), which is needed for the fair evaluation of the Arrhenius nature for the liquid-phase mass-diffusion, to determine L 0. Note that the proportionality constant is the arithmetic mean of the constants determined at j 0 and 1. When Eq. (90) is substituted into Eq. (83) and the first two terms of the following mathematical formula is used ln 1 ⫹ z ( 2 z z3 z5 ⫹ ⫹ ⫹… 3 2⫹z 3 2 ⫹ z 5 2 ⫹ z5 91 ) z2n⫺1 … ⫹ ⫹ ; 2n ⫺ 1 2 ⫹ z2n⫺1 The integral Iu in the LHS of Eq. (83) is evaluated by performing an asymptotic expansion for large values of the Zeldovich number b T~ d T~ ∞ ⫺ T~ 0 T~ ∞2 85 and retaining up to the second term in the expansion. That is, by using the relation o T~ d T~ n d 1 ⫹ a 1 ⫺ u ⫹ a2 1 ⫺ u2 ⫹ … T~ T~ ∞ " # b 3 1 2 ⫹ a2 1 ⫺ u ⫹ ⬇ ; a 4 2a we have p p 3b p exp ⫺ Iu ⬇ 4a 2b a=b 1 1 ⫹ 2a 1 ⫺ um p erfc p ⫺ erfc 2 a=b 2 a=b 86 87 where a T~ ∞ ⫺ T~ 0 ; T~ ∞ 88 Fig. 17. Temperature profile in the reaction zone, with the mixture ratio m taken as a parameter [37]. Solid curves are numerical results and dashed curves the assumed profiles to obtain the approximate analytical expressions for the mass burning rate eigenvalue L 0. 28 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 p Fig. 18. Nondimensional burning velocity L0 for Ti–C system [37]; (a) as a function of the mixture ratio m ; and (b) as a function of the degree of dilution k . The initial temperature T0 is taken as a parameter. Solid curves are numerical results and dashed curves are those with analytical expression. initial temperature T0, the mixture ratio m , and the degree of dilution k are operating parameters when the experimental conditions are specified. Keeping only the leading terms in Iu and Ij in Eqs. (87) and (92), respectively, the functional dependence of L 0 on the system parameters can be displayed as the integral Ij in the RHS of Eq. (83) is expressed as " ! 1 ⫹ um 1 f ⫺ 1 Ij ⬇ f ⫹ g 2 ⫺ ⫹ 5 f ⫹ g 2a 2 ( ! p ) p ⫺1 2 ⫺ a 1 ⫹ a 3 3 p ln p ⫺ ⫺ 2 3 tan 3 f 3a f ⫹ 2g ⫹ 1 ⫹ 6b ( ! p )# p ⫺1 2 ⫹ b b ⫺ 13 3 p × ln p ; ⫺ ⫹ 2 3 tan 3 f ⫹ 2g 3b 92 where a3 f ⫺ 1; and b3 f ⫹ 2g ⫹ 1: Then, from Eqs. (87) and (92), we have r Iu I L0 t;0 1 ⫺ Z 0 u ; rt 1 ⫹ gIj Ij 93 which is found to be surprisingly close to that obtained by a numerical calculation. This good agreement further demonstrates the importance of the describing processes well at the hot boundary, rather than the cold boundary (cf. Fig. 17). Since Eq. (93) is relatively simple, its use could facilitate analysis and interpretation. 4.6.2. General behavior An explicit form of L 0 provides useful insight into the effects of the reduced parameters (a , b , u m, f , and g ), which depend on the system parameters (fst, Td, Tm, T0, m , and k ) used in the experiments. Among these parameters, the stoichiometric mass ratio fst, the activation temperature Td in the mass diffusivity, and the melting point Tm of the lower melting-point metal are physicochemical parameters when the two components of the mixture are specified. The L0 ⬃ m 2 1 ⫺ k2 exp ⫺b=a m ⫹ fst 1 ⫹ fst ⫺ k 1 ⫺ mb 1 ⫹ um 94 As far as m and k are concerned, L0 ⬃ m2 = m ⫹ fst for 2 k 0; and L0 ⬃ 1 ⫺ kp for m 1: That is, the nondimen sional burning velocity L0 decreases with decreasing m and 1 ⫺ k: Its quantitative dependence on m and 1 ⫺ k; however, is more complicated because volumetric heat generation, which is a function of m and 1 ⫺ k; also influences a and b through T∞. The effect of T0, which determines T∞, may also be significant because T∞ can exponentially influence L 0 through a and b . Note that b also depends on the choice of the mixture components. When we compare L 0, Ti–C for Ti–C system with L 0, Ti–B for Ti–B system, (L 0, Ti–B/L 0, Ti–C) 1/2 is about 4 for m 1; k 0; and T0 450 K; because of the smaller value of b for Ti–B system. 4.6.3. Comparisons with numerical results Applicability of the analytical expression has been examined by comparing it with the numerical results [32,47] for TiC, TiB2, ZrB2, HfB2, and TiCo systems in Sections 4.4 and 5.5, and a fairpagreement has been reported in Ref. [37]. Fig. 18(a) shows L0 for Ti–C system as a function of m , with T0 taken as a parameter. Solid and dashed curves are respectively numerical and analytical results.pExcept near stoichio metry m 1; both results show that L0 decreases until m reaches the limit given by Eq. (77), and that they mostly A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 29 p Fig. 19. L0 for Ti–B system. Notation is the same as that in Fig. 18. agree well with each other. Although the trend in the analytical expression near m 1 is opposite to that of the numerical result, the quantitative comparison can be considered to be fair (⬍7% error). This discrepancy is attributed to the overestimated temperature profile given by Eq. (90), which causes an overestimation of Ij , and hence an underestimation of L 0. Fig. 18(b) shows the similar comparison with respect to k . The analytical expression again agrees well with the numerical results, to be within 7%. Figs. 19 and 20 show similar comparisons for Ti–B and Co–Ti systems. The extent of agreement is similar to that for Ti–C system. 4.6.4. Approximate expression for m ⱖ 1 When the mixture ratio is over-stoichiometric, the normalized concentration z of M-species is z⬇ Z∞j ; Z 0 ⫹ Z∞ ⫺ Z0 j 95 then, we have r I fI u L 0 t;0 Z∞ ⫺ Z0 u ; rt Ij 1 ⫹ gIj 96 where the relation has been made use of in Eq. (61). The integral Iu is the same as that in Eq. (87), while the integral Ij is expressed as [40] ! ! 1 ⫹ um f⫹g Ij ⬇ 2 f2 " 3 ⫺ 5f ⫺ 3 1 ⫺ fa2 f ⫹ g × 2 1 ⫺ a2 ⫹ ⫹ 10 f ⫹ g 3b ( ln 2 f ⫹ g f ⫹ 2g b⫺1 a⫹b !3 p ⫺2a ⫹ b p ⫺ 2 3 tan⫺1 3b p 2⫹b ⫹2 3 tan⫺1 p 3b !)# p Fig. 20. L0 for Co–Ti system. Notation is the same as that in Fig. 18. : ! (97) 30 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 p Fig. 21. L0 for Ti–C system as a function of m when T0 450 K; with k taken as a parameter. Solid curves are numerical results and dashed curves are those with analytical expressions. p Dependence of L0 for Ti–C system on m in its wide range is shown in Fig. 21, with k taken as a parameter. The extent of agreement in m ⱖ 1 is similar to that in m ⱕ 1: 4.7. Some remarks on the flame propagation in the adiabatic condition Calculated results for several ceramics and intermetallic compounds, such as Ti–C, Ti–B, Zr–B, Hf–B, Al–Ni, Co– Ti, Ni–Ti systems, using consistent sets of physicochemical parameters, respectively, show satisfactory agreement with experimental results in the literature, if we use the heterogeneous theory in the adiabatic condition. An important relation, that the burning velocity is inversely proportional to the nonmetal particle size, has been discovered. In addition, these comparisons suggest that the nonmetal (or higher melting-point metal) particles in these systems react in the diffusion-controlled limit, indicating that finite rate chemical reaction is not the controlling factor in the propagation of the bulk flame. Furthermore, it has been suggested that concentration of the molten metal is mainly affected through consumption, because of the excessively slow rate of diffusion compared to the thermometric conductivity and reaction rate, due to excessively large Lewis number in the combustion wave. These identifications further enable us to obtain an approximate expression for burning velocity, which can show clear dependence of the burning velocity on the dominant parameters. 5. Flame propagation in the nonadiabatic condition The heterogeneous theory [32] described in Section 4 is an adiabatic one, and hence is incapable of describing the experimentally observed phenomena of flame extinction due to heat loss from the high-temperature combustion process. From a practical point of view, extinction is to be properly described and hence controlled in the SHS process. Toward this aim, it is required to extend the adiabatic theory to allow for the volumetric heat loss, to properly describe flame extinction and the various parameters that characterize it. It will be emphasized here that not the reaction, but the temperature sensitive Arrhenius liquid-phase mass-diffusion is responsible for the flame extinction, because in the SHS process, combustion is basically diffusion controlled. 5.1. Governing equations and boundary conditions A one-dimensional model with heat loss has been used to discover the dominant parameters which determine the limit of steady flame propagation. Although heat loss perpendicular to the propagation direction might lead to the curvature of the combustion wave, a choice of this model simplifies the formulation and hence meets the present purpose. Therefore, regardless of the existence of heat loss, the governing equations (except for the energy conservation equation) have the same form as those in the adiabatic condition, that is Eq. (46) for N-consumption and Eq. (49) for M-conservation [64]. The energy conservation equation is now not Eq. (57), but Eq. (58). Defining the nondimensional variable and parameter Z ∞ ⫺ Z0 ma l=cL 98 s F u 2 0 x; l=c m a q T~ ∞ ⫺ T~ 0 ; in addition to those in Eqs. (66), (71) and (72), we have the governing equations as " # ~ m Le 0 rt;0 =rt exp T~ d =T dz ma ds Z0 ⫹ Z∞ ⫺ Z0 j Z∞ Z0z z⫺ ; 99 j⫹ Z∞ ⫺ Z0 Z∞ ⫺ Z0 d2 u m d u⫺j F u ⫺ ; m a ds Z ∞ ⫺ Z0 ds 2 Z ∞ ⫺ Z0 2 100 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 dj ds " x~ ~ m=m a L0 rt;0 =rt exp T~ d =T 31 # 1=3 Z∞ ⫺ Z0 1⫺ j : 1 ⫺ Z0 101 Eqs. (99)–(101), together with the mass burning rate m, are to be solved subject to the boundary conditions s0: s!∞: u u m ; z 0; j 0; 102 du F u dz dj ; ⫺ 0; 0: m=m a Z∞ ⫺ Z0 ds ds ds 103 In the same manner as that in the adiabatic flame propagation, the cold boundary difficulty is eliminated by specifying that the reaction is initiated at the melting point of M. The downstream boundary condition for u , based on convection-loss balance, requires a short domain of integration, and is preferred over the statement of T T 0 u 0; which would require an excessively large domain of integration. If we further assume the heat loss to be Newtonian, it is expressed as F u C·u; where C is called the heat loss parameter. 5.2. Characteristics in the nonadiabatic condition Values of the physicochemical parameters employed are those for Ti–C system and the Lewis number Le0 is set to be 100 [32] as described in Section 4.4. Representative flame structure for the stoichiometric mixture m 1 without dilution k 0 has already been shown in Fig. 5 by solid curves, when the nonadiabatic situation is close to the extinction state. In contrast to the adiabatic case in which temperature increases monotonically to the adiabatic flame temperature, the temperature gradually decreases to the ambient value after the maximum flame temperature. Decrease in the temperature u reduces the particle consumption rate dj /ds and hence the normalized burning-rate m/ma. The reactant consumption zone is then broadened due to heat loss. By obtaining profiles of the four terms constituting Eq. (100), which, respectively, represent diffusion d 2u /ds 2, convection ⫺ m=ma = Z∞ ⫺ Z0 du=ds; consumption m=ma = Z∞ ⫺ Z0 dj=ds; and heat loss ⫺F u= Z ∞ ⫺ Z0 2 ; a further examination has been made to identify the flame structure for the nonadiabatic case in Fig. 5. As shown in Fig. 22, the combustion wave consists of a relatively thin zone of diffusion and consumption/convection, which is followed by a much thicker zone of convection and heat loss. The existence of an extended zone of downstream heat loss is similar to the nonadiabatic homogeneous gaseous flame. Fig. 23 shows the variation of the normalized mass burning rate m/ma,st as a function of the normalized heat loss C ⴱ m a =ma;st 2 ·C with m and k taken as parameters. Fig. 22. The representative profiles of the diffusion, consumption, convection, and loss terms in the energy conservation equation, along the nondimensional distance [64]. Conditions are the same as those in the nonadiabatic situation in Fig. 5; m=ma 0:6 and C 3:347 × 10 ⫺2 : Here, ma,st is the mass burning rate for the stoichiometric mixture in the adiabatic condition. Results exhibit the characteristic extinction turning point behavior, with the upper branch being the stable solution and the turning point designating the state of extinction. With decreasing m or increasing k , the mass burning rate is gradually reduced, until extinction occurs. Since not only m/ma,st but also (ma/ ma,st) 2·C strongly depends on the mass burning rate m, the parameters m/ma and C are usually used to obtain the extinction condition; the values of (m/ma)cr corresponding to flame extinction ranges from 0.57 to 0.70, depending on m and k . Here it may be informative to note the significant implication of Eq. (81). Since extinction is frequently associated with the nonlinear, highly temperature-sensitive nature of the Arrhenius kinetics, the disappearance of its influence in the diffusion-controlled limit implies that extinction as shown in Fig. 23 is actually caused by an alternate factor, which also has a nonlinear influence on flame propagation. This factor is the liquid mass-diffusivity D D 0 exp ⫺Td =T which varies in a sensitive, nonlinear, Arrhenius manner with temperature. Thus in the SHS process, even in the diffusion-controlled limit, combustion behavior, as well as the particle regression rate, is still an Arrhenius function of temperature, with the large activation temperature for diffusion, Td, exerting the dominant temperature sensitivity. Fig. 24(a) and (b), respectively, show the effect of m and k on the critical heat loss parameter Ccr ; with the initial temperature T0 taken as a parameter. The change in Ccr is gradual when either m or k is away from the flammability limit, while Ccr decreases rapidly as this limit is approached. It is also seen that Ccr increases with increasing T0 for a given m or k . This suggests that even if a heat loss exists, flame 32 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 23. Normalized mass burning rate m/ma,st as a function of the normalized heat loss C ⴱ m a =ma;st 2 ·C; with the mixture ratio m and the degree of dilution k taken as parameters when the initial temperature T 0 450 K [64]. The values of numerically obtained (m/ma,st)cr corresponding to the turning point are indicated by the horizontal solid lines; those of analytical ones are indicated by the dashed lines. propagation can be sustained when T0 is high. Note that there exists a maximum in C cr at high T0. 5.3. Turning-point determined by the thermal theory The heterogeneous SHS flame propagation has some special features, as shown in Eqs. (81) and (82), which can be contrasted with the homogeneous gaseous flame propagation. Nevertheless, an explicit relation for the state of extinction can be derived [64] in the same manner as that for the homogeneous premixed flame [65]. In the thin diffusion consumption/convection zone, contribution of the heat loss term is very small, such that Eq. (100) becomes du m u⫺j : 104 ⬇ ma ds Z ∞ ⫺ Z0 Eliminating s from Eq. (104) with Eq. (101) and conducting integration, we have ! rt;0 T~ d exp⫺b 1 ⫺ u m 2 exp L0 ma b rt T~ ∞ × Zj 0 u⫺j 1=3 dj; Z∞ ⫺ Z0 Z ∞ ⫺ Z0 1 ⫺ j x~ 1 ⫺ Z0 105 where the relation has been used in Eq. (86), retaining up to the first term. When je ⬇ 1 at the downstream edge of the consumption zone, by taking account of the adiabatic condition, that is u j 1 and m m a ; we obtain the following relation to the lowest order: m 2 exp⫺b j e ⫺ ue : 106 ma Fig. 24. Critical heat loss parameter C cr [64]: (a) as a function of the mixture ratio m when the degree of dilution k 0; and (b) as a function of k when m 1:0: The initial temperature T0 is taken as a parameter. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 33 Fig. 25. Comparisons between numerical and analytical solutions for the critical value of the mass burning rate ratio (m/ma)cr [64]; (a) as a function of m ; and (b) as a function of k . The initial temperature T0 is taken as a parameter. On the other hand, if we integrate and evaluate Eq. (100) at the downstream edge of the consumption zone, we have V=b j e ⫺ ue ; m=m a 2 where (V /b ) is expressed as Zu e V f u du ⫹ f ue ; b u 0 107 108 by the use of the upstream relation du=ds m=m a u= Z∞ ⫺ Z0 and the downstream boundary condition (103). Combining Eqs. (106) and (108), we obtain the relation m 2 m 2 ln ⫺V; 109 ma ma from which we have m=m a cr e ⫺1=2 0:6065 at the turning point, corresponding to the state of extinction. Fig. 23 shows that this analytical extinction state fairly represents the numerical value, as the first approximation. Fig. 25(a) and (b) further show (m/ma)cr as functions of m and k , with T0 taken as a parameter. Although the numerical and analytical solutions for (m/ma)cr differ by 15%, this difference is reduced to less than 4% in determining C cr. A more precise treatment will be given in Section 5.6. 5.4. Experimental comparisons for Ti–C system There exist several sources of experimental data which can be compared with the calculated results, as in Refs.[64,66], by keeping as many parameters as fixed possible. For the heat loss, only thermal radiation is considered, because of the high temperature in the consumption zone, say, higher than 2000 K. Then the lateral heat loss for a cylindrical compact with radius r can be expressed as 4h rad L 110 T ⫺ T0 ; 2r where hrad es SB T 2 ⫹ T02 T ⫹ T0 111 is a radiative heat-transfer coefficient, e the emissivity, and s SB the Stefan–Boltzmann constant. As for the heat loss parameter C , it is expressed as # 2! " l 4es SB T 2 ⫹ T02 T ⫹ T0 R0 C : 112 rtc 2r rt c u0;a R0 2 Note here that C depends not only on the physicochemical parameters, but also on the particle radius R0 and the compact diameter 2r. It is also inferred that R 02 = 2r; which is sometimes called the “Heat Loss Index”, can be a useful parameter in correlating size effects related to the heat loss, from a practical point of view. This parameter R20 = 2r can be understood as follows, from a physical point of view; the increase in R0 decreases the total surface area of N particles in the combustion wave, thereby reducing the volumetric heat generation rate; the decrease in r increases the surface-volume ratio of the compact, thereby enhancing heat loss rate. 5.4.1. Range of flammabilty and extinction limit Predicted results are compared with experimental data for Ti–C system [64]. In addition to the values of physicochemical parameters, described in Section 4.1, use has been made of l= rt c 10⫺5 m2 =s [41], e 0:4 [41], and s SB 5:67 × 10 ⫺11 kJ= m2 :s:K4 : The compact radius is set to be r 9 mm and its relative density 0.6, in accordance with the experiments [33,43]. As for the heat loss parameter C , it is evaluated at the maximum temperature of the combustion wave. Fig. 26(a) shows the range of flammability when T 0 450 K; the abscissa is the mixture ratio m and the ordinate the particle radius R0. Extinction is anticipated to occur when C in Eq. (112) reaches the critical heat loss parameter C cr in Fig. 24(a). We see that while the range of flammability spans from 0.448 to 1 when R 0 ! 0; it contracts with 34 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 26. Nonadiabatic combustion behavior for Ti–C system [64]; solid curves are calculated from theory; data points are experimental [33,43]. (a) Range of flammability with respect to the mixture ratio m and the particle radius R0; (W) designates the steady propagation, (Z) the pulsating combustion, (K) the flame extinction during the propagation, and ( × ) the non-ignition. (b) SHS rate-constant u0·R0 as a function of m , with R0 taken as a parameter. increasing R0. It also shows that the combustion wave cannot propagate through a compact when R0 is larger than 20.9 mm. Experimental results [33,43] are also shown in Fig. 26(a). Outside the range of flammability, even if the flame propagation might be initiated by a powerful ignition source, it loses its self-sustained nature and is extinguished in the course of the propagation. Fair agreement is demonstrated for particles smaller than 7.5 mm, while discrepancy becomes large for R0 12:5 mm: This discrepancy is attributed to the effect of smaller particles, because of the existence of particle-size distribution. It is also pointed out that the extinction limit obtained here corresponds to that for the steady flame propagation. When the process occurs in an unsteady mode, due to disappearance of stability of the steady propagation, the limit of self-sustained combustion could be modified accordingly. Fig. 26(b) shows the SHS rate-constant u0·R0 in the adiabatic condition as a function of m . The agreement is again fair, as far as the trend and approximate magnitude are concerned, suggesting that u0·R0 is one of the important parameters in correlating experimental results. The range of flammability with respect to the degree of Fig. 27. Nonadiabatic combustion behavior for Ti–C system [64]. (a) Range of flammability with respect to the degree of dilution k and the particle radius R0; notation is the same as that in Fig. 26(a). (b) u0·R0 as a function of k , with R0 taken as a parameter. dilution k is shown in Fig. 27(a), and Fig. 27(b) shows u0·R0 as a function of k . The same comments as those for Fig. 26(a) and (b) can be made. Fig. 28(a) shows the range of flammability for various m and k , with R0 taken as a parameter [66]. The range of flammability becomes narrow with increasing R0. Experimental results [33,43] in Fig. 28(a) are those of R0 5 mm: Fig. 28(b) shows a similar comparison [40] for the dependence of the range of flammability on m in its wide range when R0 12:5 mm: In the range from m 1 to 2, the range of flammability with respect to k is wider than that for the stoichiometric mixture. This is attributed to the increase in u0,a·R0 (cf. Fig. 9), which reduces C , and hence retards the flame extinction. Discrepancy between the predicted and experimental results for the under-stoichiometric mixtures m ⬍ 1 is attributed to the effect of smaller particles because of the existence of particle size distribution. Discrepancy for the over-stoichiometric mixtures m ⬎ 1 might be attributed to an initiation of a radial crack, which inevitably appears in pulsating combustion [1–3] resulting in a laminated structure with separable layers. Munir and Lai [67] measured the boundaries of the SHS for Ti–C system at several ambient (or initial) temperatures. Their test specimen (19 mm in diameter and 0.65 in relative density) is composed of tens of layers (5 mm in thickness) A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 35 with increasing initial temperature T0, as expected, due to preheating. This suggests that even if there is heat loss, flame propagation can be sustained when T0 is high. It is also seen that for a given T0, the range of flammability narrows with increasing R0. Shkiro and Borovinskaya [26] reported experimental data for Ti–C system with varying degree of dilution k and compact diameter 2r. The relative density is 0.56. Fig. 30(a) shows the range of flammability as functions of k and 2r, with T0 taken as a parameter. The carbon particle size and the initial temperature are estimated to be R 0 2:5 mm and T0 300–450 K; respectively, by examining the dependence of the reported burning velocity on k and T0. It is seen that with increasing k , the compact diameter in which the combustion wave can propagate increases first gradually and then rapidly. Further, an increase in T0 expands the range of flammability. Since it has been found [66] that their experimental data correspond well to the calculated result for T0 300 K when k is less than about 0.2, while the comparison is better for T0 450 K near the limit of flammability, by examining the dependence of u0 on k and T0, further confirmation has been made by investigating the effect of the compact diameter 2r on the burning velocity u0. Fig. 30(b) shows that u0 decreases with decreasing 2r Fig. 28. Range of flammability for Ti–C system as functions of m and k , with R0 taken as a parameter: (a) for R0 5 mm [66]; (b) for R0 12:5 mm [40]. Data points are experimental [33,43] and notation is the same as that in Fig. 26(a). with a gradual increase in the degree of dilution. The average diameter of the carbon particles is reported to be 5 mm. Fig. 29 shows the limit of dilution with the radius of carbon particles, R0, taken as a parameter. The range of flammability expands Fig. 29. Range of flammability for Ti–C system as functions of k and T0, with R0 taken as a parameter [66]. Data points are experimental [67]; (W) designates the steady propagation, (L) the unsteady propagation, and ( × ) the flame extinction. Fig. 30. Nonadiabatic combustion behavior for Ti–C system [66]. (a) Range of flammability as functions of k and the compact diameter 2r, with T0 taken as a parameter, when R0 2:5 mm; data points are experimental [26] and notation is the same as that in Fig. 26(a). (b) Burning velocity u0 as a function of 2r, with k taken as a parameter; T0 450 K: 36 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 because the heat loss rate is enhanced due to an increase in the surface-to-volume ratio. Fair agreement is demonstrated between the experimental and predicted results. 5.4.2. Burning velocity under heat loss condition Here, let us investigate the effect of the heat loss on the burning velocity. As pointed out in Section 5.4, not only the physicochemical properties, but also sizes of N particles and test specimen influence C through Eq. (112), hence the burning velocity. Fig. 31(a) shows the SHS rate-constant u0·R0 as a function of 2r=R 02 ; which is the inverse of the heat loss index, for T0 450 K and k 0; with m taken as a parameter [40]. With decreasing 2r=R20 ; u0 ·R0 decreases because of the increase in the heat loss rate, and reaches the extinction limit shown by a dashed curve. We see that effects of the compact diameter 2r and the particle radius R0 on the burning velocity u0 can generally be discussed by the use of parameters 2r=R20 and u0·R0. Experimental results [33,43] for various R0 are also shown in Fig. 31(a). Fair agreement is shown, as far as the general trend and approximate magnitude are concerned. Fig. 31(b) shows a similar comparison for the stoichiometric mixture m 1; with k taken as a parameter. Again, usefulness of the SHS rate- Fig. 31. SHS rate-constant u0·R0 for Ti–C system as a function of 2r/ R02 for T0 450 K [40]: (a) with m taken as a parameter; and (b) with k taken as a parameter. Data points are experimental [33,43] for various sizes of N particles. constant u0·R0 and the heat loss index R20 = 2r is demonstrated, as the correlating parameters for experimental results. 5.5. Experimental comparisons for other systems The nonadiabatic theory can also be applied to other experimental results obtained from the synthesis of several borides and intermetallic compounds for which reliable experimental data are available. Through extensive comparisons [64,66] for the range of flammability with respect to effects of dominant parameters (m , k , T0, R0, and 2r) on the burning velocity u0, it has been demonstrated that the theory is indeed viable, thereby allowing the synthesis of the diverse set of experimental results within a coherent framework of description. In the same manner as that for the Ti–C system, only thermal radiative heat loss is considered, because of the high temperature in the combustion wave, say, higher than about 1000 K. 5.5.1. Boride synthesis Munir and Sata [68] measured the boundary of SHS in the boride synthesis of Ti, Zr, and Hf at several ambient (or initial) temperatures. The test specimen was prepared in the same manner as that explained for Fig. 29, with amorphous boron of about 1 mm diameter and crystalline boron particles whose diameter is less than 42 mm. Fig. 32(a) shows the limit of dilution for Ti–B system [64] obtained Fig. 32. Range of flammability for Ti–B system; (a) as functions of the degree of dilution k and the initial temperature T0, with the initial particle radius R0 taken as a parameter [64]; and (b) as functions of k and R0, with T0 taken as a parameter [66]. Data points are experimental [68] and notation is the same as that in Fig. 29. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 37 parameter. The same trend as that in Fig. 27(a) can be observed and the range of flammability expands with increasing T0. Experimental data with crystalline boron particles have been plotted for R0 5 mm because sieved powders of ⫺325 mesh (⬍44 mm) size are estimated to have about 10 mm in diameter. The deviation at large particles could be caused by the difference of crystallization and/or the existence of small particles, as well as the experimental determination of T0, as inferred in Section 4.5. Figs. 33 and 34 show similar comparisons for Zr–B system [45,50,68] and Hf–B system [68]. Extent of agreement is similar to that for the Ti–B system. Fig. 33. Range of flammability for Zr–B system [66]; (a) as functions of k and T0, with R0 taken as a parameter; data points are experimental [68] and notation is the same as that in Fig. 29; (b) as functions of k and R0, with T0 taken as a parameter; datum for k 0 is from Hardt and Holsinger [45]; data for k 0:3 and 0.5 are from Zenin et al. [50]. by the use of physicochemical parameters in Table 2 and the transport parameters in Table 3. The same trend as that in Fig. 29 is observed. Data points are experimental with boron particles of about 1 mm diameter. It is pointed out that the limit of flammability observed by Munir and Sata [68] could have been affected by other effects not considered in the theory such as agglomeration of the particles, as inferred in Section 4.5.2. Fig. 32(b) shows the range of flammability for Ti–B system [66] as functions of k and R0, with T0 taken as a Fig. 34. Range of flammability for Hf–B system as functions of k and T0, with R0 taken as a parameter [64]. Data points are experimental [68] and notation is the same as that in Fig. 29. 5.5.2. Synthesis of intermetallic compounds Naiborodenko and Itin [58] measured the burning velocities for stoichiometric Al–Ni system by varying the diameter of compact whose relative density is 0.4. The diameter of the Ni powder was reported to be of the order of 20 mm. Fig. 35 shows that u0 decreases with decreasing compact diameter because of the enhanced heat loss rate; the boundary of the steady propagation is shown by a dashed curve [66]. The agreement is fair for R0 15 mm: Itin et al. [53] measured the burning velocities for Co–Ti system by varying the compact diameter and the particle size of Ti. The compact for the stoichiometric mixture had a relative density of 0.6; diameters of Ti particles were from 100 to 160 mm; initial temperature was 573 K. Fig. 36(a) and (b), respectively, show the range of flammability and the burning velocity [66]. We see that the range of flammability expands with increasing compact diameter and that the burning velocities agree well for the 130 mm particles, which is representative of the average particle diameters. Fig. 35. Burning velocity u0 in the nonadiabatic condition for Al–Ni system as a function of the compact diameter 2r, with the initial particle radius R0 taken as a parameter [66] the mixture is stoichiometric and the initial temperature T0 300 K; data are from Naiborodenko and Itin [58]. 38 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 36. Nonadiabatic combustion behavior for Co–Ti system with stoichiometric mixture at T0 573 K [66]. (a) Range of flammability as functions of 2r and R0; data are from Itin et al. [53] and notation is the same as that in Fig. 29. (b) Burning velocity u0 as a function of 2r, with R0 taken as a parameter; data points are experimental [53]. Bratchikov et al. [60] measured the burning velocities for stoichiometric Ni–Ti system with a relative density of 0.6 by varying the compact diameter and the particle size of Ti. The initial temperature was 573 K. Fig. 37(a) shows the range of flammability as functions of 2r and R0 [66]. Particle size of Ti is estimated from their related paper [70] which states that Ti particles were from 100 to 160 mm in diameter. Fig. 37(b) shows that the burning velocities are correlated fairly well with the predictions for R0 100 mm: Fig. 37. Nonadiabatic combustion behavior for Ni–Ti system with stoichiometric mixture at T0 573 K [66]. (a) Range of flammability as functions of 2r and R0; data are from Bratchikov et al. [60]; datum for R0 25 mm is from Itin et al. [71]; notation is the same as that in Fig. 29. (b) Burning velocity u0 as a function of 2r, with R0 taken as a parameter; data points are experimental [60]. for the temperature, as the outer solutions, in the upstream and downstream regions where the consumption term dj /ds is exponentially small: C C a0 u 1 ⫹ 1 ⫹ 22 ⫹ … exp s s ⬍ 0; b Z∞ ⫺ Z0 b 113 C C 1 bC u 1 ⫹ 1 ⫹ 22 ⫹ … exp ⫺ s Z∞ ⫺ Z0 a 0 b b 5.6. Approximate expression of the heat loss parameter s ⬎ 0: Extinction of the combustion wave in the SHS process has been obtained numerically and hence presented parametrically. An approximate analytical expression for the extinction criterion [72] which explicitly displays its functional dependence on the various system parameters, is also obtainable by referring to an analytical criterion for the “premixed” SHS flame propagation obtained by Kaper et al. [73] by the use of matched asymptotic expansion method. The analysis and results presented herein is an extension of those described in Section 5.3. Here, b is the Zeldovich number defined in Eq. (85), 5.6.1. Asymptotic expansion analysis Following Ref. [73], we have the following expressions s (114) s m a a ; a0 ⫹ 1 ⫹ 22 ⫹ …; b ma b b 115 and C1 ; C2 ; … are constants. As for the inner solution, with the inner variables being u1⫹ s h ; b t1 t y y ⫹ 22 ⫹ …; j y 0 ⫹ 1 ⫹ 22 ⫹ …; b b b b 116 we obtain the following equations for the first-order inner A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 structure have d 2t1 a0 dy0 ⫹ 0; Z∞ ⫺ Z0 dh dh2 exp ⫺ 117 " # r dy 3b b 1 at 2 ⫺ 1 bL0 t;0 a0 0 exp ⫺ ⫺ a 2 rt dh 4a b 3 2 Z∞ ⫺ Z0 1=3 f ⫺ y 6 Z∞ ⫺ Z0 1 ⫺ Z0 0 7 7; × 1⫺ y0 ln 6 1 ⫹ 5 4 Z ⫺ Z0 1 ⫺ Z0 y0 g⫹ ∞ 1 ⫺ Z0 118 2bC a 02 39 ! a20 k 1 ⫺ M ⫹ M; 125 which has nearly the same form as that presented by Kaper et al. [73] except for the factor k. The parameter M identified in Ref. [74] is called the melting parameter because it corresponds to the energy to heat the unburned mixture from T0 to Tm. Consequently, 1 ⫺ M is the normalized driving force for the adiabatic flame propagation. The turning point corresponding to the flame extinction, then, appears at dC=da 02 0; which yields Xcr 1 ⫹ ln Xcr M; 126 where a is defined in Eq. (88) and f 1=m: Matching the outer and inner solutions in the limit of h ! ^∞; we have as well as the following relawhere ln Xcr tions for the critical parameters at extinction: a0 h ! ⫺∞ : t1 ! C1 ⫹ h; y0 ! 0; Z∞ ⫺ Z0 a0 2cr h ! ∞ : t1 ! C1 ; y0 ! 1; 119 120 where the constant C1 is given by the second-order inner problem as C1 ⫺ 2bC : a 02 121 By using the matching conditions, Eq. (117) can be integrated once to yield dt1 a0 1 ⫺ y0 : dh Z∞ ⫺ Z0 122 Eliminating h from Eq. (122) with Eq. (118) and conducting integration, we have " # Z∞ ⫺ Z0 ZC1 3b b 1 a t1 2 dt1 ⫺ exp ⫺ ⫺ 4a b a 2 bL0 rt;0 =rt a20 ⫺ ∞ Z1 0 1 ⫺ y0 2 3 dy0 : Z∞ ⫺ Z0 1=3 f ⫺ y0 7 6 Z∞ ⫺ Z0 1 ⫺ Z0 7 ln 6 1⫹ y0 1⫺ 5 4 Z∞ ⫺ Z0 1 ⫺ Z0 g⫹ y0 1 ⫺ Z0 123 If we note here that L 0 is expressed in Eq. (93), that Iu in Eq. (87), and that Ij in Eq. (92) for m ⬉ 1 and Ij in Eq. (97) for m ⱖ 1; then Eq. (123) becomes " !# 2a 02 b 1⫺ bIu exp 1 ⫹ um a Z0 C1 at dt1 ⬇ 1 ⫺ exp C1 : exp t1 1 ⫺ 1 b 124 Putting k 2= 1 ⫹ um and 1 ⫺ M bIu exp b=a; we ⫺2bC cr = a0 cr2 ; ⫺Xcr ln Xcr X ln Xcr 2 ; C cr cr : k 1 ⫺ M 2bk 1 ⫺ M 127 Note that hte maximun flame temperature is given by umax 1 ⫹ ln Xcr =b: Knowing the critical heat loss parameter C cr and the maximum flame temperature u max, the critical size of N particles and the critical compact diameter for the extinction are determined through ! C cr rt c u0;a R0 2 R 02 ; 128 2 ⫹ T 2 T 2r cr l= rt c4es SB Tmax max ⫹ T0 0 derived from Eq. (112) when only the thermal radiative heat loss from the cylindrical surface of a compact with radius r is considered [64,66]. It is informative to recall that R20 is closely related to the heat generation rate in the consumption zone and 2r the heat loss rate. In addition, the SHS rateconstant in the adiabatic condition, which is the product of the burning velocity u0,a and the initial radius R0 of N particles, is expressed as ! 1 ⫺ Z0 Iu 2 2 rM u0;a ·R0 3Le0 D0 ; 129 rN Z∞ ⫺ Z0 Ij by combining Eqs. (75) and (96). 5.6.2. Correction term for the heat loss parameter Predictions based on Eq. (126) have been found to deviate substantially from the numerical results shown in Section 5.4. The discrepancies are at least 50% for the heat loss parameter C cr, and are even more for the critical particle size and/or compact diameter. Since this discrepancy is expected to be caused by ignoring the O(b ⫺1) term in obtaining the RHS of Eq. (118), which would have modified the RHS of Eqs. (125) and (126) by an amount ⫺O 1 ⫺ M=b; we are therefore required to add the term (1 ⫺ M)/b to Eq. (126) as 1⫺M X cr 1 ⫹ ln Xcr M ⫺ : 130 b 40 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 have been used. If we further use 1 ⫹ um =2 ⬃ 1; a ⬃ 1; and T~ d T~ ∞ ⫺ T~ 0 T~ d T~ d ⬃ 2 ~ ~ ~ Z∞ ⫺ Z0 T∞ ⫺ T0 T∞ 1 ⫺ Z0 m ⫹ fst ~ T ; Z∞ ⫺ Z0 m 1 ⫺ k d b 135 we have the following relations with respect to m and k : ! ( ) 1 m e⫺2 m 1⫺ m ⱕ 1; C cr ⬃ m ⫹ fst 2eT~ d m ⫹ fst T~ d 136 Fig. 38. Critical value of the normalized heat loss, Xcr exp⫺2bC cr = a0 cr2 ; as a function of the melting parameter M, with the measure of activation energy relative to thermal energy, b , taken as a parameter [72]. Solid curves are numerical results and dashed curves are results by the explicit approximate analytical expression. With this modification, the discrepancies can be reduced to less than 20%. Furthermore, Xcr is now explicitly expressed in terms of M as Xcr p ! 1=2 1 ⫹ 2e 1 1⫺M M2; 1 ⫹ 2e M ⫺ ⫹ 1⫺ e e b 131 which yields a discrepancy of less than 1% (cf. Fig. 38), in the range of practical values of M. 5.6.3. General behavior of the heat loss parameter An explicit form of C cr provides useful insight into the effects of the reduced parameters (a , b , and u m), which depend on the system parameters (fst, Td, Tm, T0, m , and k ), in the same manner as that for the mass burning rate eigenvalue L 0 in Section 4.6.2. Keeping only the leading terms, the functional dependence of C cr on the system parameters can clearly be displayed as C cr ⬃ 1 ⫹ um 4eb 1⫺ 2a e ⫺ 1 ⫺ e : b 132 Here the following relations 1 ⫺ M bIu exp b a ⬃1⫺ 2a b 133 and Xcr ⬃ 1 e 2a ⫺ 1 1⫹ b e 134 C cr ⬃ 1 2eT~ d 1 m ⫹ fst ( 1⫺ e⫺2 T~ d ! 1 m ⫹ fst ) m ⱖ 1 137 when there is no dilution k 0; when the mixture is stoichiometric m 1; we have ! ( ) 1 1⫺k e⫺2 1⫺k : 138 1⫺ C cr ⬃ 1 ⫹ fst 2eT~ d 1 ⫹ fst T~ d That is, C cr increases with increasing m when m ⱕ 1; while it decreases with increasing k or m when m ⱖ 1; although with progressively diminishing rate. The effect of T0, which determines T∞, also appears because T∞ influences C cr through a and b . 5.6.4. Applicability of the analytical expression Applicability of the analytical expression has been examined by comparing it with the numerical results for the syntheses of TiC, TiB2, ZrB2, HfB2, and AlNi, because of fair agreement between numerical and experimental results, as shown in Sections 5.4 and 5.5. Fig. 39(a) shows the critical value of the heat loss parameter C cr for Ti–C system as a function of m , with k 0 and T0 taken as a parameter. Solid and dashed curves are respectively numerical and analytical results. Fig. 39(b) shows C cr as a function of k , with m 1 and T0 taken as a parameter. The analytical result mostly agrees with the numerical result, within 17% error. Fig. 40(a) shows the critical value of the heat loss index R 02 = 2rcr as a function of m , with k 0 and T0 taken as a parameter. The agreements between the numerical and analytical results are generally good, except near the stoichiometric state at which the analytical result shows a reverse trend, resulting in about 18% under-prediction at m 1: The discrepancy here is attributed to the overestimation in Ij in obtaining the approximate expression of L 0, as mentioned in Section 4.6. Fig. 40(b) shows R 02 = 2rcr as a function of k , with m 0:95 and T0 taken as a parameter. When m is 0.95 or less, agreement is good, with less than 14% error. Figs. 41 and 42 show similar comparisons for Hf–B and Al–Ni systems, as examples, when the degree of dilution is A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 41 Fig. 39. Critical heat-loss parameter, C cr, for Ti–C system [72]; (a) as a function of the mixture ratio m , with the initial temperature T0 taken as a parameter when the degree of dilution k 0; and (b) as a function of k , with T0 taken as a parameter when m 1: Solid curves are numerical results and dashed curves are those with the analytical expression. varied. The extent of agreement, including Ti–B and Zr–B systems, is similar to that for Ti–C system. 5.7. Some remarks on the flame propagation under heat loss condition A nonadiabatic, heterogeneous theory for the SHS process with heat loss has been formulated and the range of flammability can be determined from the characteristic extinction turning-point. A particularly significant identification is that the process responsible for its temperature sensitivity and ultimate extinguishment is the Arrhenius liquid-phase mass-diffusion. Fair agreement between predicted and experimental results for the extinction limits, in trend and in approximate magnitude, suggests that this theory has captured the essential features of the heterogeneity in the SHS process. A useful parameter called the “heat loss index”, which is defined as the square of the nonmetal particle size divided by the compact diameter, has also been identified. Furthermore, under the appropriate and realistic assumptions of diffusion-limited reaction and infinite Lewis number, an approximate criterion for the extinction has been derived, which can yield useful insight into manipulating the flame extinction. 6. Other aspects of the nonadiabatic flame propagation In this section, several other important effects, such as the bimodal particle distribution, representative length of the cross-section of a compacted medium, on the SHS flame Fig. 40. Critical heat-loss index [R02/(2r)]cr for Ti–C system [72]; (a) as a function of m , with T0 taken as a parameter when k 0; and (b) as a function of k , with T0 taken as a parameter when m 0:95: Notation is the same as that in Fig. 39. 42 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 41. Critical values of parameters for stoichiometric Hf–B system as a function of k , with T0 taken as a parameter: (a) C cr; (b) [R02/(2r)]cr. Notation is the same as that in Fig. 39. propagation and/or extinction are described. Also described here is the correspondence between the heterogeneous theory and the homogeneous premixed-flame theory, from the analytical point of view. Effects of the electric field on the flame propagation and/or extinction are also described in this section. 6.1. Bimodal particle dispersion Analyses of the heterogeneous flame, hitherto described, have assumed a monodisperse distribution of the nonmetal particles. However, there exist practical interests in having a bimodal particle distribution. For example, such a distribution would allow smaller particles to fill in part of the inter-particle spacing between large particles, hence allowing for a denser, and possibly more desirable, packing. Since not only the increase in the burning velocity but also the extension of the flammability limit is anticipated by reducing the particle size, for a fixed amount of the dispersed phase, it is considered to be desirable to have small particle size for the propagation of the combustion wave. However, preparation of the small particles can be quite energy intensive. Therefore, preparation of the bimodal mixture, with only a fraction of the N particles having small sizes, can be less energy intensive without substantially affecting the burning intensity. Further, since SHS processes have been observed to possess a variety of flame instability phenomena, which are invariably dependent on the burning intensity and hence the particle size, such instabilities could be modulated for bimodal mixtures. In view of the above considerations, Fig. 42. Critical values of parameters for stoichiometric Al–Ni system as a function of k , with T0 taken as a parameter: (a) C cr; (b) [R02/(2r)]cr. Notation is the same as that in Fig. 39. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 formulation has been extended to bimodal systems under nonadiabatic situations [75,76]. 6.1.1. Equivalent particle radius The problem of interest is basically the same as that described in Section 5.1. The bimodal system for N particles is considered to consist of small particles of density r N,S, initial radius RS,0, and number density nS,0, and large particles with the corresponding parameters r N,L, RL,0, and nL,0. In spite of the existence of the bimodal particle distribution, the governing equations and boundary conditions are identical to those for the monomodal case, if we introduce the equivalent particle radius and number density as (n0R0)E and express the conservation equation for the N-Consumption formally as [75,76] d 1 ⫺ Z 4prM D u 1 ⫺ Z 1=3 ⫺ n0 R0 E 0 x : u dx m 1 ⫺ Z0 139 Thus, our primary concern is to find out the specific form of (n0R0)E in Eq. (139), as a function of the bimodal parameters. The description for the bimodal system has first been made for the system with the same kind of N particles [75] and then extended for the system with different kinds of N particles [76]. The parameter (n0R0)E can be obtained by using the size distribution function rN G0 R rN;S nS;0 d R ⫺ RS;0 ⫹ rN;L nL;0 d R ⫺ RL;0 ; 140 where d (·) is the delta function. By substituting it into the general N-consumption equation in Eq. (45), we have " ! !2 # rN;L RL;0 n0 R0 E 1 ⫹ PG rN;S RS;0 1⫹G 1 ⫹ P 3G 1=3 n L;0 RL;0 141 where G 3 r N;S nS;0 RS;0 rN;L nL;0 R3L;0 142 is the mass ratio, P hS =RS;0 = hL =RL;0 the size ratio, and hi =Ri;0 i S; L the normalized particle size defined as !2 hi 2 Zx x dx i S; L: 143 1⫺ 2 Ri;0 Ri;0 ⫺ ∞ u If we further take account of the relation for the mass fraction of N particles as 4 3 rt;0 1 ⫺ Z0 144 prN;L 1 ⫹ G n L;0 RL;0 3 the mass burning rate eigenvalue becomes " # Z∞ ⫺ Z0 2 m2a L0 4p rM D0 l=c n0 R0 E # " Z∞ ⫺ Z0 2 1 ⫺ Z0 u20;a RL;0 =F2 1 ⫺ Z0 3Le0 D20 rM =rN;L where RL,0/F is the equivalent particle radius and ! !2 rN;L RL;0 1⫹ PG rN;S RS;0 2 F : 1 ⫹ G 2=3 1 ⫹ P3 G 1=3 43 145 146 We see that u0,a·(RL,0/F) is independent of the bimodality, and that Eq. (143) suggests P O R S;0 =RL;0 : Note that G= 1 ⫹ G is the mass fraction of small particles and can be called the degree of small particle content. As for the heat loss parameter C , it is given as #" 2 # " RL;0 =F2 l 4es SB T 2 ⫹ T02 T ⫹ T0 C : rtc 2r rt c u20;a RL;0 =F2 147 It depends not only on the physicochemical parameters but also on the equivalent particle radius RL,0/F and the compact diameter 2r. 6.1.2. Experimental comparisons for spherical carbon particles There exist several sources that can be compared with the numerical results. Makino et al. [43,77,78] using various radii of spherical carbon particles, measured the range of flammability and the burning velocity of the SHS process for Ti–C system. The representative Lewis number of Le0 25; which corresponds to l= rt c 5 × 10⫺6 m2 =s and has been determined through correlation similar to that in Section 4.4 with experimental results [43,77,78] (0.6 relative density) of monomodal system, has been used in numerical calculations [75] because of the fair correlation for various radii of spherical carbon particles and various compact diameters, as shown in Figs. 43 and 44. Fig. 43(a) shows the range of flammability for the monomodal system with respect to the mixture ratio m and the heat loss index R20 = 2r; with T0 taken as a parameter [79]. Experimental results [43,77,78] for various particle radii (1.5, 2.5, 5, 10, and 15 mm) and various compact diameter (5, 7, 10, and 18 mm) at T0 300 K show that not only the range of flammability but also the range of pulsating combustion extends with decreasing R20 = 2r: Fig. 43(b) shows the similar comparison with respect to the degree of dilution k and R20 = 2r: Fair agreement is demonstrated between the predicted and experimental results. The choice of Le0 25 for the spherical carbon particles requires further comments because Le0 100 has been used in comparisons in Section 4.4 and 5.4. As a remarkable difference, compared to the irregular particles in Refs. 44 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 43. Range of flammability for Ti–C system with monomodal particles; (a) with respect to the mixture ratio m and the heat-loss index R02/ (2r), with the initial temperature T0 taken as a parameter [79]; and (b) with respect to the degree of dilution k and the heat-loss index R02/(2r). Data points are experimental at T0 300 K [43,77,78]: (W) designates the steady-propagation; (Z) the pulsating combustion; (S) the quasisteady, multipoint combustion; (K) the extinction during the flame propagation, and ( × ) the non-ignition. [33,43], the spherical particles are reported to have smooth surfaces and very narrow size distribution [77]. This suggests that the value of Le0 100 for the irregular particles implicitly accounts for the effects of surface roughness and the relatively wide size distribution because these effects, which appear in (n0R0)E in Eq. (139), are grouped into l /c when only the particle size is known. Fig. 44(a) shows the SHS rate-constant u0·R0 for the monomodal system as a function of the inverse of the heat loss index R20 = 2r at T0 300 K; with m taken as a parameter [79]. The trend is the same as that in Fig. 31(a). Data points are experimental [43,77,78] and the agreement is again considered to be satisfactory. Fig. 44(b) shows a similar comparison, with k taken as a parameter. Having confirmed the appropriateness of the choice of several physicochemical parameters, comparisons for the bimodal system have been conducted, with the size ratio P taken to be P RS;0 =RL;0 : Fig. 45(a) shows the range of flammability for the bimodal system with respect to the degree of small particle content G= 1 ⫹ G and the degree of dilution k , for T 0 300 K; m 1; RS;0 1:5 mm; RL;0 10 mm; and 2r 18 mm: It is seen that the range of flammability, which spans from k 0 to 0.178 when G= 1 ⫹ G 0; gradually extends with increasing G /(1 ⫹ G ) and finally spans from k 0 to 0.400 when G= 1 ⫹ G 1: Extension of the range of flammability is attributed to the decrease in the equivalent particle radius RL,0/F. The limit of dilution k 0:446 without heat loss suggests that even at G= 1 ⫹ G 1: the heat loss affects the range of flammability. It is seen that as far as the trend and approximate magnitude are concerned, we can fairly predict the experimentally obtained extinction limit [43,77]. Fig. 45(b) shows u0,a in the adiabatic condition as a function of G= 1 ⫹ G for T 0 300 K and m 1; with k taken as a parameter. Owing to the decrease in the equivalent particle radius, u0,a increases first gradually and then rapidly with increasing G= 1 ⫹ G : At a certain G= 1 ⫹ G ; u0,a decreases with increasing k because of the reduced amount of chemical heat generation. The dashed curve is the extinction limit under heat loss conditions. A fair degree of agreement is shown. The appropriateness of the equivalent particle radius can further be shown in Fig. 46. It is, however, important to note that since the present formulation is based on the statistical analysis, a few tens of large particles should exist at least across the consumption zone, even for the bimodal system. Violation of this restriction results in the breakdown of the present formulation, such that discussion with the equivalent particle radius becomes less useful. 6.1.3. Experimental comparisons for diamond particles There exists another source for the bimodal system. Makino et al. [77] measured the range of flammability and the burning velocity of the SHS process for Ti–diamond system, as a basic research to fabricate the diamond embedded TiC/Ti composites by the SHS process, to achieve advanced cutting tools and/or abrasives. In this fabrication, the SHS flame must propagate through a medium of compacted particles of titanium, carbon (or graphite), and diamond, without participation of the diamond particles in the SHS process. By comparing the predictions [76] with the experimental results [43,77] it has been determined to use the physicochemical parameters, described in Sections 4.4 and 5.4.1, except for r N 3:51 × 10 3 kg=m3 : A Lewis number of Le0 100 is used because of the fair correlation with the A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 44. SHS rate-constant u0·R0 for Ti–C system with monomodal particles, as a function of the inverse of the heat loss index R02/(2r) at T0 300 K [79]; (a) with m taken as a parameter and (b) with k taken as a parameter. Data points are experimental [43,77,78]; open symbol designates the steady propagation and half-open symbol the pulsating combustion. experimental results [43,77] (0.6 relative density) for the various radii of diamond particles, as shown in Figs. 47 and 48. The compact diameter 2r is 18 mm in accordance with the experiments. Fig. 47(a) shows the range of flammability for monomodal compacts with respect to m and R0, for T0 300 K and k 0: Experimental results [43,77] for the steady propagation, quasi-steady multipoint combustion, extinction, and non-ignition show that the range of flammability extends with the decreasing R0. Fig. 47(b) shows u0,a·R0 as a function of m , with R0 taken as a parameter. Compared to the Ti–C system, the burning velocity for Ti–diamond system is reduced because of the increase in the nonmetal density r N, as shown in Eq. (76), which suggests that u0,a is inversely proportional to (r N) 1/2. The increase in r N, thus, reduces the total surface area of particles in the combustion wave, and hence, suppresses the volumetric heat-generation rate. Fig. 48 shows a similar comparison with respect 45 Fig. 45. Combustion behavior for Ti–C system with bimodal particles [75]; (a) Range of flammability with respect to the degree of small particle content G= 1 ⫹ G and the degree of dilution k , for T 0 300 K; m 1; RS;0 1:5 mm; and RL;0 10 mm; data points are experimental [43,77] and notation is the same as that in Fig. 43. (b) Burning velocity u0,a in the adiabatic condition as a function of G= 1 ⫹ G ; with k taken as a parameter; notation is the same as that in Fig. 44. to k . Data points are experimental [43] and the same comments as those for Fig. 47 can be made. Having confirmed the fair agreement for the monomodal system, it was intended to make experimental comparisons for the bimodal system. Fig. 49(a) shows the range of flammability with respect to G= 1 ⫹ G and k , and Fig. 49(b) shows u0,a as a function of G= 1 ⫹ G ; for the stroichiometric mixture m 1 at T 0 300 K with RS;0 3 mm; Fig. 46. SHS rate-constant u0,a·(RL,0/F) in the adiabatic condition as a function of the degree of dilution, with T0 taken as a parameter. Data points are experimental [43,77] and notation is the same as that in Fig. 44. 46 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 47. Combustion behavior for Ti–diamond system with monomodal particles [76]. (a) Range of flammability with respect to m and R0, for T0 300 K without dilution k 0; data points are experimental [43,77] and notation is the same as that in Fig. 43. (b) u0,a·R0 as a function of m , with R0 taken as a parameter. RL;0 15 mm; and rN;S rN;L : Extent of agreement between the predicted [76] and the experimental [77] results is similar to that of Fig. 45. The appropriateness of the equivalent particle radius RL,0/F can be further shown by comparing the experimental results of R0 8 mm for the monomodal system in Fig. 48(a) with the predicted results for the bimodal system of G= 1 ⫹ G 0:463; which corresponds to R L;0 =F 8 mm; in Fig. 49(a). As mentioned in Section 6.1.2, this discussion with the equivalent particle radius becomes less useful when a number of large particles across the combustion wave becomes too small to apply the statistical counting method. In order to further study a fundamental aspect of the practical fabrication of the diamond embedded TiC/Ti composite, the bimodal system with large diamond and small carbon particles has been used. Here, the diamond particles are large enough to prevent the SHS flame propagation, even for the stoichiometric mixture. Fig. 50(a) shows the range of flammability with respect to G= 1 ⫹ G and k , and Fig. 50(b) shows u0,a as a function of G= 1 ⫹ G ; for the stoichiometric mixture m 1 at T 0 300 K with RS;0 2:5 mm; RL;0 15 mm; rN;S 2:25 × 10 3 kg=m3 ; and rN;L 3:51 × 103 kg=m3 : The same trends as those shown in Fig. 45(a) and (b) are obtained, respectively, for the Fig. 48. Combustion behavior for Ti–diamond system with monomodal particles [76]. (a) Range of flammability with respect to k and R0, for stoichiometric mixture m 1 at T0 300 K; data points are experimental [43,77] and notation is the same as that in Fig. 43. (b) u0,a·R0 as a function of k , with R0 taken as a parameter. present bimodal system with different densities of N particles. Data points are experimental [77] and agreement is fair, as far as the trend is concerned. 6.2. Representative length of the cross-sectional area Thus the far, shape of a compacted medium has been considered to be cylindrical in evaluating the radiative heat loss from the side surface of the compacted medium. Then, the representative length of the cross-sectional area has been the compact diameter, as has appeared in the heat loss parameter C in Eq. (112). Here the choice of a representative length for the noncircular cross-sectional area is examined, with a rectangular compact taken as an example. When the radiative heat loss was introduced into the formulation in Section 5.1, the heat loss L per unit time and per unit volume was considered. When a sample specimen has a cross-sectional area S and perimeter p, the heat loss per unit time is expressed as LSd , where d is the thickness of a volume considered along the x-coordinate, say, the thickness of the combustion wave or the thickness of the consumption zone. As for the heat loss per unit time transferred from the surface to the ambience by radiation, it is A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 47 Fig. 49. Combustion behavior for Ti–diamond system with bimodal particles [76]. (a) Range of flammability with respect to G= 1 ⫹ G and k , for m 1 at T 0 300 K with RS;0 3 mm; RL;0 15 mm; and rN;S rN;L ; data points are experimental [77] and notation is the same as that in Fig. 43. (b) Burning velocity u0,a in the adiabatic condition as a function of G= 1 ⫹ G ; with k taken as a parameter; notation is the same as that in Fig. 44. Fig. 50. Combustion behavior for Ti–C–diamond system [76]. (a) Range of flammability with respect to G= 1 ⫹ G and k , for m 1 at T 0 300 K with RS;0 2:5 mm; RL;0 15 mm; rN;S 2:25 × 10 3 kg/m 3, and rN;L 3:51 × 103 kg/m 3; data points are experimental [77] and notation is the same as that in Fig. 43. (b) Burning velocity u0,a in the adiabatic condition as a function of G= 1 ⫹ G ; with k taken as a parameter; notation is the same as that in Fig. 44. expressed as es ST T 4 ⫺ T04 pd: Then, we have Fair agreement is again obtained, as far as the trend and approximate magnitude are concerned. es ST T 4 ⫺ T04 ; S=p by which the heat loss parameter is expressed as L l C rtc " 4es SB T 2 ⫹ T02 T ⫹ T0 rt c u0;a R0 2 # ! R20 : 4S=p 148 149 When the cross-sectional area is circular with radius r, we have S pr2 and p 2pr; which yields 4S=p 2r: When the cross-sectional area is rectangular with a in length and b in width, we have S ab and p 2 a ⫹ b; which yields 4S=p 2ab= a ⫹ b: It may be informative to note that this is a harmonic mean of a and b, and is called the equivalent diameter in the field of the heat-transfer engineering. Fig. 51(a) and (b) shows the range of flammability for Ti– C system as a function of m and k , respectively. Values of the physicochemical parameters are as those in Sections 4.4 and 5.4.1 Data points are experimental [80] at T0 450 K with various a and b, as well as the various radii of (irregular) carbon particles. Appropriateness of the choice of the representative length is demonstrated. Fig. 52(a) shows u0 as a function of m , with k taken as a parameter, and Fig. 52(b) shows u0 as a function of k , with m taken as a parameter. 6.3. Correspondence between the heterogeneous theory and the homogeneous theory Much of the theoretical analyses on the flame propagation in SHS, as reviewed in Ref. [24], are based on the classical treatment of the premixed flame propagation, characterized by a large activation energy Arrhenius reaction rate and large Lewis number heat and mass diffusion. Although these analyses have successfully captured the temperaturesensitive nature of the flame propagation, and consequently have been able to describe such empirically observed phenomena as abrupt flame extinction due to heat loss, and instability of the combustion wave in the form of pulsating and spinning propagation, as explained in Refs. [4–9], there are several fundamentally unsatisfactory aspects as mentioned in the previous sections. While it is pleasing that the heterogeneous theory has been successful in explaining the experimental results, especially the temperature sensitivity, one must nevertheless confront the somewhat curious result that the characterization of an Arrhenius diffusion process can be similar to that of an Arrhenius reaction process. To resolve this interesting correspondence, we shall demonstrate [81] that the governing equations for the heterogeneous theory can indeed be cast in a form which 48 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 51. Range of flammability for Ti–C system, with T0 taken as a parameter; (a) with respect to m and R20 =2ab= a ⫹ b; and (b) with respect to k and R20 =2ab= a ⫹ b: Data points are experimental [80] and notation is the same as that in Fig. 43. can be considered to be functionally identical to that describing the large activation-energy premixed-flame propagation in a condensed medium. To this aim, the N-consumption equation (46) must first be rewritten as d 1 ⫺ Z T ⫺rt B~ w exp ⫺ d ; 150 rt u dx T where B~ w ! ! 3 rM rt 1 ⫺ Z0 D0 ; rN rt;0 R20 1⫺Z 1 ⫺ Z0 Y~ M 1=3 ln 1 ⫹ A ~ YM ; 1⫹A 1⫺m 1 ⫺ Z0 ⫹ 1 ⫺ Z m : 1 ⫺ 1 ⫺ Z 151 152 153 In deriving Eq. (153), the following relation has been used Z 1 ⫹ Y~ M 1 ⫹ Y~ M;0 0 ; 154 Z derived from Eq. (49) under the assumption of the negligible diffusion effect. The particle consumption term in the RHS of Eq. (150) therefore corresponds to the “reaction” term in the homogeneous premixed-flame theory, with B~ being the “frequency” factor, w the “reactant concentration”, and Td the “activation temperature”. As for the energy conservation equation, from Eq. (58), we have d l d d T cT ⫺ rt u cT ⫹ q0 rt B~ w exp ⫺ d dx dx c dx T L: (155) The boundary conditions are as follows: x 0 : T Tm ; 1 ⫺ Z 1 ⫺ Z0 x!∞: d L d 1 ⫺ Z cT ⫺ 0: ; dx rt u dx 156 157 We see that a set of governing equations and boundary conditions is exactly the same as that for the conventional homogeneous theory (cf. Ref. [4]). Although differences are minor from a mathematical point of view, this identification is valuable for the physical interpretation of the combustion behavior. First, it should be noted that w in the “reaction” term is a function of the mass fraction of solid N particles, 1 ⫺ Z: In addition, if we note that the SHS proceeds under diffusion-controlled situation, as is the case for usual SHS, we have A= 1 ⫹ A ! 1; which suggests that the mass diffusivity in the Arrhenius fashion exerts nonlinear influence on flame propagation and/or extinction. That is, the activation temperature which can influence the nonlinear, highly temperature-sensitive nature of the SHS flame propagation is not the activation temperature Ta for the reaction, but Td for the liquid-phase mass-diffusivity. As for the effect of particle size, it has been incorporated into the frequency ~ tacitly. factor B; Since it has been confirmed that the set of the governing equations and boundary conditions in the heterogeneous theory is identical to that in the conventional homogeneous theory, results of theoretical analyses hitherto obtained by the homogeneous premixed-flame theory, as reviewed in Refs. [23,24], can be anticipated to be qualitatively valid if we take note of the differences involved. 6.4. Field activated SHS For low exothermic systems, in order to sustain flame A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 49 described in Section 5.1, except for the energy conservation equation, which is now not Eq. (58) but d dT ⫺ mc T ⫺ T0 ⫹ mq0 Z ⫺ Z0 L ⫺ P 158 l dx dx when it is expressed in dimensional form [89] where P is the energy supplied by the electric field. When we consider the situation that only N particles participate in the external heating due to electric field, the energy supply rate P is expressed as P 1 ⫺ Z scE 2 ; 2 159 where E is the electric field (V/m) and s c the electric conductivity (W ⫺1/m). It may be informative to note that the electric field induced by the alternative current must p be evaluated as the effective value with a factor of 1= 2: In addition, it is anticipated that when the temperature is lower than the melting point of M species, there is insufficient energy supply by electric current, because of the contact resistance among particles. When the combustion product, such as SiC has been produced, its electric resistance is too high to allow for electric current. By introducing a nondimensional parameter H Fig. 52. SHS rate-constant u0·R0 for Ti–C system as a function of the inverse of the heat loss index R20 =2ab= a ⫹ b at T0 450 K; (a) with m taken as a parameter; and (b) with k taken as a parameter. Data points are experimental [80] and notation is the same as that in Fig. 44. propagation, it is usual to heat a compacted medium externally by preheating and/or by applying an electric field. Since preheating in a long period of time could induce production of undesired compounds which might inhibit initiation of the combustion wave, it is reported to be preferable [82,83] for low exothermic systems to apply the electric field, which can also exert a great influence on the flame propagation [84,87] especially when the electric conductivity in the combustion wave is much higher than those in the unburned and burned states. Although there exist several analytical [88] and/or numerical [84,86] works on this subject based on the homogeneous premixed-flame theory, heterogeneity involved has not been explored. In this section, focusing on the effect of the electric field on the burning velocity, let us examine it by use of the heterogeneous theory. 6.4.1. Heat-input parameter The problem of interest is basically the same as that l=cs c E 2 ; m2a q0 160 in addition to other variables and parameters in Sections 4.2 and 5.1, Eq. (158) becomes ! d du m=m a m=ma ⫺ u⫹ j ds ds Z∞ ⫺ Z0 T~ ∞ ⫺ T~ 0 C·u H 1 ⫺ j ; ⫺ Z ∞ ⫺ Z0 2 Z ∞ ⫺ Z0 T~ ∞ ⫺ T~ 0 161 where H can be called the heat-input parameter. The other governing equations are the same as those in Eqs. (99) and (101). The boundary conditions are also the same as those in Eqs. (102) and (103). 6.4.2. Temperature profiles outside the combustion wave As mentioned in Section 5.2, in the limit of large Zeldovich number, melting, diffusion, and consumption/convection are confined to a thin layer in the neighborhood of the combustion wave. Outside this layer, the diffusion and consumption/convection terms are exponentially small. If we consider the situation that the combustion wave locates at s s f and that the electric field exerts influence only in the region 0 ⬍ s ⬍ s f ; temperature profile outside the combustion wave is expressed as s ⱕ 0 : u A u exp lu s; j 0; 162 50 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 0 ⬍ s ⬍ s f : u Ah exp lu s ⫹ B h exp ld s (170) as H Z ∞ ⫺ Z0 ⫹ ; j 0; C T~ ∞ ⫺ T~ 0 (163) s ⱖ s f : u C exp ld s; j 1; 164 where s) m=ma 4C ; lu 1⫹ 1⫹ 2 Z∞ ⫺ Z0 m=m a 2 ( 165 s) ( m=ma 4C : 1⫺ 1⫹ ld 2 Z∞ ⫺ Z0 m=m a 2 Relations among constants Au, Ah, Bh, and C are determined by use of the continuities of temperature and heat flux at s 0; respectively, the continuity of temperature at the combustion wave s s f ; and the jump condition of du / ds in the combustion wave as du du m=ma ⫺ ; 166 ds s f⫺ ds s f⫹ T~ ∞ ⫺ T~ 0 which is obtained by integrating Eq. (161) and evaluating at s s f : Then, we have Au 1 m=ma exp ⫺lu s f lu ⫺ ld T~ ∞ ⫺ T~ 0 ⫺ Ah ld H Z∞ ⫺ Z0 1 ⫺ exp ⫺lu s f ; lu ⫺ ld C T~ ∞ ⫺ T~ 0 1 m=ma ld H Z∞ ⫺ Z0 ⫹ lu ⫺ ld T~ ∞ ⫺ T~ 0 lu ⫺ ld C T~ ∞ ⫺ T~ 0 exp ⫺lu s f ; Bh ⫺ C lu H Z∞ ⫺ Z0 lu ⫺ ld C T~ ∞ ⫺ T~ 0 167 168 169 1 m=ma exp ⫺ld s f lu ⫺ ld T~ ∞ ⫺ T~ 0 ⫺ lu H Z∞ ⫺ Z0 1 ⫺ exp ⫺ld s f : lu ⫺ ld C T~ ∞ ⫺ T~ 0 T~ ∞ ⫺ T~ 0 Z∞ ⫺ Z0 1 ⫹ sf ln Z∞ ⫺ Z0 " Hs f ; Z∞ ⫺ Z0 # 1 ⫺ H Z∞ ⫺ Z0 ; T~ m ⫺ T~ 0 ⫺ H Z∞ ⫺ Z0 172 173 in the limit of C ! 0: We see that the adiabatic flame temperature increases with increasing strength of the electric field. It may be informative to note that the relation in Eq. (59), which gives a relation between the temperature and the mass fraction of fluid cannot be used any more when heat is supplied externally. 6.4.3. Effective range of electric field In order for the electric field to be effective for increasing combustion temperature, the flame location must be positive, which requires that the logarithmic term in Eq. (173) must be positive. From this requirement, the range in which the heat-input parameter H can exist is determined as 0ⱕH⬍ T~ m ⫺ T~ 0 ⬍ 1: Z∞ ⫺ Z0 174 In the same manner, when there exists heat loss, we have from Eq. (171) the following relation between C and H. ( ) 1 m=m a 2 H Z∞ ⫺ Z0 ⫺ ⫺ 2 4 T~ m ⫺ T~ 0 v ( ) u u m=ma 2 m=ma 2 H Z ⫺ Z ∞ 0 ⫺t ⫹ 8 8 T~ m ⫺ T~ 0 ) m=m a 2 H Z∞ ⫺ Z0 ⫺ 4 T~ m ⫺ T~ 0 v ( ) u u m=ma 2 m=ma 2 H Z ⫺ Z ∞ 0 ⫹t : 175 ⫹ 8 8 T~ m ⫺ T~ 0 ⬍C⬍⫺ 1 2 ( Note that when C has the value of the limit, length between the location of melting and that of the combustion wave becomes infinite. 170 If we further note that the temperature at s 0 is the melting point, we can determine the flame location s f as 3 m ld ⫹ H Z ⫺ Z ∞ 0 7 6 1 ma C 7; sf ln 6 5 4 l lu d ~ ~ T m ⫺ T 0 lu ⫺ ld ⫹ H Z ∞ ⫺ Z0 C 171 2 which is obtained from Eq. (167) by setting Au um : The flame temperature in the adiabatic condition, which is indispensable for nondimensionalization, is obtained from Eq. 6.4.4. Experimental comparisons for Si–C system Values of physicochemical parameters employed are those for Si–C system: q 0 6:23 MJ=kg; c 1 kJ= kg=K; r M 2:34 × 103 kg=m3 ; rN 2:25 × 103 kg=m3 ; WM 28:1 × 10⫺3 kg=mol; WN 12:0 × 10⫺3 kg=mol; Tm 1681 K: For the mass diffusivity, D 3:3 × 10⫺5 exp ⫺3:39 × 104 =T m2 =s is used [90]. As for the representative Lewis number, Le0 1; because experimental results [85] have been correlated for this value. Fig. 53 shows the SHS rate-constant u0·R0 for stoichiometric Si–C system, as a function of the square root of the heat-input parameter H. A solid curve is u0,a·R0 in the adiabatic condition and a dashed curve is that at the extinction A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 51 In this context, the heterogeneous theory is indeed viable in examining the SHS process. Correspondence between the heterogeneous theory and the homogeneous premixed-flame theory has further been examined. In spite of the minor differences, from the viewpoint of the mathematics, the important identification related to the physical interpretation, pointed out in the previous sections, has been reconfirmed. 7. Boundary between steady and pulsating combustion Fig. 53. SHS rate-constant u0·R0 for stoichiometric Si–C system, as a function of the square root of the heat-input parameter H [89]; a solid curve is that in the adiabatic condition and a dashed curve is that at the limit of flame propagation. Data points are experimental [85]; (S) designates data for rrel 0:52 and (W) for rrel 0:64: limit. A range surrounded by these curves corresponds to that for flame propagation under electric field. With increasing H, up to the limit determined by Eq. (174), u0·R0 increases first gradually and then rapidly. In the abscissa, for convenience, the strength of electric field is shown, obtained by use of the electric conductivity of carbon, s c 1:21 × 10 4 ⫹ 6:25 Tr ⫺ 450 (W ⫺1/m), based on the data in the literature, [91] with the representative temperature Tr taken as the arithmetic mean of the melting point of Si and the maximum combustion temperature. The initial carbon radius R0 of 2 mm is evaluated from the reported burning velocity [85] that was sustained in the lowest electric field of about 0.8 kV/m. Data points are experimental [85] and fair degree of agreement is demonstrated, as far as the trend and approximate magnitude are concerned. It is also reported that the characteristic extinction turning-point behavior described in Section 5.2 can exist when the heat-input parameter H is less than about 0.2, while it ceases to exist and the flame extinction occurs at the limit in Eq. (175) when H ⬎ 0:2: It may be informative to note that since the heat-input parameter H depends on the particle size, only the heterogeneous theory could captured the heterogeneous feature of the field activated SHS. 6.5. Some remarks on the several, other, important factors Related to the SHS flame propagation and/or extinction, other important factors, such as the bimodal particle distribution, the representative length of the cross-section, and the field activation are examined. It has been shown that these effects on the flame propagation and/or extinction are also closely related to the particle size of the nonmetal. So far, flame propagation and/or extinction in steady combustion has been examined. However, SHS flame propagation can be either steady or nonsteady, depending on the system parameters. Since nonhomogeneity in combustion products induced by nonsteady propagation is often responsible for the weaknesses in the solid materials synthesized, understanding of the mechanisms that trigger instabilities is expected to give appropriate directions for controlling the SHS process. The nonsteady flame propagation can assume two characteristic modes [3,7,9] namely pulsating and spinning combustion. In pulsating combustion, the combustion wave travels in a planar but pulsating manner as a result of the one-dimensional instability of the planar combustion wave to longitudinal perturbations. In spinning combustion, the combustion wave is nonplanar, and one or more hot spots are observed to move along a spiral path over the surface of the sample specimen as a consequence of the instability of the one-dimensional oscillatory to transverse perturbations. It has been observed that spinning combustion occurs in low exothermic systems with excess amounts of either one of the reactants, or the diluent, or the large particles, at low preheat temperatures. As for pulsating combustion, it has been observed in energy intensive systems with stoichiometric mixtures of small particles at elevated preheat temperatures. Since rapid synthesis is preferred from a practical point of view, and it can be accomplished by using small particles and/or high preheat temperatures, pulsating combustion frequently occurs and results in materials with a laminated structure [3] (cf. Fig. 3), which cannot be used as a bulk. As such, the study and control of transition from steady combustion to pulsating combustion are of greater practical relevance than those of the transition from steady combustion to spinning combustion. In order to identify and gain an understanding on the boundary governing the transition from steady to pulsating combustion, we are required to extend the steady heterogeneous theory of SHS flame propagation developed in Refs. [32,64] to allow for a nonsteady effect, as has been done in Ref. [92]. In the following, first by considering the material strength and flame instability, description for the boundary between steady and pulsating combustion is made in terms of the system parameters, which consist not only of the 52 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 physicochemical parameters, but also of the operating parameters, such as the mixture ratio, degree of dilution, particle size, and initial temperature. From this identification, operating parameters at which steady flame propagation is maintained, even at the adiabatic condition, are obtained, to elucidate the necessary condition for pulsating combustion, in order to identify the useful parameter in suppressing pulsating combustion. Finally, transverse effects related to heat loss from the side surface of the compact are examined to assess the range of steady combustion. 7.1. Criterion for the appearance of cracks Pulsating combustion is characterized by the periodic movement of the combustion wave, with the laminated structure of the combustion product being caused by the presence of cracks in the radial direction. Although it has been pointed out [51] that both thermal and mechanical effects induce pulsating combustion, the main concern has been directed to the thermal effect, and various numerical and analytical investigations have been conducted, as reviewed by Margolis [24]. However, the appearance of radial cracks in the combustion products implies that mechanical effects caused by pressure increase in pores [93,94] and/or thermal expansion, can exert strong influence on the combustion behavior. When there exists temperature distribution along the axial direction, there appears not only a normal stress corresponding to macroscopic pressure distribution along the axial direction, due to desorbed gas in pores, but also a shear stress corresponding to a difference of thermal expansions in the radial direction. When strains caused by these stresses are kept below a certain value Ksb, determined by the material fracture toughness, as well as the coefficient of thermal expansion, radius of the curvature of the pores, compact diameter, and so on, we can anticipate that radial cracks can be suppressed and steady combustion can be maintained. Since both of these stresses are related to a difference of the temperature gradients along the axial direction, and the difference is the maximum at the propagating combustion wave, the criterion that steady combustion is maintained can be expressed as [95] 2T 2T ⱕ Ksb : ⫺ 176 2x 0⫺ 2x 0⫹ In evaluating the temperature gradients, an attempt [92] has been made to apply and extend the linear stability analysis within the framework of the heterogeneous theory, by linearizing the problem about the basic solution and seeking harmonic solutions through the perturbation quantities u p(s , t ) and ap(t ) as u u 0 s ⫹ u p s; t; m=m a a0 ⫹ ap t; 177 as has been made by Margolis [24,74] and Kaper et al. [73] based on the homogeneous premixed-flame theory. Note that the perturbation quantities are O(b ⫺1), where b is the Zeldovich number. 7.2. Linear stability analysis 7.2.1. Basic solution for planar propagation Prior to the linear stability analysis, let us first confirm the governing equations and boundary conditions. By introducing the nondimensional time defined as t Z∞ ⫺ Z0 2 m2a t ; l=crt 178 in addition to the other variables and parameters introduced in Sections 4.2 and 5.1, the energy conservation equation and the N-concentration equation are expressed as 2 m=m a 2 22u F u j ⫺ u ⫹ j ⫺ u ⫹ ; 2t Z∞ ⫺ Z0 2s Z ∞ ⫺ Z0 2 2s 2 179 2j m=m a 2 Z0 ⫹ j⫹ Z∞ ⫺ Z0 2s 2t Z∞ ⫺ Z0 1=3 x~ Z∞ ⫺ Z0 1⫺ j : ~ 1 ⫺ Z0 Z∞ ⫺ Z0 L0 rt 0=rt exp T~ d =T 180 The boundary conditions are as follows: s ! ⫺∞ : u 0; j 0; 181 s0: u u max ; 182 s!∞: 2u F u 2j ⫺ 0: ; m=m a Z∞ ⫺ Z0 2s 2s 183 The mass burning rate m r t u; subject to heat loss L, would emerge from the solution as an eigenvalue, with ma and L 0 being, respectively, the mass burning rate and the mass burning rate eigenvalue in the steady, adiabatic limit. The SHS rate-constant u0·R0 is obtained from this eigenvalue. In conducting the linear stability analysis, we are first required to obtain the basic solution u 0(s ) and a0. Since the perturbation quantities are O(b ⫺1), the basic solution is exactly the same as that for the asymptotic “outer” solution in the steady state, as presented in Eqs. (113) and (114), in the limit of large Zeldovich number b . In Ref. [73], the basic solution is expressed in a slightly different form as u C exp l u s; j 0; s ⬍ 0; 184 u C exp l d s; j 1; s ⬎ 0; 185 where 1 a0 1⫹ 2 Z∞ ⫺ Z0 1 a0 1⫺ ld 2 Z∞ ⫺ Z0 lu 1 ; C 1 ; C 186 and C is the maximum temperature in the normalized form, A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 expressed as s 4C C 1 ⫹ 2 1; a0 187 which is obtained from the jump in du /ds at s 0: Note that by conducting a similar treatment in Eq. (125), we have the relation m 2 exp b C ⫺ 1 ⫺ M 188 ma k 1 ⫺ M: Having obtained the basic solution, we can now evaluate the difference of the temperature gradients in the criterion in Eq. (176). Then, we have a0 Z∞ ⫺ Z0 ⫹ O b⫺1 ⱕ a0 Z∞ ⫺ Z0 sb ⬅ l=c Ksb ma q0 =c 189 for steady combustion. We see that the appearance of cracks in pulsating combustion can be primarily controlled by a0 Z∞ ⫺ Z0 because the perturbed O(b ⫺1) term is much smaller than the first term. Note here that Z∞ ⫺ Z0 is related to the heat of combustion, which depends on the mixture ratio m , degree of dilution k , and initial temperature T0, and that a0 is related to the heat loss which depends on the particle size R0 and the compact diameter 2r. When Z∞ ⫺ Z0 is kept constant, we can anticipate that steady combustion occurs in the range a0 ⬍ a0 sb ; which is equivalent to C ⬎ Csb ; if we recall the relationship between C and a0. When the heat loss condition is kept constant, the steady combustion range is Z ∞ ⫺ Z0 ⬍ Z∞ ⫺ Z0 sb ; which yields m ⬍ msb for under-stoichiometric mixture m ⬍ 1 and m ⬎ msb for an over-stoichiometric mixture m ⬎ 1 at a certain k ; for constant m , k ⬎ ksb : Eq. (189) further suggests that when Z∞ ⫺ Z0 ⬍ Z∞ ⫺ Z0 sb ; steady combustion can be maintained, even if the perturbed O(b ⫺1) term is positive; likewise, when Z∞ ⫺ Z0 ⬎ Z∞ ⫺ Z0 sb ; steady combustion cannot be maintained even if the perturbed term is negative. In this context, the condition for no amplification, nor attenuation of the perturbed term is anticipated to provide information about the transition boundary, and then we are required to obtain the specific form of the perturbed term in Eq. (189). 7.2.2. Asymptotic solution and linear stability boundary In obtaining asymptotic solution, we should note that the first order inner solution is identical to that for the adiabatic combustion when the Zeldovich number is large, because of a higher order effect of heat loss, while the outer problem is influenced by both heat loss and unsteadiness. The outer problem is then given by [92] 2u a0 2u 2 2u C·u ⫹ ⫺ ; 2 Z∞ ⫺ Z0 2s 2t 2s Z ∞ ⫺ Z0 2 190 u 0 as s ! ^∞: 191 53 In examining the linear stability of the steady planar SHS flame, it is the standard to linearize the nonlinear problem (Eqs. (179)–(183)) regarding the basic solution (Eqs. (184), (185), (187) and (188)) and to seek harmonic solutions. Substituting the perturbation quantities in Eq. (177) into Eq. (190), we obtain 2u p 2up ap a0 du0 ⫹ ⫹ 2t Z∞ ⫺ Z0 2s Z∞ ⫺ Z0 ds 2 2up C· u p ⫺ ; 2 2s Z∞ ⫺ Z0 2 192 up 0 as s ! ^∞: 193 Here, a relation between the perturbed quantities is given as [73] ( ) B b M 1⫹ 2 a0 2 a0 k 1 ⫺ M a p Bup 0; t; 194 where use has been made of Eq. (188). Note that, since 1 ⫺ M is the normalized driving force for the flame propagation, as mentioned in Section 5.6.1, B/a0 is a measure of the energy generated in the combustion wave compared to the driving force for the nonadiabatic flame propagation, that is a measure of the driving force for instability. A normal mode solution for u p(s , t ) is then found to be a scalar multiple of " 1 iv u p s; t exp ivt ⫺ ( × exp lu s ⫹ 1 ⫹ 1 iv ! a0 Z∞ ⫺ Z0 ! a0 Z∞ ⫺ Z0 ! B l C a0 u ! ) B lu C a0 # × exp pu s ; s ⬍ 0; " (195) 1 iv a0 Z∞ ⫺ Z0 1 × exp ld s ⫹ 1 ⫹ iv # a0 Z∞ ⫺ Z0 u p s; t exp ivt ⫺ ( × exp pd s ; s ⬍ 0; where ! ! ! B l C a0 d ! ) B lC a0 d (196) 54 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 8 s 9 < = 1 a0 1 Z∞ ⫺ Z0 2 pu ⫹4 iv ; 1⫹ 2 : ; 2 Z∞ ⫺ Z0 a0 C 197 8 9 s < = 1 a0 1 Z∞ ⫺ Z0 2 pd ⫹4 iv : 1⫺ 2 ; a0 2 Z∞ ⫺ Z0 : C 198 Having obtained the perturbation quantities u p(s ,t ) and ap(t ), we can now evaluate the perturbed term in Eq. (176) as " !2 ! 1 a0 B a 0 Z∞ ⫺ Z0 exp ivt ⫺ 2 iv Z∞ ⫺ Z0 a0 ( 1 ⫹ 1⫹ 2 iv a0 Z∞ ⫺ Z0 !2 ! ) B C a0 199 v !2 # u u 1 Z ⫺ Z0 t 2 ⫹ 4 ∞ iv : a0 C Since the jump condition with respect to 2u=2s at s 0 gives iv as [73] ! 2" ( !2 ! ) 1 a0 1 B B C iv ⫺ ⫹4 ⫺ 8 Z∞ ⫺ Z0 a0 a0 C2 !v !2 ! # u 1 B u 1 B B t ^ ⫺ ⫹ C ; ⫺8 C a0 C a0 a0 200 we see that the perturbed term attenuates in the regime 0 s1 p ! 6 B 1 A 1 0ⱕ ; 1 ⱕ ⬍ ; ⬍ 2C@1 ⫹ 1 ⫹ a0 C 2 4C4 Fig. 54. Location of the neutral stability boundary [92]. where Xsb is determined by solving the neutral stability condition b X 2 X⫺M 8 9 s 2= < 2 1 2 s 1 ⫹ 1 ⫹ 1 ⫺ ln X ; 206 : ; 4 b 2 1 ⫺ ln X b with X ⬅ exp ⫺2bC=a 02 a20 k 1 ⫺ M ⫹ M: In obtaining Eq. (206), use has also been made of Eqs. (187) and (194). Fig. 55 shows the neutral stability boundary in the b –M plane, with X being taken as a parameter. Once the critical values of the heat-loss parameter C sb, and the flame temperature u sb, are obtained, the critical size of N particles and compact diameter for steady combustion 201 0ⱕ B a0 ⬍ 1 ; C 1 ⫺ C2 ! p 1 6 ⬍ ; 2 C 202 because of the negative value of the real part of iv , as shown in Fig. 54. Restricting ourselves to the neutral stability boundary on which there is neither attenuation nor amplification of the burning velocity, we have a set of critical values as a0 2sb Xsb ⫺ M ; k 1 ⫺ M 203 C sb ⫺ a0 sb2 ln Xsb ; 2b 204 usb 1 ⫹ ln Xsb b 205 Fig. 55. Neutral stability boundary in the b –M plane [92]. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 56. Ranges of flammability and instability for Ti–C system with l= rt c taken as a parameter [92]; the initial temperature T0 300 K and the compact diameter 2r 18 mm; (a) as functions of the mixture ratio m and the radius R0 of N particles, with l= rt c taken as a parameter; and (b) as functions of the degree of dilution k and R0. In the region below the dashed curve, pulsating combustion proceeds. In the region between the solid (or dotted) and dashed curves, stable and steady propagation occurs. Data points are experimental [43,77] and notation is the same as that in Fig. 43. are determined through ! C sb rt c u0;a R0 2 R 02 2r sb l= rt c4es SB Tsb2 ⫹ T02 Tsb ⫹ T 207 Note that the SHS rate-constant u0,a·R0 in the adiabatic condition is given by Eq. (129). 7.2.3. Experimental comparisons for the boundary In order to demonstrate the appropriateness of the transition boundary determined above, a comparison has been made with Ti–C system taken as an example. Values of the physicochemical parameters are the same as those described in Sections 4.4 and 5.5, while the Lewis number Le0 25 in accordance with experiments with spherical carbon particles. Fig. 56(a) shows the ranges of flammability and instability with respect to the mixture ratio m and the radius R0 of N particles for an initial temperature T0 of 300 K. In accordance with the experiments [77] the compact diameter 55 2r is 18 mm. The solid and dotted curves are respectively the numerically [64] and analytically [72] obtained extinction limits. The dashed curves are the neutral stability boundaries, with l /(r tc) taken as a parameter. Results show that, except near stoichiometry m 1; the range of instability with respect to R0 is reduced with decreasing mixture ratio. When m ⬍ 0:673; instability is suppressed, even for R0 ! 0; because of the decrease in the volumetric heat-generation with decreasing m . We also see that the instability can be suppressed with increasing l= rt c: This is attributed to the fact that heat transferred from the consumption zone increases with increasing l= rt c: Likewise, a reduction of T0 is also expected to suppress instability, although the range of flammability is reduced. The slight increase in the range of instability with decreasing m near stoichiometry is attributed to the overestimation in Ij in Eq. (92), as noted in Section 4.6. Experimental results [77] also show that the range of pulsating combustion extends with decreasing R0. The decrease in R0 increases the total surface area for surface reactions in the combustion wave, and hence, makes the mixture more energy intensive. Since instability is invariably dependent on the burning intensity, mixtures with small particles have wider ranges for pulsating combustion. Experimental results suggest that the limit of instability predicted by the present analytical result is fair. Fig. 56(b) shows the similar comparison with respect to the degree of dilution k . We see that the range of instability with respect to R0 is narrowed with increasing k and that instability is suppressed even at R 0 ! 0 when k ⬎ 0:280: Again, fair agreement is demonstrated with the experimental results [43,77] in the trend and the approximate magnitude. Fig. 56(a) and (b) show that the predicted limit of instability with l= rt c 5 × 10 ⫺6 m2 =s fairly represents the experimental results, while the limit of extinction has been obtained with l= rt c 2:5 × 10⫺6 m2 =s: This difference may be caused by the different temperatures at which these two phenomena occur; that is, the temperature at which the pulsating combustion ceases to occurs is about 200 K higher than that for flame extinction. Fig. 57 shows the ranges of flammability and instability with respect to m and k , with R0 1:5 mm at T0 300 K: The solid curve is the limit of flammability [64] the dashed curve the boundary for R0 1:5 mm obtained by numerically solving Eq. (206) with l= rt c 5 × 10⫺6 m2 =s; and the dotted curve the boundary for the adiabatic condition to be mentioned in the next section. It is seen that the pulsating combustion can be suppressed with decreasing m and/or increasing k because of the decrease in the volumetric heat-generation. Fair agreement is again demonstrated between the predicted and experimental [77] results for spherical carbon particles (3 mm in diameter). Since the condition for these experimental results is considered to 56 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 57. Ranges of flammability and instability for Ti–C system as functions of m and k , for T0 300 K and 2r 18 mm [95]. The solid curve is the limit of flammability [64], the dashed curve the numerically obtained boundary [92] for R0 1:5 mm; and the dotted curve the predicted result [95] in the adiabatic condition. Data points are experimental [77] for R0 1:5 mm and notation is the same as that in Fig. 43. be close to the adiabatic condition, as shown in Fig. 56(a) and (b), heat-loss effect on this boundary is relatively small. 7.2.4. Boundary at the adiabatic condition In the SHS process, reduction of particle size, which reduces heat-loss effect, facilitates the occurrence of pulsating combustion when the system parameters, such as the physicochemical and operating parameters in Sections 4.6.2 or 5.6.3, are given. However, even if we use small particles, there is a possibility that steady combustion is maintained because appearance of pulsating combustion also depends on the operating parameters, such as the mixture ratio m , degree of dilution k , and initial temperature T0, through the heat of combustion when the physicochemical parameters for a specific system are given. Such that, operating parameters at which steady combustion proceeds even at the adiabatic condition (i.e. X 1) has been obtained in order to elucidate the necessary condition for pulsating combustion, under which the only way effective in suppressing the appearance of pulsating combustion is increasing size of N particles. In order for Xsb to be unity, the critical values of b and M must satisfy p b=2 2 ⫹ 5; 208 1 ⫺ M sb which has already been identified in the field of stability analysis. If we recall the expression 1 ⫺ M bIu exp b=a in Section 5.6.1 and Iu in Eq. (87), we can obtain T∞ ⫺ T0 sb because a and b are defined by use of T∞ ⫺ T0 : Note that when the normalized melting-point u m is smaller than about 0.7, the following equation fairly represents the relation between a and b . p p b=a 5 ⫹ 2 p exp y2 erfc y ; y⬅ p a⬇ : 209 y 2 2 Since T∞ ⫺ T0 is related to the system parameters as [40] 8 m 1 ⫺ k > > m ⱕ 1 > < m ⫹ fst T~ ∞ ⫺ T~ 0 Z∞ ⫺ Z0 ; 210 > 1 ⫺ k > > m ⱖ 1 : m ⫹ fst when there is no external heating, we can obtain the critical values for these operating parameters. Fig. 58(a) shows the boundary between steady combustion and pulsating combustion in the adiabatic condition, as functions of the mixture ratio m and the initial temperature T0, when there is no dilution k 0: The solid curve is the limit of flammability numerically obtained [64] and the ~ 0 is the boundary obtained with Eq. dashed curve for k (208) [95]; the chain curve for k~ k~lower is the boundary when the transverse effect is the largest, which will be examined in the next section. We see that the pulsating combustion regime expands with increasing T0, and that the steady combustion regime contracts. Experimental results [33,43,77] for small carbon particles, with diameters less than 3 mm, are also shown because these data are considered to be influenced very little by the heat loss (cf. Fig. 57). Fair agreement is demonstrated as far as the general trend is concerned. Fig. 58(b) shows the transition boundary as function of k and T0, when the system is stoichiometric. The extent of agreement is similar to that in Fig. 58(a). 7.3. Transverse effects on the range for steady combustion The transition from steady combustion to pulsating combustion is in general accompanied by the appearance of nonplanar spatial perturbations. Through various analyses for examining the transition boundary by use of the multidimensional media with cylindrical [74,96,97] or rectangular [98–100] shapes, it is reported that this kind of A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 57 Fig. 58. Ranges of flammability and instability for Ti–C system [95]; (a) as functions of m and T0; (b) as functions of k and T0. The solid curve is the limit of flammability [64] numerically obtained, while the dashed curve for k~ 0 and the chain curve for k~ k~lower are the numerically obtained boundaries [95] between steady and pulsating combustion; k~ 0 means that there is no transverse effect, while k~ k~lower means the transverse effect is the largest. Data points are experimental [33,43,77] when R0 is less than 1.5 mm; notation is the same as that in Fig. 43. perturbation can be properly described by introducing overall transverse wave-number k~ into the linear stability analysis. When this effect is taken into account, the boundary under heat-loss conditions is given by the following relation: [73,74] 3 1 b X C⫺ C 2 X⫺M 2 1 1 b X ⫹ 2 ⫹ 4k~2 ⫹ ⫺4C C ⫺ C 2 X⫺M C 1 1 1 b X ⫹ 4C C ⫺ ⫺ ⫹ 4k~2 2 C C 2 X⫺M C 2 1 ⫺ ⫹ 4k~2 0: (211) C2 given by solving b X 4 s ; 2 X⫺M 2 1 ⫺ ln X b 213 ~ which is specified as k~lower : For the appears at a certain k; stoichiometric Ti–C system, we have b=2 2:29 and M 0:240 at T0 300 K: In this system, Xsb is 0.5792 when ~ is given, the When a set of parameters (b , a , M, and k) critical value of Xsb can be determined. As reported in Refs. [73,74] and as shown in Fig. 59, with an increasing k~ for fixed Xsb, b /2 first decreases, reaches a minimum, and then increases. The minimum value of b /2 exists at k~ determined by solving 4k~2 2 ⫺ 1 : C2 212 We see in Fig. 59 that b /2 decreases with decreasing Xsb when k~ is fixed. It may be common in practical situations that the system parameters such as b , a , and M are given by specifying combinations of metal and nonmetal species, initial temperature, and so on. Thus, it is of interest to examine ~ with b /2 taken as a changes in Xsb due to an increase in k; ~ constant. With increasing k for fixed b /2, Xsb determined from Eqs. (211) and (212) first increases, reaches a maximum, and then decreases. The maximum of Xsb, which is Fig. 59. Neutral stability boundary for stoichiometric Ti–C system, ~ the half of the Zeldovich number as functions of the wave-number k; b , and the normalized heat loss Xsb at the critical condition. The initial temperature T0 is 300 K and there is no dilution. 58 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 there are no effects of the transverse wave-number k~ 0; ~ k~lower 0:4489; as shown in Fig. while Xsb 0:6412 for k 59. This result suggests that the steady combustion regime is the broadest when k~ k~lower : The boundary between the steady and pulsating combustion determined with k~ k~lower without heat loss X 1 is also shown in Fig. 58(a) and (b). Although k~ depends on the compact diameter and the transverse mode of the combustion wave, because it is influenced by the boundary conditions in the radial direction such as the heat-loss condition on the side surface of the compact, we see that the difference between the boundary for k~ 0 and that for k~ k~lower is relatively small. Fig. 60(a) shows the ranges of flammability and instability for Ti–C system at T0 300 K as functions of m and R20 = 2r [95]. The solid curve is the limit of flammability numerically obtained [64] the dotted curve is that analytically obtained [72]; the dashed curve is the boundary between the steady and pulsating combustion for k~ 0; and the chain curve is that for k~ k~lower ; by use of Eqs. (206) and (211), respectively. Data points are experimental [43,77,78] for various particle radii (1.5, 5, 10, 15 mm) and compact diameters (3, 5, 7, 10, 18 mm). We see that the range of instability is extended with decreasing R20 = 2r because the effect of heat loss is reduced due to the increase in the volume-to-surface ratio. Fair agreement between the predicted and the experimental results suggests that the heat loss index R02/(2r) is useful in correlating experimental results. Fig. 60(b) shows the similar plot for k and R20 = 2r: Finally, it may be informative to note that bifurcation [74,96,97,99] of pulsating and spinning combustion, and chaotic combustion [98,100] can occur at a certain con~ as shown in dition, because of the trend between b and k; Fig. 59. Topics about bifurcation and the chaotic combustion, which are related to the nonlinear stability, are not covered here; readers are recommended to refer Ref. [24]. 7.4. Some remarks on the boundary of steady combustion Use of small particles enhances the SHS flame propagation speed and necessarily induces pulsating combustion, such that the identification obtained here offers the limit of the use of small particles when other system parameters are kept unchanged. As for the effects of transverse wave number, which are related to heat loss on the side surface of the compacted medium on the location of the transition boundary, it has been found to be relatively small. Further, appropriateness of the transition boundary, which has been demonstrated by comparing the predictions with experimental results in the literature, suggests that the heterogeneous theory has captured the essential, heterogeneous feature of the SHS flame propagation and that the results are of practical importance in controlling the combustion behavior to be steady. Fig. 60. Ranges of flammability and instability for Ti–C system for T0 300 K [95]; (a) as functions of m and R20 = 2r; and (b) as functions of k and R20 = 2r: Data points are experimental [43,77,78] for various particle radii and various compact diameter; notation is the same as that in Fig. 43. The solid curve is the limit of flammability numerically obtained [64], the dotted curve is that analytically obtained [72], and the dashed curve for k~ 0 and the chain curve for k~ k~lower are the numerically obtained boundaries [95]. 8. Initiation of the combustion wave Besides flame propagation, initiation of the combustion wave (which is indispensable for the SHS process) has also been investigated extensively because the energy that must be supplied externally in the SHS process is basically that for initiating the combustion wave. Although much of the theoretical accomplishment about the initiation of the combustion wave has been reviewed by Barzykin et al. [101] and Barzykin [102] all are based on the homogeneous premixed-flame theory, and hence the heterogeneity has not A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 59 over a short period when the activation energy (or temperature) is large. Let tc denote an ignition delay time, at which temperature rises rapidly. During most of time up to the ignition, t ⬍ tc ; the specimen experiences only inert heating because heat-generation term is exponentially small due to large activation energy, and reactant consumption can be neglected. Then, Eq. (214) becomes as Fig. 61. Model used for analyzing initiation of the SHS combustion wave in the specimen, induced by another SHS flame that has passed through the igniter [103]. aj 2T~ 2T~ 1 4hrad ~ ⫹ T ⫺ T~ 0 ; 2t rt j c·2r 2x2 j ig; sp; 216 been clarified. The heterogeneous theory, which has succeeded in describing dependence of the flame-propagation process on the particle size, is also expected to be applied for examining ignition phenomena. First, we shall consider the initiation of the combustion wave induced by an igniter [103] in order to examine heterogeneity that has not been examined in a series of analyses [104–110]. After that, initiation of the combustion wave by use of the radiative heat flux will be examined. where T~ cT=q0 ; a l= rt c: The boundary conditions at x 0 are as follows: ~ ig T ~ sp ; T l 8.1. Flame initiation induced by igniter One-dimensional, planar, heterogeneous flame initiation has been considered in a semi-infinite medium, which is heated by an SHS flame passed through the other semi-infinite medium, called an igniter; cf. Fig. 61. It is supposed that the region x ⬎ 0 is the igniter with thermophysical properties of l ig, (r t)ig, and cig, and that the region x ⬍ 0 is the medium to be ignited, called a specimen, with l sp, (r t)sp, and csp. The contact surface is set to be x 0 and the moment that the SHS flame reaches the surface is t 0: The igniter and the specimen are compacted media respectively, consisting of particles of nonmetal (or higher meltingpoint metal) N, lower melting-point metal M, and an inert I that can be the product P of the reaction between N and M according to nM M ⫹ nN N ! nP P: Monodisperse size distribution is considered for N particles, with an initial radius R0 and number density n0. Governing equations for these media, from Eqs. (179) and (180), are expressed in a dimensional form as follows. Energy: rt 2 22 T 2Z cT l 2 ⫹ q0 rt ⫺ L: 2t 2t 2x N-consumption: rt 2Z rt 4prN n0 R0 x 2t rt;0 ! 1⫺Z 1 ⫺ Z0 : 215 8.1.1. Inert stage Ignition processes can usually be characterized by a gradual increase in temperature followed by a rapid increase 2T~ 2x ! l ig 2T~ 2x ! : 218 sp As the initial conditions, we have " x ⬍ 0 : T~ ⫺ T~ 0 T~ max ⫺ T~ 0 exp # mig x ; Z Z0 ; l=csp 219 x ⬎ 0 : T~ ⫺ T~ 0 " 1 T~ max ⫺ T~ 0 exp ⫺ mig # 4hrad x ; Z Z∞ ; c·2r 220 which are obtained for the steady SHS flame propagation (e.g. Ref. [73]), under an assumption of large activation energy for the rate-limiting process, by noting that melting, diffusion, and consumption/convection are confined to a thin layer in the combustion wave. Putting s Z ∞ ⫺ Z0 sp ma sp x Z∞ ⫺ Z0 sp2 ma 2sp t ; t ; l=csp l=csp rt sp u ⫺ T~ 0 T~ ⫺ T~ 0 T~ ; us max ; T~ ∞ ⫺ T~ 0 sp T~ ∞ ⫺ T~ 0 sp 214 1=3 217 mig = ma sp C k~ ; w~ ; ~ ∞ ⫺ Z0 sp2 Z∞ ⫺ Z0 sp k Z C a sp 4hrad rt csp u0 ·R0 sp2 ! R20 l ; L p 2r a 60 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 and using the Laplace transformation, we have ! us s2 ~ uI ⫺ w~ k t exp ⫺ 1 ⫹ Lsp =Lig 4t ( × L sp 1 1⫹ 2 Lig " ! diffusive region near the contact surface, stretched variables c b u ⫺ u I ; X ⫺bKz s ⫺ s c ; are introduced, where b is the Zeldovich number defined as ! 2# p s p ⫹ k~ t 2 t ! ! p L sp s 1 ~ p ⫹ k t ⫺ 1⫺ 2 2 t Lig × erfc " b exp p s ⫺ p ⫹ k~ t 2 t ! p s ~ × erfc ⫺ p ⫹ k t 2 t ⫹ exp d2 c D ⫺ t ec⫺X ; 2 dX 2 ! 2 x~ c Dt ; bKz2 Z∞ ⫺ Z0 L0 rt;0 =rt exp T~ d =T~ c ! 2# × erfc which is subject to dc c X0 0; 0: dX X!∞ s! 2 # a ig s ⫺ p ⫹ w~ t 2 t asp s!) a ig s ⫺ p ⫹ w~ t : 2 t asp (221) Here, L ig and L sp are the heat penetration coefficients. In the following, temperature rise in the inert, heating stage is designated with the subscript I. Note that temperature response for a finite size of the igniter can also be obtained in the same manner, as shown by Strunia et al. [106] 8.1.2. Transition stage In analyzing the transition from inert heating to vigorous reaction, temperature deviation must be taken into account because the heat generation term becomes important as temperature rises. Since departure from the inert heating is anticipated to occur at time and position such that uI ⫺ uI c is close to zero, we have the lowest order solution as uI uI c ⫹ Kt t ⫺ tc ⫹ Kz s ⫺ s c ; 222 by expanding Eq. (221), where Kt uI =utc ; Kz 2uI =2s c : Note that (·)c designates the condition corresponding to the rapid increase in temperature, which occurs at t tc and s s c ; uI c is the maximum temperature in the inert heating. Since temperature at a certain position in the specimen first increases and then decreases, the most plausible condition for the ignition is that for 2uI =2t 0; at which the position has its maximum temperature. The inner problem for the asymptotics is then as 2u 2 2u x~ : ⫹ ~ 2t 2s z2 Z∞ ⫺ Z0 L0 rt0 =rt exp T~ d =T T~ d Z∞ ⫺ Z0 : T~ c2 225 Retaining only nonvanishing terms on the exponential, we have × exp " 224 223 Here, reactant consumption and heat loss are ignored, as is the case for usual ignition phenomena. Identification of Eq. (223) infers that the present problem parallels the earlier development [104–110]. In analyzing the reactive– Eq. (226) can be solved analytically and we have " ( !2 )# 1 ⫺1 ⫹ C e ⫺X ; c X ⫹ ln 1⫺ Dt 1 ⫹ C e⫺X p 1 ⫺ 1 ⫺ Dt p C 1 ⫹ 1 ⫺ Dt 226 227 228 229 (e.g. Ref. [111]). Since there is no solution for Dt ⬎ 1; the critical condition is as follows: ! 2 x~ c Dt ⬅ 1; 230 bKz2 Z∞ ⫺ Z0 L0 rt;0 =rt exp T~ d =T~ c where 2uI 2s Kz ( ! c us exp ⫺w~ k~tc 1 ⫹ Lsp =Lig × p k~ exp k~ 2 tc erfc k~ tc ⫺ w~ ! s s!) aig aig aig : exp w~ 2 tc erfc w~ t asp asp asp c (231) Recalling definitions of parameters, we see that the ignition delay time can be expressed as functions of not only physicochemical parameters, but also operating parameters characteristic to the SHS process, such as the mixture ratio, degree of dilution, initial temperature, and particle radius, through the eigenvalue L 0 and Z∞ ⫺ Z0 : It may be informative to note that the criterion used by Averson et al. [104] and/or Strunia et al. [106] was Dt 2 which was obtained by Enig [112] by the use of the so-called integral method by equating a transient–reactive relation at the flame initiation and a transient–diffusive relation in the inert heating. Such that, compared to the criterion in A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 61 We see that the ignition delay time is proportional to the square of the particle size, which has not been captured in previous theories. Fig. 62. Ignition delay time as a function of the mixture ratio m ig in the igniter, with the emissivity e taken as a parameter [103]. Data points are experimental [113]. Eq. (230) by the asymptotics, the ignition delay time is overestimated by a factor of 2 when k~ ! ∞ and w~ 0: 8.1.3. Limiting solution It may be informative to consider a limiting situation for k~ ! ∞ and w~ 0; 232 which corresponds to the flame initiation caused by a spontaneous contact with a hot substance in adiabatic condition. The critical condition yields explicit dependence of the ignition delay time on other parameters as tc R20 exp T~ d =T~ c T~ d T~ max ⫺ T~ 0 2 : 2 2 ~ 2p1 ⫹ Lsp =Lig T c 3D0 1 ⫺ Z0 sp rM =rN x~ c 233 8.1.4. Experimental comparisons for the ignition delay time In numerical calculations, the situation considered is that both the igniter and the specimen consist of mixtures of titanium and carbon particles, although their mixture ratios and degrees of dilution are different. In accordance with experiments [113] the initial temperature before arrival of the SHS flame is set to be 450 K, the compact diameter is 18 mm, and particle radius of carbon particles in the igniter is 5 mm. The Lewis number (Le0)ig for the igniter is set to be 100 [32] while that for the specimen, (Le0)sp, is 25 [74] because different kinds of carbon powder are used in an experiment; that in the igniter has an irregular shape, while that in the specimen has a spherical shape. Ratio of the heat penetration coefficients L sp/L ig is set to be 0.5, accordingly. Fig. 62 shows the ignition delay time tc as a function of the mixture ratio m ig in the igniter, with the emissivity e taken as a parameter, which is equivalent to varying heat loss rate. The particle radius (R0)sp in the specimen is 10 mm. With decreasing m ig, tc increases first gradually, then rapidly, and reaches the ignition limit, due to the decrease in heat supplied from the igniter. We see that the decrease in the heat loss makes tc shorter, because the extent of heat used for initiating the combustion wave becomes large. A further decrease in the heat loss brings about a situation at which the ignition limit locates beyond the extinction limit for the combustion wave in the igniter, because of the heat loss during the flame propagation [64]. Data points are experimental for R0 sp 10 mm [113]. We see that the ignition delay times are fairly correlated with e 1:0 when m ig is higher than about 0.8, while they are correlated with e 0 as m ig decreases. This is attributed to the Fig. 63. Ignition delay time tc for e 1, with (R0)sp in the specimen taken as a parameter [103]; (a) as a function of the mixture ratio m sp; and (b) as a function of the degree of dilution k sp. Data points are experimental [113]. 62 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 2 Fig. 64. Ignition delay time tc as a function of (R0)sp , with e taken as a parameter [103]. maximum temperature of the SHS flame that has passed through the igniter because influence of heat loss can be strongly suppressed as the flame temperature decreases by reducing m ig in the igniter. Note that tc becomes short as R0 sp in the specimen is reduced, because a decrease in the particle size increases the total surface area which participates in the reaction. Fig. 63(a) shows tc for e 1:0 as a function of m sp in the specimen with R0 sp taken as a parameter. The mixture ratio m ig in the igniter is 0.8. With decreasing m sp, tc for R0 sp 10 mm increases first gradually, then rapidly, and reaches the ignition limit because of a decrease in the heat generated during the flame initiation. As R0 sp is reduced, tc becomes short. In addition, with decreasing m sp, tc gradually increases and before reaching the ignition limit, encounters the flame extinction limit [64] corresponding to R0 sp ; beyond which establishment of the combustion wave is impossible, even though initiated. Data point are experimental [113] and a fair degree of agreement is demonstrated. Fig. 63(b) shows the similar comparison with respect to k sp. It is also reported that the same trends can be observed as those in Fig. 63(a) and (b) by reducing the heat loss, instead of the particle size. As for the effects of particle size on the ignition phenomena, little attention has been paid in the previous theoretical works, as was reviewed [101,102]. However, particle size is expected to exert great influences in the flame initiation no less than in the flame propagation and/or extinction, because it is closely related to the total surface area which participates in reaction [32,64]. Fig. 64 shows tc as a function of R0 2sp ; with e taken as a parameter when mig 0:8 and msp 1: As shown in Eq. (233), tc is proportional to ~ say, R0 2sp when e 0; because of the large value of k; larger than about 10. With increasing heat loss, tc becomes large, because the extent of heat used for initiating combustion wave becomes small. At a certain heat loss, with increasing R0 2sp ; tc increases first gradually, then rapidly, and reaches the ignition limit. The increase in tc is attributed to the decrease in the total surface area for the reaction, caused by the increase in R0 sp : Note that the flame extinction limit with respect to the particle radius is 20.9 mm when the compact diameter 2r is 18 mm [64]. Fig. 65(a) shows the ignitable range as functions of m sp and R20 = 2r: When there is no heat loss e 0; the ignitable range coincides with the flammable range because ignition is limited only by the flame extinction after the establishment of the combustion wave. The ignitable range is contracted as m sp is reduced. In addition, it narrows drastically as the heat loss increases, because of the appearance of the ignition limit. Experimental results [113] in Fig. 65(a) also represent a fair agreement. Fig. 65(b) shows the similar Fig. 65. Ignitable range with heat loss taken as a parameter [103]; (a) as functions of the mixture ratio m sp and the heat loss index R20 = 2rsp ; and (b) as a functions of the degree of dilution k sp and the heat loss index R20 = 2rsp : Data points are experimental [113]. (W) designates the flame initiation and ( × ) the non-ignition. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 comparison with respect to k sp and again demonstrates a fair degree of agreement. 8.1.5. Expression for the ignition energy Concerning the ignition delay time tc, fair agreement has been demonstrated between the predicted and experimental results. Now, let us evaluate the ignition energy [114] for establishing combustion wave in the specimen, by integrating heat flux through the contact surface [104,106,108]. Note here that the ignition energy consists of two parts: one is the energy supplied after arrival of the SHS flame, which has passed through the igniter, at the contact surface; and the other is supplied during the flame propagation in the igniter before arrival of the SHS flame at the contact surface. By the use of the temperature profile in the inert stage and the ignition delay time, the energy Ec, supplied after the arrival of the SHS flame, is obtained as E~ c us 1 ⫹ Lsp =Lig " 1 ~ tc } erfc k~ptc exp{ k~2 ⫺ w~ k k~ ⫺ w~ s p q # aig =asp k~ ⫺1 ⫹ {1 ⫺ erfc w~ k~tc } ⫺ w~ w~ aig =asp ⫺ k~ " ( ! ) s! aig 2 aig ~ × exp w~ ⫺ w~ k tc erfc w~ t ⫺1 asp asp c s! r! cig 4w~ 4w~ ⫺ 1⫹ 1⫹ 1⫹ 1⫹ csp k~ k~ig s ; CA ⫺ cig 4w~ 2 1⫹ csp k~ig 237 s! r! cig 4w~ 4w~ 1⫹ 1⫹ 1⫺ 1⫹ ⫺ csp k~ k~ig s CB ; cig 4w~ 2 1⫹ csp k~ig l=csp ~ k~ig k: l=cig 238 In deriving Eq. (236), a temporal variation of temperature in the specimen is first obtained as 0 q 1 ~ ⫹ k~ 2 ⫹ 4w~ k~ k B C 239 s ⱕ 0 : u B u exp @ sA; 2 B u 2us ( D s! cig 4w~ 1 1⫹ ; csp k~ig D q k~ig ⫹ k~ig2 ⫹ 4w~ k~ig ! × exp 235 s!) cig 4w~ ⫺ exp 1⫹ 1⫹ csp k~ig It may be informative to note that the expression of Ec obtained by Strunia et al. [106] is slightly different from that of Eq. (234) because their expression was derived under an assumption that there is no temperature distribution in the compacted medium, although they have taken the temperature distribution into account in obtaining the ignition delay time. As for the energy E0, supplied before the arrival of the SHS flame, although it has not been taken into account in the literature [106] is expressed as s! 1 rt ig 4w~ us E~ 0 1⫹ 1⫹ 2 rt ig k~ CB " ! ( s!⫺1 )# n ∞ X CA 2 4w~ 1 ⫹ 2n ⫹ 1 1 ⫹ ; ⫺ × CB k~ig k~ig n0 236 240 s! s!) cig 4w~ 4w~ ⫺ 1⫺ 1⫹ 1⫹ 1⫹ ~k csp k~ig 234 where Ztc 2t ma sp E~ c ⬅ Ec ⫺ du: 0 2s s0 l=csp q rt sp 0 where where s q #! w~ aig ⫹ {1 ⫺ erfc w~ k~tc } ; k~ asp 63 2 ( sf ⫺ s! 4w~ 1⫹ 1⫹ k~ q k~ig ⫺ k~ig2 ⫹ 4w~ k~ig 2 ! sf ; 241 when the propagating SHS flame in the igniter locates at s s f ⬎0: After that by assuming constant speed of the flame propagation, the following relation has been used s f ⫺k~ rt sp t rt ig 242 which is equivalent to xf ⫺ uig t: 8.1.6. Results for the ignition energy In numerical calculations, the situation considered is the same as that in Section 8.1.4. Fig. 66 shows the ignition energy E as a function of m ig in the igniter, with e taken as a parameter. It is seen that with decreasing m ig, the ignition energy E increases first gradually and then rapidly until the ignition limit, and that the contribution of E0 supplied before arrival of the SHS flame is relatively large. In addition, in spite of the decrease in the heat of combustion in the igniter, the energy E0 increases with decreasing m ig. This 64 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 trend is similar to that for the ignition delay time tc in Fig. 63(a). Fig. 67(b) shows the similar results with respect to k sp. The same trends as those in Fig. 67(a) and (b) can be observed by reducing heat loss, instead of reducing particle size. Fig. 68(a) shows E as a function of R0 sp2 ; with e taken as a parameter when mig 0:8 and msp 1: The general trend is similar to that for tc in Fig. 64. Fig. 68(b) shows the similar results for E at mig 0:65: The same trend as that in Fig. 68(a) is observed although the ignition energy is increased and the ignitable range is contracted. This is attributed to the decrease in the maximum temperature in the SHS flame that has passed through the igniter, due to the decrease in m ig in the igniter. 8.2. Flame initiation by use of radiative heat flux Fig. 66. Ignition energy E as a function of m ig in the igniter, with e taken as a parameter [114]. The specimen is stoichiometric with R0 sp 10 mm: Energy E0 is that supplied before the arrival of the SHS flame in the igniter. is attributed to the gradual heating of the specimen due to slow burning velocity of the SHS flame in the igniter. As for the energy supplied after the arrival of the SHS flame, it increases first gradually and then rapidly, with decreasing m ig, due to a decrease in the maximum temperature of the SHS flame in the igniter. It is also seen that at a certain m ig, a decrease in the heat loss makes the ignition energy smaller, because extent of heat used for initiating combustion wave becomes large. A further decrease in the heat loss brings about a situation at which the ignition limit locates beyond the extinction limit. Fig. 67(a) shows E for e 1:0 as a function of m sp in the specimen, with R0 sp taken as a parameter. The general Combustion wave in the SHS process can also be initiated by use of radiative heat flux, instead of using igniters. Although numerical results based on the heterogeneous theory are reported [115] it is considered that an extension of the heterogeneous theory for the ignition phenomena should be conducted by including analytical results related to the ignition of reactive condensed materials, such as explosives and solid propellants [116–120]. For this aim, it is considered the situation that the front surface of a compacted medium experiences partial, stepwise heating by the radiative heat flux [121–123]. In analyzing the inert stage, use has been made of the formulation presented by Niioka and Hasegawa [124]. In the analysis for the transition stage, related to the initiation of the combustion wave, results of the asymptotics [116] have been introduced into the heterogeneous theory. Comparisons between predicted and experimental results are also conducted. Fig. 67. Ignition energy E under heat loss condition e 1; with (R0)sp in the specimen taken as a parameter [114]; (a) as a function of m sp in the specimen; (b) as a function of k sp in the specimen. The mixture ratio m ig in the igniter is 0.8. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 65 Fig. 69. Model used for analyzing the SHS flame initiation by use of radiative heat flux [124]. with large activation energies, the compacted medium experiences only inert heat conduction during most of the time that t ⬍ tc because the heat generation term, related to reactant consumption, is exponentially small due to the large activation energy. Then, the equation of heat conduction in the axisymmetric coordinate becomes as 2u 22 u 22 u 1 2u ⫹ ⫹ ⫺ fu; 2t s r 2s r 2s z2 2s r2 243 which is subject to the initial and boundary conditions as t 0; u 0; I:C: B:C: Fig. 68. Ignition energy E as a function of R0 2sp ; with e taken as a parameter [114]; (a) the mixture ratio m ig in the igniter is 0.8; (b) mig 0:65: ~ s z 0; 兩s r 兩 ⱕ a; 2u ~ ⫺Q; 2s z 245 ~ s z 0; 兩s r 兩 ⬎ a; 2u 0; 2s z 246 s z ! ∞; s r ! ∞; u 0; 247 where si 8.2.1. Model definition and inert stage The problem of interest is the radiative heating of the front surface of an axisymmetric semi-infinite compacted medium (Fig. 69), originally consisting of a mixture of N particles, metal M, and an inert I that can be the product P of the reaction nM M ⫹ nN N ! nP P. The front surface of the compacted medium is subjected to a partial, stepwise heating with a constant heat flux Q for time t ⬎ 0; the diameter of heating area is 2a. Monodisperse size distribution is considered for N particles, with an initial radius R0 and number density of n0. As described in Section 8.1.1, in the ignition processes 244 a~ Z∞ ⫺ Z0 ma xi Q ; i z; r; Q~ ; l=c Z∞ ⫺ Z0 2 q0 ma Z∞ ⫺ Z0 ma a ; l=c C ; Z ∞ ⫺ Z0 2 # 2! " l 4es SB T 2 ⫹ T02 T ⫹ T0 R0 C : rtc 2r rt c u0;a R0 2 f Noted that the inert problem described has a known solution 66 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 and can be written as [123,125] ( q ! Q~ a~ Z∞ ~ exp ⫺ s z l 2 ⫹ f uI J ls r J1 la 2 0 0 q! q! sz 2 p ⫺ e t l ⫹ f ⫺ exp s z l 2 ⫹ f × erfc 2 t q!) sz dl p ⫹ l 2 ⫹ f p ; × erfc (248) 2 t l2 ⫹ f variables are to be introduced. where Jn is the Bessel function of the first kind of order n. Note that temperature rise in the inert, heating stage is designated with the subscript I. where b is the Zeldovich number defined in Eq. (225). By retaining only the first nonvanishing terms on the exponential, the energy conservation Eq. (223) yields 8.2.2. Transient stage In the same manner as that described in Section 8.1.2, the transition stage can be analyzed. The only difference is the position at which departure from the inert heating occurs; that is, s z s r 0: Therefore, Eq. (248) is expanded as uI uI c ⫹Kt t ⫺ tc ⫺ Kz s z;c ; 249 where ( p Q~ uI c p 2 ⫺ 2erfc ftc 2 f p ⫺ exp ⫺a~ f erfc p a~ p ⫺ ftc 2 tc ! p p a~ ⫹ exp a~ f erfc p ⫹ ftc ; 2 tc Kt 2uI 2t !) (250) 251 ⫺Kz ! ~ ⫺Q: 253 s bKt t ⫺ tc ( ! !) 1 x~ T~ d p 2 ⫹ ln ; exp ⫺ bKz Z∞ ⫺ Z0 L0 rt;0 =rt T~ c 254 X bKz s z;c ; d 2c 1 ⫺ p exp c ⫺ X ⫹ s; b dX 2 which is subject to dc c ∞; s c X; ⫺∞ 0: dX X0 255 256 257 The condition as s ! ⫺∞ is obtained from matching to the inert stage. By using c ⫺ X as the dependent variable, Eq. (256) is integrated to produce s dc 2 1 ⫺ 1 ⫹ p {exp c 0 ⫹ s ⫺ exp c ⫺ X ⫹ s}; dX b 258 ( !) Q~ a~ 2 exp ⫺ftc ; p 1 ⫺ exp ⫺ ptc 4tc c 2uI 2s z c b u ⫺ u I ; 252 c Terms relevant to s r can be ignored because temperature distribution in the radial direction is flat around s r 0: Here, (·)c designates the condition corresponding to t tc and s z s r 0: Since the independent variables in the transition stage are just t and s z, the use of a set of one-dimensional, transient governing equations in the heterogeneous theory [92,95] can be justified. If we further ignore the heat loss effect, as is the case for usual inner problems related to ignition, the energy conservation equation coincides with Eq. (223). This identification infers that the present problem closely parallels the earlier development [116] for the ignition of solid propellants. In analyzing the reactive-diffusive region near the surface of the compacted medium, new stretched where c 0 c 0; s: This solution must match to that of the transient-diffusive region which has the spatial variable p p bKt v bK t s z X: 259 bKz Application of matching and introduction of the time variable Kz s~ s ⫹ ln p 260 Kt results in the transient–diffusive problem as 2c 2 2c 2c ; ⫺exp c 0; s~ ⫹ s~ ; 2 2~s 2 v 2v v0 261 c ∞; s~ c v; ⫺∞ 0; which has already been solved by Lin˜a´n and Williams ÿ [116]. It is reported that the thermal runaway, that is c 0; s~ ! ∞; occurs at a finite time s~ ⫺0:431: Since it is defined that t t c at ignition, Eqs. (254) and (260) yield an implicit formula for the ignition delay time as ! ! 1 x~ T~ p exp ⫺ d e ⫺0:431 Kz bKt Z∞ ⫺ Z0 L0 rt0 =rt T~ c 0:650: (262) A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 67 We see that the ignition delay time can be expressed as functions of not only Td, Q, and a, as pointed out by Niioka and Hasegawa [124] but also the operating parameters characteristic to the SHS process, such as the mixture ratio m , the degree of dilution k , the initial temperature T0, and the initial particle radius R0. This can be seen from the definition of t , which is expressed as ! rt;0 rM t 263 1 ⫺ Z0 t 3D0 L0 rt rN R20 by using Eq. (71), such that tc ⬃ tc R20 for a given stoichiometry and/or a degree of dilution, once t c is obtained by solving Eq. (262). Note also that 1 ⫺ Z0 m 1 ⫺ k= m ⫹ fst ; shown in Eq. (62). The dependence of the nondimensional parameters on the operating parameters, in the same manner as that in Eq. (263), is represented as follows. For Q~ and a~ we have Q~ QR0 q0 r t;0 D0 Z∞ ⫺ Z0 {3Le0 L0 1 ⫺ Z0 rM =rN }1=2 ⬃ QR0 ; a~ 1=2 3L0 r a a ⬃ ; 1 ⫺ Z0 M R0 R0 Le0 rN 264 265 which suggest that for a given stoichiometry and/or a degree of dilution, both Q~ and a~ depend on the particle radius R0 even when the heat flux Q and the radius a of the heating area are fixed, as are the cases for usual experimental conditions. The energy supplied per unit area in the heating surface, required for initiating the combustion wave, is given by 1=2 1 3L0 r Qtc Q~ tc 0 1 ⫺ Z0 M rN R0 q rt Z∞ ⫺ Z0 Le0 Qtc ⬃ ; R0 266 while the total energy supplied is simply ! Qtc a 2 ~ Qtc pa~ ⬃ : R30 267 It is clear that the particle radius R0 is an essential parameter affecting initiation of the combustion wave. 8.2.3. Results and experimental comparisons In calculations, use has been made of physicochemical parameters for Ti–C system [32,64] as described in Sections 4.4 and 5.4.1; a Lewis number of Le0 25 is used [75] in accordance with an experiment. Let us first examine the situation without heat loss, and then with heat loss. 8.2.3.1. Situation without heat loss Fig. 70(a) shows the ignition delay time tc as a function of the mixture ratio m , Fig. 70. Ignition delay time tc for Ti–C system with R0 5 mm; with the radiative heat flux Q taken as a parameter [121]; (a) as a function of the mixture ratio m ; (b) as a function of the degree of dilution k . Data points are experimental [126]; open symbol designates the ignition and flame establishment, while solid symbol designates the ignition but extinguishment. with the radiative heat flux Q taken as a parameter. The initial radius R0 of N particles in the compacted medium is set to be 5 mm. We see that the ignition delay time increases with decreasing m , due to a decrease in heat generated during the flame initiation. A further decrease in m makes the combustion wave impossible to propagate, for the same reason as that in Fig. 63(a). As for the effect of Q, we see that tc increases with decreasing Q, because its decrease prolongs the heating period required for the front surface to be heated to the ignition temperature. Data points are experimental [126] for different heat fluxes (3.3 and 2.57 MW/m 2) and a fair degree of agreement is demonstrated. Here, it may be informative to note that in general, tc for the radiative heating is much longer than that for the igniter. This may be attributed to the fact that heat supply by the SHS flame with high temperature can be made much more rapidly than by the radiative heat flux. Fig. 70(b) shows a similar comparison with respect to the degree of dilution k . The same comment can be made as that in Fig. 70(a). Fig. 71 shows tc for m 1 as a function of R0, with Q taken as a parameter. We see that tc increases with increasing R0, in the same reason as that in Fig. 64. It is important to emphasize that such a significant parametric dependence cannot be described by the homogeneous premixed-flame theory. The decrease in Q causes tc longer, because of the prolonged heating period of time. Fig. 72 shows tc for m 1 with R0 5 mm; as a function of the radius a of the heating area, with Q taken as a parameter. With decreasing a, tc becomes longer, first gradually and then rapidly, due to a decrease in the heating area at the front surface. Fig. 73(a) and (b) shows tc as functions of m and k , 68 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 71. Ignition delay time tc for Ti–C system of the stoichiometric mixture m 1; as a function of R0, with Q taken as a parameter [121]. Data points are experimental [126]. respectively, with Q and a taken as parameters. Because of the strong effect in decreasing the heating area, as shown in Fig. 72, tc becomes longer. In addition, this effect becomes remarkable as m decreases or k increases, because the increase in m or decrease in k reduces the heat generated during the flame initiation. Fig. 74 shows tc for m 1 as a function of R0, with Q and a taken as parameters. It is further confirmed that the effect of the heating area on tc is strong, as well as the effect of R0. 8.2.3.2. Situation with heat loss Now, let us examine effects Fig. 73. Ignition delay time tc for Ti–C system with R0 5 mm; as a function of k , with Q and a taken as parameters [122]; (a) as a function of m ; and (b) as a function of k . Data points are experimental [126]. Fig. 72. Ignition delay time tc for Ti–C system of m 1 with R0 5 mm; as a function of the radius a of the heating area, with Q taken as a parameter [121]. Data points are experimental [126]. of heat loss on the ignition delay time. Since the ignition delay time t c is determined by solving the criterion in Eq. (262), dependence of the LHS of Eq. (262) on the heat loss f is to be examined, with t c taken as a parameter. As shown in Fig. 75, with increasing f , the LHS of the criterion, designated by F, first decreases slightly and then increases A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 69 which initiation of the combustion wave cannot be realized. The state of the turning point can be determined by dF=df 0; which yields duc ⫺ df t=2 ⬃ 0: Z∞ ⫺ Z0 b⫹ T~ c 268 It is informative to note that as a first approximation, p a~ ft p 2 t Fig. 74. Ignition delay time tc for Ti–C system of m 1 with R0 5 mm; as a function of R0, with Q and a taken as parameters [122]. Data points are experimental [126]. rapidly. When tc 2:4317; F for f 0 is 0.650, which corresponds to the value for the thermal runaway in the adiabatic limit. Another solution for tc 2:4317 exists when f is about 0.093, at which heat loss is quite large and it is considered to be unrealistic. It is further seen that for F 0:650; there exist two solutions for 2:4317 ⱕ tc ⬍ 2:7252; one solution at tc 2:7252; and there is no solution for tc ⬎ 2:7252: Thus, there exists an ignition limit, beyond 269 can be a solution for determining the state of the turning point. Fig. 76 shows the ignition delay time tc as functions of the mixture ratio m and degree of dilution k , with the radiative heat flux Q and the heat loss taken as parameters. The conditions are the same as those in Fig. 70. Note that the emissivity e represents the extent of the heat loss. It is seen that, as the emissivity e increases, which results in an increase in the heat loss, the ignitable range becomes narrow although the ignition delay time itself is not influenced very much by the heat loss. Fig. 77 shows tc for the stoichiometric mixture as a function of the radius R0, with the radiative heat flux Q and the emissivity e taken as parameters. The results clearly demonstrate that the system ignitability can be profoundly affected by the particle size in the presence of heat loss. Since heat loss is invariably present, the existence of the loss induced ignition limit is particularly significant in assessing the global performance of a synthesis system. 8.3. Some remarks on the flame initiation In the SHS process, the energy required is basically for initiating SHS flame, such that effects of various parameters Fig. 75. Behavior of the thermal runaway criterion on the heat loss [123], for the stoichiometric mixture of Ti–C system with R0 5 mm; heated by the radiative heat flux of 3.3 MW/m 2, with 10 mm in heating diameter. 70 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Fig. 76. Ignition delay time tc for Ti–C system, with the radiative heat flux Q and the emissivity e taken as parameters [123]; (a) as a function of the mixture ratio m ; (b) as a function of the degree of dilution k . Conditions are the same as those in Fig. 70. on the flame initiation are indispensable for practical utility. Investigation based on the heterogeneous theory suggests that the ignitable range is limited by the heat loss during flame initiation when the particle size is large and the compact diameter is small; while it is the heat loss during flame propagation that confines the ignitable range when the particle size is small and the compact diameter is large. This identification provides useful insight in controlling ignition phenomena in the SHS process by varying particle size, which is expected to make a contribution in applying the SHS process for near-net-shape fabrications by use of the SHS process. 9. Concluding remarks In the following the current status of SHS research on flame propagation, extinction, and initiation by use of the heterogeneous theory is summarized, and areas which require further research are suggested. 9.1. Summary of the present survey From a phenomenological consideration of flame propagation in the SHS process, it may appear in the combustion wave as a slurry-like configuration, which consists of the higher melting-point nonmetal particles and the molten metal. This consideration leads to the important identification about the SHS process that, although the flame propagation in the bulk appears like that of the premixed flame, this bulk flame propagation is eventually supported by the nonpremixed reaction of the dispersed nonmetal particles in the molten metal. This nonpremixed mode has not been incorporated in formulations in the previous theoretical works based on the theory for the Fig. 77. Ignition delay time tc for Ti–C system of the stoichiometric mixture m 1; as a function of the particles radius R0, with Q and e taken as parameters [123]. premixed flame in the homogeneous medium. So that heterogeneous features of the combustion wave have been unresolved in previous theoretical works. Recognizing the limitation of a homogeneous premixedflame theory in describing the heterogeneous nature related to the SHS process (especially on the influence of particle size), the existing theory on flame propagation in fuel sprays has been extended to describe such processes. Specific modifications included are finite rate of reaction at the particle surface and Arrhenius mass diffusion in the liquid phase of molten metal, so that the particle consumption rate controlled by both diffusion and reaction can be incorporated into the formulation, and prediction and characterization of such important bulk flame propagation as the flame propagation speed, the range of flammability, and the flame extinction limit, have successfully been made in relation to the heterogeneity in the combustion wave. Calculated results for several representative ceramics and intermetallic compounds, such as Ti–C, Ti–B, Zr–B, Hf–B, Al–Ni, Co–Ti, Ni–Ti systems, using consistent sets of physicochemical parameters, respectively, show satisfactory agreement with experimental results in the literature. An important relation that the burning velocity is inversely proportional to the nonmetal particle size has been confirmed. These comparisons further suggest that the nonmetal (or higher melting-point metal) particles in these systems might react in the diffusion-controlled limit, indicating that finite rate chemical reaction is not the controlling factor in the propagation of bulk flame. Under such situations, the Arrhenius liquid-phase mass-diffusivity exerts the dominant temperature sensitivity on the flame propagation in the SHS process. In addition, because of the excessively slow rate of diffusion compared to the thermometric A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 conductivity and reaction rate (suggested by the excessively large Lewis number in the combustion wave), it has been clarified that concentration of the molten metal is mainly affected through consumption. A nonadiabatic, heterogeneous theory for the SHS process with volumetric heat loss, has further been formulated and shown that the characteristic extinction turning point exists in the same manner as that in the premixed flame. It has also been found that the range of flammability under heat loss conditions can well be described, in terms of its dependence on stoichiometry, dilution, initial temperature, particle size, and compact diameter. A particularly significant identification is that the process responsible for its temperature sensitivity and ultimate extinguishment is the Arrhenius liquid-phase mass-diffusion. The fair agreement between predicted and experimental results for the extinction limits in trend and in approximate magnitude suggests that this theory has captured the essential features of the nonadiabatic heterogeneous flame propagation in the SHS process. A useful parameter called the heat loss index, which is defined as the square of the nonmetal particle size divided by the compact diameter, has also been found. Based on these identifications, not only an approximate expression for burning velocity in the adiabatic condition, but also for extinction criterion in the nonadiabatic condition, has been derived under the appropriate and realistic assumptions of diffusion-limited reaction and infinite Lewis number. Effects of the dominant parameters on the burning velocity and/or extinction criterion have been clarified, yielding useful insight into manipulating the SHS flame propagation and extinction. Since a fair degree of agreement with numerical results has been demonstrated by examining applicability of these expressions, these expressions are expected to have general, qualitative and quantitative utility, in each case. Bimodal particle dispersion for the nonmetal particles has also been taken into account in the nonadiabatic heterogeneous theory. It has resulted that the mathematical problem can be rendered identical to the monomodal case by introducing an equivalent particle radius expressed as a function of the bimodal parameters. This identification has further provided useful insight into manipulating the flame propagation and/or extinction by making the system bimodal, especially in recognizing that because of the increase in the total surface area of particles in the combustion wave, the reduced equivalent particle size in the bimodal system increases the burning velocity and broadens the range of flammability. Furthermore, flame instability, usually caused by the high energy-release rate in the combustion wave when the preparation is made with small particles, can even be suppressed by increasing the equivalent particle size in the bimodal system. The heterogeneous theory which has succeeded in capturing the essential feature of the heterogeneity in the SHS process has further been examined to elucidate differences from the homogeneous premixed-flame theory. Although 71 differences are minor from the mathematical point of view, important identification which has already been pointed out for the physical interpretation has been reconfirmed. So that qualitative validity of theoretical results based on the homogeneous premixed-flame theory has been shown, if the differences involved are noted. The field activated SHS, which is effective especially for low exothermic systems to sustain flame propagation, can also be examined by use of the heterogeneous theory, by introducing an additional term due to external heating by electric current. It has been shown that the external heating enhances the burning velocity and extends the range of flammability, as expected. Significant identifications obtained are that extent of these effect is strongly influenced by the particle size, as well as the physicochemical parameters, and that the range of flammability is still restricted even under external heating conditions, neither of which have been captured by the homogeneous premixed-flame theory. In this context, the heterogeneous theory is indeed viable in examining the field activated SHS. The study for the transition boundary from steady to pulsating combustion by use of the neutral stability analysis, has also been conducted and it has been shown that there exists a limit in the use of small particles, which is an important suggestion for near-net-shape fabrication by use of the SHS process. As for the effects of transverse wave number, which are related to heat loss on the side surface, on the location of the boundary, it has been found to be relatively small from a practical point of view. In addition to the flame propagation and/or extinction, initiation of the combustion wave has been examined because energy required is basically that for initiating SHS flame and effects of various parameters on the flame initiation are indispensable for practical utility. It has been found that the ignitable range is limited by heat loss during flame initiation when the particle size is large and/or compact diameter is small; while it is heat loss during flame propagation that confines the ignitable range when the particle size is small and/or compact diameter is large. This identification provides useful insight to controlling ignition phenomena in the SHS process by varying particle size, which is expected to make various contributions in practical applications. 9.2. Area for further research In view of the recent interest to produce Functionally Graded Materials (FGMs) by use of the SHS process, a candidate for research is the examination of flame propagation in a medium with distributed reactant concentrations, not only in the longitudinal direction but also in the lateral directions. In this situation, contour of the combustion wave is no longer planar and the combustion wave spreads two-dimensionally, because of the distributed heat generation rate and/or two-dimensional temperature distribution, caused by the distributed reactant concentrations. 72 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 Since flame strength is not constant along the curved combustion wave, the concept of flame stretch might be of great use in fundamental understanding of this flame propagation and its related extinction. Another candidate for research related to producing FGMs is that of examining flame propagation for multicomponent system. Since both diffusion-controlled combustion for some components and reaction-controlled combustion for other components can proceed simultaneously in the combustion wave, recognition of dominant parameters for each component and/or each competing reaction might be indispensable for conducting theoretical works and fundamental understanding. This identification would further facilitate exploration of such combustion mechanism as bulk flame propagation, which proceeds prior to the other bulk flame propagation, observed for multicomponent system. In order to gain fundamental understanding for the SHS flame propagation, it might be of great importance to incorporate results of microscopic observations into formulations of theoretical works, because information indispensable for constructing the heterogeneous theory is that of microscopic observations. In this context, further accumulation of experimental results through microscopic observations is strongly advised, in order to refine the heterogeneous theory. In addition to microscopic observations, experiments in the gravity free environment might be of great value in obtaining further fundamental understanding, although this experiments cannot be easily conducted on earth. For the SHS process, natural convection would exert little influence in removing heat from the combustion wave, compared to that of thermal radiation, but density difference between the nonmetal particles and the molten metal in the combustion wave (by yielding local change of stoichiometry, thickness of the combustion wave, etc.), could exert influences, because of the existence of gravity. Although not specifically covered in the present review, the field activated SHS might also prove a fertile area of research because this is indispensable for low exothermic systems to initiate and/or sustain combustion wave. In addition to the effects of external heating, if influences of electric and/or magnetic fields on profiles of reactant concentrations could be utilized in producing advanced materials, useful and multifunctional characteristics would be generated by inducing heterogeneity in composition and/or texture. Finally, it may be informative to point out the similarity between spray combustion and the SHS process, as has been mentioned in the present review. This similarity also suggests that the SHS process might be used for the basic research of spray combustion, especially for identifying effects of dominant parameters on the flame propagation, extinction, and/or initiation, because influence of thermal expansion which usually conceal those of the other parameters is relatively weak in the SHS process. Acknowledgements It is a pleasure to acknowledge the continuing interest and support that the author has received for this research on flame propagation, extinction, and/or initiation of the combustion wave in the SHS process, through a series of the Grant-in-aid for Scientific Research of the Ministry of Education, Science, Sports, and Culture, the Iwatani Naoji Foundation’s Research Grant, the Research Grant of Tokai Foundation for Promoting Industrial Technologies, and the Mechanical Industry Development and Assistance Foundation’s Research Grant, Japan. References [1] Merzhanov AG, Borovinskaya IP. Dokl Akad Nauk SSSR 1972;204:366 (in Russian). [2] Merzhanov AG, Filonenko AK, Borovinskaya IP. Dokl Akad Nauk SSSR 1973;208:892 (in Russian). [3] Merzhanov AG, Borovinskaya IP. Combust Sci Technol 1975;10:195. [4] Merzhanov AG. Archivum Combustionis 1981;1:23. [5] Frankhouser WL, Brendley KW, Kieszek MC, Sullivan ST. Gasless combustion synthesis of refractory compounds. Park Ridge: Noyes, 1985. [6] Munir ZA. Ceram Bull 1988;67:342. [7] Munir ZA, Anselmi-Tamburini U. Mater Sci Rep 1989;3:277. [8] Merzhanov AG. Self-propagating high-temperature synthesis: twenty years of search and findings. In: Munir ZA, Holt JB, editors. Combustion and plasma synthesis of hightemperature materials, New York: VCH Publishers, 1990. p. 1–53. [9] Varma A, Lebrat J-P. Chem Engng Sci 1992;47:2179. [10] ISMAN, Self-propagating high-temperature synthesis of inorganic compounds. In: ISMAN leaflet, Institute of Structural Macrokinetics, Russian Academy of Science, Chernogolovka, 1993. [11] Merzhanov AG, Dvoryankin AV, Strunina AG. Dokl Akad Nauk SSSR 1982;267:869 (in Russian). [12] Strunina AG, Dvoryankin AV, Merzhanov AG. Combust Explos Shock Waves 1983;19:158. [13] Novozhilov BV. Dokl Akad Nauk SSSR 1961;141:151 (in Russian). [14] Merzhanov AG. Combust Flame 1969;13:143. [15] Maksimov EI, Shkadinskii KG. Combust Explos Shock Waves 1971;7:392. [16] Aldushin AP, Martem’yanova TM, Merzhanov AG, Khaikin BI, Shkadinskii KG. Combust Explos Shock Waves 1973;9:531. [17] Vol’pert VA, Vol’pert AI, Merzhanov AG. Combust Explos Shock Waves 1983;19:435. [18] Khaikin BI, Khudyaev SI. Dokl Phys Chem, Acad Sci USSR 1979;245:225. [19] Aldushin AP, Khaikin BI, Shkadinskii KG. Combust Explos Shock Waves 1976;12:725. [20] Firsov AN, Shkadinskii KG. Combust Explos Shock Waves 1987;23:288. A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 [21] Boddington T, Cottrell A, Laye PG. Combust Flame 1989;76:63. [22] Margolis SB, Williams FA. Flame propagation in solids and high-density fluids with Arrhenius reactant diffusion. In: Sandia Report, SAND90-8508, Sandia National Laboratories, 1990. [23] Merzhanov AG, Khaikin BI. Prog Energy Combust Sci 1988;14:1. [24] Margolis SB. Prog Energy Combust Sci 1991;17:135. [25] Dunmead SD, Readey DW, Semler CE, Holt JB. J Am Ceram Soc 1989;72:2318. [26] Shkiro VM, Borovinskaya IP. Study of combustion mechanism for mixtures of titanium with carbon. In: Merzhanov AG, editor. Combustion processes in chemical technology and metallurgy. Akad. Nauk SSSR, OIKhF, 1975. p. 253–8 (in Russian). [27] Nekrasov EA, Maksimov YuM, Ziatdinov MKh, Shteinberg AS. Combust Explos Shock Waves 1978;14:575. [28] Deevi SC. J Mater Sci 1991;26:2662. [29] Hardt AP, Phung PV. Combust Flame 1973;21:77. [30] Armstrong R. Combust Sci Technol 1990;71:155. [31] Merzhanov AG. Combust Sci Technol 1995;105:295. [32] Makino A, Law CK. Heterogeneous flame propagation in the self-propagating high-temperature synthesis (SHS) process: theory and experimental comparisons. In: The 24th Symposium (International) on Combustion. The Combustion Institute, Pittsburgh. 1992. p. 1883–91. [33] Makino A, Araki N, Kuwabara T. Trans JSME Ser B 1992;58:1925 (in Japanese); JSME Int’l J Ser. B (English Translation) 1994;37:576. [34] Williams FA. Theory of laminar flames. Combustion theory. 2nd ed. Menlo Park: Benjamin/Cummings, 1985. p. 130–81. [35] Williams FA. Spray combustion. Combustion theory. 2nd ed. Menlo Park: Benjamin/Cummings, 1985. p. 446–84. [36] Makino A. Combust Flame 1992;90:143. [37] Makino A, Law CK. Combust Flame 1995;101:551. [38] Kanury AM. Metall Trans A 1992;23:2349. [39] Huque A, Kanury AM. Combust Sci Technol 1993;89:27. [40] Makino A. Ann Chim Fr 1995;20:143. [41] Touloukian YS. Thermophysical properties of high temperature solid materials. Thermophysical properties research center, Purdue University, vol. 5, New York: MacMillan, 1967. [42] Kirdyashkin AI, Maksimov YuM, Merzhanov AG. Combust Explos Shock Waves 1981;17:591. [43] Makino A, Kita K, Araki N. Nensho no Kagaku to Gijutsu (Combust Sci Technol, Jpn) 1993;1:165 (in Japanese). [44] Bloshenko VN, Bokii VA, Borovinskaya IP, Merzhanov AG. Combust Explos Shock Waves 1984;20:676. [45] Hardt AP, Holsinger RW. Combust Flame 1973;21:91. [46] Azatyan TS, Mal’tsev VM, Merzhanov AG, Seleznev VA. Combust Explos Shock Waves 1977;13:156. [47] Makino A, Law CK. J Am Ceram Soc 1994;77:778. [48] Touloukian YS, Powell RW, Ho CY, Nicolaou MC. Thermal diffusivity, Thermophysical properties of matter, The TPRC Data Series, vol. 10. New York: IFI/Plenum Press, 1973. [49] Kirdyashkin AI, Maksimov YuM, Merzhanov AG. Combust Explos Shock Waves 1981;17:591. [50] Zenin AA, Merzhanov AG, Nersisyan GA. Combust Explos Shock Waves 1981;17:63. [51] Borovinskaya IP, Merzhanov AG, Novikov NP, Filonenko AK. Combust Explos Shock Waves 1974;10:2. 73 [52] Novikov NP, Borovinskaya IP, Merzhanov AG. Combust Explos Shock Waves 1974;10:157. [53] Itin VI, Bratchikov AD, Merzhanov AG, Maslov VM. Combust Explos Shock Waves 1981;17:293. [54] Itin VI, Bratchikov AD, Merzhanov AG, Doronin VN. Combust Explos Shock Waves 1982;18:536. [55] Shcherbakov VA, Pityulin AN. Combust Explos Shock Waves 1983;19:631. [56] Maksimov YuM, Merzhanov AG, Raskolenko LG, Pak AT, Lepakova OK. Dokl Akad Nauk SSSR 1986;286:911 (in Russian). [57] Maslov VM, Borovinskaya IP, Merzhanov AG. Combust Explos Shock Waves 1976;12:631. [58] Naiborodenko YuS, Itin VI. Combust Explos Shock Waves 1975;11:293. [59] Naiborodenko YuS, Itin VI. Combust Explos Shock Waves 1975;11:626. [60] Bratchikov AD, Merzhanov AG, Itin VI, Khachin VN, Dudarev EF, Gyunter VE, Maslov VM, Chernov DB. Sov Powder Metall Met Ceram 1980;19:5. [61] Itin VI, Khachin VN, Bratchikov AD, Gyunter VE, Chernov DB. Inorg Mater 1980;16:1337. [62] Samsonov GW, Vinitskii IM. Handbook of refractory compound. New York: Plenum Press, 1980. [63] Graham D. Summary of impurity diffusion in the beta phase of titanium. In: Wheeler Jr. JA, Winslow FR, editors. Diffusion on body-centerd cubic metals. Metals Park: American Society for Metal, 1965. p. 27–34. [64] Makino A, Law C K. Propagation and diffusion-limited extinction of nonadiabatic heterogeneous flame in the SHS process. In: The Twenty-Fifth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, 1994, p. 1659–67. [65] Williams FA. Ignition. Extinction, and flammability limit. in combustion theory. 2nd ed. Menlo Park: Benjamin/ Cummings, 1985. p. 265–93. [66] Makino A, Law CK. J Am Ceram Soc 1996;79:3097. [67] Munir ZA, Lai W. Combust Sci Tech 1992;88:201. [68] Munir ZA, Sata N. Int J SHS 1992;1:355. [69] Touloukian YS, DeWitt DP. Thermophysical properties of matter, vol. 8. New York: IFI/Plenum Press, 1972. [70] Itin VI, Bratchikov AD, Doronin VN, Pribytkov GA. Sov Phys J (English translation) 1981;24:1134. [71] Itin VI, Naiborodenko YuS, Bratchikov AD, Butkevich NP, Korostelev SV, Sholokhova LV. Sov Phys J (English translation) 1975;18:408. [72] Makino A, Law CK. Int J SHS 1995;4:25. [73] Kaper HS, Leaf GK, Margolis SB, Matkowsky BJ. Combust Sci Technol 1987;53:289. [74] Margolis SB, SIAM. J Appl Math 1983;43:351. [75] Makino A, Law CK. Combust Sci Technol 1995;106:193. [76] Makino A. SHS flame propagation and extinction for titanium and diamond system: theory and experimental comparisons. In: Chan SH, editor. Transport phenomena in combustion, vol. 1. Washington: Taylor and Francis, 1996. p. 678–88. [77] Makino A, Yamada T, Araki N. Nensho no Kagaku to Gijutsu (Combust Sci Technol, Jpn) 1995;3:41 (in Japanese). [78] Makino A, Sugiyama K, Araki N. Nensho no Kagaku to Gijutsu (Combust Sci Technol, Jpn) 1996;4:93 (in Japanese). [79] Makino A. Energy Convers Mgmt 1997;38:1043. 74 A. Makino / Progress in Energy and Combustion Science 27 (2001) 1–74 [80] Makino A, Izumi H, Araki N. Nensho no Kagaku to Gijutsu (Combust Sci Technol, Jpn) 1998;6:65 (in Japanese). [81] Makino A, Law CK. Submitted for publication. [82] Knyazik VA, Merzhanov AG, Solomon VB, Shteinberg AS. Combust Explos Shock Waves 1985;21:333. [83] Yamada O, Miyamoto Y, Koizumi M. J Mater Res 1986;1:275–9. [84] Feng A, Munir ZA. Metall Mater Trans B 1995;26:581. [85] Feng A, Munir ZA. Metall Mater Trans B 1995;26:587. [86] Feng A, Munir ZA. J Am Ceram Soc 1996;79:2049. [87] Munir ZA. Int J SHS 1997;6:165. [88] Filimonov IA, Kidin NI. SHS-process in external electric fields. In: The 24th Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, 1992. p. 1893–8. [89] Makino A. Investigation of effects of electric field on the combustion behavior in the SHS process by use of the heterogeneous theory. In SHS-99, Moscow, Russia, 1999. [90] Thermophysical Properties Handbook. Japan Society of Thermophysical Properties, editor. Tokyo: Yokendo, 1990. p. 232 (in Japanese). [91] Touloukian YS. Thermophysical properties of high temperature solid materials, vol. 1. Thermophysical properties research center, Purdue University. New York: Macmillan, 1965. p. 85 [92] Makino A, Law CK. Pulsating instability in the nonadiabatic heterogeneous SHS flame: theory and experimental comparisons. In: The 26th Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1996. p. 1867– 74. [93] Shkadinsky KG, Shkadinskaya GV, Matkowsky BJ, Volpert VA. Combust Sci Technol 1992;88:271. [94] Butakov AA, Vaganov DA, Leont’ev SN. Combust Sci Technol 1995;106:137. [95] Makino A, Law CK. On the transition boundary from steady to pulsating combustion in SHS flames. In: The 27th Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1998. p. 2469–76. [96] Margolis SB, Kaper HG, Leaf GK, Matkowsky BJ. Combust Sci Technol 1985;43:127. [97] Margolis SB. Matkowsky. SIAM J Appl Math 1988;48:828. [98] Margolis SB. Proc R Soc London A 1991;433:131. [99] Margolis SB, SIAM. J Appl Math 1991;51:693. [100] Margolis SB. Combust Sci Technol 1993;88:223. [101] Barzykin VV, Merzhanov AG, Strunina AG. Ignition of heterogeneous systems containing condensed reaction products. In: The 23rd Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, 1990. p. 1725–31. [102] Barzykin VV. Pure Appl Chem 1992;64:909. [103] Makino A. Dependence of the ignition delay time on various parameters in the SHS process initiated by another SHS flame. In: Proceedings of the 5th ASME/JSME Joint Thermal Engineering Conference, AJTE99-6321, San Diego, CA, 1999. [104] Averson E´A, Barzykin VV, Merzhanov AG. Dokl Phys Chem, Acad Sci USSR 1968;178:1. [105] Strunina AG, Martem’yanova TM, Barzykin VV, Ermakov VI. Combust Explos Shock Waves 1974;10:449. [106] Strunina AG, Vaganova NI, Barzykin VV. Combust Explos Shock Waves 1977;13:707. [107] Strunina AG, Ermakov VI, Barzykin VV. Combust Explos Shock Waves 1978;14:441. [108] Strunina AG, Ermakov VI, Averson E´A. Combust Explos Shock Waves 1979;15:484. [109] Strunina AG, Ermakov VI, Kostin SV, Barzykin VV. Combust Explos Shock Waves 1981;17:351. [110] Strunina AG, Firsov AN, Kostin SV. Combust Explos Shock Waves 1981;17:500. [111] Law CK. Int J Heat Mass Transfer 1978;21:1363. [112] Enig JW. Proc R Soc A 1968;305:205. [113] Yamada T, Makino A, Araki N. Measurement of ignition energy in the SHS process for Ti–C system. In: Proceedings of the 32nd Japanese Symposium on Combustion. The Japanese Section of the Combustion Institute, Sendai, 1994. p. 579–81. [114] Makino A. Initiation of SHS flame induced by another SHS flame—determination of the ignition energy. In: Proceedings of the Second International Symposium on Advanced Energy Conversion Systems and Related Technologies, RAN98, 1998. p. 242–3. [115] Kanury AM, Hernandez-Guerrero A. Combust Sci Technol 1995;104:181. [116] Lin˜a´n A, Williams FA. Combust Sci Technol 1971;3:91. [117] Linan A, Williams FA. Combust Flame 1972;18:85. [118] Lin˜a´n A, Crespo A. Combust Sci Technol 1972;6:223. [119] Niioka T, Williams FA. Combust Flame 1977;29:43. [120] Niioka T. Combust Sci Technol 1978;18:207. [121] Makino A. Initiation of the heterogeneous flame in the selfpropagating high-temperature synthesis (SHS) process induced by radiative heat flux. In: ASPACC99, Taiwan, 1999. [122] Makino A. Initiation of the heterogeneous SHS by radiative heat flux. In: Proceedings of the 37th Japanese Symposium on Combustion. The Japanese Section of the Combustion Institute, Chiba, 1999. p. 395–6. [123] Makino A, Law CK. Transient radiative initiation of the heterogeneous flame in SHS: theory and experimental comparisons. In: The 28th International Symposium on Combustion, The Combustion Institute, Edinburgh, 2000 (submitted for presentation). [124] Niioka T, Hasegawa S. Ignition time of locally heated reactive materials. In: Rep. Inst. Fluid Sci., vol. 1, Tohoku University, 1990. [125] Carslaw HS, Jaeger JC. Conduction of Heat in Solids. 2nd ed. Oxford: Clarendon Press, 1959. p. 264. [126] Kaise T, Makino A, Araki N. Effects of Various parameters on the ignition delay time in the SHS process for Ti–C system. In Proceedings of the 34th JSME Conference in Tohoku Section, Sendai, 1999. p. 127–8.
© Copyright 2024 ExpyDoc