United States Environmental Protection Agency Industrial Environmental Research Laboratory Cincinnati OH 45268 Research and Development EPA-600/S2-84-138 Oct. 1984 % f\ \v Project Summary Determination of the Thermal Decomposition Properties of 20 Selected Hazardous Organic Compounds Barry Dellinger, Juan L. Torres, V\ Graham Laboratory-determined thermal dec position profiles and kinetic data for a list of 20 selected hazardous organic compounds are reported. All data were obtained in flowing air at mean jias- phase, high-temperature zone reside nee times ranging from one to six secoi The extrapolated temperatures requ for 99.99% destruction of the pa compound at two seconds mean resi- dence time, T9999(2), ranged fiom 600°C for 1,1,1-trichloroetham to 950°C for acetonitrile. The processes and parameters potentially controlling incineration efficiency are discussed, and four previously proposed methods of ranking compound incinerability are critically reviewed. The possible chemical mechani for destruction of hazardous org compounds are examined and use explain trends in the experimen ayne A. Rubey, Douglas L. Hall, and John L. ed ent ms nic to ally determined thermal decomposition data. It is proposed, through proper application of the principles of organic chemistry, kinetics, and physics that laboratory, gas-phase thermal decompo- sition data generated under controlled conditions can be incorporated into models of full-scale incineration, can serve as a viable ranking of waste incinerability, and can be used to predict the formation of products of incomplete combustion. This Project Summary was developed by EPA's Industrial Environmental Re- search Laboratory, Cincinnati, OH, to announce key findings of the research project that is fully documented in a separate report of the same title (see Project Report ordering information at back). ds. Introduction The ultimate goal of hazardous waste incineration is to destroy the waste with as high a destruction efficiency (DE) as possible. Under the Resource Conserva- tion and Recovery Act (RCRA) of 1976, an incinerator operator must show that the facility can adequately destroy those hazardous waste constituents which are most difficult to incinerate. In theory, the permit writer selects compounds within the mixture that are of sufficient con- centration and thermal stability to be designated as principal organic hazardous constituents (POHCs). It must then be shown, possibly by trial burn, that the designated POHCs can be destroyed or removed by the particular incineration system to a destruction and removal efficiency (ORE) of 99.99 percent. Further- more, the specific operating conditions must be established under which the 99.99 percent ORE is achieved. The development of a ranking of the incinerability for compounds that are candidates for POHC selection is of obvious utility. The U.S. Environmental Protection Agency (EPA) is currently using a ranking based on chemical com- pound heat of combustion per gram. This method has received considerable criti- ------- cism, and the development of an alter- native ranking scheme is of a very high priority to the EPA. Experimentally determined gas-phase thermal stability under controlled labo- ratory conditions has been proposed as an alternative ranking method.This report presents the results of the laboratory determination of the gas-phase thermal decomposition properties of twenty haz- ardous organic compounds. The com- pounds were selected by EPA based on their frequency of occurrence in hazard- ous waste streams, apparent prevalence in the stack effluent, and representative- ness of the spectrum of hazardous organic waste materials. Thermal decomposition profiles were determined at various reactor residence times in an atmosphere of flowing air, which enabled the calculation of global thermal decomposition kinetic param- eters. The results of these laboratory measurements are then compared with previously proposed ranking procedures and an attempt is made to correlate this data with various parameters which may relate to gas-phase thermal stability. The chemical mechanisms potentially respons- ible for thermal decomposition of the twenty test compounds and their poten- tial use in explaining trends in the data and predicting the relative thermal stabil- ity of untested compounds are discussed. Experimental Procedures All of the experimental data pre- sented in this report were generated on the thermal decomposition unit-gas chro- matographic (TDU-GC) system designed and built with funding provided by the EPA (Cooperative Agreement No. 807815- 01 -0). Samples of the twenty compounds were prepared and introduced into the system by several procedures depending upon their physical state and vapor pressure. As seen in Figure 1, the TDU-GC system is a closed, in-line system. In- stream instrumentation measures pres- sure and flow accurately. Most of the instrumentation controls are located in a console from which test functions can be continuously monitored. To initiate a test, the sample is intro- duced into the system and gradually vaporized in a flowing gas stream (i.e., nitrogen, air, or nitrogen/oxygen mix- tures). The vaporized sample passes through a controlled, high-temperature tubular reactor where it undergoes ther- mal decomposition. The products of the thermal decomposition of the compound and the remaining parent compound are swept into a Varian VISTA 4600 high- resolution gas chromatograph for anal- ysis. The sample insertion chamber, the reactor, and the entire transport system are fabricated of fused quartz to minimize interaction with the sample. The rate at which molecules are admit- ted into the high-temperature reactor is important in thermal decomposition stud- ies. In the TDU-GC system, the sample is deposited into a sample insertion cham- ber packed with quartz wool. The chamber is initially kept at or below room temper- ature. The chamber is slowly heated to 250-300°C by applying a linear temper- ature program (10-20 °C/min). The sam- ple molecules are thermally desorbed and gradually swept into the thermal reactor. One of the most important components of the TDU-GC system is the high-tem- perature reactor. The major portions of the reactor consists of a narrow-bore nominal 1 mm ID quartz tube flowpath in a race-track configuration (3.5 cycles, 1 meter in length). The all quartz construc- tion and racetrack configuration mini- mizes the possibility of wall reactions while simultaneously providing a narrow residence time distribution and square wave exposure temperature profile. In addition, the reactor design has fine bore entrance and exit tubes to transport the sample rapidly into and out of the central portion of the reactor. This design also minimizes non-ideal temperature pro- files. The quartz tube reactor assembly fits into a high-temperature three-zone Lindberg furnace designed for continuous operation at temperatures up to 1,200°C with control to ± 1 °C. The effluent from the high-temperature reactor zone is swept through a heated transfer line toward a Varian 4600 GC with a Vista CDS 401 dedicated data terminal. A 30:1 splitter between the furnace and the GC directs only a small portion (—3%) of the effluent sample to the capillary column in order to prevent column overloading. The sample emerging from the reactor is trapped at the head of the chromato- graphic column because the GC oven is maintained at a cryogenictemperature(~ minus 30°C). A fused silica capillary column was used in the majority of the investigation. It was 15 meters in length and contained a dimethylsilicone chemi- cally-bonded stationary phase. A flame- ionization detector was used in this investigation. Throughout the investiga- tion the signal/noise ratio for peak de- tection was four (4). For the ethane analyses, Tedlar bags were used to capture the reactor effluent at the splitter. The captured gas was analyzed using a Varian Aerograph Series 1800 gas chromatograph equipped with one meter x 4.0 mm ID packed column (5 X molecular sieve, 45/60 mesh) operated isothermally at 150°C. This system was utilized to insure resolution Of the three Cz hydrocarbons. Sample handling and preparation was performed in the TDU-GC system's glove- box compartment. Samples of each test compound were prepared according to individual characteristics. Use of solvents, were avoided whenever possible to facil- itate observation of products of incom- plete combustion (PICs). However, for solid samples requiring preparation in solution, Eastman Spectra ACS Grade cyclohexane was selected as the solvent. Cyclohexane was chosen for its low thermal stability and high volatility rela- tive to the specific solutes. In the case of 1,2,3,4-tetrachlorobenzene, the sample was prepared in a methylene chloride solution. For some samples, the probe was removed from the insertion chamber and solutions were injected into the quartz wool portion of the probe. After the solvent was evaporated, the probe with the deposited sample was returned to the insertion chamberwhichwasthen heated at a programmed rate. High vapor pres- sure liquid-phase samples were prepared at concentrations of 1Oppm(v/v) in air for direct gas-phase injection into the inser- tion chamber. Results For each of the twenty test compounds, the fraction of the feed material unde- stroyed at a given set of temperatures and mean reactor residence times (tr) was determined. This resulted in the genera- tion of what may be termed thermal decomposition profiles, i.e., a plot of logarithms of the fraction remaining (fr) vs. the reactor temperature (°C) at con- stant residence time. Thermal decomposi- tion profiles were generated in flowing air at four mean residence times, tr=1.0,2.0, 4.0, and 6.0 seconds. An example of the determination of this family of profiles is given in Figure 2 for chloroform. The thermal decomposition profile for chloro- form is representative of the great major- ity of the compounds tested and serves to illustrate several features. Below400°C, no measurable decompo- sition occurs for chloroform (the flat portion of the curve). Between 400°C and 525°C the rate of decomposition begins to increase. Above 550°C, it increases rapidly, resulting in an apparently linear region of steep slope. The chloroform has ------- Thermal Decomposition Unit Capture of Effluent Products \ Controlled High Temperature Exposure Sample Insertion and Vaporization Pressure and Flow Regulation Compressed Gas and Purification High Temperature Transfer Multifunctional Gas Chromatographic Instrumentation Containment or Destruction of Effluent Products Figure 1. Block diagram of the thermal decomposition unit-gas Chromatographic system. 100 t .£ (0 CD 10 0.1 0.01 Figure 2. CHLOROFORM O tr • 1.0 D i, -Z.O A tr -4.0 O V -6.0 -I 100 400 500 600 Exposure Temperature, °C * 700 Thermal decomposition profiles for chloroform in flowing air at mean residence times of 1.0. 2.0. 4.0, and 6.0 seconds. been destroyed with 99.99% efficiency by 620°C at a residence time of 2.0 seconds. with 99.99% efficiency by 620°C at a residence time of 2.0 seconds. The data for the test compounds have been summarized in Table 1 with entries for the temperature for the onset of decomposition, Tonset (2) (°C), the inter- polated temperature for 99% destruction, T99 (2) (°C), and the extrapolated temper- ature for 99.99% destruction, T9999 (2) (°C). All these values are for tr = 2.0 seconds in flowing air. With only the data presented in this table, the thermal decomposition profile for the compounds may be approximately reconstructed. The table lists the compounds in order of decreasing temperature required for 99% destruction efficiency. A slight reordering occurs if T9999(2) is used for the ranking. However, the numerical differences for the reordered compounds are small. For the conditions possibly encountered during gas-phase thermal decomposition in an incinerator (600°C to 1,400°C and oxygen levels of 0.1 to 21 percent), two possibly global decomposition pathways predominate. The first is pyrolysis, for which the rate of decomposition of the parent species is independent of the oxygen concentration. The second is oxidation, for which the decomposition of the parent is dependent both on the oxygen concentration and susceptibility of the parent species to attack by oxygen or other oxidizing species. The global expressions for these two reaction schemes are where: ki and k2 are the global rate constants for pyrolysis and oxidation, respectively, and a, a, and b, are the reaction order for the decomposition of species A with respect to A and 02. The time dependence is included in this expression. The temperature dependence is included in the rate constants for the two processes. This may be expressed by the Arrhenius equation: k = A exp (-Ea/RT) where: Ea is the activation energy for the process, cal mole"1 A is the Arrhenius coefficient, s-1 R is the universal gas constant, 1.99 cal mole"1 °K"1. For each of the test compounds, when the thermal decomposition reaction oc- curred in an atmosphere with a large ------- Table 1 . Summary of Thermal Decomposition Data Compound Acetonitrile Tetrachloroethylene Acrylonitrile Methane Hexachlorobenzene 1, 2,3,4- Tetrachlorobenzene Pyridine Dichloromethane Carbon Tetrachlonde Hexachlorobutadiene 1,2,4- Trichlorobenzene 1 ,2-Dichlorobenzene Ethane Benzene Aniline Monochlorobenzene Nitrobenzene Hexachloroethane Chloroform 1, 1 , 1 -Trichloroethane Empirical Formula 0^3/V C2C/4 CM CH< CaC/s CeHiCU CM CH2C/2 ecu C4C/6 CeH3C/3 CeHtC/2 C^6 Ce/Ye CeHjN Cer/sCl CeHs/vOz C2C/6 CHC/3 C^C/3 excess of molecular oxygen relative to the concentration of the waste material, the decomposition equation could be simpli- fied to an expression that is first order in tKA f*nnr*antratir\n rtf tho camnln %Afhiis*h 'onset (2) 760 660 650 660 650 660 620 650 600 620 640 630 500 630 620 540 570 470 410 390 Taa(2) 900 850 830 830 820 800 770 770 750 750 750 740 735 730 730 710 670 600 590 570 Taa 99 (2) ~ 950 920 860 870 880 850 840 780 820 780 790 780 785 760 750 780 700 640 620 600 pression for the required temperature for a given level of destruction in an atmos- phere of flowing air: T - nm p ir, / ~l'A \ «.-, 4.7x10? 2.6x1 0s 1.3x106 3.5x10s 2.5x1 0s 1.9x10e I.JxW5 3.0x10'3 2.8 x JO6 6.3xW'2 2.2x1 0s 3.0x10s 1.3x10* 2.8x10» 9.3xW'5 8.0x10" 1.4xW'5 1.9x10? 2.9x1012 1.9x1 0s fa (kcal/mole) 40 33 31 48 41 30 24 64 26 59 39 39 24 38 71 23 64 29 49 32 Calculated Taa(2}(°C) 908 900 910 874 845 834 767 796 824 763 789 766 830 757 726 810 672 64 J 606 60 f line with the slope equal to -Ea/R and an intercept of 1 n A. Regression analyses of this type have been performed on each of the twenty tact r+f\mr\nt mrlc In oil /"«aeao t Ka f i rot resulted in the integrated pseudo-first order rate expression: f, = exp(-k2t,) where: f, is the fraction of the parent species remaining, and kz = ka[02] is the pseudo-first order rate constant This expression may be combined with the Arrhenius Equation to yield an ex- where: TDE is the temperature required for a given DE, °K, Ea has units of kcal mole"1 and the other variables are as previously defined. A plot of In fr vs. t, for the four residence times will yield the rate constant for the reaction at a given temperature. A plot In k vs. 1/T for the four experimental temperatures should then yield a straight order kinetic plots yielded far better fits than zeroth or second order plots. The measured kinetic parameters, A and Ea, along with the calculated temperatures for 99% destruction are included in Table 1. Discussion Several methods have been proposed for ranking the relative incinerability of hazardous organic compounds. Research- ers at the Mitre Corporation, in conjunc- tion with EPA, have proposed a scale ------- based on the heat of combustion per gram molecular weight (Hc/gram of the pure compound). Researchers at IT Enviro- science have proposed a method based on the laboratory determination of auto- ignition temperatures (AIT) of the pure compounds. Researchers at the National Bureau of Standards(NBS) have proposed a purely theoretical approach based on the kinetics of flame-mode thermal de- composition. Researchers at the Univer- sity of Dayton and Union Carbide have proposed scales based on laboratory- determined rates of gas-phase thermal decomposition of pure organic com- pounds in flowing air. Other parameters might be appropriate as a basis for such a scale. The correlation of one scale of incinerability with another would be additional evidence of its validity, al- though correlation with a broad range of incineration data will eventually be re- quired before a proposed ranking will receive universal acceptability. Thus, an attempt was made to correlate laboratory- generated gas-phase thermal decomposi- tion data reported in the previous section with the previously proposed ranking scales. This comparison was complicated by lack of overlap of the compounds investigated. For the heat-of-combustion scale, the only readily discerned agreement was the general increase in thermal stability with decreasing heat of combustion for the chlorinated benzenes. Other than for this group, there was no discernable trend of agreement, either within a class or be- tween classes of compounds. There appeared to be a general positive correlation for both T99 (2) and T99.99 (2) with AIT for those compounds with an AIT below 550°C. Above this temperature, the gas-phase thermal stability appeared to vary little with AIT. In essence there appears to be a positive correlation for the compounds which have a significant fuel value as manifested through their low autoignition temperature. Direct comparison of thermal decompo- sition data with the NBS scale was hampered because only four of the twenty compounds were ranked by the group at NBS. However, the thermal stability of the chlorobenzenes ranked by NBS fell in the same order as suggested by the T99 99 (2) based ranking. Chloroform was pre- dicted to be rather unstable due to the relatively low-energy carbon-chlorine bond. This thermal instability is evident from the gas-phase thermal decompo- sition results, and it is probably even more fragile than predicted by the NBS ranking. Only five of the compounds were studied at Union Carbide. The most significant disagreement was for acrylo- nitrile, the least stable of the five com- pounds based on the Union Carbide calculated T99.99 (2), although the data from this study indicate that its thermal stability rivals that of methane. Ethane is slightly less stable than monochloroben- zene, as measured by Union Carbide; this trend is reversed in the present data, although the difference is small. With the exception of acrylonitrile, the two rank- ings are similar, although the data from this study predict the compounds to be typically more stable by 40°C than that predicted by Union Carbide. No single proposed ranking scheme or molecular parameter was identified that correlated with all our thermal decompo- sition data, although trends were observ- able in homologous subgroups. Using the principles of chemical reactions, it is possible to explain the behavior of each of the twenty compounds on a relative basis and identify mechanisms that might be used to extrapolate this limited data to other compounds not studied in the laboratory. The twenty test compounds may be divided into five subclasses which are discussed in the following paragraphs. Methane, Dichloromethane, Chloroform, and Carbon Tetrachloride The observed trend in this group is decreasing thermal stability with increas- ing chlorine substitution, except for car- bon tetrachloride, which is intermediate in thermal stability between methane and dichloromethane. The carbon-hydrogen bond in methane has a bond dissociation energy (BDE) of 104 kcal/mole. The carbon-chlorine bonds in the other three compounds are 79,77, and 70 kcal/mole, respectively, and the carbon-chlorine bonds weaken with increased chlorine substitution. This trend disagrees with the experimental observations. The data agree with a mechanism based on abstraction of a hydrogen, probably by OH. Since the carbon-hydro- gen BDE decreases with increasing chlo- rine substitution up to chloroform, one would predict decreasing thermal stabil- ity, which is in fact observed. However, carbon tetrachloride contains no hydro- gens and thus would not be susceptible to this mode of attack. It would instead be expected to decompose by bond rupture. The implication of this data set is that H abstraction reaction rates may be faster at these temperatures than previously expected. Benzene, Monochlorobenzene, 1,2-Dichlorobenzene, 1,2,4-Trichlorobenzene, 1,2,3,4-Tetrachlorobenzene, Hexachlorobenzene, Pyridine, An/line, Nitrobenzene The observed trend is toward increasing thermal stability with increasing chlorine substitution. Pyridine is more stable, nitrobenzene less stable, and aniline is about as stable as benzene. All of the bonds in benzene, the chlorinated ben- zenes, and pyridine are probably in excess of 90 kcal/mole, and one would expect electrophilic addition to be the predomi- nant reaction path. The chlorines and the nitrogen in pyridine are more electro- negative than hydrogen or carbon, which leads to a destabilization of the electron- deficient intermediate resulting from OH addition; thus, the rate of decomposition is reduced and a greater thermal stability results than for benzene. The stabiliy of aniline and nitrobenzene relative to benzene can also be explained by electrophilic attack by a radical such as OH. However, the effect of resonance interaction is somewhat different than in normal electrophilic attack by a cation. The radical intermediate formed by nitro- benzene is actually stabilized by reso- nance. This is opposed to the norm for cationic electrophilic addition, where nitro substitution results in destabiliza- tion because a positive charge is placed on the electronegative oxygen. This sta- bilization of the intermediate again leads to a less stable molecule relative to benzene. Also in nitrobenzene, the nitro- gen carbon BDE is 70 kcal/mole which may be easily broken and does represent an altenative mode of decomposition. The radical intermediate formed by OH attack on aniline would not receive signif- icant resonance stabilization due to lack of an octet on the normally very stable resonance structure involving the lone pair on the nitrogen. On this basis, one would expect aniline to be about as stable as benzene, which is observed. The nitrogen-hydrogen BDE is only 80 kcal/ mole and may have some role in the decomposition; however, the similarity in stability, as opposed to that of benzene, indicates that electrophilic addition is the predominant mode of destruction. Ethane, 1,1,1- Trichloroethane, Hexachloroethane One might expect ethane to be de- stroyed by unimolecular decomposition through rupture at the weakest bond. The ------- carbon-carbon BDE in ethane is 88 kcal/mole compared to the carbon-hydro- gen BDE of 104 kcal/mole in methane. Thus, one would predict significantly lower stability than for methane but still moderate thermal stability for ethane. Although these compounds are similar in structure, one would expect the path- way of decomposition for 1,1,1 -trichloro- ethane to be by concerted elimination of HCI, which is a very low-energy process. Therefore, 1,1,1 -trichloroethaneisoneof the least stable compounds studied. Hexachloroethane would have to elimi- nate CU to proceed by a concerted path- way. This process is more endothermic than the elimination of HCI and the decomposition of hexachloroethane would instead be expected to proceed through carbon-chlorine or carbon-hydrogen bond rupture, both BDEs being approximately 73 kcal/mole. Based on these considera- tions, hexachloroethane is observed to be intermediate in stability between ethane and 1,1,1-trichloroethane. Tetrachloroethylene and Hexachlorobutadiene Both of these compounds are quite stable. This is probably due to the large BDEs caused by sp2 hybridization of all carbon atoms and the lack of hydrogen atoms available for abstraction by OH or formation of OH. Carbon-carbon BDEs would be expected to decrease in the order ethylene > hexachlorobutadiene > ethane. This may be used to explain the relative stability of the series. Electrophilic addition of OH, however, may be the predominant mode of attack in an incinerator. The prediction is the same, since butadiene can form an allyl radical intermediate known to be quite stable. Acetonitrile and Acrylonitrile These compounds are very stable, and all BDEs are 93 kcal/mole or greater. Based on the mechanism of bond rupture, one might predict acetonitrile to be less stable than acrylonitrile for the carbon- hydrogen bonds would certainly be stronger in acrylonitrile. This reasoning would also apply to hydrogen abstraction reactions. Carbon-nitrogen triple bonds are ex- pected to be much less reactive toward electrophilic addition than double bonds. Apparently, the only mode of decomposi- tion for acetonitrile is loss of a hydrogen through bond rupture or abstraction. These are both high-energy processes which account for the stability of aceto- nitrile. Acrylonitrile may, on the other hand, be susceptible to addition at the carbon-carbon double bond. The stability of the resulting intermediate may be expected to be somewhat greater than in tetrachloroethylene through resonance stabilization. This would account for the fact that acrylonitrile is less stable than either tetrachloroethylene or acetonitrile. Conclusions When this program was initiated in April 1982, very limited information ex- isted on thermal decomposition properties of organic compounds commonly subject- ed to incineration. The data generated in this program provide a consistent initial data base for the development of a concept of incinerability. Preliminary calculations indicate that "fault" or "failure" modes of incinerator operation or, equivalently, the "extremes" of operational parameter distribution functions, may well control measured incineration efficiency for full-scale units. Following this line of reasoning, one would conclude that the incinerator ef- fluent would only contain undecomposed feed material and products of incomplete combustion which were formed under these conditions of failure. Thus, "fault" modes are essentially worst-case condi- tions and appear to have a dominating effect on the composition of the inciner- ator effluent. This suggests that the majority of future studies should identify and address these failure conditions. Although these conditions are not pres- ently well defined, an atmosphere con- taining 1% oxygen and a residence time of 0.25 seconds might be considered representative. Furthermore, the temperature at which this study was conducted is probably representative of "fault" modes. The experimental laboratory temperature range covered 0 to 99.9% destruction of the feed material, which is typically several hundred degrees below mean temperatures quoted for hazardous waste incineration. If a given incinerator does not meet the 99.99% destruction effi- ciency requirement, yet has a high mean operating temperature, then a likely pos- sibility for its failure is that a fraction of the waste feed experiences temperatures somewhat lower than the mean (where destruction efficiency is low) i.e., the destruction efficiency and temperature range measured in laboratory studies. Thus, one might expect the actual PICs emitted from the incinerator to be the same as those formed under the condi- tions studied in the laboratory. Further- more, this reasoning suggests that the relative thermal stability of hazardous wastes should be compared at "fault" mode temperatures since only this frac- tion of the waste is escaping incineration. Because selecting a suitable temperature for comparison of every compound is difficult and still somewhat arbitrary, a ranking based on the temperature re- quired for 99% destruction at 2 seconds mean residence time is proposed. One could just as well select 90% or 99.9%, but examination of the data shows that the rankings over this range are essen- tially identical. Products of incomplete combustion have not been considered in the full report. PIC determination would remove much of the speculation in the discussion of reaction mechanisms. In previous research, the formation of numerous PICs from a wide variety of organic compounds has been observed. These PICs have, in some cases, been produced in as much as 50% yields and have been as hazardous as or more hazardous than the parent com- pound. The determination of PICs should be an integral part of future research. The reported research has addressed non-flame, high-temperature, gas-phase reaction chemistry. Extension of this work to include so-called flame mode studies represents a special challenge due to the difficulties in scaling results. Of all incineration processes which must be modeled to full scale, gas-phase chemical kinetics is the easiest and most successfully performed. The temporal and spatial distributions present in small laboratory- or bench-scale flames are not easily scaled to the turbulent, poorly- defined flames present in full-scale sys- tems. Consequently, an elementary chem- ical kinetic approach to determining flame mode destruction efficiencies might prove most effective. The importance of the hydroxyl radical in flames is well documented. Thus, an experimental program to determine the rate of attack of hydroxyl radicals on hazardous wastes would produce easily scaled kinetic results. Kinetic data of this type, in combination with measurements or estimates of hydroxyl radical concen- trations in full-scale systems would allow simple scaling of laboratory results to full-scale. This data over different tem- perature ranges would be applicable to both flame and non-flame modes of destruction. ------- Recommendations • Studies including more thorough ki- netic investigations and identification of products of incomplete combustion should be conducted on a limited number of selected compounds to identify the dominant mechanisms of destruction of hazardous organic com- pounds as well as the formation of toxic products of incomplete combus- tion. • Further surveys of thermal stability of hazardous organic wastes should be conducted under "fault" modes of an incinerator, as calculations indicate that these modes control the incinera- tion efficiency. A representative "fault" mode might be 1 % oxygen (or less for destruction systems other than incin- erators) and a residence time of 0.25 seconds. • Determining the relative importance of various "fault" modes such as reduced gas-phase residence time, low levels of oxygen, and low exposure temperature should be emphasized. • Studies should be performed to ad- dress the effect of changing the compo- sition of the organic fraction of the reaction atmosphere, since this has the potential of modifying the reaction pathway, thereby affecting thermal stability and product formation. • A laboratory study to determine the rate of OH attack on hazardous organic compounds at incineration tempera- tures should be undertaken in light of the importance of this reaction inferred from this study. • A round-robin test program should be conducted using a well-defined waste sample. The waste should be evaluated by each proposed incinerability ranking scheme and predictions made concern- ing the organic composition (both POHCs and PICs) of the stack effluent. The results of a trial burn of this waste at the EPA's Combustion Research Facility could be used as the basis for evaluating the test program results. Barry Dellinger, Juan L. Torres, Wayne A. Rubey, Douglas L Hall, and John L. Graham are with University of Dayton Research Institute, 300 College Park, Dayton, OH 45469. Richard A. Carnes is the EPA Project Officer (see below). The complete report, entitled "Determination of the Thermal Decomposition Properties of 20 Selected Hazardous Organic Compounds," (Order No. PB 84-232 487; Cost: $19.00, subject to change) will be available only from: National Technical Information Service 5285 Port Royal Road Springfield, VA 22161 Telephone: 703-487-4650 The EPA Project Officer can be contacted at: Industrial Environmental Research Laboratory U.S. Environmental Protection Agency Cincinnati, OH 45268 •&U. S. GOVERNMENT PRINTING OFFICE: 1984/559-111/10718 ------- United States Environmental Protection Agency Center for Environmental Research Information Cincinnati OH 45268 BULK RATE POSTAGE & FEES PAI EPA PERMIT No. G-35 Official Business Penalty for Private Use $300 ps 990052? 50 S DEARBORN STREET HICAGO IL &06oa ------- |