Real-Time Thermal Imaging of Fast Exothermic Reactions Involving

Jul 23, 2012 - ABSTRACT: The vigorous, exothermic decomposition of methyl ethyl ketone peroxide (MEKPO) can be stimulated by acids, metals, or heat, a...
0 downloads 4 Views 3MB Size
Article pubs.acs.org/IECR

Real-Time Thermal Imaging of Fast Exothermic Reactions Involving the Hazardous Combination of Methyl Ethyl Ketone Peroxide and Inorganic Acids Yan-Fu Lin,* Yi-Huan Wu, and Syu-Ming Lai Department of Chemistry, National Chung Hsing University, 250, Kuo-Kuang Road, Taichung 40227, Taiwan, Republic of China S Supporting Information *

ABSTRACT: The vigorous, exothermic decomposition of methyl ethyl ketone peroxide (MEKPO) can be stimulated by acids, metals, or heat, and the initial stage is critical to runaway reactions. Real-time thermal data are important in the evaluation of the hazards of fast reactions during the initial stage. We report the first real-time observation of the reaction of MEKPO with inorganic acids using a thermal imaging camera (TIC), which we used to determine the rate of heat evolution and to recognize the degree of the hazard present during accidents. In addition, ab initio computational methods were used to estimate the thermokinetic parameters for the proposed reaction mechanisms. The results indicated that the temperature of the solution, the concentration of the acid, and the dehydration ability were the primary factors that affected the thermal reaction. This work provides a new method to study thermal chemistry and to evaluate thermal hazards.



INTRODUCTION Investigations of the stability of thermal hazards are important because runaway reactions that result from the combination of incompatible molecules in a laboratory often result in fires and potentially in explosions.1−5 Runaway reactions have resulted in numerous casualties due to the unexpected behavior of chemical mixtures. 6,7 Methyl ethyl ketone peroxide (MEKPO) is a hazardous material because of its unstable structure, and it has been implicated in many accidents around the world.8−11 MEKPO has caused many thermal explosions in Taiwan, Japan, Korea, and China, which makes it one of the most hazardous materials in Asia (Table 1).12,13 These

necessary. Many researchers have employed calorimetric techniques and instruments, such as differential scanning calorimetry (DSC), thermal activity monitor III (TAM III), and vent sizing package 2 (VSP2), to evaluate thermokinetic parameters, including the activation energy, the frequency factor, the heat of reaction, and the reaction order.15−21 These instruments cannot record real-time thermal data during the initial stages of the reaction because of the time interval between the mixing of reactants and the measurement by the calorimeter. Therefore, direct and real-time detection is important for monitoring the reactions between incompatible molecules, especially for highly sensitive peroxides. The work presented here represents the first analysis of the reaction between MEKPO and inorganic acids using a thermal imaging camera (TIC) and ab initio computational methods. The TIC used in this study can record a maximum of 60 frames/s, which has allowed us to probe the critical stages of the reaction within the first 300 s. In the present work, we attempted to observe the real-time mixing and reaction behavior when MEKPO (31% by mass) and an inorganic acid were mixed. We demonstrated that higher concentrations of sulfuric acid (H2SO4), nitric acid (HNO3), and hydrochloric acid (HCl) could induce the decomposition of MEKPO and thereby release a substantial amount of heat. When the concentration of H2SO4 reached 98 mass %, H2SO4 caused a more vigorous exothermic reaction than did the other acids. The ab initio computational analysis of the reaction mechanism between MEKPO and the acid revealed that the dehydrating agents serve to accelerate the decomposition rate of the protonated MEKPO intermediates. The affinity of sulfuric acid for water is high, and this acid will

Table 1. Selected Thermal Explosion Incidents Caused by MEKPO in Asia12,13 date

location

no. of injuries/deaths

hazard

1964 1984 1996 2000 2001 2003

Tokyo (Japan) Taoyuan (Taiwan) Taoyuan (Taiwan) Yosu (Korea) Jiangsu (China) Zhejiang (China)

114/19 55/5 47/10 11/3 2/4 3/5

explosion (storage) explosion (reactor) explosion (tank) explosion explosion explosion

accidents were usually caused by the inappropriate use of MEKPO. For example, the heat of dilution that was generated by the mixing of acid wastes resulted in a MEKPO explosion and fire in Yoshitomi, Fukuoka, Japan, in 1998.14 Currently, MEKPO is manufactured via the slow addition of hydrogen peroxide to methyl ether ketone (MEK) under acid-catalyzed conditions, which results in a series of reactions that increase the possibility of accidents.8 Because inorganic acids are extensively used in common industrial plants, MEKPO could be accidentally mixed during storage or transportation. Therefore, a detailed investigation of the runaway reaction that occurs when MEKPO is mixed with inorganic acids is © 2012 American Chemical Society

Received: Revised: Accepted: Published: 10778

May 15, 2012 July 16, 2012 July 23, 2012 July 23, 2012 dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research

Article

Figure 1. Instrumental setup used for TIC experiments.

with incompatible materials both at room temperature and in the presence of mild external heating. To further identify possible reasons for the different effects of different inorganic acids, we used ab initio computational methods to estimate the bond dissociation energy (BDE), the activation energy (Ea), and the heat of reaction (ΔH) for each of the proposed reaction mechanisms. The calculation results indicated that the dehydrating agents accelerated the decomposition reaction. Heat of Solution of MEKPO plus Acidic Aqueous Solutions. The heat of solution of MEKPO plus water was evaluated first, and no apparent heat generation was observed because MEKPO and water are immiscible. However, after we replaced pure water with an acidic aqueous solution, the mixture of MEKPO and the acidic aqueous solution generated the heat of solution, as shown in Figure 2. The primary reason for this behavior is that hydrogen bonding formed between MEKPO and the hydronium ion, which increased the solubility of MEKPO in the acidic aqueous solution and thereby produced an endothermic heat of solution. When the concentration of the acids was increased, the reaction from the addition of acidic aqueous solution to MEKPO absorbed

remove H2O from the protonated MEKPO intermediates and cause a runaway reaction. The TIC and ab initio computational methods can be used to understand runaway reactions and to generate information that can be used to prevent future accidents from occurring in high-risk operations, such as the storage, transportation, or manufacture of reactive compounds. To reduce the hazards associated with handling reactive chemicals, the results of this study will be provided to relevant manufacturing plants to help them better understand the hazards associated with MEKPO.



EXPERIMENTAL SECTION Materials and Methods. MEKPO (31% by mass) was purchased directly from Aldrich and stored in a refrigerator at 4 °C. MEKPO was placed in a fume hood for at least 3 h before the experiments to allow it to reach room temperature. Solutions with different concentrations of H2SO4, HNO3, and HCl were prepared at room temperature using distilled water. Solutions of D2SO4/D2O were prepared using commercially available 96−98 mass % D2SO4 (Cambridge Isotope Laboratories) mixed with deuterium oxide, D2O (99.9 atom % D, Aldrich). Thermal Imaging of MEKPO Reactions. A TIC was used to evaluate the evolution of heat during reactions of MEKPO with incompatible materials both at room temperature and in the presence of mild external heating. Briefly, MEKPO and the acid solutions were mixed under ambient conditions; in some tests, the temperature of the mixtures was gradually increased using an external heating source after the solutions were mixed. The experimental data were collected using a TIC (TVS500EX, NEC) equipped with a field of vision (FOV) of 19.4° (horizontal) × 14.6° (vertical) with a standard 22 mm lens that can capture temperature changes every 1/15 s with a temperature resolution of less than 0.05 °C. The experimental setup is shown in Figure 1. Ab Initio Computational Methods. The Gaussian 09 package was used throughout this work.22 Second-order Møller−Plesset perturbation theory calculations using the 6311++G(d,p) basis set were performed for all of the geometry optimizations.23−25 All of the calculated equilibrium structures (local minima and transition states) were characterized using harmonic vibration frequency calculations at the same levels. The intrinsic reaction coordinate (IRC) calculations were performed with the same levels when the reaction path of the desired decompositions was followed.26−28



RESULTS AND DISCUSSION The sophisticated TIC allowed us to acquire crucial thermal data during our experiments. We attempted to observe the evolution of the system in real time when MEKPO was mixed

Figure 2. Temperature differences during thermal imaging of solutions of MEKPO with different concentrations of H2SO4, HCl, and HNO3. 10779

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research

Article

more thermal energy. Different reaction rates were observed in the experiments where the D2SO4/D2O solution was added to MEKPO and where the H2SO4/H2O solution was added to MEKPO because the hydrogen bond formed with deuterium is stronger than that formed with ordinary hydrogen,29,30 as shown in Figure 3. According to these experimental results, the

Figure 3. Comparison of addition of D2SO4/D2O solution to MEKPO with addition of H2SO4/H2O solution to MEKPO.

proton played a key role in the reaction rate, which was demonstrated in the reactions when the incompatible MEKPO was mixed with inorganic acids. Influence of Different Inorganic Acids. After the concentrated solutions of H2SO4, HCl, and HNO3 were individually mixed with MEKPO at room temperature, the thermal data were recorded using a TIC (Figure 4). The acids

Figure 5. Rates of temperature increase for solutions of MEKPO and different concentrations of H2SO4, HCl, and HNO3. aFrame A: TIC frame at the start of the experiment. bFrame B: TIC frame when the temperature experienced the fastest change. cFrame C: composite TIC frame created by subtracting frame A from frame B.

Mixing MEKPO with sulfuric acid is very dangerous, and if it occurs, a fire will inevitably result. Effects of Acid Concentration. Different concentrations of H2SO4, HCl, and HNO3 were mixed with MEKPO at room temperature, and the thermal data were recorded using a TIC, as shown in Figures 6−8. Higher concentrations of the inorganic acids induced a larger release of heat during the decomposition of MEKPO. The rate of temperature increase at the beginning of the reaction was determined when MEKPO was combined with 6, 12, 18, 24, and 36 N solutions of H2SO4. The rates of temperature change were approximately −0.07, −0.03, 0.03, 5.66, and 30.89 °C/s, respectively, as listed in Figure 5. The rate of the increase in temperature when sulfuric acid was added to MEKPO increased as the concentration of sulfuric acid was increased. When the H2SO4 concentration was

Figure 4. Comparison of solutions that contain different concentrated inorganic acids mixed with MEKPO.

caused the heterolytic decomposition of the peroxide molecule.31 Sulfuric acid was more dangerous than the other inorganic acids. The combination of 36 N H2SO4 with MEKPO resulted in a substantial release of heat at a rate of 30.89 °C/s, as shown in Figure 5. The primary reason for this heat release is that H2SO4 is a good dehydrating agent and will accelerate the heterolytic decomposition, which triggers a runaway reaction. 10780

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research

Article

less than 12 N, the reactions did not significantly affect the mixture temperature and would not result in a hazardous situation under the tested conditions. In general, more dilute acid solutions had smaller effects on the reaction rate. Based on Le Chatelier’s principle, water plays a crucial role by inhibiting the heterolytic decomposition reaction and reducing the effect of the acid on the reaction rate. When MEKPO was mixed with more dilute acid solutions, the temperature of the mixture decreased, which reduced the probability of a runaway reaction. The 6 N H2SO4, 12 N H2SO4, 4 N HCl, 5 N HNO3, and 10 N HNO3 acid solutions resulted in temperature change rates of −0.07, −0.03, −0.23, −0.27, and −0.09 °C/s, respectively. This type of thermal data is difficult to collect unless a TIC is used to record the temperature data in real time with a fast frame rate. Influence of Increasing Temperature. Runaway reactions can be induced either by hot spots or by insufficient heat removal. We observed that the temperature of the mixture decreased during the initial stages of the reaction when MEKPO was mixed with some of the lower-concentration acid solutions at room temperature. This phenomenon was detected for the first time with the aid of a TIC. We also examined situations where the temperature of the mixture gradually increased through the application of an external thermal source. In these cases, the MEKPO and acid mixtures quickly released heat and a runaway reaction occurred, as shown in Figure 9.

Figure 6. Temperature profiles for solutions that contain different concentrations of H2SO4 and MEKPO.

Figure 7. Temperature profiles for solutions that contain different concentrations of HCl and MEKPO.

Figure 9. Temperature profiles for solutions that contain different concentrations of H2SO4, HCl, and HNO3 mixed with MEKPO and for a solution of MEKPO without acids exposed to an external thermal source.

The results indicate that accumulated heat will cause thermal hazards; therefore, sufficient heat removal plays an important role in reducing the hazards associated with the mixing of incompatible materials. Determining the Thermokinetic Data from ab Initio Computational Methods. Although seven different types of MEKPO exist, only the monomer and dimer forms are prevalent in industry,32 and both the monomer and dimer of MEKPO generate monomer radicals during the initial stage of decomposition.11 Therefore, the MEKPO monomer was used as a target molecule for computational simulations. The vigorous, exothermic decomposition of MEKPO can be initiated by the acids, which results in a heterolytic or ionic decomposition of the peroxide molecule. This decomposition accompanies the homolytic cleavage of MEKPO to form free

Figure 8. Temperature profiles for solutions that contain different concentrations of HNO3 and MEKPO.

10781

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research

Article

Figure 10. Proposed mechanisms for the homolytic cleavage of the O−O bond of MEKPO (path 1) and the heterolytic decomposition of protonated MEKPO (path 2).

Deduction of Reaction Evolution. On the basis of the combination of thermal imaging and ab initio computational methods, we determined that the evolution of the fast reaction of MEKPO with the acid during the initial stage proceeds through the following three steps. The first step is the dissolution of MEKPO in the acidic aqueous solution:

radicals, which initiates additional reactions and causes a runaway situation. The proposed mechanisms are presented in Figure 10. Ab initio computational methods were used to determine the O−O BDE for the homolytic cleavage of MEKPO and the Ea and ΔH for the heterolytic decomposition of the protonated MEKPO.33 The results are shown in Figure 11. The geometric structures of all stationary points in the XYZ

MEKPO + H 2O → MEKPO(aq) (endothermic step)

(1)

The second step is the protonation of MEKPO: ROOH + H+ → ROOH2+

(exothermic step)

(2)

The third step is the heterolytic decomposition of the protonated MEKPO: ROOH 2+ → (RO)+ + H 2O

(3)

The first step (eq 1) is an endothermic reaction. When acidic solutions are added to MEKPO, the temperature will decrease, except when highly concentrated acids are used, as shown in Figures 6−9. The second step (eq 2) is an exothermic process, and more heat will be produced as the concentration of the acid is increased. The generated heat compensates for the heat absorbed during the first step. The temperature reduction caused by the mixture of MEKPO with the acid decreases as the concentration of the acid is increased. When MEKPO is mixed with a highly concentrated acid, the temperature drop is no longer observed, as shown in Figure 4. The heat emission rate is primarily associated with the third step (eq 3). For example, H2SO4 is a strong dehydrating agent that can accelerate the dehydration reaction of the third step. Therefore, this acid is capable of inducing a runaway reaction, which results in the rapid emission of heat. Among the investigated mixtures, the mixture of MEKPO and 36 N H2SO4 ignites. The reaction between MEKPO and HNO3 is mild because HNO3 does not completely dissociate: even at a concentration of 6 N HNO3, 20% of the nitric acid remains in the undissociated form.34 Under these conditions, the second step does not easily compensate for the first step; therefore, the second step is the critical step. As the concentration of the acid is increased, more energy will be generated to compensate for the first step. This result initiates the third step and releases a significant amount of heat.

Figure 11. Energy profiles for the homolytic cleavage of MEKPO (left) and the heterolytic decomposition of protonated MEKPO (right).

coordinate plane are presented in the Supporting Information. The O−O BDE for the homolytic cleavage of MEKPO was calculated to be 41.62 kcal/mol. The Ea for the heterolytic decomposition of the protonated MEKPO was calculated to be 5.62 kcal/mol, and the ΔH was calculated to be −89.75 kcal/ mol. Because the Ea is lower than the O−O BDE, the heterolytic decomposition pathway is the dominant process during the initial stage of the reaction when MEKPO is mixed with acids. Because the ΔH for the heterolytic decomposition of protonated MEKPO is so large, the heat will cause the chain reaction to release more thermal energy and induce a runaway reaction. 10782

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research



Our approach using a TIC for thermal hazard studies provides new insights through the observation of the initial stage of the reaction of MEKPO and inorganic acids, which could not be monitored via the traditional calorimetric techniques used in previous reports. A decrease in the exothermic onset temperature (To) and an increase in ΔH were simultaneously observed during the decomposition of MEKPO with inorganic acid catalysts. In addition, HNO3 and H2SO4 produced a more serious reaction with MEKPO than did HCl.13 The majority of our results are consistent with previous findings, except for the MEKPO reaction with higher concentrations of acid catalysts, such as 8 N HCl and 12 N HCl, which resulted in a rapid increase in temperature that was observed using the TIC (Figures 4 and 7). We also observed that more heat was released when MEKPO was mixed with 16 N HNO3 (Figure 9), which may be due to the existence of HNO3 in its undissociated form at concentrations greater than 6 N; therefore, the hydroxyl radical from MEKPO can react with HNO3 and generate the NO3 radical. This NO3 radical undergoes a subsequent chain reaction, which results in additional heat release.11 Therefore, high concentrations of inorganic acids should be kept separate from MEKPO during transport and storage. In addition, efficient removal of the heat is important when MEKPO is mixed with diluted inorganic acids because the temperature decreases the most during the initial stage of the reaction of MEKPO with diluted inorganic acids. Careful maintenance of a low temperature can prevent a runaway reaction during the subsequent stages of the reaction.

Article

ASSOCIATED CONTENT

S Supporting Information *

The structures of all stationary points are provided in XYZ coordinates. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Tel.: +886-4-22840411, ext 422. Fax: +886-4-22862547. Email: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are indebted to Prof. Jyi-Sy Yang at the Department of Chemistry, National Chung Hsing University, for providing the TIC and technical support during this work and to Mr. ChenChang Wu from the Department of Chemistry, National Chung Hsing University, for his friendly assistance. We would also like to thank the National Center for High-Performance Computing for computer time and facilities.



REFERENCES

(1) Wang, Y. W.; Shu, C. M. Calorimetric Thermal Hazards of tertButyl Hydroperoxide Solutions. Ind. Eng. Chem. Res. 2010, 49, 8959− 8968. (2) Lian, P.; Ng, D.; Mejia, A. F.; Cheng, Z.; Mannan, M. S. Study on Flame Characteristics in Aerosols by Industrial Heat Transfer Fluids. Ind. Eng. Chem. Res. 2011, 50, 7644−7652. (3) Yun, G.; Ng, D.; Mannan, M. S. Key Observations of Liquefied Natural Gas Vapor Dispersion Field Test with Expansion Foam Application. Ind. Eng. Chem. Res. 2011, 50, 1504−1514. (4) Liaw, H. J.; Chen, C. C.; Chang, C. H.; Lin, N. K.; Shu, C. M. Model To Estimate the Flammability Limits of Fuel−Air−Diluent Mixtures Tested in a Constant Pressure Vessel. Ind. Eng. Chem. Res. 2012, 51, 2747−2761. (5) Tseng, J. M.; Lin, Y. F. Evaluation of a tert-Butyl Peroxybenzoate Runaway Reaction by Five Kinetic Models. Ind. Eng. Chem. Res. 2011, 50, 4783−4787. (6) Lu, Y.; Ng, D.; Mannan, M. S. Prediction of the Reactivity Hazards for Organic Peroxides using the QSPR Approach. Ind. Eng. Chem. Res. 2011, 50, 1515−1522. (7) Kossoy, A. A.; Hofelich, T. Methodology and Software for Assessing Reactivity Ratings of Chemical Systems. Process Saf. Prog. 2003, 22, 235−240. (8) Zhang, J.; Wu, W.; Qian, G.; Zhou, X. G. Continuous Synthesis of Methyl Ethyl Ketone Peroxide in A Microreaction System with Concentrated Hydrogen Peroxide. J. Hazard. Mater. 2010, 181, 1024− 1030. (9) Lin, Y. F.; Tseng, J. M.; Wu, T. C.; Shu, C. M. Effects of Acetone on Methyl Ethyl Ketone Peroxide Runaway Reaction. J. Hazard. Mater. 2008, 153, 1071−1077. (10) Chen, J. R.; Tseng, J. M.; Lin, Y. F.; Wang, S. Y. S.; Shu, C. M. Adiabatic Runaway Studies for Methyl Ethyl Ketone Peroxide with Inorganic Acids by Vent Sizing Package 2. Korean J. Chem. Eng. 2008, 25, 419−422. (11) Tseng, J. M.; Shu, C. M.; Gupta, J. P.; Lin, Y. F. Evaluation and Modeling Runaway Reaction of Methyl Ethyl Ketone Peroxide Mixed with Nitric Acid. Ind. Eng. Chem. Res. 2007, 46, 8738−8745. (12) Chang, R. H.; Tseng, J. M.; Jehng, J. M.; Shu, C. M.; Hou, H. Y. Thermokinetic Model Simulations for Methyl Ethyl Ketone Peroxide Contaminated with H2SO4 or NaOH by DSC and VSP2. J. Therm. Anal. Calorim. 2006, 83, 57−62. (13) Tseng, J. M.; Chang, R. H.; Horng, J. J.; Chang, M. K.; Shu, C. M. Thermal Hazard Evaluation for Methyl Ethyl Ketone Peroxide



CONCLUSIONS The combination of real-time thermal imaging experiments and ab initio computational methods was used for the first time to estimate the thermokinetic parameters of exothermic reactions. These techniques provide a new, quick, and green method to evaluate thermal hazards and explore the evolution of reactions, especially those of highly sensitive peroxides. MEKPO plays an important role in industrial processes, and it is involved in shelter production, polymer production, and many other manufacturing processes. Accidents caused by MEKPO typically occur when it comes into contact with incompatible materials, such as inorganic acids. This study is the first to examine runaway reactions that involve mixtures of MEKPO and inorganic acids in an open system using a TIC. The TIC is a useful tool because it can precisely collect temperature data during the initial stage of the reaction because of its fast frame rate. The results presented in this study could provide a more realistic evaluation of the hazards associated with MEKPO systems, such as the temperature change, the rate of the temperature increase, and the evolution of the reaction. The ab initio computational methods are able to facilitate use of the TIC by providing thermal data for the thermal hazard study. Indeed, our green method using ab initio computational methods to obtain the thermokinetic parameters of MEKPO mixed with inorganic acids obeyed the 12 principles of green chemistry (originally defined by Anastas and Warner)35 related to the elimination of product waste (point 1). The results from the TIC observation provided useful information about the hazards associated with handling MEKPO with inorganic acids for accident prevention (point 12). Therefore, this work demonstrates an economical and green method for the study of thermal hazards. 10783

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784

Industrial & Engineering Chemistry Research

Article

Mixed with Inorganic acids. J. Therm. Anal. Calorim. 2006, 85, 189− 194. (14) Arai, M.; Tamura, M. Explosion and fire caused due to mixing of waste acids of different concentrations in a waste acid tank. JST Failure Knowledge Database, 1998. http://www.sozogaku.com/fkd/en/cfen/ CC1200070.html. (15) Kossoy, A. A.; Koludarova, E. Specific Features of Kinetics Evaluation in Calorimetric Studies of Runaway Reactions. J. Loss Prev. Process Ind. 1995, 8, 229−235. (16) Iizuka, Y.; Surianarayanan, M. Comprehensive Kinetic Model for Adiabatic Decomposition of Di-tert-butyl Peroxide Using BatchCAD. Ind. Eng. Chem. Res. 2003, 42, 2987−2995. (17) Li, X.; Koseki, H.; Iwata, Y.; Mok, Y. S. Decomposition of Methyl Ethyl Ketone Peroxide and Mixtures with Sulfuric Acid. J. Loss Prev. Process Ind. 2004, 17, 23−28. (18) Tseng, J. M.; Chang, Y. Y.; Su, T. S.; Shu, C. M. Study of Thermal Decomposition of Methyl Ethyl Ketone Peroxide Using DSC and Simulation. J. Hazard. Mater. 2007, 142, 765−770. (19) Liang, Y. C.; Jhu, C. Y.; Wu, S. H.; Shen, S. J.; Shu, C. M. Evaluation of Adiabatic Runaway Reaction of Methyl Ethyl Ketone Peroxide by DSC and VSP2. J. Therm. Anal. Calorim. 2011, 106, 173− 177. (20) Tseng, J. M.; Liu, M. Y.; Chen, S. L.; Hwang, W. T.; Gupta, J. P.; Shu, C. M. Runaway Effects of Nitric Acid on Methyl Ethyl Ketone Peroxide by TAM III Tests. J. Therm. Anal. Calorim. 2009, 96, 789− 793. (21) Graham, S. R.; Hodgson, R.; Vechot, L.; Essa, M. I. Calorimetric Studies on the Thermal Stability of Methyl Ethyl Ketone Peroxide (MEKP) Formulations. Process Saf. Environ. Prot. 2011, 89, 424−433. (22) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision A.02; Gaussian, Inc.: Wallingford, CT, 2009. (23) Møller, C.; Plesset, M. S. Note on an Approximation Treatment for Many-electron Systems. Phys. Rev. 1934, 46, 618−622. (24) Head-Gordon, M.; Pople, J. A.; Frisch, M. J. MP2 Energy Evaluation by Direct Methods. Chem. Phys. Lett. 1988, 153, 503−506. (25) Frisch, M. J.; Head-Gordon, M.; Pople, J. A. Direct MP2 Gradient Method. Chem. Phys. Lett. 1990, 166, 275−280. (26) Hratchian, H. P.; Schlegel, H. B. Accurate Reaction Paths Using a Hessian Based Predictor-Corrector Integrator. J. Chem. Phys. 2004, 120, 9918−9924. (27) Hratchian, H. P.; Schlegel, H. B. Finding minima, transition states, and following reaction pathways on ab initio potential energy surfaces. Theory and Applications of Computational Chemistry: The First 40 Years; Elsevier: Amsterdam, 2005; pp 195−249. (28) Hratchian, H. P.; Schlegel, H. B. Using Hessian Updating to Increase the Efficiency of a Hessian Based Predictor-corrector Reaction Path Following Method. J. Chem. Theory Comput. 2005, 1, 61−69. (29) Katz, J. J. Chemical and biological studies with deuterium. Annual Priestley Lectures; Pennsylvania State University: University Park, PA, 1965; Vol. 39, pp 1−110. (30) Blake, M. I.; Crespi, H. L.; Katz, J. J. Studies with Deuterated Drugs. J. Pharm. Sci. 1975, 64, 367−391. (31) Class, J. B. A Review of the Fundamentals of Crosslinking with Peroxides. Rubber World 1999, 220, 35−39.

(32) Yeh, P. Y.; Shu, C. M.; Duh, Y. S. Thermal Hazard Analysis of Methyl Ethyl Ketone Peroxide. Ind. Eng. Chem. Res. 2003, 42, 1−5. (33) Bach, R. D.; Su, M. D.; Schlegel, H. B. Oxidation of Amines and Sulfides with Hydrogen Peroxide and Alkyl Hydrogen Peroxide. The Nature of the Oxygen-Transfer Step. J. Am. Chem. Soc. 1994, 116, 5379−5391. (34) Katsumura, Y.; Jiang, P. Y.; Nagaishi, R.; Oishi, T.; Ishigure, K.; Yoshida, Y. Pulse Radiolysis Study of Aqueous Nitric Acid Solutions Formation Mechanism, Yield, and Reactivity of NO3 Radical. J. Phys. Chem. 1991, 95, 4435−4439. (35) Anastas, P.; Warner, J. Green Chemistry: Theory and Practice; Oxford University Press: New York, 1998.

10784

dx.doi.org/10.1021/ie301280c | Ind. Eng. Chem. Res. 2012, 51, 10778−10784