Next Article in Journal
SERVQUAL to Determine Relationship Quality and Behavioral Intentions: An SEM Approach in Retail Banking Service
Next Article in Special Issue
Risk Assessment Method for Spontaneous Combustion of Pyrophoric Iron Sulfides
Previous Article in Journal
The European Solar Communication—Will It Pave the Road to Achieve 1 TW of Photovoltaic System Capacity in the European Union by 2030?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effects of Acid and Hydrogen Peroxide Stabilizer on the Thermal Hazard of Adipic Acid Green Synthesis

Jiangsu Key Laboratory of Hazardous Chemicals Safety and Control, College of Safety Science and Engineering, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(8), 6530; https://doi.org/10.3390/su15086530
Submission received: 12 March 2023 / Revised: 30 March 2023 / Accepted: 31 March 2023 / Published: 12 April 2023
(This article belongs to the Special Issue Risk Assessment and Management in the Process Industries)

Abstract

:
The synthesis of adipic acid, which is formed by the reaction of cyclohexene oxidized by hydrogen peroxide (H2O2), is hazardous because of the highly exothermic nature of this reaction and H2O2 decomposition. The objective of this comprehensive study was to investigate and illustrate the effects of sulfuric acid (H2SO4) and H2O2 stabilizer (EDTA) on the thermal hazard of H2O2 decomposition and the green synthesis of adipic acid, which also provided a reference to reduce the risk of the reactions. Various calorimetry techniques were carried out to characterize the exothermic behavior of the reactions. An HPLC device was used to characterize the yield of adipic acid and the conversion rate of the raw materials, cyclohexene and H2O2. Meanwhile, density functional theory calculations were performed to understand the reaction mechanism and the associated energies of H2O2 decomposition catalyzed by sodium tungstate dihydrate (Na2WO4·2H2O). Finally, combined with the calorimetry results, the risk of the adipic acid synthesis reaction was assessed using the intrinsic control index method (ITHI). The results show that the addition of H2SO4 and EDTA can reduce the exothermic heat of the H2O2 decomposition reaction and the green synthesis reaction of adipic acid. The yield of adipic acid was also increased. The hazard level of stage A was IV, and to remove more reaction heat, it was recommended to enhance the reflux cooling of stage A. The hazard level of stage B was I, which was very low and no further measures could be taken.

1. Introduction

As the most practical dicarboxylic acid, adipic acid has important application values in the organic synthesis of plastics, synthetic resins, food, and other fields. Adipic acid is mainly used as an intermediate in the manufacture of nylon 66, polyurethane, plasticizers, and other materials [1].
At present, the two main adipic acid synthesis processes are the cyclohexane [2] and the cyclohexene methods [3], both of which have low atomic utilization and the serious pollution problem of “the three wastes”. The oxidant used in these two methods is nitric acid, which generates a large amount of NOx exhaust and acidic wastewater [2]. The catalyst is cuprum and vanadium, which generates waste residue. Therefore, the research and development of new technologies for the green synthesis of adipic acid has received extensive attention from experts and scholars at home and abroad. Among these new technologies, the hydrogen peroxide (H2O2) oxidation method could align with the requirements of green chemistry, with its mild conditions and simple operations, and which can ensure a high conversion rate, high yield, and high selectivity while maintaining clean production [4,5,6]. Jiang et al. [7] investigated the thermal hazards of adipic acid synthesis achieved by the oxidation of cyclohexene with H2O2, combined with reaction calorimetry and mechanism analysis, and found that the process could be divided into two reaction stages, both of which were assessed as a class 5 risk according to the Stoessel criticality diagram method, indicating that the process had a high risk of explosion.
Typically, H2O2 is unstable with highly exothermic hazards. The initial decomposition temperature of H2O2 ranges from 47 to 81 °C, and the decomposition heat is about 192–1079 J/g [8,9]. Wu et al. [10] suggested that the sulfuric acid (H2SO4) at a low concentration could slow down the decomposition rate of H2O2, and the rate could be increased with an increase in H2SO4 concentration. Jung et al. [11] found that the addition of C8H6O4 as a stabilizer to a Fenton reagent or Fenton-like reagent could extend the lifetime of H2O2 and reduce the rate of H2O2 autolysis. Wu et al. [12] studied the effects of inorganic salts and organic acids on the thermal runaway behavior of hydrogen peroxide, and found that the order of catalytic effect from large to small was Fe2+ > Fe3+ > Cu2+, and the order of risk of thermal runaway from large to small was H2O2/HCOOH > H2O2 > H2O2/CH3COOH. Zhang et al. [13] investigated the effects of free manganese ions and manganese on H2O2 decomposition in the presence of precipitated lignin under typical pulp bleaching conditions, and found that adding only Ethylene Diamine Tetraacetic Acid (EDTA) was the most effective method for reducing the decomposition of H2O2.
Both high temperatures and the catalysts used can accelerate the ineffective decomposition of H2O2 during the oxidation of cyclohexene by H2O2 for the preparation of adipic acid, resulting in the concentration of H2O2 decreasing, and the heat generated by the H2O2 decomposition reaction increasing. Consequently, the thermal hazard of the process is greatly increased. Therefore, it is necessary to select an effective H2O2 stabilizer to inhibit the decomposition reactions of excessive H2O2, so that the H2O2 can fully and effectively participate in the adipic acid synthesis.
In this work, various calorimetric and analytical techniques were carried out to investigate the effects of H2SO4 concentration and the addition of EDTA on the thermal hazard of the adipic acid synthesis reaction: (1) Differential scanning calorimetry (DSC) was implemented to obtain the thermal decomposition parameters of H2O2 under different catalytic conditions. (2) An automatic reaction calorimeter (EasyMaxTM102) was employed to measure the enthalpy of the H2O2 decomposition reaction and the adipic acid synthesis reaction. Meanwhile, high-performance liquid chromatography (HPLC) was used to assess the conversion rate of the substances used in the reaction and the yield of adipic acid. (3) An adiabatic calorimeter (Phi-TEC II) was implemented to simulate the adiabatic decomposition behavior and parameters of H2O2 and the reaction products obtained by calorimetric experiments. Finally, the reaction risks of the target reactions were assessed using the ITHI method [14].

2. Materials and Methods

2.1. Materials

Hydrogen peroxide (30 wt.%) and H2SO4(98 wt.%, AR) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). The cyclohexene and sodium tungstate dehydrate (Na2WO4·2H2O) used in this work were of analytical grade and made by Shanghai Macklin Biochemical (Shanghai, China). Adipic acid (HPLC, Shanghai Aladdin Bio-Chem Technology, Shanghai, China), methanol, (HPLC, Shanghai Aladdin Bio-Chem Technology, Shanghai, China), acetonitrile (HPLC, Shanghai Aladdin Bio-Chem Technology, Shanghai, China), phosphoric acid (AR, Shanghai Macklin Biochemical, Shanghai, China), potassium dihydrogen phosphate (AR, Shanghai Macklin Biochemical, Shanghai, China), sodium dodecyl sulfate (SDS) (AR, Sinopharm Chemical Reagent, Shanghai, China), Ethylene Diamine Tetraacetic Acid (EDTA) (AR, tcichemicals) were used as supplied. All reagents were used without further purification.

2.2. Methods

The reaction equation for the catalytic oxidation of cyclohexene by H2O2 to synthesize adipic acid is shown in Scheme 1.
During the synthesis of adipic acid, a total of four oxidation and two hydrolysis reactions occur. At 73 °C for two hours, cyclohexene is oxidized and hydrolyzed to produce 1,2-cyclohexanediol. The 1,2-cyclohexanediol continues to be oxidized to 2-hydroxycyclohexanone, which is recorded as stage A. At 90 °C for five hours, 2-hydroxycyclohexanone undergoes two steps of oxidation and hydrolysis to produce the final product, adipic acid, and the first three hours of this period are recorded as stage B [15].

2.2.1. HPLC Experiment

An HPLC device (Agilent-1260, Agilent Technologies Co., Ltd., Santa Clara, CA, USA) was used to characterize the yield of adipic acid and the conversion rate of the raw materials, cyclohexene and H2O2 [16]. The HPLC conditions for the adipic acid testing were as follows: The injection volume was 20 μL. The mobile stage was phosphoric acid-potassium dihydrogen phosphate buffer (50 mmol/L) and methanol. The elution rate was 0.3 mL/min. The column temperature was 30 °C. The chromatographic conditions for the cyclohexene and H2O2 tests were as follows: The injection volume was 5 μL. The mobile stage was acetonitrile and ultrapure water. The elution rate was 0.8 L/min. The column temperature was 30 °C and the UV detection wavelength was 209 nm [17].

2.2.2. Calorimetric Experiments

An EasyMaxTM102 calorimeter (Mettler Toledo Ltd., Greifensee, Switzerland) was used to determine the exothermic characteristics of adipic acid synthesis and the decomposition of H2O2 using a 100 mL atmospheric pressure glass reactor. The reactor temperature (Tr), jacket temperature (Tj), and heat release rates (qr) were all tracked automatically to analyze the reaction enthalpy, adiabatic temperature rises ( T ad ), and other thermodynamic parameters.

2.2.3. DSC Experiment

DSC is among the most used techniques for thermal analysis [18]. In this experiment, DSC 250 (TA Instruments, Eden Prairie, MN, USA) was carried out to investigate the exothermic behavior of H2O2, the intermediates, and the products in stages A and B under Na2WO4·2H2O, H2SO4, and EDTA conditions. The scanning temperature range was 40~180 °C and 40~300 °C. The ramp-up rate was 5 °C/min. Nitrogen was used for protection and the flow rate was 50 mL/min.

2.2.4. Adiabatic Tests

An adiabatic calorimeter (Phi-TEC II, HEL Limited, Hertford, UK) was used to identify the adiabatic decomposition characteristics of the reactants. Samples (~1 g) of the stage A product, the stage B product, and the raw H2O2 mixtures from the Calorimetric experiments were loaded into a 10 mL Hastelloy test cell. The temperature ranged from 30 to 300 °C with a step-upward trend of 5 °C. The temperature and pressure data from the adiabatic experiments were tracked to record any exotherms. The detection sensitivity was 0.02 °C/min [19].

3. Results and Discussion

3.1. The Working Curves Obtained from the HPLC Experiments

Three sets of adipic acid, cyclohexene, and H2O2 solutions with concentrations ranging from 1 mg/g to 50 mg/g were prepared and used to obtain the working curves using HPLC (mg/g is the ratio of the mass of the substance to the mass of the prepared standard solution). The experimental results are shown in Figure 1, Figure 2 and Figure 3.

3.2. The Decomposition Mechanism of H2O2 Catalyzed by Na2WO4·2H2O

Typically, in addition to participating in the catalytic oxidation of cyclohexene to adipic acid, H2O2 can also be auto catalyzed for decomposition. This leads to the reduced utilization of H2O2 during the whole process, and the heat released from the decomposition increase the thermal hazard of the reaction. Hence, it is necessary to identify the decomposition mechanism of H2O2 to reduce the thermal hazard of the process and to improve the utilization of H2O2.
Nardello V et al. [20] investigated the disproportionation reaction of H2O2 catalyzed by Na2WO4·2H2O to form singlet oxygen (1O2). The mono-, di-, and tetraperoxotungstate intermediates W O 2 n O 4 n 2 (n = 1, 2, 4) have been characterized using UV and 183W NMR spectroscopies. The diperoxo species was proposed as the precursor of 1O2. Yoshiro O et al. [21] carried out the kinetics for the oxidation of dimethyl sulfoxide (DMSO) by H2O2 using iodometry. Meanwhile, the major protons were formed at a solution pH of 3.5–5.0 acidity (H2WO5 and H2WO8) with using Na2WO4·2H2O as the catalyst. The effect of acidity on the rate of oxidation and the rate determining mechanism of peroxotungstic acid were also discussed.
According to Nardello V et al. and Yoshiro O et al., the decomposition of H2O2 catalyzed by sodium tungstate was studied because of the presence of H2O2, Na2WO4·2H2O, and H2SO4 in the oxidative synthesis of adipic acid. In this work, as shown in Figure 4, a possible decomposition mechanism of H2O2 catalyzed by Na2WO4·2H2O was proposed on the basis of existing research. First, W O 4 2 is hydrolyzed to produce H W O 4 , after which H W O 4 is gradually oxidized by H2O2 to form H W O 3 ( O 2 ) and H W O 2 O 2 2 . O 2 is generated through H W O 2 O 2 2 , and then H W O 4 is produced, leading a catalytic decomposition cycle. As the pH of the solution decreases, H W O 2 O 2 2 and W ( O 2 ) 4 2 are in dynamic equilibrium in the presence of 2 mol of H2O2.

3.3. Reaction Calorimetry Results and Analysis

3.3.1. The Exothermic Results and Analysis of the Catalytic Decomposition Reaction of H2O2

In reaction condition 1, the molar ratio of H2O2 to Na2WO4·2H2O was 4.4:0.06. In reaction condition 2 the molar ratio of H2O2 to Na2WO4·2H2O was 4.4:0.06 and the H2SO4 concentration was 0.22 mol/L. In reaction condition 3, the molar ratio of H2O2 to Na2WO4·2H2O was 4.4:0.06, the H2SO4 molar concentration was 0.22 mol/L, and the EDTA mass concentration was 18.3 g/L.
The experimental tests were all carried out at room temperature with stirring at 300 rpm, then Tr was heated to 73 °C and held for 120 min. Subsequently, the reactions were maintained for 240 min at 90 °C. Samples were taken at every 60 min intervals and analyzed the conversion of H2O2 by HPLC.
As depicted in Figure 5, when Tj was raised to about 60 °C, Tr increased rapidly, indicating that the reaction was violently exothermic, and a large number of bubbles were also generated during the process. After temperature stabilization, the decomposition of H2O2 was basically complete. The maximum temperature difference between Tr and Tj during the process was about 59 °C. Thus, if a cooling failure occurred, the reaction temperature could rise sharply, resulting in runaway accidents.
The amount of H2SO4 was increased under reaction conditions 2 to reduce the potential of hydrogen (pH) of the solution. Hence, the process of catalytic decomposition could be inhibited under the dynamic equilibrium of H W O 2 O 2 2 and W ( O 2 ) 4 2 in the presence of 2 mol of H2O2. The decomposition rate of H2O2 slowed down as is shown in Figure 6. The decomposition rate of H2O2 at the isothermal stage of 90 °C is significantly higher than at the isothermal stage of 73 °C. This indicates that the addition of acid effectively reduced the catalytic decomposition of H2O2, while the high temperature still led the H2O2 to decompose completely; the conversion rate reached 100% after 240 min of reaction time at the 90 °C isothermal stage. Therefore, it is essential to add H2O2 stabilizers into H2O2 solutions to inhibit the ineffective decomposition of H2O2 during the oxidation reaction.
EDTA, which is commonly used as a stabilizer of H2O2 in the field of paper bleaching, is rarely used in the chemical industry. In test 3, EDTA was added and the other conditions were kept consistent with those of test 2. As depicted in Figure 7, after 240 min of reaction time at the 90 °C isothermal stage, only about 41.24% of the H2O2 was converted, the results indicate that the addition of EDTA can improve the thermal stability of H2O2 Therefore, the application of EDTA in the process of catalytic oxidation of cyclohexene to adipic acid with H2O2 was relatively feasible.

3.3.2. Calorimetry and Theoretical Calculation of Catalytic Decomposition of H2O2

Gaussian is a professional calculation software in the field of quantum chemistry, which can be used to calculate thermodynamic parameters such as enthalpy change ΔH and free energy change ΔG of chemical reactions, and the calculated values are very close to the experimental values, thus accurately predicting the thermodynamic properties of reactions. Based on the mechanism of the catalytic decomposition of hydrogen peroxide, density functional theory (DFT) calculations of the enthalpy of reaction for the catalytic decomposition of H2O2 solution were performed. The standard molar enthalpies of production for each step of the reaction are shown in Table 1. H W O 2 O 2 2 gives   H W O 4 after releasing O 2 and H W O 2 O 2 2 is oxidized to W ( O 2 ) 4 2 by 2 mol of hydrogen peroxide as a heat-absorbing reaction. The standard molar enthalpy of production for the entire hydrogen peroxide decomposition reaction is 61.98 kJ/mol.
The heat of decomposition and T a d of the H2O2 decomposition reaction were obtained using test 1. The whole reaction heat was 19.04 kJ and T a d was 170.77 °C. As shown in Figure 8, the trend of qr curve showed two severe peaks of exothermic heat. The maximum heat release rate was about 55.72 W. This indicates that the decomposition rate of H2O2 catalyzed by Na2SO4·2H2O was great and acute. The large amount of gas produced could easily trigger an accidental explosion. The mass of the 30 wt.% H2O2 used in the experiment was 36.434 g, so the enthalpy of reaction was 59.5 kJ/mol. The theoretical calculation gives a standard molar enthalpy of production of 61.98 kJ/mol. The theoretical value is close to the calculated value.

3.3.3. Adipic Acid Green Synthesis Reaction Results and Analysis

Sodium dodecyl sulfate (SDS) is an anionic surfactant with good emulsification. In the present work, to investigate the effect of H2SO4 concentration and the addition of EDTA on the reaction process of adipic acid synthesis, nine experiments were conducted on the synthesis of adipic acid using calorimetric experiments, and the yield of adipic acid was analyzed using HPLC. The molar ratio of the H2O2, cyclohexene, and Na2WO4·2H2O used in the tests was all 4.4:1:0.06. Different concentrations of H2SO4 (CH2SO4), SDS(ωSDS), and EDTA(ωEDTA) were used in the different tests, as shown in Table 2.
In experiment 1, the incomplete yield of the entire synthesis reaction, without the addition of SDS, was only 8.67%. Since the process is a liquid–liquid non-homogeneous reaction, both the lack of organic solvents and the ultimate stirring speed could lead to the poor mass transfer of the reaction. The mass transfer could be enhanced by SDS emulsification, and the yields of adipic acid were significantly increased, as depicted in experiments 2 and 3. The comparison of experiments 2, 3, 4, and 5 indicated that the addition of EDTA could reduce the total heat release while increasing the yield of adipic acid. By comparing experiments 4 and 7, it was found that the decrease of H2SO4 concentration also reduced the total heat release and increased the yield of adpic acid. Experiment 7, which had the lowest total heat release and temperature difference, the H2SO4 concentration of 0.12 mol/L, the SDS concentration of 9.15 g/L, and the EDTA concentration of 18.3 g/L, was determined to be the optimum process for adipic acid synthesis after a comprehensive comparison of experiments 1–9.
In order to accurately assess the thermal hazard of the adipic acid synthesis process, the entire reaction process was divided into two stages, A and B, as shown in Figure 9.
The severity of the reaction runaway is usually determined using T a d due to the exothermic of the target reaction. Since the yield of adipic acid was 82.78%, the accurate adiabatic temperature rise T a d , r should be corrected by the yield Y [22]. The equations for calculating T a d and T a d , r are listed in Equations (1) and (2). The calorimetric experimental data and results are shown in Table 3, where H is the heat of the oxidation reaction, kJ; m r is the total mass of the oxidation system, g; and C p is the specific heat capacity of the system after the oxidation reaction J/(K·g).
T a d = H r C p m r
T a d , r = H r C p m r Y
T c f = T p + H r 0 t q r t d t C p M t
T c f , r = T p + H r Y 0 t q r t d t C p M t
In Equations (3) and (4), T p is the process operating temperature, which depends on the process conditions, where T p is the temperature in the kettle during the synthesis process, °C. M t is the total mass of the oxidized reaction substance. T c f is the temperature that can be reached by the oxidation reaction after cooling failure, and the maximum value of T c f is MTSR [23].   T c f should also be corrected by the yield Y , which is T c f , r , and q r is the heat release rate of the reaction, W. The changes in T c f , r and T c f , r are shown in Figure 10 and Figure 11. The stage A MTSRr was 192.33 °C and the stage B MTSRr was 161.95 °C.

3.4. DSC Experimental Results and Analysis

DSC was used to investigate the effect of H2SO4 and EDTA conditions on the catalytic decomposition characteristics of 30 wt.% H2O2 and the thermal stability of the products from stages A and B. Sample 1 was a mixture of 30 wt.% H2O2 and tungsten Na2WO4·2H2O. Sample 2 was a mixture of 30 wt.% H2O2, Na2WO4·2H2O, and H2SO4. Sample 3 was a mixture of 30 wt.% H2O2, Na2WO4·2H2O, H2SO4, and EDTA. The proportion of each substance in the mixture followed the process parameters of experiment 7. Samples 4 and 5 were the products from stages A and B. The experimental results are shown in Figure 12 and Figure 13.
As indicated in Table 4, the mixed system became more acidic with the addition of H2SO4, and slowed the decomposition rate of H2O2. The rate of heat generation and the heat of liberation could be decreased, although the initial decomposition temperature ( T o n s e t ) was advanced. The autolytic reaction of H2O2 was effectively inhibited by the addition of the H2O2 stabilizer, EDTA. T o n s e t was delayed and the heat released from decomposition was decreased. The product of stage A was exothermic due to the presence of unfinished H2O2 and intermediate products. However the product of stage B, was mostly adipic acid, and consequently there was no exothermic heat during the test.

3.5. Adiabatic Tests Results and Analysis

In this work, the adiabatic tests of the stage A and B products and sample 3 were executed using Phi-TEC II, as depicted in Figure 14. No exothermic heat was detected in the stage B products. The characteristic parameters of the adiabatic decomposition reaction of the stage A products and sample 3 are shown in Table 5.
In order to assess the thermal hazard of the adipic acid green synthesis process, T M R a d and T D 24 were calculated for the stage A products to access the possibility and reaction risk of the target process. T M R a d is the time required to reach the maximum rate under adiabatic conditions [24], calculated using Equation (5), and T D 24 is the temperature at which T M R a d equals 24 h [25].
T M R a d = R T 2 A T f T T n T e x p E R T E
d T d t = A T f T T n T e x p E R T
Here, Tf is the maximum temperature obtained by self-heating under adiabatic conditions, K; R is the gas constant, 8.314 J/(mol·K); E is the apparent activation energy, kJ/mol; A is pre-exponential, s−1; and n is the number of reaction stages.
According to Equation (6), the thermodynamic parameters were obtained by non-linear fitting. The correlation coefficient (R2) was 0.954 and the fit was essentially the same as the experimental data. E was 98.215 kJ/mol, A was 7.688 × 1012 s−1, and n was 1.238, as depicted in Figure 15. As indicated in Figure 16, the curve of T M R a d vs. Temperature was fitted from the adiabatic kinetic data, and T M R a d was 34.8 min and T D 24 was 27.95 °C.

3.6. ITHI Method for Assessing the Thermal Hazards of Adipic Acid Synthesis Reactions

The ITHI method, an intrinsic control index method that considers the combination of substance hazard and reaction heat hazard, was chosen to provide a comprehensive and accurate assessment of the hazards of adipic acid synthesis [12]. The two stages, A and B, of the adipic acid synthesis process were assessed, respectively.
As depicted in Table 6, Table 7 and Table 8, the material coefficients (MF) of stages A and B were 1.25 and 1.188.
The probability of runaway (P) in stages A and B were 10 and 4, respectively, and the severity of runaway (S) is the sum of (IH,rx,IΔTad,rx)max and (IH,dec,IΔTad,dec)max. The accident consequence severity S in stage A was 5, and the post-accident severity S in stage B was 3.
The runaway consequence severity (RI) is the product of S and P. The ITHI value is the product of MF and RI. Thus, the ITHI of stages A and B were 62.5 and 13.42, respectively. The risk level of stage A was IV, since stage A was carried out under the reflux period. Consequently, the cooling of reflux should be increased to remove more of the heat generated from stage A and to more steadily control the temperature. The risk level of Stage B was I, which indicated the risk was low and no action should be taken.

4. Conclusions

We used various calorimetric equipment and material analysis instruments to conduct a thermal hazard analysis of the catalytic decomposition reaction of hydrogen peroxide and the green synthesis reaction of adipic acid. The conclusions are as follows:
(1)
The ineffective decomposition of hydrogen peroxide during the green synthesis reaction of adipic acid not only wastes raw materials, but also greatly increases the thermal risk of the reaction. It was found that the addition of sulfuric acid and the hydrogen peroxide stabilizer EDTA improved the thermal stability of hydrogen peroxide, and reduced the ineffective decomposition of hydrogen peroxide.
(2)
The addition of sulfuric acid and EDTA to the DSC experiment delayed the onset of decomposition temperature (Tonset) and reduced the heat separation.
(3)
In the green synthesis process of adipic acid, by varying the concentration of sulfuric acid, SDS, and EDTA, the process conditions were optimized to 0.12 mol/L sulfuric acid, 9.15 g/L SDS, and 18.3 g/L EDTA, resulting in the lowest overall heat release and temperature difference, and an adipic acid yield of 82.78%.
(4)
A thermal hazard assessment of the optimized adipic acid green synthesis reaction was carried out using the ITHI method for stages A and B. The hazard level of stage A is Ⅳ, and improvement of the heat transfer capacity of the reflux unit is recommended. The hazard level of stage B is I, and no additional measures are needed due to the low hazard level.

Author Contributions

Methodology, H.Y.; Software, H.Y.; Validation, Y.L.; Formal analysis, J.S.; Data curation, J.S.; Writing—original draft, Y.L.; Writing—review & editing, Z.C.; Project administration, L.N.; Funding acquisition, L.N. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the support of the National Natural Science Foundation of China (no. 52274209).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cao, H.G.; Zhu, B.R.; Yang, Y.F.; Xu, L.; Yu, L.; Xu, Q. Recent advances on controllable and selective catalytic oxidation of cyclohexene. Chin. J. Catal. 2018, 39, 899–907. [Google Scholar] [CrossRef]
  2. Schneider, L.; Lazarus, M.; Kollmuss, A. Industrial N2O Projects under the CDM: Adipic Acid—A Case of Carbon Leakage; Working Paper WP-US-1006; Stockholm Environmental Institute: Stockholm, Sweden, 2010. [Google Scholar]
  3. Castellan, A.; Bart, J.C.J.; Cavallaro, S. Synthesis of adipic acid via the nitric acid oxidation of cyclohexanol in a two-step batch process. Catal. Today 1991, 9, 285–299. [Google Scholar] [CrossRef]
  4. Sato, K.; Aoki, M.; Takagi, J.; Noyori, R. Organic solvent-and halide-free oxidation of alcohols with aqueous hydrogen peroxide. J. Am. Chem. Soc. 1997, 119, 12386–12387. [Google Scholar] [CrossRef]
  5. Sato, K.; Aoki, M.; Noyori, R.A. “Green” route to adipic acid: Direct oxidation of cyclohexene with 30 percent hydrogen peroxide. Science 1998, 281, 1646–1647. [Google Scholar] [CrossRef]
  6. Van de Vyver, S.; Román-Leshkov, Y. Emerging catalytic processes for the production of adipic acid. Catal. Sci. Technol. 2013, 3, 1465–1479. [Google Scholar] [CrossRef] [Green Version]
  7. Jiang, W.; Ni, L.; Jiang, J.C.; Chen, Q.; Chen, Z.Q. Thermal hazard and reaction mechanism of the preparation of adipic acid through the oxidation with hydrogen peroxide. AIChE J. 2021, 67, e17089. [Google Scholar] [CrossRef]
  8. Wu, L.K.; Lin, C.M.; Shu, C.M.; Kao, C.S. An evaluation on thermokinetic parameters for hydrogen peroxide at various concentrations by DSC. J. Therm. Anal. Cal. 2006, 85, 87–89. [Google Scholar]
  9. Lin, Y.J.; Lin, Z.S.; Wang, Y.W. Thermal runaway evaluation using DSC1, VSP2, and kinetics models on Cu etchant and its waste in high-tech etching process. J Therm. Calorim. 2020, 144, 285–294. [Google Scholar] [CrossRef]
  10. Wu, L.K.; Chen, K.Y.; Cheng, S.Y.; Lee, B.S.; Shu, C.M. Thermal decomposition of hydrogen peroxide in the presence of sulfuric acid. J. Therm. Anal. Calorim. 2008, 93, 115–120. [Google Scholar] [CrossRef]
  11. Oh, H.S.; Kim, J.J.; Kim, Y.H. Stabilization of hydrogen peroxide using tartaric acids in Fenton and fenton-likeoxidation. Korean J. Chem. Eng. 2016, 33, 885–892. [Google Scholar] [CrossRef]
  12. Wu, D.J.; Qing, X.M.; Liu, L.Q.; Zang, N. Eeffect of inorganic salt and organic acid on the thermal runaway of hydrogen peroxide. J. Loss Prev. Proc. 2019, 57, 34–40. [Google Scholar] [CrossRef]
  13. Zhang, X.J.; Wang, L.J.; Li, Y.M.; Lei, L.R. Decomposition of hydrogen peroxide by manganese in the presence of lignin. Appita J. 2016, 69, 241–246. [Google Scholar]
  14. Jiang, J.C.; Wei, D.; Ni, L.; Fu, G.; Pan, Y. Inherent thermal runaway hazard evaluation method of chemical process based on fire and explosion index. J. Loss Prev. Proc. 2020, 64, 104093. [Google Scholar]
  15. Peng, J.; Zhao, Z.; Wei, D.; Tang, M.; Wang, X. Influence of reaction conditions on product distribution in the green oxidation of cyclohexene to adipic acid with hydrogen peroxide. Catal. Today 2011, 175, 619–624. [Google Scholar]
  16. Jin, F.M.; Moriya, T.; Enomoto, H. Oxidation reaction of high moleculer weight carboxylic acids in supercritical water. Environ. Sci. Technol. 2003, 37, 3220–3231. [Google Scholar] [CrossRef]
  17. Mouheb, L.; Dermeche, L.; Mazari, T.; Benadji, S.; Essayem, N.; Rabia, C. Clean adipic acid synthesis from liquid-stage oxidation of cyclohexanone and cyclohexanol using (NH4)(x)A(y)PMo(12)O(40) (A:Sb,Sn,Bi) mixed heteropolysalts and hydrogen perpxide in free solvent. Cata Lett. 2018, 148, 612–620. [Google Scholar] [CrossRef]
  18. Wang, J.Y.; Sun, Y.; Gao, X.D.; Mannan, M.S.; Wilhite, B. Experimental study of electrostatic hazard inside scrubber column using response surface methodology. Chem. Eng. Sci. 2019, 200, 46–48. [Google Scholar] [CrossRef]
  19. Yong, Z.; Ni, L.; Jiang, J.C. Themal hazard analyses for the synthesis of benzoyl peroxide. J. Loss Prev. Proc. 2016, 43, 35–41. [Google Scholar]
  20. Nardello, V.; Marko, J.; Vermeersch, G.; Aubry, J.M. 183W NMR Study of Peroxotungstates Involved in the Disproportionation of Hydrogen Peroxide into Singlet Oxygen (1O2, 1Δg) Catalyzed by Sodium Tungstate in Neutral and Alkaline Water. Inorg. Chem. 1998, 37, 5418–5423. [Google Scholar] [CrossRef]
  21. Ogata, Y.; Tanaka, K. Kinetics of the oxidation of dimethyl sulfoxide with aqueous hydrogen peroxide catalyzed. Can. J. Chem. 1980, 59, 718–722. [Google Scholar] [CrossRef]
  22. Zang, N.; Qian, X.M.; Liu, Z.Y.; Shu, C.M. Thermal hazard evaluation of cyclohexanone peroxide synthesis. J. Therm. Anal. Calorim. 2016, 124, 1131–1139. [Google Scholar] [CrossRef]
  23. Abele, S.; Schwaninger, M.; Fierz, H.; Shcmidet, G.; Funel, J.A.; Stoessel, F. Safety assessment of Diels-Alder reactions with highly reactive acrylic monomer. Org. Process Res. Dev. 2012, 16, 2015–2020. [Google Scholar] [CrossRef]
  24. Zhao, X.Q.; Wu, W.Q.; Li, H.B.; Guo, Z.C.; Chen, W.H.; Chen, L.P. Thermal hazards of benzaldehyde oxime: Based on decomposition products and kinetics analysis by adiabatic calorimeter. Process Saf. Environ. Prot. 2021, 153, 527–536. [Google Scholar] [CrossRef]
  25. Wu, X.J.; Zhu, Y.; Chen, L.P.; Chen, W.H.; Guo, Z.C. Effect of reaction type on TMRad, TD24 and other data obtained by adiabatic calorimetry. J. Therm. Anal. Calor. 2022, 147, 5883–5896. [Google Scholar] [CrossRef]
Scheme 1. The synthetic scheme of adipic acid.
Scheme 1. The synthetic scheme of adipic acid.
Sustainability 15 06530 sch001
Figure 1. The working curve of adipic acid standardization solution.
Figure 1. The working curve of adipic acid standardization solution.
Sustainability 15 06530 g001
Figure 2. The working curve of cyclohexene standardization solution.
Figure 2. The working curve of cyclohexene standardization solution.
Sustainability 15 06530 g002
Figure 3. The working curve of hydrogen peroxide standardization solution.
Figure 3. The working curve of hydrogen peroxide standardization solution.
Sustainability 15 06530 g003
Figure 4. Possible decomposition mechanism of hydrogen peroxide catalyzed by sodium tungstate.
Figure 4. Possible decomposition mechanism of hydrogen peroxide catalyzed by sodium tungstate.
Sustainability 15 06530 g004
Figure 5. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test 1.
Figure 5. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test 1.
Sustainability 15 06530 g005
Figure 6. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test 2.
Figure 6. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test 2.
Sustainability 15 06530 g006
Figure 7. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test.
Figure 7. The Tr, Tj, and conversion of H2O2 curves of the decomposition reaction test.
Sustainability 15 06530 g007
Figure 8. The Tr, Tj, and qr curves in 30 wt.% hydrogen peroxide decomposition reaction.
Figure 8. The Tr, Tj, and qr curves in 30 wt.% hydrogen peroxide decomposition reaction.
Sustainability 15 06530 g008
Figure 9. The Tr, Tj, and qr curves of the adipic acid synthesis reaction.
Figure 9. The Tr, Tj, and qr curves of the adipic acid synthesis reaction.
Sustainability 15 06530 g009
Figure 10. MTSR and Tcf curves of oxidation reaction in stage A.
Figure 10. MTSR and Tcf curves of oxidation reaction in stage A.
Sustainability 15 06530 g010
Figure 11. MTSR and Tcf curves of oxidation reaction in stage B.
Figure 11. MTSR and Tcf curves of oxidation reaction in stage B.
Sustainability 15 06530 g011
Figure 12. DSC curves of 30 wt.% H2O2 decomposition reaction.
Figure 12. DSC curves of 30 wt.% H2O2 decomposition reaction.
Sustainability 15 06530 g012
Figure 13. DSC curves of adipic acid synthesis reaction product.
Figure 13. DSC curves of adipic acid synthesis reaction product.
Sustainability 15 06530 g013
Figure 14. Temperature vs. time curves of exothermic section in Phi-TEC II experiment.
Figure 14. Temperature vs. time curves of exothermic section in Phi-TEC II experiment.
Sustainability 15 06530 g014
Figure 15. Comparison of kinetic model simulation for stage A.
Figure 15. Comparison of kinetic model simulation for stage A.
Sustainability 15 06530 g015
Figure 16. The calculation results of TMRad with Temperature under the adiabatic conditions of the stage A product.
Figure 16. The calculation results of TMRad with Temperature under the adiabatic conditions of the stage A product.
Sustainability 15 06530 g016
Table 1. The standard enthalpy of decomposition of H2O2 catalyzed by sodium tungstate.
Table 1. The standard enthalpy of decomposition of H2O2 catalyzed by sodium tungstate.
Reaction StepsStandard Molar Enthalpy of Production (kJ/mol)
2 H 2 O 2 2 H 2 O + O 2 61.98
W O 4 2 + H 2 O H W O 4 + O H 168.18
H W O 4 + H 2 O 2 H W O 3 ( O 2 ) + H 2 O 68.55
H W O 3 ( O 2 ) + H 2 O 2 H W O 2 O 2 2 + H 2 O 59.48
H W O 2 O 2 2 H W O 4 + O 2 −66.05
H W O 2 O 2 2 + 2 H 2 O 2 W ( O 2 ) 4 2 + H 3 O + + H 2 O −1025.97
Table 2. Thermal chemical parameters under different reaction conditions for adipic acid synthesis reaction.
Table 2. Thermal chemical parameters under different reaction conditions for adipic acid synthesis reaction.
RunCH2SO4 (mol/L)ωSDS (g·L−1)ωEDTA (g·L−1) H A (kJ) H B (kJ) H r (kJ) ΔTA (°C)ΔTB (°C)YAA (%)
10.2200-----8.67
20.2218.3022.63513.13435.7710.423.1772.31
30.2218.39.1524.413.18527.59514.25−0.3575.21
40.229.1518.321.2743.40624.6811.96−0.5778.25
50.229.15022.4157.30729.7227.872.2373.62
60.129.159.1522.4575.23427.697.126.5381.25
70.129.1518.313.2297.80921.0385.3486.06782.78
80.129.15015.5675.89921.4665.456.5476.65
90.129.159.1521.2314.700325.9316.624.9581.83
Table 3. Thermal parameters of oxidation reaction in EasyMaxTM102 experiment.
Table 3. Thermal parameters of oxidation reaction in EasyMaxTM102 experiment.
Stage H r (kJ) C p (J/(K·g)) T a d , r (°C) M T S R , r (°C)
A13.2292.88122.82192.33
B7.8092.8872.5161.95
Table 4. Thermal decomposition parameters for different products.
Table 4. Thermal decomposition parameters for different products.
Sample Sequence T o n s e t (°C) T peak (°C) H d (J/g)
183.6189.71825.72
272.2093.84508.80
3106.08115.6455.92
483.54123.3587.3
582.2, 135.781.1, 114.4−1143.24
Table 5. Thermal characteristic parameters of the first decomposition reaction of stage A products and hydrogen peroxide in the Phi-TEC II experiment.
Table 5. Thermal characteristic parameters of the first decomposition reaction of stage A products and hydrogen peroxide in the Phi-TEC II experiment.
Thermal Characteristic ParametersStage A ProductsSample 3
T 0 , d , 1 (°C)50.0341.27
T m a x , 1 (°C)93.190.44
T a d , d , l (°C)43.0749.17
d T d t m a x , 1 (°C/min)0.182.97
φ 3.413.28
( T a d , d , 1 ) correct (°C)146.87161.28
( d T d t m a x , 1 ) correct (°C)0.629.74
Table 6. The MF of the adipic acid synthesis process in stages A and B.
Table 6. The MF of the adipic acid synthesis process in stages A and B.
Working Conditions T onset (°C) I Tonset H d (J/g) m T m a x (°C/min) T a d (°C) M P D (W/mL) I M P D MF
Stage A106.82451.359.74161.2031.4721.25
Stage B83.543422.980.62146.871.9011.188
Table 7. The probability of runaway adipic acid synthesis process in stages A and B.
Table 7. The probability of runaway adipic acid synthesis process in stages A and B.
Working Conditions T p (°C) M T T (°C) M T S R (°C) T D 24 (°C) I C C (°C) T M R a d (h) I T M R P
Stage A40–7873–75192.3327.9550.58510
Stage B90–96100161.953003>2414
Table 8. The severity of runaway adipic acid synthesis process in stages A and B.
Table 8. The severity of runaway adipic acid synthesis process in stages A and B.
Working Conditions H r x (kJ/kg) I H , r x T a d , r x (°C) I T a d , r x H r , d e c (kJ/kg) I H , d e c T a d , d e c (°C) I T a d , d e c
Stage A292.832122.822422.983161.202
Stage B172.84272.520101
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Y.; Ni, L.; Yao, H.; Su, J.; Cheng, Z. The Effects of Acid and Hydrogen Peroxide Stabilizer on the Thermal Hazard of Adipic Acid Green Synthesis. Sustainability 2023, 15, 6530. https://doi.org/10.3390/su15086530

AMA Style

Liu Y, Ni L, Yao H, Su J, Cheng Z. The Effects of Acid and Hydrogen Peroxide Stabilizer on the Thermal Hazard of Adipic Acid Green Synthesis. Sustainability. 2023; 15(8):6530. https://doi.org/10.3390/su15086530

Chicago/Turabian Style

Liu, Yinshan, Lei Ni, Hang Yao, Jimi Su, and Zhen Cheng. 2023. "The Effects of Acid and Hydrogen Peroxide Stabilizer on the Thermal Hazard of Adipic Acid Green Synthesis" Sustainability 15, no. 8: 6530. https://doi.org/10.3390/su15086530

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop