Next Article in Journal
How to Train an Artificial Neural Network to Predict Higher Heating Values of Biofuel
Next Article in Special Issue
CFD-DEM Simulation of Particle Fluidization Behavior and Glycerol Gasification in a Supercritical Water Fluidized Bed
Previous Article in Journal
Molecular Dynamics Simulation of Thermophysical Properties and the Microstructure of Na2CO3 Heat Storage Materials
Previous Article in Special Issue
Study on Two-Phase Permeation of Oxygen and Electrolyte in Lithium Air Battery Electrode Based on Digital Twin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Hydrogen Production by Supercritical Water Gasification of Unsymmetrical Dimethylhydrazine under Multi-Parameters

1
International Institute for Innovation, Jiangxi University of Science and Technology, Ganzhou 341000, China
2
State Key Laboratory of Multiphase Flow in Power Engineering (SKLMF), Xi’an Jiaotong University, Xi’an 710049, China
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(19), 7081; https://doi.org/10.3390/en15197081
Submission received: 25 August 2022 / Revised: 12 September 2022 / Accepted: 14 September 2022 / Published: 27 September 2022
(This article belongs to the Special Issue Challenges and Research Trends of Multiphase Flow)

Abstract

:
Unsymmetrical dimethylhydrazine (UDMH) is very toxic and hard to decompose in traditional ways. In this paper, the gasification of unsymmetrical dimethylhydrazine (UDMH) in supercritical water was studied in a batch reactor under different conditions. The hydrogen production process of supercritical water gasification of UDMH in metal containers is a multiphase reaction process. The effects of reaction temperature, alkaline catalysts, residence time, and oxidation on gasification were systematically studied. COD and ammonia nitrogen of the residual liquid were tested. Results showed that the maximum molar fraction and yield of hydrogen were 87.0% and 97.9 mol/kg, respectively, with KOH at 600 °C, 23 MPa. The COD removal efficiency in relation to alkaline catalysts was in the following order: NaOH > Na2CO3 > KOH > K2CO3. The highest COD removal efficiency (up to 95%) can be obtained at the temperature of 600 °C, 23 MPa with NaOH as the catalyst, and a residence time of 20 min. Ammonia nitrogen can be decreased by adding an oxidant. The COD and ammonia nitrogen of the residual liquid can meet the requirement of the Chinese emission standard of water pollution for space propellants. In addition, the organic compounds formed under different conditions were also identified.

1. Introduction

Unsymmetrical dimethylhydrazine (UDMH) combined with dinitrogen tetroxide (N2O4) serves as a liquid propellant that was employed for rocket fuel (CZ-2, CZ-3, and CZ-4) and missiles in China. Although it owns the advantages of high specific impulses, high thrust levels, better thrust control [1], and easy storage at reasonable temperatures and pressures, UDMH has gradually been substituted with other propellants due to its high toxicity. It has a high potential to contaminate the environment and do great harm to human beings [2,3,4]. Nowadays, there is still a large amount of UDMH wastewater that needs to be processed in an environmentally friendly way.
In our last article [5], we proposed that supercritical water gasification (SCWG) technology could be an alternative to degrade wastewater containing UDMH, with many advantages such as full degradation in a short time, no carcinogen and secondary pollution formed, etc. The properties of supercritical water are quite different from those of liquid water under ambient conditions. The number and persistence of hydrogen bonds are both diminished, and the dielectric constant is much lower, which is one of the reasons why organic wastes enjoy complete miscibility in supercritical water. On the other hand, the dissociation constant (Kw) for supercritical water is much higher than it is for ambient liquid water, which makes it an ideal solvent for organic compounds [6,7,8,9,10].
Both resource utilization and harmless treatment were realized in our last work. What is more, a reaction pathway and kinetics were also proposed based on the results obtained in the quartz reactor. However, in the process of industrialization, UDMH wastewater is processed in a metallic reactor. Hence, a study on the behavior of UDMH wastewater processed in the reactor with a metallic wall should be conducted.
As a matter of fact, research concerning the effect of the metallic wall on the chemical reaction in supercritical fluid has been studied by many researchers. In P. Christian’s group [11], they implemented an experiment on the radical dispersion polymerization of methyl methacrylate in supercritical carbon dioxide. They found that the exposed metal surface appears to terminate the growing polymer and inhibit polymer formation. This might be related to the sensitivity of the free radical polymerization of methyl methacrylate in supercritical CO2 to the condition of the internal surface of the high-pressure autoclave. When they control the initiator concentration and deposit coating on the walls of the autoclave, the coating appears to prevent metal termination. Yu et al. [12] studied supercritical water gasification of glucose at 600 °C, 34.5 MPa, a reaction time of 30 s with different reactors (Hastelloy and Inconel reactor). They found that when the reactant concentration is 0.8 M, the gasification efficiency is 85% in the new Hastelloy reactor while 68% in the Inconel reactor, corresponding with different catalytic mechanisms of different metal reactors. Zhu et al. [13] investigated the supercritical gasification of glycerol and glucose in a quartz reactor made of SiO2 and a tubular reactor made of Hastelloy C-276. They found that the gasification of the feedstocks could be promoted by the catalytic effects of the metallic reactor wall. All these experiments demonstrate that the metallic wall of the reactor has a certain effect on the gasification of the reactants in supercritical water.
In this paper, the gasification of UDMH in supercritical water was systematically studied in a batch reactor under different conditions. The effect of different parameters including reaction temperature, alkaline catalysts, residence time, and oxidation on gasification was discussed. We hope that our results could provide some references for industrialization.

2. Experiment

2.1. Apparatus and Methods

The simulated UDMH wastewater is a mixture of UDMH (98 wt%) and a certain amount of deionized water. The mass ratio of the UDMH to water is 1:999. The COD concentration of the simulated UDMH is 1213 mg/L in accordance with that of UDMH wastewater, which was obtained from the Sixth Research Institute of China Aerospace Science and Technology Corporation. The reagents KOH, K2CO3, NaOH, and Na2CO3 were AR grade purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).
The experiment was carried out in a batch reactor which is designed for temperatures up to 750 °C and pressure up to 30 MPa. The reactor is made of Inconel 625 and the volume of the reactor chamber is 10 mL. More details about the batch reactor can be found in [14].
Before each experiment, UDMH (98 wt%) was diluted with deionized water to the desired concentration and 1.6 g of the simulated wastewater was loaded into the reactor. The reactor was placed into an electric furnace after it had been sealed. Before heating, the sealed chamber was swept out of the air with high purity argon three times and pressurized to about 5 MPa. The initial pressure could change depending on the final temperature. Then, the reactor was heated from ambient temperature to the desired temperature with a heating rate of about 70 K/min in the electric furnace [14]. When the temperature reached the desired value, the corresponding pressure was 23 MPa. After a certain reaction time, the reactor was firstly cooled down to 300 °C by the air and then by water to the ambient temperature. The gaseous product was collected with a gas bag and the gas composition was tested by the gas chromatograph. The gas yield was measured by a wet gas flow meter. Each experiment under the same condition was repeated at least three times and the average value was calculated.

2.2. Analysis Methods

The composition of the gaseous product was detected by Agilent 7890A gas chromatograph (GC) equipped with a thermal conductivity detector (TCD) and capillary column C-2000 purchased from Lanzhou Institute of Chemical Physics in China. The carrier gas was argon (99.999%) with a flow rate of 5 mL/min. A standard gas mixture of H2, CO, CO2, and CH4 was used for calibration. COD and ammonia nitrogen concentrations were quantitatively analyzed by a multi-parameter controller (Lov-ibond-ET99722). The liquid residual was extracted by dichloromethane and qualitatively detected by gas chromatography/mass spectrometry (GC/MS, Agilent 6890A-5973N MSD).

3. Results and Discussion

3.1. Effect of Temperature

The thermodynamic equilibrium model based on the minimum Gibbs free energy methodology (using the RGIBBS block) was calculated with Aspen software (all values were obtained at 450 °C, 23 MPa) to predict the yield of different gaseous products. This method also has been employed by other researchers [15,16,17]. The result calculated by the RGIBBS block of ASPEN is shown in Table 1. The state of equilibrium can only be achieved under ideal conditions. Actually, our goal is to find a way to approach the equilibrium state. A small amount of methane, carbon monoxide, and ammonia, which were at least two orders of magnitude lower than nitrogen, appeared in the calculated result and was omitted in Equation (1). The equation is as follows:
C 2 H 8 N 2 + 4 H 2 O 8 H 2 + 2 CO 2 + N 2       Δ H = 172.1 kJ / mol
However, the gas obtained from our experiment contained not only H2 and CO2, but also CO and CH4. The deviation indicated that the reaction under our conditions was far from chemical equilibrium. As a matter of fact, once the gaseous products were generated, reactions between different gases may occur [18]. Relative reactions are below:
CO + H 2 O CO 2 + H 2       Δ H = 34.5 kJ / mol
CO + 3 H 2 CH 4 + H 2 O             Δ H = 226.5 kJ / mol  
The effect of temperature on UDMH gasification in supercritical water (23 MPa, residence time 20 min) is shown in Figure 1. As can be seen, the COD removal efficiency increased from 90.3% to 96.0% as the temperature increased from 450 °C to 600 °C. The gas yield increased from 32.2 mol/kg to 42.4 mol/kg as the temperature elevated from 450 °C to 600 °C, in which H2 accounted for 56.5% to 65.8%. The YH2 and the molar fraction of hydrogen increased from 19.2 mol/kg, 58.1 % to 28.6 mol/kg, 66.1%, respectively, indicating that high temperature favored the degradation of UDMH. From Equation (1), it could be seen that the reaction of UDMH with supercritical water is endothermic. Hence, increasing temperature could promote the degradation of UDMH and it is an important way to increase hydrogen production because every one-unit consumption of UDMH can obtain eight-unit hydrogen. A theory proposed by Bühler’s group [19] is that there are two competing reaction pathways in supercritical water: the preference for the free radical reaction at higher temperature and lower densities and the preference for the ionic reaction at lower temperature and higher densities. When the temperature is above the critical point, the decrease in water density causes the drop of ion product, indicating that the free-radical mechanism is preferable [20]. On the other hand, gas is formed by a free-radical reaction [21]. As a result, high temperatures can promote the degradation of UDMH and gas yield. However, from the point of view of industrialization, the energy loss will increase with temperature, but decrease with volume [22]. Therefore, catalyst deserves to be added to improve the gas yield and carbon efficiency.

3.2. Effect of Catalyst

In theory, catalysts can promote the reaction rate but cannot change the final equilibrium state [23,24]. In Equation (1), it could be seen that the main composition of the gaseous products is hydrogen. It is reported that the addition of alkali catalysts can drive the molar fraction of gaseous products closer to the equilibrium state [25] and promote hydrogen production [26]. So, the effect of KOH, K2CO3, NaOH, and Na2CO3 on the UDMH wastewater gasification was performed at 550 °C, residence a time of 20 min, 23 MPa. The mass of the catalysts added was equal to the UDMH. The results are shown in Figure 2. As can be seen, the total gas yield was increased to varying degrees. The maximal gas yield was obtained with KOH and the maximal CE with Na2CO3. According to Figure 2, the effect of gasification is promoted with different catalysts. With the presence of KOH, the molar fraction of H2 and YH2 were 87.0% and 97.9 mol/kg, increasing by 30.4% and 258.6%, respectively, and the total gas yield was 112.5 mol/kg, increasing by 174.4%. It was unpredicted that CE was enhanced with Na2CO3 and K2CO3, which seemed to be inconsistent with other literature [25,27]. They believed that alkaline catalysts had an insignificant influence on carbon gasification efficiency. This phenomenon could be explained from two aspects: firstly, the difference in waste concentration, which is at least 4 wt% in their works while 0.1 wt% in this work, may result in catalysts having a greater influence on the gaseous products. Secondly, the carbonate anion might be decomposed and be dedicated to the increase in CE. As a result, catalysts may have a greater influence on gaseous products. As we can see from Figure 2a that CE was promoted in the presence of carbonate. It is generally accepted that water–gas shift reaction can be accelerated by alkali metal [18,27,28,29], which can explain the decline in CO molar fraction and the increase in H2 molar fraction. CH4 molar fraction declined, indicating the reverse Equation (3) was promoted. To sum up, considering all the factors above, the alkaline catalyst can increase gas yield and has a better selectivity for hydrogen.

3.3. Effect of Residence Time

The effect of residence time was studied at 650 °C, 23 MPa. The result is shown in Figure 3. It could be seen in Figure 3a that the total gas yield is significantly enhanced with the prolonging of the reaction time. The gas yield obtained at 30 min is 2.6 times as much as that obtained at 0 min, illustrating that reaction time is an important factor affecting the gas yield. The amount of CO2 decreases with the addition of KOH and NaOH because CO2 is absorbed by basic catalysts [29]. The COD removal efficiency was increased by prolonging the residence time. It also should be noted that the COD removal efficiency had already been above 90% when the residence time was 0 min and it increased only by 1% as the residence time increased every 10 min. The reason is that the temperature was high enough to degrade the UDMH wastewater and it was difficult to go further. The hydrogen molar fraction increased from 61.1% to 71.1%, increasing by 16.4% and the YH2 increased from 16.5 mol/kg to 51.0 mol/kg, increasing by 209.1% when the residence time increased from 0 min to 30 min.
Compared with the chemical equilibrium mentioned above, we concluded that the distribution of different gases was approaching the thermodynamic equilibrium with the increase in the residence time. However, the ammonia nitrogen concentration was still too high to be discharged into the environment.

3.4. Effect of Oxidant and ER

In order to decrease the concentration of ammonia nitrogen and obtain the optimum amount of oxidant, different amounts of oxidant were added. Oxidation of UDMH has been implemented by many researchers [30,31,32,33] and many types of oxidants [30,32,34] have been attempted to realize effective degradation and safe discharge.
The effect of oxidant and ER on UDMH gasification was studied at different temperatures, 23 MPa, with a residence time of 20 min. The result is shown in Figure 4.
When oxidant is added, the following reactions may occur:
2 C 2 H 8 N 2 + 3 H 2 O 2 8 H 2 + 2 CO 2 + 2 CO + 2 NH 3 + N 2  
2 H 2 + O 2 2 H 2 O       Δ H = 506.3   kJ / mol  
CH 4 + 1 / 2 O 2 CO + 2 H 2         Δ H = 23.4   kJ / mol  
CO + 1 / 2 O 2 CO 2       Δ H = 284.6   kJ / mol  
2 H 2 O 2 2 H 2 O + O 2       Δ H = 168.9   kJ / mol  
4 NH 3 + 3 O 2 6 H 2 O + 2 N 2       Δ H = 1284.1   kJ / mol  
In this paper, hydrogen peroxide was used as the oxidant. As can be seen, CE in the presence of oxidant is higher. The yield and molar fraction of H2 and CH4 tends to decrease when ER increases from 0.2 to 0.4. This result is predictable because H2 and CH4 yields at the chemical equilibrium state decrease with the increasing amounts of oxygen addition [35]. On the other hand, most of the nitrogen appears as N2 in the reaction products [6]. It could be assumed that the declined hydrogen yield resulted from the oxidation of reduced hydrogen atoms. Meanwhile, more carbon dioxide could be produced by the oxidation of carbon monoxide and other intermediate products containing carbon. It should be noted that the ammonia nitrogen of the residual liquid was decreased in the presence of the oxidant. When ER is 0.4, ammonia nitrogen of the residual liquid can meet the requirement of the Chinese emission standard of water pollution for space propellant (GB14374-93), which is 25 mg/L.

3.5. Reaction Mechanism

To have a deeper understanding of the degradation mechanism of UDMH processed in supercritical water in the batch reactor, GC/MS analysis was implemented. Considering that the residual liquid with low organic concentration may not be detectable comprehensively, UDMH solution with a concentration of 50 wt% was loaded into the reactor. Relative percentage content at a reaction time of 1 min and a certain temperature was used to investigate the major organic intermediate products. Through the trend of the relative percentage content at different temperatures (400 °C, 450 °C, and 500 °C), we could deduce the reaction mechanism and determine which products are hard to be degraded. Table 2 shows the identified organic compounds along with their detected information. Figure 5 gives the photo comparison of residual liquid extracted by dichloromethane at different temperatures.
In general, the color becomes lighter as the temperature increases, indicating that the residual liquid contains fewer organic compounds. Trimethylamine (TMA) accounts for a major part of all three temperatures and has a downward trend with the increase in temperature, illustrating that high temperature can promote the degradation of TMA. N-containing heterocyclic compounds have an increasing proportion and seem hard to degrade even at high temperatures. These compounds including pyridine, indole, and carbazole are essentially unreactive under supercritical water conditions [36]. The addition of formic acid or sodium formate results in the degradation of the aromatic rings [6]. The reason why the result is different from our previous work may result from the difference in the concentration of reactants. From the micro perspective, high reactant concentration leads to a high collision probability of UDMH molecules, thus the probability of forming a heterocyclic molecule is higher.

4. Conclusions

Supercritical water gasification technology used for processing UDMH wastewater proved to be a harmless, resourceful utilization technology. In this paper, gasification of unsymmetrical dimethylhydrazine in supercritical water was studied in the range of 450 °C to 650 °C, 23 MPa for the first time. The main conclusions could be drawn as:
(1)
Gas efficiency can be promoted with the addition of an alkaline catalyst and KOH owned the highest gas efficiency. The maximum molar fraction and yield of hydrogen reached 87.0% and 97.9 mol/kg, respectively, with KOH at 550 °C, 23 MPa, and a residence time of 20 min.
(2)
The COD removal efficiency reached 95% at 600 °C, 23 MPa, no catalyst added, and a residence time of 30 min. When ER was 0.4, the ammonia nitrogen concentration of the residual liquid can meet the requirement of the Chinese emission standard of water pollution for space propellant, which is 25 mg/L.
(3)
When the concentration of UDMH solution is up to 50 wt%, the major degradation product under low temperature is trimethylamine (TMA) whose relative amount has a reduction tendency as the temperature increases.

Author Contributions

Conceptualization, L.Y. and J.C.; methodology, Z.L. (Zhigang Liu); validation, H.C. and D.L.; formal analysis, Z.L. (Zheng Liu); investigation, L.Y.; writing—original draft preparation, L.Y.; writing—review and editing, B.C.; supervision, B.C.; project administration, L.Y.; funding acquisition, L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Jiangxi University of Science and Technology, grant number 205200100572.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CEthe mass of carbon in gaseous product/the mass of carbon in UDMH ×100, %
YH2the molar number of produced hydrogen/the mass of UDMH, mol/kg
Gas fractionthe molar number of a certain gas product/the summation of the molar number of all the gaseous product ×100, %
ERthe mass of oxidation added/the mass of oxidant for complete oxidation of UDMH by stoichiometry calculation

References

  1. Salvador, C.A.V.; Costa, F.S. Vaporization lengths of hydrazine fuels burning with NTO. J. Propuls. Power 2006, 22, 1362–1372. [Google Scholar] [CrossRef]
  2. Fedorov, L.A. Liquid missile propellants in the former Soviet Union. Environ. Pollut. 1999, 105, 157–161. [Google Scholar]
  3. Tamura, T.; Shibutani, M.; Toyoda, K.; Shoda, T.; Takada, K.-I.; Uneyama, C.; Takahashi, M.; Hirose, M. Tumor-promoting activities of hydroquinone and 1,1-dimethylhydrazine after initiation of newborn mice with 1-methyl-1-nitrosourea. Cancer Lett. 1999, 143, 71–80. [Google Scholar] [CrossRef]
  4. Carlsen, L.; Kenesova, O.A.; Batyrbekova, S.E. A preliminary assessment of the potential environmental and human health impact of unsymmetrical dimethylhydrazine as a result of space activities. Chemosphere 2007, 67, 1108–1116. [Google Scholar] [CrossRef]
  5. Yi, L.; Guo, L.; Jin, H.; Kou, J.; Zhang, D.; Wang, R. Gasification of unsymmetrical dimethylhydrazine in supercritical water: Reaction pathway and kinetics. Int. J. Hydrogen Energy 2018, 43, 8644–8654. [Google Scholar] [CrossRef]
  6. Savage, P.E. Organic Chemical Reactions in Supercritical Water. Chem. Rev. 1999, 99, 603–622. [Google Scholar] [CrossRef]
  7. Watanabe, M.; Sato, T.; Inomata, H.; Smith, R.L.; Arai, K.; Kruse, A.; Dinjus, E. Chemical reactions of C-1 compounds in near-critical and supercritical water. Chem. Rev. 2004, 104, 5803–5821. [Google Scholar] [CrossRef]
  8. Loppinet-Serani, A.; Aymonier, C.; Cansell, F. Current and Foreseeable Applications of Supercritical Water for Energy and the Environment. Chemsuschem 2008, 1, 486–503. [Google Scholar] [CrossRef]
  9. Guo, L.J.; Jin, H. Boiling coal in water: Hydrogen production and power generation system with zero net CO2 emission based on coal and supercritical water gasification. Int. J. Hydrog. Energy 2013, 38, 12953–12967. [Google Scholar] [CrossRef]
  10. Guo, L.; Jin, H.; Ge, Z.; Lu, Y.; Cao, C. Industrialization prospects for hydrogen production by coal gasification in supercritical water and novel thermodynamic cycle power generation system with no pollution emission. Sci. China-Technol. Sci. 2015, 58, 1989–2002. [Google Scholar] [CrossRef]
  11. Christian, P.; Giles, M.R.; Howdle, S.M.; Major, R.C.; Hay, J.N. The wall effect: How metal/radical interactions can affect polymerisations in supercritical carbon dioxide. Polymer 2000, 41, 1251–1256. [Google Scholar] [CrossRef]
  12. Yu, D.H.; Aihara, M.; Antal, M.J. Hydrogen-Production by Steam Reforming Glucose In Supercritical Water. Energy Fuels 1993, 7, 574–577. [Google Scholar] [CrossRef]
  13. Zhu, C.; Wang, R.; Jin, H.; Lian, X.; Guo, L.; Huang, J. Supercritical water gasification of glycerol and glucose in different reactors: The effect of metal wall. Int. J. Hydrogen Energy 2016, 41, 16002–16008. [Google Scholar] [CrossRef]
  14. Lan, R.; Jin, H.; Guo, L.; Ge, Z.; Guo, S.; Zhang, X. Hydrogen Production by Catalytic Gasification of Coal in Supercritical Water. Energy Fuels 2014, 28, 6911–6917. [Google Scholar] [CrossRef]
  15. Dileo, G.J.; Neff, M.E.; Savage, P.E. Gasification of guaiacol and phenol in supercritical water. Energy Fuels 2007, 21, 2340–2345. [Google Scholar] [CrossRef]
  16. DiLeo, G.J.; Savage, P.E. Catalysis during methanol gasification in supercritical water. J. Supercrit. Fluids 2006, 39, 228–232. [Google Scholar] [CrossRef]
  17. Resende, F.L.P.; Savage, P.E. Expanded and Updated Results for Supercritical Water Gasification of Cellulose and Lignin in Metal-Free Reactors. Energy Fuels 2009, 23, 6213–6221. [Google Scholar] [CrossRef]
  18. Guo, S.; Guo, L.; Yin, J.; Jin, H. Supercritical water gasification of glycerol: Intermediates and kinetics. J. Supercrit. Fluids 2013, 78, 95–102. [Google Scholar] [CrossRef]
  19. Bühler, W.; Dinjus, E.; Ederer, H.; Kruse, A.; Mas, C. Ionic reactions and pyrolysis of glycerol as competing reaction pathways in near- and supercritical water. J. Supercrit. Fluids 2002, 22, 37–53. [Google Scholar] [CrossRef]
  20. Guo, Y.; Wang, S.; Xu, D.; Gong, Y.; Ma, H.; Tang, X. Review of catalytic supercritical water gasification for hydrogen production from biomass. Renew. Sustain. Energy Rev. 2010, 14, 334–343. [Google Scholar] [CrossRef]
  21. Lu, Y.; Jin, H.; Guo, L.; Zhang, X.; Cao, C.; Guo, X. Hydrogen production by biomass gasification in supercritical water with a fluidized bed reactor. Int. J. Hydrogen Energy 2008, 33, 6066–6075. [Google Scholar] [CrossRef]
  22. Kruse, A. Hydrothermal biomass gasification. J. Supercrit. Fluids 2009, 47, 391–399. [Google Scholar] [CrossRef]
  23. Tang, H.Q.; Kitagawa, K. Supercritical water gasification of biomass: Thermodynamic analysis with direct Gibbs free energy minimization. Chem. Eng. J. 2005, 106, 261–267. [Google Scholar] [CrossRef]
  24. Cao, C.; Guo, L.; Jin, H.; Guo, S.; Lu, Y.; Zhang, X. The influence of alkali precipitation on supercritical water gasification of glucose and the alkali recovery in fluidized-bed reactor. Int. J. Hydrogen Energy 2013, 38, 13293–13299. [Google Scholar] [CrossRef]
  25. Guo, S.M.; Guo, L.; Cao, C.; Yin, J.; Lu, Y.; Zhang, X. Hydrogen production from glycerol by supercritical water gasification in a continuous flow tubular reactor. Int. J. Hydrogen Energy 2012, 37, 5559–5568. [Google Scholar] [CrossRef]
  26. Weijin, G.; Binbin, L.; Qingyu, W.; Zuohua, H.; Liang, Z. Supercritical water gasification of landfill leachate for hydrogen production in the presence and absence of alkali catalyst. Waste Manag. 2018, 73, 439. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, Y.; Guo, L.; Cao, W.; Jin, H.; Guo, S.; Zhang, X. Hydrogen production by sewage sludge gasification in supercritical water with a fluidized bed reactor. Int. J. Hydrogen Energy 2013, 38, 12991–12999. [Google Scholar] [CrossRef]
  28. Kruse, A. Supercritical water gasification. Biofuels Bioprod. Biorefining 2008, 2, 415–437. [Google Scholar] [CrossRef]
  29. Jin, H.; Lu, Y.; Guo, L.; Zhang, X.; Pei, A. Hydrogen Production by Supercritical Water Gasification of Biomass with Homogeneous and Heterogeneous Catalyst. Adv. Condens. Matter Phys. 2014, 2014, 160565. [Google Scholar] [CrossRef]
  30. Lunn, G.; Sansone, E.B. Oxidation of 1,1-dimethylhydrazine (UDMH) in aqueous solution with air and hydrogen peroxide. Chemosphere 1994, 29, 1577–1590. [Google Scholar] [CrossRef]
  31. Tuazon, E.C.; Carter, W.P.L.; Winer, A.M.; Pitts, J.N. Reactions of hydrazines with ozone under simulated atmospheric conditions. Environ. Sci. Technol. 1981, 15, 823–828. [Google Scholar] [CrossRef]
  32. Pestunova, O.P.; Elizarova, G.L.; Ismagilov, Z.R.; Kerzhentsev, M.A.; Parmon, V.N. Detoxication of water containing 1, 1-dimethylhydrazine by catalytic oxidation with dioxygen and hydrogen peroxide over Cu-and Fe-containing catalysts. Catal. Today 2002, 75, 219–225. [Google Scholar] [CrossRef]
  33. Mathur, M.A.; Sisler, H.H. Oxidation of 1,1-dimethylhydrazine by oxygen. Inorg. Chem. 1981, 20, 426–429. [Google Scholar] [CrossRef]
  34. Abilev, M.; Kenessov, B.; Batyrbekova, S.; Grotenhuis, T. Chemical oxidation of unsymmetrical dimethylhydrazine transformation products in water. Prog. Brain Res. 2015, 148, 321–328. [Google Scholar] [CrossRef]
  35. Guo, L.J.; Lu, Y.J.; Zhang, X.M.; Ji, C.M.; Guan, Y.; Pei, A.X. Hydrogen production by biomass gasification in supercritical water: A systematic experimental and analytical study. Catal. Today 2007, 129, 275–286. [Google Scholar] [CrossRef]
  36. Katritzky, A.R.; Shipkova, P.A.; Allin, S.M.; Barcock, R.A.; Siskin, M.; Olmstead, W.N. Aqueous High-Temperature Chemistry. 24. Nitrogen-Containing Heterocycles in Supercritical Water at 460 °C. Energy Fuels 1995, 9, 580–589. [Google Scholar] [CrossRef]
Figure 1. Effect of temperature on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (23 MPa, residence time 20 min).
Figure 1. Effect of temperature on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (23 MPa, residence time 20 min).
Energies 15 07081 g001
Figure 2. Effect of catalyst on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (550 °C, 23 MPa, residence time 20 min).
Figure 2. Effect of catalyst on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (550 °C, 23 MPa, residence time 20 min).
Energies 15 07081 g002
Figure 3. Effect of residence time on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (650 °C, 23 MPa).
Figure 3. Effect of residence time on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (650 °C, 23 MPa).
Energies 15 07081 g003
Figure 4. Effect of oxidant and ER on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (23 MPa, residence time 20 min).
Figure 4. Effect of oxidant and ER on gasification characteristic: (a) gas yield and CE (b) gas fraction (c) COD removal efficiency and ammonia nitrogen (23 MPa, residence time 20 min).
Energies 15 07081 g004
Figure 5. Residual liquid of 400 °C, 450 °C, and 500 °C (from left to right) extracted by dichloromethane.
Figure 5. Residual liquid of 400 °C, 450 °C, and 500 °C (from left to right) extracted by dichloromethane.
Energies 15 07081 g005
Table 1. Product distribution, calculated by RGIBBS block of ASPEN.
Table 1. Product distribution, calculated by RGIBBS block of ASPEN.
ComponentFeedProduct
UDMH0.00029997 10
H2O0.99970.9967211
H200.00237643
CO02.98 × 10−7
CH401.59 × 10−6
CO200.000596985
C2H403.77 × 10−18
C2H606.62 × 10−15
N200.000295299
NH308.27 × 10−6
N2O01.22 × 10−26
NO201.08 × 10−34
1 The unit of the value in this table is mole fraction.
Table 2. Identities and relative amounts of compounds in residual liquid.
Table 2. Identities and relative amounts of compounds in residual liquid.
RT (min)NameChemical FormulaStructural Formula400 °C450 °C500 °C
Area %
1.430TrimethylamineC3H9N Energies 15 07081 i00135.18416.39617.477
1.527Ethanamine, N,N-dimethyl-C4H11N Energies 15 07081 i0024.4111.1332.504
2.168Aziridine, 1,2,3-trimethyl-, trans-C5H11N Energies 15 07081 i0032.753--
5.465PropanenitrileC3H5N Energies 15 07081 i0041.0161.7844.049
5.642Propanal, 2-methyl-, dimethylhydrazoneC6H14N2 Energies 15 07081 i0051.0988.1281.327
7.071Butanenitrile, 2-methyl-C5H9N Energies 15 07081 i0060.7640.8260.466
8.9631H-Pyrrole, 1-methyl-C5H7N Energies 15 07081 i0072.1181.060-
10.272PyridineC5H5N Energies 15 07081 i0081.3481.5790.542
11.6261H-Pyrrole, 2,5-dimethyl-C6H9N Energies 15 07081 i0092.1861.7542.262
13.1241H-Pyrazole, 1-methyl-C4H6N2 Energies 15 07081 i0108.7788.2425.871
14.627Pyridine,3-methyl-C6H7N Energies 15 07081 i0118.65911.19114.148
17.130Pyrazole, 1,4-dimethyl-C5H8N2 Energies 15 07081 i01212.79010.6535.218
19.931Pyridine, 3,5-dimethyl-C7H9N Energies 15 07081 i0138.98313.13513.581
33.9561H-Pyrrole, 2,3,4,5-tetramethyl-C8H13N Energies 15 07081 i0149.91324.11732.555
“-”: not detect.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yi, L.; Chen, J.; Liu, Z.; Chen, H.; Liu, D.; Liu, Z.; Chen, B. Study on Hydrogen Production by Supercritical Water Gasification of Unsymmetrical Dimethylhydrazine under Multi-Parameters. Energies 2022, 15, 7081. https://doi.org/10.3390/en15197081

AMA Style

Yi L, Chen J, Liu Z, Chen H, Liu D, Liu Z, Chen B. Study on Hydrogen Production by Supercritical Water Gasification of Unsymmetrical Dimethylhydrazine under Multi-Parameters. Energies. 2022; 15(19):7081. https://doi.org/10.3390/en15197081

Chicago/Turabian Style

Yi, Lei, Jingwei Chen, Zhigang Liu, Huiming Chen, Daoxiu Liu, Zheng Liu, and Bin Chen. 2022. "Study on Hydrogen Production by Supercritical Water Gasification of Unsymmetrical Dimethylhydrazine under Multi-Parameters" Energies 15, no. 19: 7081. https://doi.org/10.3390/en15197081

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