Next Article in Journal
Mathematical Modelling and CFD Simulation for Oxygen Removal in a Multi-Function Gas-Liquid Contactor
Previous Article in Journal
Geochemical Characteristics and Development Model of the Coal-Measure Source Rock in the Kuqa Depression of Tarim Basin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Superheated Surface on the Deposition Behavior of Na2SO4 in Supercritical Water

1
State Key Laboratory of Chemical Engineering, School of Chemical Engineering, East China University of Science and Technology, Shanghai 200237, China
2
Dalian Research Institute of Petroleum and Petrochemicals, SINOPEC, Dalian 116000, China
*
Author to whom correspondence should be addressed.
Processes 2023, 11(6), 1779; https://doi.org/10.3390/pr11061779
Submission received: 27 April 2023 / Revised: 5 June 2023 / Accepted: 9 June 2023 / Published: 11 June 2023
(This article belongs to the Section Environmental and Green Processes)

Abstract

:
The reduced solubility of inorganic salts in supercritical water has a significant impact on the stable operation of desalination facilities as it may lead to surface fouling due to salt deposition. In this study, the solubility of Na2SO4 was experimentally determined to be 0.04–15.34 mmol/kg water at 23–25 MPa and 390–420 °C. To investigate the precipitation behavior of Na2SO4 in supercritical water, a reactor with a heating bar was designed and the deposition effect of salt on the superheated surface in an autoclave was tested at a temperature of 390 °C and a pressure of 23 MPa. Then, the deposition mechanism of salt in the autoclave was analyzed and the temperature field in the reactor was simulated using CFD commercial software. The experimental results showed that Na2SO4 was present on both the heating rod and the bottom of the autoclave with a loose salt layer. The simulation results indicated that the temperature near the heating rod was significantly higher than the bulk fluid temperature and it provided the temperature condition where the inorganic salt preferentially nucleated and precipitated. Nickel foam was chosen as the porous media carrier to investigate the selective precipitation of salt on different superheated surfaces. The results showed that nickel foam could collect a large amount of salt on the heating rod and change the state of the compact salt layer on the kettle bottom wall, providing a technical choice for salt recovery under supercritical conditions.

1. Introduction

Supercritical water is a fluid with a temperature greater than 374 °C and a pressure greater than 22.1 MPa [1]. It is characterized by a lower dielectric constant and density compared to those of water at a normal temperature and pressure. Specifically, the dielectric constant reduces from 78 (T = 25 °C, P = 0.1 MPa) to 6 (at critical point), whereas the density decreases from 997 kg/m3 to 280 kg/m3 [2]. Therefore, it has properties similar to non-polar solvents, resulting in the enhanced solubility of inorganic substances and reduced solubility of inorganic salts [3]. Due to its unique properties, supercritical water has found widespread applications in numerous processes, including supercritical water oxidation (SCWO) [4,5,6,7], supercritical water desalination (SCWD) [8,9,10], supercritical water gasification (SCWG) [11,12,13], supercritical hydrothermal synthesis (SHS) [14,15,16], etc. SCWO [4] is a homogeneous oxidation reaction that uses supercritical water as the reaction medium and oxygen and air as the oxidant to rapidly convert organic matter into harmless small molecules such as carbon dioxide and water. SCWD [8] exploits the non-polar characteristic of supercritical water to separate salts from wastewater and achieve zero liquid discharge of saline wastewater. SHS is commonly used for the synthesis of nanomaterials with the advantages of fast reaction rates and high product crystallinity. The process of SHS is particularly promising for the generation of high-performance nanoparticles, which can be applied in various fields, such as enhancing the functionality of ion-selective electrodes [17].
Conventional methods for treating wastewater include membrane separation, biodegradation, chemical precipitation, and incineration [18,19,20]. However, these methods have limited application in treating high-salt organic wastewater. Membrane separation techniques, such as reverse osmosis and nanofiltration, are ineffective at removing organic matter from wastewater. Moreover, the membranes are susceptible to contamination during the treatment process, leading to reduced efficiency and service life and high costs [18]. High concentrations of salt in wastewater inhibit the survival of biological cells, thereby limiting the applicability of biological methods for large-scale wastewater treatment. Incineration methods are associated with secondary pollution [21]. In comparison to the traditional methods of treating high-salinity organic wastewater, SCWO technology has several advantages, such as a shorter reaction time, no secondary pollution, and environmental friendliness [21]. However, the solubility of inorganic salts is significantly reduced due to the decrease in the dielectric constant and density of supercritical water. For instance, the solubility of NaCl is only 1.44–87 mmol/kg at the temperature range of 380–400 °C and the pressure range of 18–23.5 MPa [22]. Similarly, the solubility of Na2SO4 decreases to 0.025–69.7 mmol/kg at a temperature range of 380–400 °C and a pressure of 25 MPa [23]. As a result, when the salt concentration in wastewater is high, inorganic salts precipitate and settle rapidly, forming a sedimentary layer that can block pipelines and interfere with heat transfer [24].
The application of SCWO has been limited due to issues of corrosion and blockage. These issues arise from the deposition of salts originating from the raw material solution or the oxidative decomposition of organic matter. To mitigate salt deposition during the SCWO process, the existing research has mainly focused on controlling salt concentration in the material to reduce deposition or on conducting one-step pretreatment to remove a portion of salt before entering the reactor [25]. Alternatively, reactor design can be employed to reduce salt deposition on the walls. Samuel et al. [8] designed a secondary separation process for the desalination of seawater, utilizing the non-polar properties of SCW to remove salt from brine in a separator.
In summary, reducing inorganic salt deposition in applications is typically achieved through reactor design and raw material pretreatment. However, methods to provide crystalline carriers have not been extensively studied. Dell’Orco et al. [26] added γ-Al2O3 spheres to provide the surface area for non-homogeneous precipitation when determining the solubility of NaNO3 at 24.8–30 MPa and 450–525 °C, but the effect of γ-Al2O3 spheres on inorganic salts was not described. Moreover, few studies have explored the deposition characteristics of inorganic salts in the reactor, although it is suggested that superheated surfaces are more likely to be preferential sites for salt nucleation to occur [27]. Porous metals may provide a better medium to induce directed salt deposition due to their low density, large specific surface area, and good thermal conductivity.
In this study, a reactor with an in-built suspended heating rod was designed. Taking typical type II Na2SO4 salt as a representative, the deposition of salt in the autoclave was tested at the temperature of 390 °C at 23 MPa, and the induction of deposition of the inorganic salt by nickel foam was also tested. The effect of gravity and different heating walls on salt crystallization was comprehensively investigated. Additionally, temperature and velocity field distributions in the autoclave were obtained through CFD simulations to analyze the results of salt deposition.

2. Experiment

2.1. Chemicals

Sodium sulfate (analytically pure, ≥99%) was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China), and deionized water was purchased from Shanghai Huazhen Technology Co., Ltd. (Shanghai, China). The details of the chemical reagents are listed in Table 1. Considering the corrosion effect under supercritical conditions, nickel foam was selected as the primary experimental material. In this study, two different sizes of nickel foam were used: 50 × 50 × 5 (mm3), and 25 × 25 × 5 (mm3), both with a porosity of 20 PPI (pores per inch).

2.2. Device Process

Figure 1 illustrates the process of the device, in which the autoclave is made of Inconel625 alloy to resist corrosion in the supercritical state. Figure 2 displays a diagram of the experimental setup. The prepared test material was placed at the bottom of the autoclave or fixed to the heating rod with nickel wire. A specific concentration of salt solution was prepared by dissolving a weighed amount of the sample salt in deionized water and pumping it into the preheater with a high-pressure pump (10–100 mL/min). After preheating to 200–300 °C, the solution entered the reactor, where the brine was heated to the supercritical temperature (390–450 °C) using an electric heating furnace. Under the supercritical condition, salt particles were precipitated from the supercritical water. The desalinated solution was filtered by a filter (SS10TF-MM-2, Swagelok, OH, USA), and cooled to ambient temperature by a condenser. The overall pressure of the plant was controlled at 23 MPa by the back pressure valve (R42LG-BBG-11-11-P, Amflo, Shanghai, China).
The device operation and data measurements were controlled and acquired, respectively, by a computer. The relationship between the concentration of the inorganic salt solution and the conductivity was described by a linear equation, and the salt concentration in the solution was calculated from the conductivity values. The relationship between salt concentration and conductivity can be expressed as follows:
S = 0.0055∙σ − 0.03826
where S is the salt concentration in mmol/L; σ is the conductivity in μS/cm. The correlation coefficient R2 = 0.99932 indicated that the relationship between concentration and conductivity was linear over the experimental concentration range. Figure 3 depicts the concentration of the Na2SO4 solution as a function of conductivity.
To investigate the impact of the superheated surface on salt deposition in the supercritical state, a heating rod (Figure 4) was designed for the autoclave, and the upper temperature of it could reach 600 °C. After the brine was injected for a certain time, when the conductivity value was maintained within a certain range, the solution was stopped. To avoid the residual gas in the kettle from redissolving the precipitated salt particles, a section of the nitrogen booster was placed at the front end of the reactor. An air compressor was used to pressurize nitrogen into the autoclave and bring out the residual water vapor. After depressurizing and waiting for the temperature to drop to a normal temperature, the carrier material was removed and the experimental phenomena were observed.

3. Model and Theory

In recent decades, various mathematical and numerical methods have been developed to study different supercritical fluid processes and analyze phenomena such as thermodynamics, hydrodynamics, crystal nucleation and growth, etc. These methods allow the prediction of nucleation times, temperature field distributions and supersaturation field distributions that are difficult to obtain experimentally. Computational fluid dynamics (CFD) software has been successfully applied to simulate flow, heat transfer and reactions under various supercritical conditions [28,29,30,31]. In a numerical simulation study of supercritical water desalination, Voisin et al. [32,33] used all the experimentally measured data as fitting parameters for numerical modeling and simulated the nucleation and growth of Na2SO4 particles under supercritical conditions, without considering crystal fragmentation and aggregation, and the simulation results were in good agreement with the experimental data. Therefore, based on previous studies, CFD software was selected to simulate the temperature field inside the reactor to provide a better illustration of the effect of the heating rod on the temperature field inside the reactor and salt deposition.

3.1. Physical Model

The reactor comprises inlet piping, outlet piping, internal heating rods and an external heating furnace. To simplify the model, the flow paths inside the reactor were modeled, and the heating rod was approximated to the inner wall surface. Similarly, the external heating furnace was simplified to the outer wall’s surface. Figure 5a,b shows the simplified physical model and the mesh of the model after meshing, respectively. The total number of cells was 872,842 and the specific information is summarized in Table 2 below.

3.2. Grid Sensitivity Test

The grid sensitivity test was performed at a brine flow rate of 10 mL/min and an inlet temperature of 350 °C, under the heating rod surface temperature of 600 °C. Five grid sizes were compared: 5 mm, 3 mm, 1.5 mm, 1 mm, and 0.5 mm.
From the simulation results (Figure 6), it can be seen that the grid size of 0.5 mm did not improve the accuracy over that of 1 mm, so the grid size of 1 mm was used in the simulation.

3.3. Fluid Dynamics

Numerical calculations were carried out using the commercial software ANSYS FLUENT. Under the conditions studied in this work, the pressure was above the critical pressure. To simplify the calculation, it was assumed that the water at the entrance vaporized instantaneously. The vaporization was too fast to absorb heat from the surrounding environment, it would lead to the cooling of itself to provide heat of vaporization, while converting itself into water vapor. Furthermore, it was assumed that the average temperature of the autoclave was 390 °C, and the state of water at the entrance (350 °C; 22.9 MPa) was liquid with the enthalpy of 1632 kJ/kg.
The inlet flow rates were set at 10, 20 and 50 mL/min, with an inlet temperature of 350 °C. The temperature of the heating rod was set at 600 °C and 900 °C, while the autoclave wall temperature was set at 350 °C. An implicit pressure-based solver was selected for the calculation and the coupling method of pressure and velocity was calculated using the SIMPLE algorithm. The residual convergence value was set at 10−3.
The continuity equation is defined as follows:
ρ t +   ρ v = 0
For turbulence modeling, the standard k-ε model with standard wall functions is used as a near-wall treatment. In this model, the turbulent kinetic energy, k, and its dissipation rate, ε, are obtained from the following transport equations:
t ρ k + x i ρ k u i = x j μ + μ t σ k k x j + G k + G b ρ ε Y M
t ρ ε + x i ρ ε u i = x j μ + μ t σ ε ε x j + c ε 1 ε k G k + c ε 3 G b c ε 2 ρ ε 2 k
where G k denotes the turbulent kinetic energy due to the mean velocity gradient, G b represents the turbulent kinetic energy due to buoyancy, and Y M represents the effect of fluctuating expansion in compressible turbulence on the turbulent kinetic energy. c ε 1 , c ε 2 and c ε 3 are constants with values of 1.44, 1.92 and 0.09, respectively; σ k and σ ε are Prandtl numbers with values of 1.0 and 1.3, respectively.
The energy balance (equation) was used to consider the effect of heat transfer on fluid dynamics.
t ρ E +   v ρ E + p =   k e f f T j h j J j + τ ̿ e f f v
where k e f f is the effective thermal conductivity and J j is the diffusive flux of the substance. The terms on the right-hand side indicate the energy transfer, diffusion and viscous dissipation caused by the conducting substance.

4. Results and Discussion

4.1. Solubility Data for Na2SO4

The solubility data for Na2SO4 in water at 380–420 °C and 23–25 MPa are shown in Table 3. The solubility range is 0.04–15.34 mmol/kg. The relationship between the experimental solubility data and temperature/pressure is illustrated in Figure 5.
The solubility of Na2SO4 in supercritical water gradually decreases with increasing temperature and decreasing pressure. As shown in Figure 7, the solubility of Na2SO4 under supercritical conditions is strongly influenced by temperature and decreases significantly with increasing temperature, but its decreasing trend slows down when the temperature reaches 410 °C. After the temperature reached 390 °C, the solubility value was less than 2 mmol/L at any pressure (23–25 MPa). The effect of pressure on the solubility is more significant in the low-temperature zone, where the value falls with decreasing pressure. The influence of temperature and pressure on the solubility of salt is essentially an effect of the density of water. Thus, only in the lower temperature range, when the pressure has a greater impact on the density of water, will there be a significant change in the solubility of salt. After a temperature increase to a certain point, the water enters the low-density zone again, at which point the impact of pressure on solubility is no longer apparent.

4.2. Results of Na2SO4 Salt Deposition Experiments

The experiments on salt deposition were conducted mainly at 390 °C due to the extremely low solubility of Na2SO4 under supercritical conditions (removal rate up to 90%). Table 4 lists a series of experimental information showing that there was no nickel foam on the heating rod and the bottom of the autoclave. The experiments of salt deposition with different heating rod temperatures and amounts of solution were conducted. The experimental results show that the superheated surface had a greater effect on the deposition form of Na2SO4, and loose salt particles could be seen on the heating rod and the bottom of the kettle (Figure 8a,b). The condition of the kettle bottom after the removal of the loose salt is shown in Figure 8c,f,i. It can be observed that at low heater rod temperatures, a thick compact layer of salt developed at the bottom of the kettle and that this layer had high mechanical strength making it difficult to remove, seriously affecting heat transfer efficiency and blocking. A comparison in Figure 8a,d indicates that when the heating rod temperature is higher, the kettle deposits more of the loosely shaped salt rather than the dense salt layer shown in Figure 8b, which is undoubtedly more beneficial for industrial operations.
As shown in Figure 8d, a dense salt layer is more likely to form on the heating rod when the amount of solution is low. This is probably due to the fact that when the salt concentration entering the reactor is low, the salt particles tend to nucleate on the heating rod. At this point, the nucleation rate is much higher than the growth rate, resulting in the formation of a salt film that covers the heating rod. However, the nucleation sites on the heating rod are covered as the solution is continuously introduced. Additionally, salt particles grow in aggregates on the formed salt nuclei, which are deposited at the bottom after being continuously detached by gravity, thus forming a loose salt lump.
To further verify the effect of nickel foam on salt deposition on the heating rod and at the bottom of the autoclave, several sets of experiments were conducted by varying the amount of brine passed (Table 5). When no material was placed on the heating rods, salt also appeared on the heating rods but did not accumulate as much as it did when the material was available (Figure 8g). The salt was mainly concentrated in the bottom space of the kettle. After arranging the nickel foam, it was evident that the salt was deposited preferentially on top of the nickel foam. When the nickel foam was set up close to the bottom wall, the deposited particles covered the material, in contrast to the case in the absence of material, when the dense salt layer disappeared close to the wall. The presence of nickel foam effectively inhibited the formation of a dense salt layer on the bottom of the autoclave. As the amount of solution passing through increases, the morphology of the salt deposited on the bottom tends to be more and more loose.
It is noteworthy that the salt deposited on the material on the heating rod was primarily concentrated on the feed side, with salt particles also appearing on the reactor wall in the opposite position. Comparing the results on the heating rods (Figure 8g and Figure 9g), it is evident that nickel foam effectively enhanced the induced crystallization of Na2SO4 on the heating rods.

4.3. Numerical Simulation Results

In order to better visualize the velocity distribution in the autoclave, the velocity field at different flow rates was simulated under a heating rod temperature of 600 °C and 900 °C. Figure 10 and Figure 11 show the velocity field and the streamlines at different flow rates, respectively under a heating rod temperature of 600 °C, while the velocity field and the streamlines at different flow rates under a heating rod temperature of 900 °C are shown in Figure 12 and Figure 13, respectively. At the three flow rates, the velocity in the autoclave tended to decrease gradually from the inlet to the inside of the kettle and formed a low-velocity zone in most of the area between the bottom of the heating rod and the bottom of the autoclave.
Figure 14 and Figure 15 shows the temperature distribution at different flow rates when the heating rod temperature was 600 °C and 900 °C. It can be seen that despite the overall similarity in temperature distribution across the three flow rates, the temperature in proximity to the heating rod wall was significantly higher. Additionally, the region of the highest temperature within the autoclave was predominantly situated in the middle area of the heating rod. This phenomenon may explain the initial deposition of inorganic salts at the inlet following the placement of nickel foam on the heating rod. Considering the influence of temperature on the precipitation of inorganic salts under supercritical conditions, the heating rod surface serves as a preferential site for inorganic salt nucleation and precipitation, with sodium sulfate undergoing heterogeneous nucleation on the heating rod. At the same time, due to the phase characteristics of sodium sulfate, after nucleation, it continues to grow and agglomerate and deposits at the bottom of the autoclave under the effect of gravity. This accounts for our observation of loosely packed large particle salt blocks at the bottom of the autoclave.

4.4. Analysis of the Mechanism of Salt Deposition

The nucleation mechanisms of crystals can be classified into two categories: homogeneous and non-homogeneous. Homogeneous nucleation occurs in the main fluid, while non-homogeneous nucleation occurs mainly on various surfaces. When the concentration of the brine in the autoclave is higher than the solubility under supercritical conditions, spontaneous homogeneous nucleation occurs in the solution [21]. The primary cause of salt precipitation is the significant alteration in the properties of the solvent water near the critical point. At the same time, the higher the temperature, the greater the supersaturation of the salt solution, and the faster the rate of salt crystallization and nucleation. As the autoclave is mainly heated by an external heater, the temperature of the wall is much higher than that of the fluid. Therefore, a large temperature gradient is formed near the walls, making the properties of the fluid preferentially change near the hot surface. The hot surface itself can also provide more adsorption sites for salt ions, resulting in a greater tendency for salt to nucleate on the high-temperature surface in non-homogeneous nucleation.
Figure 16b shows the binary phase diagram of Na2SO4 under a pressure of 25 MPa. Voisin [34] showed that aqueous Na2SO4 solution has a phase change under supercritical conditions. With an increase in temperature and concentration, Na2SO4 can precipitate readily as a solid phase (Figure 16b), instead of forming a gas–liquid coexistence state as it does in aqueous sodium chloride solutions (V + L region in Figure 16a). Thus, Na2SO4, as a type II salt, is anticipated to preferentially crystallize in the region where supersaturated concentrations and high temperatures exist. Thereafter, rapid growth and aggregation are driven by concentration gradients and settling at the bottom is driven by gravity (Figure 17).
Foam metal exhibits better thermal conductivity than inorganic porous materials such as molecular sieves do due to its metallic nature. Direct contact with the superheated wall surface can extend the range of high-temperature regions near the wall. It can also provide a larger surface area, which facilitates the heterogeneous nucleation of Na2SO4. Due to the characteristics of its porous medium, Na2SO4 crystals have limited space to grow in it, making it difficult for the crystals to grow stably in a single direction. This may explain its effectiveness at inhibiting the formation of compact salt layers at the bottom of the autoclave (Figure 18).

5. Conclusions

In this study, solubility data for Na2SO4 were initially obtained at 23–25 MPa and 380–420 °C. The removal rate of Na2SO4 from brine can reach more than 90% under supercritical conditions. Subsequently, a continuous apparatus with a heating rod was utilized to investigate the deposition characteristics of Na2SO4 in an autoclave at 23 MPa and 390 °C. The effect of superheated surfaces on the preferential nucleation growth of salt under supercritical conditions was examined. The findings indicated that Na2SO4 was deposited on the surface of the heating rod as well as on the bottom of the autoclave. This is related to the phase behavior of the salt; under supercritical conditions, the Na2SO4 separated into a gas phase and a solid phase, which was preferentially nucleated by the heater bar and grew along the inlet on it before being deposited due to the effect of gravity. Simulations using CFD software were conducted to obtain the temperature distribution in the reactor. The results showed that the temperature near the heating rod wall was higher than that of the main fluid, which was where the inorganic salts preferentially nucleated under supercritical conditions.
In addition, nickel foam was utilized as a porous medium to provide a large surface area for the superheated surface. As a carrier, it could accumulate significant amounts of salt on the heating rod, altering the state of the compact salt layer on the bottom of the autoclave. This presents a novel approach for the industrial recovery of inorganic salts under supercritical conditions.
However, hypersaline wastewater typically contains more than just Na2SO4, and for the supercritical water oxidation process, there is the oxidative transformation of organic matter as well. Therefore, it is crucial to investigate the phase behavior of other types of inorganic salts, the deposition phenomena, and whether or not the oxidation process of organic matter affects the deposition of inorganic salts in future work.

Author Contributions

Conceptualization, Z.C.; methodology, Z.L. and Q.Z.; investigation, Q.Z., Z.L. and Y.S.; supervision, Y.T., T.Y. and Z.C.; data curation, Q.Z. and Z.L.; writing—original draft preparation, Q.Z.; writing—review and editing, Q.Z. and Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data collected in this study are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leusbrock, I.; Metz, S.J.; Rexwinkel, G.; Versteeg, G.F. Quantitative approaches for the description of solubilities of inorganic compounds in near-critical and supercritical water. J. Supercrit. Fluids 2008, 47, 117–127. [Google Scholar] [CrossRef] [Green Version]
  2. Deguchi, S.; Tsujii, K. Supercritical water: A fascinating medium for soft matter. Soft Matter 2007, 3, 797–803. [Google Scholar] [CrossRef] [PubMed]
  3. Marrone, P.A. Supercritical water oxidation—Current status of full-scale commercial activity for waste destruction. J. Supercrit. Fluids 2013, 79, 283–288. [Google Scholar] [CrossRef]
  4. Schmieder, H.; Abeln, J. Supercritical Water Oxidation: State of the Art. Chem. Eng. Technol. 1999, 22, 903–908. [Google Scholar] [CrossRef]
  5. Marrone, P.A.; Hodes, M.; Smith, K.A.; Tester, J.W. Salt precipitation and scale control in supercritical water oxidation—Part B: Commercial/full-scale applications. J. Supercrit. Fluids 2004, 29, 289–312. [Google Scholar] [CrossRef]
  6. Koda, S. Oxidation reactions of solid carbonaceous and resinous substances in supercritical water. J. Supercrit. Fluids 2009, 47, 400–406. [Google Scholar] [CrossRef]
  7. Loppinet-Serani, A.; Aymonier, C.; Cansell, F. Supercritical water for environmental technologies. J. Chem. Technol. Biotechnol. 2010, 85, 583–589. [Google Scholar] [CrossRef]
  8. Odu, S.O.; van der Ham, A.G.J.; Metz, S.; Kersten, S.R.A. Design of a Process for Supercritical Water Desalination with Zero Liquid Discharge. Ind. Eng. Chem. Res. 2015, 54, 5527–5535. [Google Scholar] [CrossRef]
  9. van Wyk, S.; van der Ham, A.G.J.; Kersten, S.R.A. Analysis of the energy consumption of supercritical water desalination (SCWD). Desalination 2020, 474, 114189. [Google Scholar] [CrossRef]
  10. van Wyk, S.; van der Ham, A.G.J.; Kersten, S.R.A. Potential of supercritical water desalination (SCWD) as zero liquid discharge (ZLD) technology. Desalination 2020, 495, 114593. [Google Scholar] [CrossRef]
  11. Feng, W.; van der Kooi, H.J.; de Swaan Arons, J. Biomass conversions in subcritical and supercritical water: Driving force, phase equilibria, and thermodynamic analysis. Chem. Eng. Process. Process Intensif. 2004, 43, 1459–1467. [Google Scholar] [CrossRef]
  12. Abdpour, S.; Santos, R.M. Recent advances in heterogeneous catalysis for supercritical water oxidation/gasification processes: Insight into catalyst development. Process Saf. Environ. Prot. 2021, 149, 169–184. [Google Scholar] [CrossRef]
  13. Matsumura, Y.; Minowa, T.; Potic, B.; Kersten, S.; Prins, W.; Vanswaaij, W.; Vandebeld, B.; Elliott, D.; Neuenschwander, G.; Kruse, A. Biomass gasification in near- and super-critical water: Status and prospects. Biomass Bioenergy 2005, 29, 269–292. [Google Scholar] [CrossRef]
  14. Hayashi, H.; Hakuta, Y. Hydrothermal Synthesis of Metal Oxide Nanoparticles in Supercritical Water. Materials 2010, 3, 3794–3817. [Google Scholar] [CrossRef] [Green Version]
  15. Ye, N.; Yan, T.; Jiang, Z.; Wu, W.; Fang, T. A review: Conventional and supercritical hydro/solvothermal synthesis of ultrafine particles as cathode in lithium battery. Ceram Int. 2018, 44, 4521–4537. [Google Scholar] [CrossRef]
  16. Yoko, A.; Seong, G.; Tomai, T.; Adschiri, T. Continuous Flow Synthesis of Nanoparticles Using Supercritical Water: Process Design, Surface Control, and Nanohybrid Materials. Kona Powder Part J. 2020, 37, 28–41. [Google Scholar] [CrossRef] [Green Version]
  17. Pietrzak, K.; Krstulovic, N.; Blazeka, D.; Car, J.; Malinowski, S.; Wardak, C. Metal oxide nanoparticles as solid contact in ion-selective electrodes sensitive to potassium ions. Talanta 2022, 243, 123335. [Google Scholar] [CrossRef]
  18. Amy, G.; Ghaffour, N.; Li, Z.Y.; Francis, L.; Linares, R.V.; Missimer, T.; Lattemann, S. Membrane-based seawater desalination: Present and future prospects. Desalination 2017, 401, 16–21. [Google Scholar] [CrossRef]
  19. Lefebvre, O.; Moletta, R. Treatment of organic pollution in industrial saline wastewater: A literature review. Water Res. 2006, 40, 3671–3682. [Google Scholar] [CrossRef]
  20. Goh, P.S.; Zulhairun, A.K.; Ismail, A.F.; Hilal, N. Contemporary antibiofouling modifications of reverse osmosis desalination membrane: A review. Desalination 2019, 468, 114072. [Google Scholar] [CrossRef]
  21. Zhang, Y.; Wang, S.; Li, Y.; Zhang, J.; Xu, D.; Yang, C.; Yang, J.; Li, J.; Xu, T. Inorganic salts in sub-/supercritical water—Part A: Behavior characteristics and mechanisms. Desalination 2020, 496, 114674. [Google Scholar] [CrossRef]
  22. Leusbrock, I.; Metz, S.; Rexwinkel, G.; Versteeg, G. Solubility of 1:1 Alkali Nitrates and Chlorides in Near-Critical and Supercritical Water. Chem. Eng. 2009, 54, 3215–3223. [Google Scholar] [CrossRef] [Green Version]
  23. Rogak, S.N.; Teshima, P. Deposition of sodium sulfate in a heated flow of supercritical water. AIChE J. 1999, 45, 240–247. [Google Scholar] [CrossRef]
  24. Yang, J.; Wang, S.; Li, Y.; Tang, X.; Wang, Y.; Xu, D.; Guo, Y. Effect of salt deposit on corrosion behavior of Ni-based alloys and stainless steels in supercritical water. J. Supercrit. Fluids 2019, 152, 104570. [Google Scholar] [CrossRef]
  25. Xu, D.; Huang, C.; Wang, S.; Lin, G.; Guo, Y. Salt deposition problems in supercritical water oxidation. Chem. Eng. J. 2015, 279, 1010–1022. [Google Scholar] [CrossRef]
  26. Dell’Orco, P.; Eaton, H.; Reynolds, T.; Buelow, S. The solubility of 1:1 nitrate electrolytes in supercritical water. J. Supercrit. Fluids 1995, 8, 217–227. [Google Scholar] [CrossRef]
  27. Yang, J.Q.; Wang, S.Z.; Li, Y.H.; Xu, D.H. Under-deposit corrosion of Ni-based alloy 825 and Fe-Ni based alloy 800 in supercritical water oxidation environment. Corros. Sci. 2020, 167, 108493. [Google Scholar] [CrossRef]
  28. Li, W.G.; Yu, Z.B.; Wang, Y.; Li, Y.L. Heat transfer of supercritical carbon dioxide in a tube-in-tube heat exchanger-a CFD study. J. Supercrit. Fluid. 2022, 181, 105493. [Google Scholar] [CrossRef]
  29. Demoisson, F.; Ariane, M.; Leybros, A.; Muhr, H.; Bernard, F. Design of a reactor operating in supercritical water conditions using CFD simulations. Examples of synthesized nanomaterials. J. Supercrit. Fluid. 2011, 58, 371–377. [Google Scholar] [CrossRef]
  30. Nadimpalli, N.K.V.; Bandyopadhyaya, R.; Runkana, V. A coupled CFD-PBM and thermodynamic analysis of continuous supercritical hydrothermal synthesis of nanoparticles. J. Supercrit. Fluid. 2018, 136, 164–179. [Google Scholar] [CrossRef]
  31. Leybros, A.; Piolet, R.; Ariane, M.; Muhr, H.; Bernard, F.; Demoisson, F. CFD simulation of ZnO nanoparticle precipitation in a supercritical water synthesis reactor. J. Supercrit. Fluid. 2012, 70, 17–26. [Google Scholar] [CrossRef]
  32. Voisin, T.; Erriguible, A.; Philippot, G.; Ballenghien, D.; Mateos, D.; Cansell, F.; Iversen, B.B.; Aymonier, C. Investigation of the precipitation of Na2SO4 in supercritical water. Chem. Eng. Sci. 2017, 174, 268–276. [Google Scholar] [CrossRef]
  33. Voisin, T.; Erriguible, A.; Aubert, G.; Aymonier, C. Aggregation of Na2SO4 Nanocrystals in Supercritical Water. Ind. Eng. Chem. Res. 2018, 57, 2376–2384. [Google Scholar] [CrossRef]
  34. Voisin, T.; Erriguible, A.; Aymonier, C. Influence of multiphasic systems on salt(s) solubility in supercritical water: The case of NaCl and NaCl-Na2SO4. J. Supercrit. Fluids 2019, 152, 104567. [Google Scholar] [CrossRef]
  35. Hodes, M.; Marrone, P.A.; Hong, G.T.; Smith, K.A.; Tester, J.W. Salt precipitation and scale control in supercritical water oxidation—Part A: Fundamentals and research. J. Supercrit. Fluids 2004, 29, 265–288. [Google Scholar] [CrossRef]
Figure 1. Flow chart of experimental device.
Figure 1. Flow chart of experimental device.
Processes 11 01779 g001
Figure 2. The experimental setup. 1—computer; 2,3—N2 valve; 4—N2 pump; 5—power supply; 6—liquid pump; 7—preheater; 8—pressure gauge; 9—valve; 10—back pressure valve; 11—bypass valve; 12—conductivity display; 13—autoclave; 14—tank.
Figure 2. The experimental setup. 1—computer; 2,3—N2 valve; 4—N2 pump; 5—power supply; 6—liquid pump; 7—preheater; 8—pressure gauge; 9—valve; 10—back pressure valve; 11—bypass valve; 12—conductivity display; 13—autoclave; 14—tank.
Processes 11 01779 g002
Figure 3. The concentration of Na2SO4 as a function of conductivity.
Figure 3. The concentration of Na2SO4 as a function of conductivity.
Processes 11 01779 g003
Figure 4. Schematic of high-pressure autoclave.
Figure 4. Schematic of high-pressure autoclave.
Processes 11 01779 g004
Figure 5. Physical model (a) and mesh for the modeling (b).
Figure 5. Physical model (a) and mesh for the modeling (b).
Processes 11 01779 g005
Figure 6. Grid sensitivity test. Grid size: (a) 5 mm, (b) 3 mm, (c) 1.5 mm, (d) 1 mm, and (e) 0.5 mm.
Figure 6. Grid sensitivity test. Grid size: (a) 5 mm, (b) 3 mm, (c) 1.5 mm, (d) 1 mm, and (e) 0.5 mm.
Processes 11 01779 g006
Figure 7. Solubility of Na2SO4 at different temperatures and pressures.
Figure 7. Solubility of Na2SO4 at different temperatures and pressures.
Processes 11 01779 g007
Figure 8. Deposition results of Na2SO4 without materials. Heating rod (a,d,g). Bottom of the reactor (b,e,h). Results affter removing the material and salt (c,f,i).
Figure 8. Deposition results of Na2SO4 without materials. Heating rod (a,d,g). Bottom of the reactor (b,e,h). Results affter removing the material and salt (c,f,i).
Processes 11 01779 g008
Figure 9. Deposition results of Na2SO4 with materials on the heating rod and bottom of the kettle. Heating rod: (a,d,g). Bottom of the reactor: (b,e,h). After removing the material and salt: (c,f,i).
Figure 9. Deposition results of Na2SO4 with materials on the heating rod and bottom of the kettle. Heating rod: (a,d,g). Bottom of the reactor: (b,e,h). After removing the material and salt: (c,f,i).
Processes 11 01779 g009
Figure 10. The contour of velocity in the reactor at different brine flow rates under a heating rod temperature of 600 °C: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 10. The contour of velocity in the reactor at different brine flow rates under a heating rod temperature of 600 °C: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g010
Figure 11. The streamlines in the reactor under a heating rod temperature of 600 °C at different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 11. The streamlines in the reactor under a heating rod temperature of 600 °C at different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g011
Figure 12. The contour of velocity in the reactor at different brine flow rates under a heating rod temperature of 900 °C: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 12. The contour of velocity in the reactor at different brine flow rates under a heating rod temperature of 900 °C: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g012
Figure 13. The streamlines in the reactor under a heating rod temperature of 900 °C at different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 13. The streamlines in the reactor under a heating rod temperature of 900 °C at different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g013
Figure 14. Temperature distribution at 600 °C of heating rod surface under different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 14. Temperature distribution at 600 °C of heating rod surface under different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g014
Figure 15. Temperature distribution at 900 °C of heating rod surface under different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Figure 15. Temperature distribution at 900 °C of heating rod surface under different brine flow rates: (a) 10 mL/min, (b) 20 mL/min, and (c) 50 mL/min.
Processes 11 01779 g015
Figure 16. Temperature–concentration projections of phase diagrams at 25 MPa. (a) NaCl-H2O and (b) Na2SO4-H2O. L = liquid, V = vapor, S = solid (reconstructed from [35], copyright© (2004) Elsevier).
Figure 16. Temperature–concentration projections of phase diagrams at 25 MPa. (a) NaCl-H2O and (b) Na2SO4-H2O. L = liquid, V = vapor, S = solid (reconstructed from [35], copyright© (2004) Elsevier).
Processes 11 01779 g016
Figure 17. Schematic diagram of salt deposition mechanism in autoclave.
Figure 17. Schematic diagram of salt deposition mechanism in autoclave.
Processes 11 01779 g017
Figure 18. Schematic diagram of the effect of nickel foam on Na2SO4 salt deposition.
Figure 18. Schematic diagram of the effect of nickel foam on Na2SO4 salt deposition.
Processes 11 01779 g018
Table 1. Chemical reagents used in this work.
Table 1. Chemical reagents used in this work.
Chemical ReagentSourceMass Fraction PurityPurification MethodCAS No.
Sodium sulfate anhydrousAladdin (Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China)AR, 99%None7757-82-6
deionized waterShanghai Huazhen Technology Co., Ltd., Shanghai, ChinaConductivity < 1.75 μS/cmNone7732-18-5
Table 2. Information for the mesh.
Table 2. Information for the mesh.
CellsFacesNodesPartitions
872,8421,798,513184,79112 cell zones, 9 face zones
Table 3. Solubility data for Na2SO4 at various temperatures and pressures.
Table 3. Solubility data for Na2SO4 at various temperatures and pressures.
P/MPaT/°CSolubility of Na2SO4 (mmol/L)
233800.18
3900.07
4000.06
4100.05
4200.04
243801.02
3900.45
4000.35
4100.32
4200.26
2538015.34
3901.58
4001.00
4100.44
4200.33
Table 4. Experimental information of Na2SO4 with no materials.
Table 4. Experimental information of Na2SO4 with no materials.
RunP/MPaT1 a/°CT2 b/°CT3 c/°CAmount of Solution/L
122.9390500655–7361.0
222.9390600622.3–710.30.5
322.9390600628.1–695.11.0
a T1 is the temperature in the kettle during the experiment. b T2 is the temperature of the heating rod. c T3 is the temperature of the heating furnace.
Table 5. Experimental information of Na2SO4 with materials on the heating rod and bottom of the kettle.
Table 5. Experimental information of Na2SO4 with materials on the heating rod and bottom of the kettle.
RunP/MPaT1 a/°CT2 b/°CT3 c/°CAmount of Solution/L
122.9390600608–6900.5
222.9390600637.8–6730.75
322.9390600627.7–706.81.0
a T1 is the temperature in the kettle during the experiment. b T2 is the temperature of the heating rod. c T3 is the temperature of the heating furnace.
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

Zhu, Q.; Li, Z.; Song, Y.; Tong, Y.; Yang, T.; Cheng, Z. Effects of Superheated Surface on the Deposition Behavior of Na2SO4 in Supercritical Water. Processes 2023, 11, 1779. https://doi.org/10.3390/pr11061779

AMA Style

Zhu Q, Li Z, Song Y, Tong Y, Yang T, Cheng Z. Effects of Superheated Surface on the Deposition Behavior of Na2SO4 in Supercritical Water. Processes. 2023; 11(6):1779. https://doi.org/10.3390/pr11061779

Chicago/Turabian Style

Zhu, Qiao, Zhe Li, Yafei Song, Yujun Tong, Tao Yang, and Zhenmin Cheng. 2023. "Effects of Superheated Surface on the Deposition Behavior of Na2SO4 in Supercritical Water" Processes 11, no. 6: 1779. https://doi.org/10.3390/pr11061779

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