Next Article in Journal
Clustered Distributed Data Storage Repairing Multiple Failures
Previous Article in Journal
Coded Distributed Computing Under Combination Networks
Previous Article in Special Issue
The Application of Pinch Technology to a Novel Closed-Loop Spray Drying System with a Condenser and Reheater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation of Effect of Nozzle Upper Divergent Angle on Asymmetric Rectangular Section Ejector

Institute of Marine Engineering and Thermal Science, Marine Engineering College, Dalian Maritime University, Dalian 116026, China
*
Author to whom correspondence should be addressed.
Entropy 2025, 27(3), 312; https://doi.org/10.3390/e27030312
Submission received: 7 February 2025 / Revised: 3 March 2025 / Accepted: 12 March 2025 / Published: 17 March 2025
(This article belongs to the Special Issue Thermal Science and Engineering Applications)

Abstract

:
Ejectors, as widely utilized devices in the field of industrial energy conservation, exhibit a performance that is significantly affected by their structural parameters. However, the study of the influence of nozzle geometry parameters on asymmetric ejector performance is still limited. In this paper, the effect of the nozzle upper divergent angle on the operating characteristics of an asymmetric rectangular section ejector was comprehensively investigated. The results indicated that the entrainment ratio gradually decreased with an increase in the nozzle upper divergent angle, and the maximum decrease could be 20%. At the same time, the relationship between the upper and lower divergent angles was closely linked to the trend of change in the secondary fluid mass flow rate. The analysis of flow characteristics found that the deflection of the central jet was caused by the pressure difference between the walls of the upper and lower divergent sections of the nozzle. Additionally, quantitative analysis of the development of the mixing layer showed that the mass flow rate of the secondary fluid inlet was related to the development of the mixing boundary. Shock wave analysis demonstrated that the deterioration in ejector performance was due to the reduction in the shock wave strength caused by Mach reflection and the increase in the Mach stem height.

1. Introduction

With the expansion of global large-scale social production, the consumption rate of non-renewable energy is increasing. This serious situation has prompted the innovation and rapid development of various energy-saving technologies. The main sections of an ejector are the nozzle, suction chamber, mixing chamber, constant-area mixing section, and diffuser. It is recognized as an efficient energy-saving device characterized by its simple structure and no additional power consumption. Therefore, the ejector is favored in the fields of seawater desalination [1], refrigeration [2], waste heat recovery [3], and adiabatic compressed air energy storage [4,5,6]. As computational fluid dynamics (CFD) continues to progress, many scholars [7,8,9,10,11] have adopted the CFD technique to carry out in-depth numerical studies on ejector performance.
The core of a supersonic jet predominantly occurs in the downstream region of the nozzle divergent section. Consequently, the structural parameters of this section significantly affect the ejector performance. Fu et al. [12] conducted a detailed investigation into the influence of the nozzle divergent length on the ejector performance. Their findings indicated that the entrainment ratio (ω) initially increases and then decreases with an increase in the nozzle divergent section length. Wang et al. [13] not only numerically investigated the nozzle divergent section length, but also optimized the nozzle divergent angle. They revealed that there are optimal parameters for both to obtain the best ejector performance. Similarly, Yan et al. [14] emphasized the importance of the nozzle divergent angle in improving ejector performance. Their research revealed that, under ideal operating conditions, the optimal angle can improve ω by 16.7%. Furthermore, Li et al. [15] used single-factor and multi-factor analysis methods to identify five critical structural parameters influencing ejector performance and sorted them according to their sensitivity. Among these parameters, the nozzle divergent angle was recognized as one of the most significant factors.
The internal mixing process between primary and secondary fluids is the main factor influencing an ejector. Numerous scholars have carried out investigations into this issue and achieved a series of remarkable research results. According to the standard one-dimensional theoretical model, the mixing process starts until the secondary flow is accelerated to the sonic level, and the two fluids flow independently in the first stage [16,17]. The mixing process in ejectors is the subject of ongoing research. Ariafar et al. [18] investigated the velocity vector distribution in the mixing chamber using numerical modeling. They discovered that the mixing layer first formed at the nozzle exit, where the interaction between primary and secondary fluids started. Subsequently, the mixing layer gradually thickened along the flow direction, and its thickness growth trend was approximately linear. Based on this discovery, they explored the relationship between the ω and the growth rate of the mixing layer. It was found that, with an increase in ω, the growth rate of the mixing layer accelerated accordingly. This is important for the optimization of ejector design [19]. Recently, Tang et al. [20] achieved the discrimination and tracing of primary and secondary fluids by introducing the species transport model into numerical simulations. Based on their results, they categorized the process of mixing layer growth into the following two stages: a fluctuating growth stage and an exponential growth stage. They also studied the influence law of operating conditions on the growth of the mixing layer and found that the influence of these conditions was more obvious during the exponential growth stage. Additionally, Tang et al. [21] provided a comprehensive comparative analysis of the evolution of the mixing layer from the perspectives of mass, momentum, and energy. This multi-dimensional perspective provides a new way to deeply understand the mechanism of the mixing process.
Complex processes, including shock wave formation and boundary layer separation, occur alongside supersonic fluid flow inside an ejector. These phenomena have a significant impact on ejector performance and have become the focus of research in this field. An analytical model was proposed by Zhu and Jiang [22] to predict the length of the first shock wave after the nozzle exit. They established a quantized relationship between the first shock length and ω, as follows: as the length of the first shock increased, the ω decreased. Similarly, Chen et al. [23] also focused on researching the shape parameter of the first shock wave inside the ejector. They found that the dimensionless height of the first shock wave increased with the primary fluid pressure and decreased with the secondary fluid pressure. This result can be attributed to the direct influence of shock wave height on the flow circulation of the secondary fluid. Arun et al. [24] studied a rectangular section ejector using numerical modeling in three dimensions. They noticed two different types of shock wave reflection laws occurring at the boundary layer within the mixing chambers. This discovery deepens researchers’ understanding of shock boundary layer reflection.
Although studies on ejectors have been widely explored, the current research trend is still mainly focused on symmetric structures. However, the existing studies show that under the same conditions, an asymmetric ejector is often superior to a symmetric ejector in some key indexes [25,26]. However, there are still relatively few studies on asymmetric ejectors, especially on key structural parameters such as the nozzle divergent angle. Current studies mostly focus on the effects of the pressure ratio, compression ratio, and temperature ratio of the primary and secondary flow on performance [27]. As for structural parameter studies of asymmetric rectangular section ejectors, particularly the impact of the nozzle upper divergent angle on their operating characteristics, these have not yet received sufficient attention. Therefore, this paper establishes a three-dimensional computational model of an asymmetric rectangular section ejector and employs the species transport equation to differentiate and calculate the mass fraction distribution of the fluid. In addition, the numerical Schlieren technique is used to present the specific form of the shock wave. This paper investigates the influence of the nozzle divergent angle on the performance, flow characteristics, mixing characteristics, and shock wave characteristics of the asymmetric rectangular section ejector. This paper reveals the mechanisms by which the nozzle upper divergent angle influences the ejector performance, thereby providing a theoretical basis for ejector design. By gaining a better understanding of how nozzle geometry parameters affect ejector performance, engineers can more effectively select and adjust nozzle designs to meet specific industrial requirements. Therefore, this study not only enhances the theoretical foundation of ejectors, but also offers valuable guidance for technological advancements and applications in related industries.

2. Methods

2.1. Geometry Model

Figure 1 displays the asymmetric rectangular section ejector studied in this paper. The ejector has two secondary fluid inlets, the upper secondary fluid inlet and the lower secondary fluid inlet. The mass flow of the secondary fluid is the sum mass flow rate of the lower secondary fluid inlet (ms2) and the upper secondary fluid inlet (ms1). The mass flow rate of the primary fluid inlet is (mg). Therefore, the entrainment ratio (ω) of the asymmetric rectangular section ejector in this paper is defined as follows.
ω = m s 1 + m s 2 m g
The axis of the ejector is taken as the reference line. As illustrated in Figure 2, the nozzle divergent angle consists of an upper divergent angle (θ1) and a lower divergent angle (θ2). The essence of the asymmetric rectangular section ejector described in this paper is to adjust the upper divergent angle only on the basis of the symmetrical structure, so as to give the ejector asymmetric characteristics. θ1 ranges from 9° to 30°, while θ2 remains constant at 18°. The structural characteristics of the asymmetric rectangular section ejector are referenced from Ref. [28]. The detailed dimensions are shown in Table 1.

2.2. Numerical Method and Physical Model

The following presumptions are introduced to simplify the numerical computation process:
(1)
The fluid in the ejector is a compressible ideal gas.
(2)
The walls of the ejector are non-slip and adiabatic.
(3)
Throughout the entire operation, the temperature variations brought on by the gas supersonic movement are disregarded.
(4)
The constant pressure principle governs the mixing process.
(5)
At all inlets, the fluid velocity is disregarded.
The governing equation in the solving process was as follows.
Continuity equation:
ρ t + x i ρ u i = 0
Momentum equation:
t ρ u i + x i ρ u i u j = P x i + τ i j x j
Energy equation:
t ρ E + x i [ u i ρ E + P ] = α e f f T x i + u j ( τ i j ) e f f
τ i j = μ e f f u i x j + u j x i 2 3 μ e f f u k x k δ i j
ρ = P R T
Many scholars have acknowledged the results of k-ω SST turbulence model simulations for supersonic ejectors [29,30]. Especially in numerical simulation calculations of rectangular section ejectors with air as the medium, k-ω SST model results are more consistent with experimental results compared to other turbulence models [24]. Therefore, the k-ω SST turbulence model is selected for calculation.
k-ω SST turbulence model equation:
ρ u i u j ¯ = μ t u i x j + u j x i 2 3 ρ k + μ t u k x k δ i j
t ρ k = x j Γ k k x j + G K Y K
t ρ ω = x j Γ ω ω x j + G ω Y ω + D ω
The diffusion energy source model is enabled in the species transport model to mark primary and secondary flows for an intuitive study of mixing processes. The mass diffusion coefficient is set as 0 because the two types of air have identical characteristics. The species transport equation is as follows.
ρ Y i u = ρ D i , e f f Y i + S i
where Y N = 1 i = 1 N 1 Y i , D i , e f f = D i , m + μ ρ S c t , S c t = 0.7 , S i = ω ˙ M w , i .
In this paper, Fluent is used as the solver of the CFD model, and the nonlinear control equations are discretized using a pressure-based solver. The coupled algorithm is chosen by the flow field iterative solution method to speed up the solution and improve its accuracy. The second-order upwind scheme is selected to complete the spatial discretization. The diffuser outlet is set as a “pressure outlet”, while the “pressure inlets” condition is applied to the inlets of primary and secondary fluids. The inlet pressure of the primary fluid is set as 500 kPa, the inlet pressure of the secondary fluid is set as 80 kPa, and the outlet pressure is set as 70 kPa to ensure that the ejector works during the critical working conditions. The standard temperature is 293 K. As the working medium of the ejector, air satisfies the gas state equation. Therefore, the fluid can be calculated as a continuous medium. The calculation is considered convergent when the residual value of the energy equation is less than 10−6 and the residual value of other variables is less than 10−4 [24].

2.3. Validation of Grid Independence

The three-dimensional grid of the asymmetric rectangular section ejector is divided using ICEM CFD 2021 R1, as shown in Figure 3. The grid in the downstream region of the nozzle exit is locally refined to accurately capture the mixing process. To validate the grid independence, point A (the nozzle exit position at the axis) and point B (the center of the mixing chamber entrance at the axis) are chosen as the monitor data. Table 2 shows the calculation results and deviations of the two monitor points. It is evident from the table that the deviation of these points decreases with the grid number increasing from 280,636 to 609,964. Considering the calculation accuracy and calculation time, the number of grids used in this paper is not less than 430,396.

3. Results and Discussion

Figure 4 shows that the performance of the asymmetric rectangular section ejector varies with a varying θ1. As θ1 increases, ω decreases from 0.575 to 0.46, representing a reduction of 20%, as shown in Figure 4a. Notably, there is minimal variation in ω when θ1 increases from 9° to 12°. However, beyond this range, ω experiences a rapid decline. The most significant decrease in ω is observed when θ1 transitions from 24° to 30°. Figure 4b shows the change in the mass flow rate at different secondary fluid inlets of the asymmetric rectangular section ejector with θ1. From the figure, it can be seen that the variation trends of the mass flow rates of the two secondary fluid inlets of the asymmetric rectangular section ejector are different. Specifically, as θ1 increases, ms1 continues to decrease, while ms2 experiences an increase. With θ1 increasing from 9° to 18°, ms1 is more than ms2, while, with θ1 increasing from 18° to 30°, ms2 is more than ms1. In order to further study the variation trend of the entrainment performance of the asymmetric rectangular section ejector, the flow characteristics, mixing characteristics, and shock wave characteristics are analyzed in this paper.

3.1. Flow Characteristic Analysis

Figure 5 shows the velocity contours on the XOY plane of the asymmetric rectangular section ejector with different θ1. With an increase in θ1, the length of the central jet is gradually shortened and fluid separation gradually occurs. The velocity of the central jet of the primary fluid can reach a maximum of about 570 m/s. Additionally, the over-expansion of the primary fluid at the nozzle exit becomes more obvious with the increase in θ1, which restricts the ability of the primary fluid to entrain the secondary fluid. More importantly, the center jet of the primary fluid is deflected so that the flow area of the upper and lower secondary fluid in the constant-area section is changed. When θ1 is less than θ2, the central jet of the primary fluid is deflected downward, thereby making the flow area of the upper secondary fluid larger than that of the lower secondary fluid. Conversely, when θ1 is greater than θ2, the central jet is deflected upward, resulting in a reduced flow area for the upper secondary fluid relative to the lower secondary fluid. When these observations are considered alongside the results presented in Figure 4, it becomes evident that the mass flow rate of the secondary fluid is positively correlated with the flow area for the secondary fluid.
From Figure 5, it also can be observed that the deflection of the central jet originates from the nozzle divergent section. As the primary fluid enters this segment, the pressure energy is quickly transformed into kinetic energy. The primary fluid expands rapidly due to the sudden change in geometry, causing the pressure inside the nozzle divergent section to drop noticeably. Consequently, this study analyzes the wall pressure distribution in the nozzle divergent section. Figure 6 illustrates the wall pressure distribution along the nozzle divergent section with different θ1. When θ1 is smaller than θ2, the wall pressure on the nozzle upper divergent section (P1) is higher than that of the lower divergent section (P2), and then P1 is lower than P2 near the nozzle exit. Yet, where θ1 is greater than θ2, P1 is lower than P2, and P1 becomes higher than P2 near the nozzle exit. Thus, it can be assumed that the deflection of the central jet is caused by a pressure difference between the walls of the upper and lower divergent section close to the nozzle throat. The greater difference between θ1 and θ2, the larger pressure difference between the walls of the upper and lower divergent section near the nozzle throat, which results in a greater deflection of the central jet. Additionally, the wall pressure in the nozzle divergent section increases as it approaches the outlet, indicating that the primary and secondary fluids have begun to mix at this location.

3.2. Mixing Characteristic Analysis

The mixing characteristics of an ejector are a critical consideration in its design and application, especially in improving its efficiency and performance. Figure 7 presents a schematic diagram of the formation of the mixing layer within the ejector. Following its flow from the nozzle exit, the primary fluid interacts with the secondary fluid in the mixing chamber. Due to the large difference in the thermodynamic properties of the primary fluid and secondary fluid, when they come into contact, energy is exchanged, and eventually the two fluids mix completely and achieve equilibrium. This mixing process requires a transitional region in space, known as the mixing layer. The region outside of the mixing layer is referred to as the non-mixing region. The methodology used by Tang et al. [20] is adopted so that the mixing process can be comprehensively analyzed. Contours corresponding to primary fluid mass fractions of 0.9 and 0.1 are defined as the mixing boundaries between the primary fluid and secondary fluid. Based on the structural characteristics of the asymmetric rectangular section ejector, the mixing layers are categorized into the upper mixing layer and lower mixing layer to facilitate a more intuitive investigation of the mixing characteristics, as illustrated in Figure 8. The mass fraction of the secondary fluid boundary is 0.1, while the mass fraction of the primary fluid boundary is 0.9. The primary fluid boundary develops along the central axis of the ejector in a closed state, whereas the secondary fluid boundary gradually develops along the ejector wall.
The development of the mixing boundaries with different θ1 is illustrated in Figure 9. The end positions of both the upper and lower mixing layer boundaries change with an increase in θ1. Specifically, as θ1 increases, the end positions of the primary fluid boundaries of the upper and lower mixing layers gradually shift back. This suggests that the primary fluid and secondary fluid are mixing less and less, that is, the primary fluid carries less and less secondary fluid. This is consistent with the results in Figure 4a. It is important to note that the regions closed by the primary fluid boundaries are not symmetrical. Specifically, when θ1 is smaller than θ2, the regions defined by the primary fluid boundaries are observed to be deflected downward. The primary fluid occupies the flow channels in the lower mixing layer, thus providing more regions for the flow channels in the upper mixing layer. Conversely, when θ1 is greater than θ1, the regions bounded by the primary fluid are deflected upward. In addition, as θ1 increases, the end position of the secondary fluid boundaries in the upper mixing layer is gradually shifted forward and the end position of the secondary fluid boundaries in the lower mixing layer is gradually shifted backward. For instance, the boundary of the secondary fluid in the upper mixing layer ends at X = 80.2 mm, while the boundary of the secondary fluid in the lower mixing layer ends at X = 38.8 mm when θ1 is set to 9°. Conversely, when θ1 is increased to 30°, the boundary of the secondary fluid in the upper mixing layer ends at an earlier position of X = 36.8 mm and the boundary of the secondary fluid in the lower mixing layer defers to a later position of X = 75.0 mm.
In order to further analyze the relationship between the mass flow rate and the secondary fluid boundaries, the following two evaluation indexes are introduced: the relative length of the secondary fluid boundary (λ) and the relative mass flow rate of the secondary fluid (φ). Since the asymmetric structure of the ejector is studied in this paper, it is necessary to analyze it in two cases, as follows.
λ 1 = l 1 l f
λ 2 = l 2 l f
where λ1 and λ2 represent the relative development lengths of the secondary fluid boundaries in the upper and lower mixing layers, respectively. l1 and l2 represent the lengths of the secondary fluid boundary from the nozzle exit to the ejector wall in the upper and lower mixing layers, respectively. lf is the length of the secondary fluid boundary from the nozzle exit to the ejector wall when θ1 is 18°.
φ 1 = m s 1 m f
φ 2 = m s 2 m f
where φ1 and φ2 represent the relative mass flow rates of the upper and lower secondary fluid inlets, respectively. ms1 and ms2 represent the mass flow rates of the upper and lower secondary fluid inlets. mf is the mass flow rate of the secondary fluid inlet when θ1 is 18°.
Figure 10 shows the relationship between the variables φ and λ. The data presented indicate that φ increases as λ rises. Furthermore, a strong linear correlation is observed between φ1 and λ1, as well as between φ2 and λ2. The Pearson’s correlation coefficients for these relationships are 0.97806 and 0.98775, respectively. The linear fitting function is provided as follows.
φ 1 = 1.48432 λ 1 0.36044
φ 2 = 1.27806 λ 2 0.18492
This significant linear correlation suggests that the full development of the mixing boundary has a critical influence on the mass flow rate of the secondary fluid inlet. The more adequate flow area for the secondary fluid facilitates a more complete development of the mixing boundary, thus allowing more secondary fluid to be entrained. This phenomenon can be interpreted as a higher value of λ indicating an increased contact area between the secondary and primary fluid, which subsequently enhances the probability of the secondary fluid being entrained by the primary fluid. Consequently, the value of φ increases in proportion to the value of λ.

3.3. Shock Wave Characteristic Analysis

The properties of the shock wave have a large impact on the ejector performance. This research examines the shock wave structure of the ejector using the numerical schlieren technique. The method effectively captures variations in the flow field density gradient, based on the principle that the density of the flow exactly correlates with the gradient of the refractive index. The orientation of the shock wave structure has a significant impact on the schlieren image quality. To enhance this image quality, Quirt [31] proposed using the absolute value of the density gradient. The mixing chamber and the downstream of the nozzle exit are identified as the two most critical locations for identifying shock waves. Therefore, these regions are locally refined during the meshing process, leading to a reduction in the grid cell volume in these areas. Based on this consideration, the present research refines Quirt’s formula for calculating shock wave intensity (I) by incorporating the grid cell volume (β) as a correction factor to improve the accuracy of the Schlieren image calculations. Here, I is defined as follows.
I = β ρ x 2 + ρ y 2
Ernst Mach’s investigations were the first to identify the occurrence of shock wave reflection [32]. The flow field properties of the shock wave are altered as a result of reflection. Therefore, the shock wave reflection significantly impacts the performance of the ejector. As illustrated in Figure 11, the shock wave reflection can be mainly categorized into the following two types: regular reflection and Mach reflection. Figure 11a shows the structure of a regular reflection shock. It consists of the incident shock waves (I1 and I2), the reflected shock waves (R1 and R2), the intersection point (T) of the incident shock wave and the reflected shock waves, and the boundary layer (S). Figure 11b shows the shock wave structure of Mach reflection. It consists of incident shock waves (I1 and I2), reflected shock waves (R1 and R2), and Mach stem (M), where S1 and S2 are the slip lines and S3 and S4 are the boundary layers.
Figure 12 presents numerical schlieren images with different θ1. Two compression waves as the first shock wave are observed after the nozzle exit for all structures. The difference is that these two compression waves are asymmetrically distributed due to the asymmetry of the nozzle divergent section, except for the structure with a θ1 of 18°. These two asymmetric shock waves intensify the deflection of the central jet. When θ1 is 9°, the two compression waves at the nozzle exit intersect to produce another compression wave. After, the compression waves pass through the fluid and are reflected as the expansion waves by the boundary layers. In this case, the first shock wave is characterized by regular reflection. As θ1 increases, this regular reflection turns into Mach reflection. As θ1 increases to 21°, the Mach stem becomes apparent. As θ1 is increased further, the height of the Mach stem steadily rises and the gap between the two slip lines widens. The region where the shock wave is located represents the flow region of the primary fluid. The increase in the height of the Mach stem leads to an increase in the region of the primary fluid in the Y-axis direction at the nozzle exit, thereby reducing the flow region of the nearby secondary fluid and restricting the secondary fluid being entrained. Moreover, the density gradient in the flow field decreases after Mach reflection. Especially when θ1 is 30°, there are no obvious compression and expansion waves in the numerical schlieren images. Combined with Figure 9, it can be seen that there is almost no fluctuation in the boundary of the primary fluid, which is not conducive to the mass transfer of the two fluids when θ1 is 30°. These two factors both lead to a deterioration in the ejector performance to some extent.

4. Conclusions

A three-dimensional mathematical model of an asymmetric rectangular section ejector was established in this paper. The effect of a nozzle upper divergent angle from 9° to 30° on the asymmetric rectangular section ejector performance was investigated. The reasons for the variation in performance of the asymmetric rectangular section ejector were explained in terms of flow characteristics, mixing characteristics, and shock wave characteristics. The following conclusions can be drawn:
(1)
The nozzle upper divergent angle had a significant effect on the ω. With an increase in θ1, ω decreased gradually. The ω decreased from 0.575 to 0.46, a decrease of 20%. However, the variation trend of ms1 and ms2 was opposite. When θ1 was smaller than θ2, ms1 was greater than ms2. When θ1 was greater than θ2, ms1 was less than ms2.
(2)
The deflection of the central jet of the primary fluid was caused by the pressure difference between the walls of the upper and lower expansion section of the nozzle.
(3)
As λ increased, there was a corresponding increase in φ. There was a high degree of linear correlation between λ and φ.
(4)
As θ1 increased, the reflection type of the first shock wave transitioned from regular reflection to Mach reflection. In the condition of Mach reflection, the increase in θ1 caused a rise in the height of the Mach stem, accompanied by a decrease in the intensity of the shock wave. These alterations collectively deteriorated the ejector performance to a certain extent.

Author Contributions

Conceptualization, investigation, writing and editing, project administration, funding acquisition: J.D.; validation, investigation, writing and editing: M.L.; methodology: C.F. and S.S.; investigation: M.Z. and R.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 51979022.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The research data supporting this publication are provided within this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wen, C.; Gong, L.; Ding, H.B.; Yang, Y. Steam ejector performance considering phase transition for multi-effect distillation with thermal vapour compression (MED-TVC) desalination system. Appl. Energy. 2020, 279, 115831. [Google Scholar] [CrossRef]
  2. Dong, J.M.; Yu, M.Q.; Wang, W.N.; Song, H.; Li, C.L.; Pan, X.X. Experimental investigation on low-temperature thermal energy driven steam ejector refrigeration system for cooling application. Appl. Therm. Eng. 2017, 123, 167–176. [Google Scholar] [CrossRef]
  3. Zhang, J.G.; Wang, X.G.; Pourranjbar, D.; Dykas, S.; Li, H.; Chen, J.H. The comprehensive analysis of the relationship between the latent heat, entrainment ratio, and ejector performance under different superheating degree conditions considering the non-equilibrium condensation. Appl. Therm. Eng. 2022, 200, 117701. [Google Scholar] [CrossRef]
  4. Cao, Z.; Zhou, S.H.; He, Y.; Xu, Y.J.; Chen, H.S.; Deng, J.Q. Numerical study on adiabatic compressed air energy storage system with only one ejector alongside final stage compression. Appl. Therm. Eng. 2022, 216, 119071. [Google Scholar] [CrossRef]
  5. Zhou, S.H.; He, Y.; Chen, H.S.; Xu, Y.J.; Deng, J.Q. Performance analysis of a novel adiabatic compressed air energy system with ejectors enhanced charging process. Energy 2020, 205, 118050. [Google Scholar] [CrossRef]
  6. Qing, S.W.; Wang, Y.; Wen, X.K.; Zhong, J.L.; Gou, X.L.; Tang, S.L. Optimal working-parameter analysis of an ejector integrated into the energy-release stage of a thermal-storage compressed air energy storage system under constant-pressure operation: A case study. Energy Convers. Manag. 2021, 247, 114715. [Google Scholar] [CrossRef]
  7. Dong, J.M.; Hu, Q.Y.; Yu, M.Q.; Han, Z.T.; Cui, W.B.; Liang, D.L.; Ma, H.B.; Pan, X.X. Numerical investigation on the influence of mixing chamber length on steam ejector performance. Appl. Therm. Eng. 2020, 174, 115204. [Google Scholar] [CrossRef]
  8. Tang, Y.Z.; Liu, Z.L.; Li, Y.X.; Yang, N.; Wan, Y.D.; Chua, K.J. A double-choking theory as an explanation of the evolution laws of ejector performance with various operational and geometrical parameters. Energy Convers. Manag. 2020, 206, 112499. [Google Scholar] [CrossRef]
  9. Croquer, S.; Poncet, S.; Galanis, N. Comparison of ejector predicted performance by thermodynamic and CFD models. Int. J. Refrig. 2016, 68, 28–36. [Google Scholar] [CrossRef]
  10. Han, J.Q.; Pang, Z.H.; Feng, J.M.; Besagni, G.; Mereu, R.; Inzoli, F.; Peng, X.Y. Experimental and numerical study on the ejector containing condensable species in the secondary flow for PEM fuel cell applications. Appl. Therm. Eng. 2023, 232, 121091. [Google Scholar] [CrossRef]
  11. Tan, J.G.; Zhang, D.D.; Lv, L. A review on enhanced mixing methods in supersonic mixing layer flows. Acta Astronaut. 2018, 152, 310–324. [Google Scholar] [CrossRef]
  12. Fu, W.N.; Li, Y.X.; Liu, Z.L.; Wu, H.Q.; Wu, T.R. Numerical study for the influences of primary nozzle on steam ejector performance. Appl. Therm. Eng. 2016, 106, 1148–1156. [Google Scholar] [CrossRef]
  13. Wang, L.; Yan, J.; Wang, C.; Li, X.B. Numerical study on optimization of ejector primary nozzle geometries. Int. J. Refrig. 2017, 76, 219–229. [Google Scholar] [CrossRef]
  14. Yan, J.; Cai, W.J.; Li, Y.Z. Geometry parameters effect for air-cooled ejector cooling systems with R134a refrigerant. Renew. Energy 2012, 46, 155–163. [Google Scholar] [CrossRef]
  15. Li, C.; Sun, B.G.; Luo, Q.H. Effect of Structural Parameters and Operational Characteristic Analysis on Ejector Used in Proton Exchange Membrane Fuel Cell. Sustainability 2022, 14, 9205. [Google Scholar] [CrossRef]
  16. Munday, J.T.; Bagster, D.F. A new ejector theory applied to steam jet refrigeration. Ind. Eng. Chem. Process 1977, 16, 442–449. [Google Scholar] [CrossRef]
  17. Huang, B.J.; Chang, J.M.; Wang, C.P.; Petrenko, V.A. A 1-D analysis of ejector performance. Int. J. Refrig. 1999, 22, 354–364. [Google Scholar] [CrossRef]
  18. Ariafar, K.; Buttsworth, D.; Al-Doori, G.; Malpress, R. Effect of mixing on the performance of wet steam ejectors. Energy 2015, 93, 2030–2041. [Google Scholar] [CrossRef]
  19. Ariafar, K.; Buttsworth, D.; Al-Doori, G.; Sharifi, N. Mixing layer effects on the entrainment ratio in steam ejectors through ideal gas computational simulations. Energy 2016, 95, 380–392. [Google Scholar] [CrossRef]
  20. Tang, Y.Z.; Liu, Z.L.; Li, Y.X.; Huang, Z.F.; Chua, K.J. Study on fundamental link between mixing efficiency and entrainment performance of a steam ejector. Energy 2021, 215, 119128. [Google Scholar] [CrossRef]
  21. Tang, Y.Z.; Liu, Z.L.; Li, Y.X.; Zhao, F.; Fan, P.Y.; Chua, K.J. Mixing process of two streams within a steam ejector from the perspectives of mass, momentum and energy transfer. Appl. Therm. Eng. 2021, 185, 116358. [Google Scholar] [CrossRef]
  22. Zhu, Y.H.; Jiang, P.X. Experimental and analytical studies on the shock wave length in convergent and convergent–divergent nozzle ejectors. Energy Convers. Manag. 2014, 88, 907–914. [Google Scholar] [CrossRef]
  23. Chen, W.X.; Hou, Y.Y.; Zheng, J.T.; Chong, D.T.; Yan, J.J. Studies on Shock Wave Structures in Supersonic Ejector and Its Influence on Ejector Performance. Heat Transf. Eng. 2024, 45, 631–641. [Google Scholar] [CrossRef]
  24. Arun, K.M.; Tiwari, S.; Mani, A. Three-dimensional numerical investigations on rectangular cross-section ejector. Int. J. Therm. Sci. 2017, 122, 257–265. [Google Scholar] [CrossRef]
  25. Su, Y.F. Investigation on Pumping Performance of Non-Axisymmetric Supersonic Ejector. Master’s Thesis, Nanjing University of Aeronautics and Astronautics, Nanjing, China, 2020. [Google Scholar]
  26. Leng, Z.M.; Zhou, J.H. Numerical Investigation for Performance of Non-Axisymmetric Front Variable Area Bypasses Injector. J. Propuls. Technol. 2015, 36, 1465–1473. [Google Scholar]
  27. Zhang, Y.; Dong, J.M.; Song, S.Y.; Pan, X.X.; He, N.; Lu, M.F. Numerical Investigation on the Effect of Section Width on the Performance of Air Ejector with Rectangular Section. Entropy 2023, 25, 179. [Google Scholar] [CrossRef]
  28. Su, Y.F.; Shan, Y.; Wang, L. Experimental Investigation on Non-Axisymmetric Supersonic Ejector. J. Chongqing Univ. Technol. (Nat. Sci.) 2020, 34, 91–98. [Google Scholar]
  29. Besagni, G.; Inzoli, F. Computational fluid-dynamics modeling of supersonic ejectors: Screening of turbulence modeling approaches. Appl. Therm. Eng. 2017, 117, 122–144. [Google Scholar] [CrossRef]
  30. Ruangtrakoon, N.; Thongtip, T.; Aphornratana, S.; Sriveerakul, T. CFD simulation on the effect of primary nozzle geometries for a steam ejector in refrigeration cycle. Int. J. Therm. Sci. 2013, 63, 133–145. [Google Scholar] [CrossRef]
  31. Quirk, J.J. A contribution to the great Riemann solver debate. Int. J. Numer. Methods Fluids 2010, 18, 555–574. [Google Scholar] [CrossRef]
  32. Mach, E.G.E. Ueber den Verlauf der Funkenwellen in der Ebene und in Raume. J. Phys. Theor. Appl. 1879, 8, 94–100. [Google Scholar]
Figure 1. Three-dimensional structure diagram of the asymmetric rectangular section ejector.
Figure 1. Three-dimensional structure diagram of the asymmetric rectangular section ejector.
Entropy 27 00312 g001
Figure 2. Structure diagram of the asymmetric rectangular section ejector on the XOY plane.
Figure 2. Structure diagram of the asymmetric rectangular section ejector on the XOY plane.
Entropy 27 00312 g002
Figure 3. Three-dimensional grid diagram of the asymmetric rectangular section ejector.
Figure 3. Three-dimensional grid diagram of the asymmetric rectangular section ejector.
Entropy 27 00312 g003
Figure 4. Variation in the asymmetric rectangular section ejector performance with different θ1: (a) entrainment ratio and (b) mass flow rate.
Figure 4. Variation in the asymmetric rectangular section ejector performance with different θ1: (a) entrainment ratio and (b) mass flow rate.
Entropy 27 00312 g004
Figure 5. Velocity contours on the XOY plane with different θ1.
Figure 5. Velocity contours on the XOY plane with different θ1.
Entropy 27 00312 g005
Figure 6. Wall pressure distribution in the nozzle divergent section with different θ1.
Figure 6. Wall pressure distribution in the nozzle divergent section with different θ1.
Entropy 27 00312 g006
Figure 7. Schematic diagram of the formation of mixing layer within the ejector.
Figure 7. Schematic diagram of the formation of mixing layer within the ejector.
Entropy 27 00312 g007
Figure 8. Schematic diagram of the upper and lower mixing layers of the rectangular section ejector.
Figure 8. Schematic diagram of the upper and lower mixing layers of the rectangular section ejector.
Entropy 27 00312 g008
Figure 9. Development diagram of mixing boundaries with different θ1.
Figure 9. Development diagram of mixing boundaries with different θ1.
Entropy 27 00312 g009
Figure 10. The function relationship between φ and λ: (a) (φ1, λ1) and (b) (φ2, λ2).
Figure 10. The function relationship between φ and λ: (a) (φ1, λ1) and (b) (φ2, λ2).
Entropy 27 00312 g010
Figure 11. Schematic diagram of shock wave reflection: (a) regular reflection and (b) Mach reflection.
Figure 11. Schematic diagram of shock wave reflection: (a) regular reflection and (b) Mach reflection.
Entropy 27 00312 g011
Figure 12. Numerical Schlieren images with different θ1.
Figure 12. Numerical Schlieren images with different θ1.
Entropy 27 00312 g012
Table 1. Structure parameters of the asymmetric rectangular section ejector.
Table 1. Structure parameters of the asymmetric rectangular section ejector.
ParameterSymbolValueUnit
Constant-area mixing section heightHc6mm
Diffuser outlet heightHd22.3mm
Nozzle entrance heightHe14mm
Nozzle throat heightHt1mm
Constant-area mixing section lengthLc42mm
Diffuser lengthLd46.22mm
Nozzle throat lengthLt2mm
Nozzle upper divergent section lengthL13.4mm
Nozzle lower divergent section lengthL23.4mm
Ejector section widthWs6mm
Nozzle upper divergent angleθ19, 12, 15, 18, 21, 24, 27, 30°
Nozzle lower divergent angleθ218°
Nozzle exit positionNXP11.41°
Table 2. Grid independence test.
Table 2. Grid independence test.
Grid NumbersPressure (kPa)Deviation (%)Velocity (m/s)Deviation (%)
Point A280,63621.4970595.470
339,60021.0342.2012594.120.2272
389,84020.7471.3833593.4580.1116
430,39620.6120.6549593.2760.0307
522,14820.7120.4828593.4320.0263
609,96420.6560.2711593.5210.0145
Point B280,63678.1240364.5840
339,60079.9112.2362363.3140.3500
389,84080.9891.3310362.9220.1080
430,39680.2470.9246362.7640.04355
522,14879.9980.3113362.7400.0066
609,96479.8410.1966362.7220.0050
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

Lu, M.; Dong, J.; Feng, C.; Song, S.; Zhang, M.; Wang, R. Numerical Investigation of Effect of Nozzle Upper Divergent Angle on Asymmetric Rectangular Section Ejector. Entropy 2025, 27, 312. https://doi.org/10.3390/e27030312

AMA Style

Lu M, Dong J, Feng C, Song S, Zhang M, Wang R. Numerical Investigation of Effect of Nozzle Upper Divergent Angle on Asymmetric Rectangular Section Ejector. Entropy. 2025; 27(3):312. https://doi.org/10.3390/e27030312

Chicago/Turabian Style

Lu, Manfei, Jingming Dong, Chi Feng, Shuaiyu Song, Miao Zhang, and Runfa Wang. 2025. "Numerical Investigation of Effect of Nozzle Upper Divergent Angle on Asymmetric Rectangular Section Ejector" Entropy 27, no. 3: 312. https://doi.org/10.3390/e27030312

APA Style

Lu, M., Dong, J., Feng, C., Song, S., Zhang, M., & Wang, R. (2025). Numerical Investigation of Effect of Nozzle Upper Divergent Angle on Asymmetric Rectangular Section Ejector. Entropy, 27(3), 312. https://doi.org/10.3390/e27030312

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