Next Article in Journal
Equivalent Impedance Calculation Method for Control Stability Assessment in HVDC Grids
Previous Article in Journal
Pore-Scale Characterization and PNM Simulations of Multiphase Flow in Carbonate Rocks
Previous Article in Special Issue
Influence of Clearance Flow on Dynamic Hydraulic Forces of Pump-Turbine during Runaway Transient Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation on the Performance of Two-Throat Nozzle Ejectors with Different Mixing Chamber Structural Parameters

1
College of Food Science and Technology, Shanghai Ocean University, Shanghai 201306, China
2
National Experimental Teaching Demonstration Center for Food Science and Engineering, Shanghai 201306, China
3
Shanghai Professional Technology Service Platform on Cold Chain Equipment Performance and Energy Saving Evaluation, Shanghai 201306, China
*
Author to whom correspondence should be addressed.
Energies 2021, 14(21), 6900; https://doi.org/10.3390/en14216900
Submission received: 30 August 2021 / Revised: 10 October 2021 / Accepted: 19 October 2021 / Published: 21 October 2021

Abstract

:
In this study, the effects of the mixing chamber diameter (Dm), mixing chamber length (Lm) and pre-mixing chamber converging angle (θpm) were numerically investigated for a two-throat nozzle ejector to be utilized in a CO2 refrigeration cycle. The developed simulated method was validated by actual experimental data of a CO2 ejector in heat pump water heater system from the published literature. The main results revealed that the two-throat nozzle ejectors can obtain better performance with Dm in the range of 8–9 mm, Lm in the range of 64–82 mm and θpm at approximately 60°, respectively. Deviation from its optimal value could lead to a poor operational performance. Therefore, the mixing chamber structural parameters should be designed at the scope around its optimal value to guarantee the two-throat nozzle ejector performance. The following research can be developed around the two-throat nozzle geometries to strengthen the ejector performance.

1. Introduction

Sustainable economy and societal development are easily limited by energy shortages and environmental pollution [1]. In terms of refrigeration systems, the use of traditional CFCs and HCFCs refrigerants has leaded to the ozone layer being destroyed and greenhouse effects [2]. For the betterment of the environment, the natural refrigerant CO2 with zero ODP and low GWP is chosen as a desired substitution for these refrigerants by virtue of its environmental friendliness [3]. Currently, research works on CO2 are being performed extensively in areas such as supermarkets [4,5], heat pump units [6,7,8], vehicles [9,10], light commercial refrigeration [11,12], tumble dryers [13,14], chillers [15]. CO2 systems are widely promoted for their incomparable advantages. The CO2 vapor compression refrigeration cycle unavoidably works under trans-critical conditions because of its low critical temperature. The utilization of an expansion valve in trans-critical CO2 refrigeration systems would cause a larger irreversibly throttling loss due to the greater pressure drop compared with that of other traditional refrigeration systems [16]. For the sake of enhancing the performance CO2 refrigeration system, ejectors began to be used as promising devices to reduce throttling losses [17].
Ejectors have remarkable advantages of structural simplicity, low cost, no moving parts, being basically maintenance-free devices that employ environmentally friendly refrigerants such as CO2 [18,19]. The basic structure of an ejector consists of a primary nozzle, pre-mixing chamber, mixing chamber and diffuser. They work as follows: the primary flow (PF) is depressurized and accelerated inside the nozzle. Next, the PF with high velocity and low-pressure flows into pre-mixing chamber to form a low pressure area, where the secondary flow (SF) is entrained. The PF and SF mix with each other in the mixing chamber for the function of the exchange of kinetic energy. In the end, the kinetic energy of the mixed flow starts to be transformed to pressure potential energy in the diffuser. Part of energy can be recovered during the throttling process by using an ejector.
In 2008, the Japanese company DENSO firstly advanced the two-throat nozzle ejector with two Laval nozzles in series. Subsequently, Kang et al. [20] experimentally confirmed the smaller nozzle distance and first nozzle divergence angle at 8° provided better performance for two-throat nozzle ejectors in a ejector refrigeration system. Figure 1 presents the operating principles of the two-throat nozzle and Laval nozzle. The conventional ejector nozzle is a section of a Laval nozzle, where the refrigerant can only be vaporized at the end of the nozzle. This causes the problem that the droplets in the middle part are large and the flow speed cannot be increased well. The two-throat nozzle has two Laval nozzles in series. The refrigerant is depressurized and bubbles are created at the first section of the nozzle. Subsequently, the bubbles are ruptured due to the increase in area and pressure and bubbles nucleus are formed. The second section of the nozzle promotes rapid expansion and makes the droplets smaller to make the refrigerant be a homogeneous flow. This homogeneous flow is similar to a single-phase flow, which can improve the nozzle efficiency.
Ejectors exert considerable influence on ejector refrigeration cycles due to their entrainment performance [21]. Although there is an increasing trend of using ejectors in refrigeration systems, a lot of works need to be done on the optimization of ejector performance [22]. There exist some studies using unconventional nozzles to improve ejector performance by modify features such as the shape, quantity or with swirl generators. Chang and Chen [23] developed a petal nozzle to improve the pressure lift and entrainment performance of ejectors. Yang et al. [24] numerically implemented a comparison among five nozzle structures under the same working conditions including conical, elliptical, square, rectangular and cross-shaped nozzles. The results showed that the structure of the nozzle affected the Er and back pressure greatly, and the cross-shaped nozzle increased Er by 9.1% compared with a standard circular nozzle. Their work found the conical nozzle performed better compared with the other nozzle shapes as far as critical back pressures were concerned. Wang et al. [25] found that the nozzle exit position and the nozzle throat diameter exerted marked effects on the efficiency of novel two-phase ejectors. Banu and Mani [26] added a swirl generator structure to the ejector nozzle, and the simulated outcomes showed that the Er was increased by 15% compared with a traditional ejector, however the nozzle with swirl generator has a complex structure making it difficult to machine. Zhou et al. [27] developed a new type of ejector with double nozzle to optimize refrigeration systems by comparing the new cycle with the conventional ejector refrigeration system and the standard refrigeration system using a mathematical model. The results indicated that the COP of the new cycle was increased by 22.9–50.8% in contrast to the standard refrigeration system, and 10.5–30.8% in comparison to the conventional ejector refrigeration system. Rao et al. [28] increased Er by up to 30% with the aid of tip ring supersonic nozzles and elliptic sharp tipped shallow lobed nozzles in supersonic ejectors. Although significant numerical studies were carried out for the optimization of nozzles, the works were rarely developed experimentally.
The mixing chamber is one of the main objects of ejector design optimization. The optimization of conventional ejector mixing chamber geometric parameters has been conducted numerically and experimentally to enhance ejector performance and ejector refrigeration system performance [29,30,31]. Kim et al. [32] and Wang et al. [33] revealed that the mixing chamber geometries strongly affected ejector performance. Hou et al. [34] and Nakagawa et al. [35] obtained the optimal mixing chamber length (Lm) to make ejectors achieve better pressure lift and Er. Zhang et al. [36] and Wu et al. [37] numerically analyzed the ejector performance by changing Lm and other geometries and studied the flow phenomena inside the ejector. Dong et al. [38,39] investigated the effects of Lm with the assistance of Mach numbers in an ejector refrigeration system. However, their research works failed to deeply reveal the two-phase flow mechanism with different mixing chamber geometries.
Summarizing the above studies, developing the nozzle is an effective way to improve the performance of an ejector. In previous works the nozzles were rarely improved from the perspective of two-phase flow. Although the two-throat nozzle structure may make the fluid flow from the nozzle more homogeneous and improve ejector performance, the research on two-throat nozzle ejector is still in a preliminary stage. Given the inadequacy of reports on the mixing chamber geometries of two-throat nozzle ejectors, in this study, the performance of two-throat nozzle ejectors varied with different mixing chamber structural parameters including mixing chamber diameter (Dm), Lm and pre-mixing chamber convergence angle (θpm) was analyzed under different operation conditions by a computational fluid dynamics (CFD) method. The simulated method was verified by using experimental data from the published literature [40] and displayed acceptable accuracy. The results were summarized to provide guidance for the design of mixing chamber structures for two-throat nozzle ejectors.

2. Numerical Calculation Model

2.1. Physical Model

This study aimed to investigate the performance of two-throat nozzle ejectors employed in a vessel refrigeration system, which was installed between the intercooler and gas cooler to execute the first throttling process. The elementary structural parameters for two-throat nozzle ejector were computed based on the gas dynamic function method under certain design conditions (Pp of 9.3 MPa, Tp of 35 °C, Ps of 4.1 MPa, Ts of 12 °C, Pc of 4.7 MPa). On the basis of the previous design work by Kang et al. [20], the same parameter as first throat diameter was applied for the second throat diameter. A two-throat nozzle ejector includes a two-throat nozzle, suction chamber, pre-mixing chamber, mixing chamber and diffuser. The main structural diagram and parameter details for the two-throat nozzle ejector are presented in Figure 2 and Table 1, respectively.

2.2. Governing Equations

To simplify the numerical calculation, on basis of the characteristics of the flow inside the ejector and the aims of this study, the following assumptions were made:
(1)
The expansion process of flow kept isentropic.
(2)
The ejector inner wall surface was smooth and adiabatic.
(3)
The flow eventually kept steady.
(4)
The state of secondary inlet flow was ideal-gas.
(5)
Both flows were saturated.
On account of above assumptions, the governing equations including mass, momentum and energy utilized to a control cell in the computational domain are presented on the below:
Mass equation:
ρ t + x i ( ρ u i ) = 0
Momentum equation:
( ρ u i ) t + x i ( ρ u i u j ) = τ i j x j p x i
Energy equation:
( ρ E ) t + x i ( u i ( ρ E + P ) ) = · ( α e f f T x i + u j ( τ i j ) )
where the stress tensor τ i j in Equations (2) and (3) can be expressed as:
τ i j = μ e f f ( u i x j + u j x i 2 3 u k x i δ i j )
where ρ is the density, P is the static pressure, u is the vector velocity, t is the time, τ is the viscous stress, E is the total energy, T is the static temperature, α e f f is the effective heat conductivity, i and j are the space vector directions, μ e f f is the effective dynamic viscosity and δ i j is the Kronecker delta function.

2.3. Turbulence Model

The standard, the standard realizable and renormalization-group (RNG) k-ε models on the basis of Boussinesq hypothesis are employed to handle above governing equations:
ρ u i ' u j ' ¯ = μ t ( u i u j + u j x i ) 2 3 ( ρ k + μ t u i x j ) δ i j

2.4. Simulation Settings

The setting for boundary conditions was that the primary and secondary flow inlet were pressure inlets, the ejector outlet was the pressure outlet, and the wall was no-slip and an adiabatic solid wall. The refrigerant CO2 was chose as working fluid. The primary flow properties parameters of CO2 were confirmed based on NIST REFPROP libraries and the secondary flow was assumed as an ideal gas.
A hexahedral grid mesh was generated using ICEM CFD 19.0 for the two-throat nozzle ejector 3D model. As shown in Figure 3, the mesh model was divided into around 260,000 cells, and the generated mesh around the nozzle and pre-mixing chamber was intensified to make a more precise prediction for the complex flows. The simulated computation was executed by Fluent 19.0, which was based on a finite volume approach. The nonlinear control equations were discretized by the solver in the way of pressure-based, the turbulent flows were controlled by standard k-ε model, and the standard wall function was applied for the near-wall treatment. The second-order upwind scheme was implemented to discretize convective terms. The two-phase flow was assumed to be in homogeneous equilibrium. The residuals for all the equations were below 1 × 10−6. The specific simulation settings are listed in Table 2.

2.5. Grid Independence Test and Simulation Method Validation

The grids of five different quantities were generated for the calculation model, and the results were calculated using same structural ejector model and solution. Figure 4 presents the relationship between Er and grid quantities. From the figure, it is observed that the Er increased with the enlargement of grid quantities firstly, and when the grid quantities were more than 250 thousand, Er nearly kept steady for the increment in grid quantities. For best accuracy and least computational time, the grid division quantity of the calculation model was chosen to be approximately 250,000.
To validate the simulation method, an ejector calculation model possessing the same size as an experimental ejector described in the literature [40] was established, where Figure 5 presents its structural dimensions. The referenced ejector was utilized in a CO2 heat pump water heater system, where the experimental data including primary mass flow rates (ωp) and secondary mass flow rates (ωs) were recorded with the increase of Ps under fixed conditions (Pp of 9.3 MPa, Tp of 35 °C, Ts of 12 °C, Pc of 4.7 MPa). The measured quantity uncertainties of the literature are listed in Table 3. The range uncertainties of the Er are approximately from ±0.023 kg·s1 to ±0.034 kg·s1. A comparison for Er between CFD method and experimental data outcomes is shown in Figure 6. The data indicates that the maximum error of Er between simulation and experiment is only 8.9%, indicating that the accuracy of the simulation method is satisfactory.

3. Results and Discussion

The ejector design process aim is to acquire a high-level mass Er at a settled pressure recovery value to ensure the high-efficiency operation of ejector [41]. Er, as a significant indicator for ejector performance evaluation, reflects the entrainment capacity. The Er is defined as the ωs to the ωp:
E r = ω s ω p
The main purpose of this study is to investigate the influence of the mixing chamber structural parameters including Lm, Dm and θpm on the two-throat nozzle ejector performance under different operation conditions. The primary flow inlet pressures (Pp) were determined to vary from 8.9 to 9.5 MPa, the back pressure (Pc) was fixed at 4.7 MPa, and the secondary flow inlet pressure (Ps) was fixed at 4.1MPa. Table 4 presents the details of the ejector sizes changes in this study.

3.1. Effects of Mixing Chamber Diameter on the Performance of Two-Throat Nozzle Ejector

As presented in Figure 7, the variations of Er in response to the changes in Dm are summarized under different Pp values. It is obvious from the figure that the Er over the entire range of Pp increases initially and then decreases Dm increases. A similar trend in Er was summarized by Fu et al. [1]. In the simulated pressure ranges, the curves representing Er reach their peak values from 0.85 to 0.97 as Dt increases from 8 mm to 9 mm. The broken line shows there exists a tendency that the optimum Dm increases for the rise of Pp. The figure reveals that the similar value around 0.55 in Er among each case as the Dm is 6 mm initially, then the differences in Er are getting more and more obvious for the increment of Dm from 6 mm to 10 mm, which means that the effects of Pp on Er are also affected by Dm. The outcomes indicate that the parameter Dm is of great significance to the two-throat nozzle ejector performance and needs to be cautiously designed within the optimum range.
To further explain the phenomena illustrated in Figure 7, the variations for axial static pressures at Pp of 9.3 MPa are illustrated in Figure 8. It is observed that there exists almost no distinction for axial static pressures among all the cases before the nozzle outlet. The differences of axial static pressures among the cases can be seen when the PF flows into the pre-mixing chamber. With increasing Dm, the axial static pressures for all cases fall firstly and then rise to a different degree inside the pre-mixing chamber. From the pre-mixing chamber section, it is noticed that the rates of increase for axial static pressures decline in response to the enlargement of Dm, which indicates that a better mixing between PF and SF has occured. The reason may be that the enlargement of Dm leads to the increase of flow area, resulting in more SF flooding into the pre-mixing chamber to exchange energy with PF. As the Dm increases more than 8.5 mm, the pre-mixing chamber fails to form a relatively lower pressure area to entrain more SF, then the entrainment performance of the two-throat nozzle ejector becomes worse.
Figure 9 depicts the partial fields of velocity in radial direction for simulated ejectors as Dm changes from 6 mm to 10 mm at Pp of 9.3 MPa. The broken line marks the location of the nozzle outlet. The figure displays the similarity in velocity fields for all the cases inside the two-throat nozzle. It is can be seen that there are not many differences in the velocity fields of the primary flows inside the pre-mixing chamber over the entire cases, but dramatic differences in SF occur on both sides. It is observed that the distance of relative high-speed flow on both sides of SF lengthens initially and then shortens with the further enlargement in Dm, which means that the SF of middle size Dm is well accelerated as compared to a smaller or bigger one. This phenomenon exactly proves the trend of Er in Figure 7. In addition, the figure shows that the PF of bigger Dt maintains high velocity for a shorter distance at mixing chamber, which is confirmed by the pressure differences inside the mixing chamber shown in Figure 8. A possible reason behind this variation could be that different levels of mixing occur between PF and SF.

3.2. Effects of Mixing Chamber Length on the Performance of Two-Throat Nozzle Ejector

The variations of Er as the effects of Lm under different Pp are presented in Figure 10. It is obvious from the figure that the trend lines for Er over the whole range of Pp display a similar tendency. Er increases for the augment in Lm, and reaches a peak value at the Lm from 64 mm to 82 mm, after which the Er gradually diminishes at a relatively slower pace. A similar phenomenon was also discussed by Nakagawa et al. [35].
In terms of Pp at 9.3 MPa, the least Er of 0.891 is recorded for 52 mm, while the greatest Er of 0.924 is recorded for 76 mm. The percentages of increase in Er by varying Lm from 52 mm to 94 mm for four sets of Pp are 2.6%, 3.2%, 3.7% and 3.9%, respectively. The broken line in the figure shows the relationship between optimal Lm and Pp, to be specific, that the optimal Lm of each curve representing Er increases with the rise of Pp. The broken line also indicates that the optimal Lm is not overly sensitive to the change of Pp, and the optimal Lm is recorded from 64 mm to 82 mm.
To clarify the phenomenon of Er summarized in Figure 10, the variations of axial static pressures and the partial fields of velocity in radial direction at Pp of 9.3 MPa are presented in Figure 11 and Figure 12, respectively. It is observed from the Figure 11 that there is almost no distinction for axial static pressures among all the cases before the mixing chamber inlet. The velocity fields before the mixing chamber inlet among all the cases seen in Figure 12 are almost indistinguishable. This demonstrates that the effects of Lm on flow fields before mixing chamber inlet are minimal. The differences of axial static pressures among all the cases start to be noted inside the mixing chamber. It is observed from the local enlarged image around the entrance of mixing chamber in Figure 11 that the axial static pressure representing a Lm of 76 mm is the lowest, while that of 52 mm is the highest, which exactly proves the trend of Er in Figure 10. That is to say, the changes of Pp inside the mixing chamber are primarily responsible for the variations of Er. There are almost no deviations of the velocity fields inside the mixing chamber in any of the cases due to the tiny differences in pressure in Figure 12. To summarize, the relationship between Er and Lm should be taken into account when designing a two-throat nozzle ejector.

3.3. Effects of Pre-Mixing Chamber Converging Angle on the Performance of Two-Throat Nozzle Ejector

As presented in Figure 13, the relationship between the Er and θpm can summarized under different Pp. It can be initially perceived that there is a sharp increment of Er for an enlargement in θpm from 49.7° to 53.5° over the entire range of Pp. Er rises at a relatively lower speed along with the increase of θpm up to around 60° and then attains its peak value. After 60°, the Er starts to fall slightly. The trend of Er is similar to that described in [42]. There exists little difference in the Er trend for Pp at 9.5 MPa. In terms of Pp at 9.3 MPa, the least Er of 0.863 is recorded for 49.7°, while the maximum Er of 0.918 is recorded for 60.4°. The percentages of increment in Er by changing θpm from 49.7° to 69.0° for four sets of Pp are 9.2%, 7.9%, 6.3% and 6.0%, respectively, which indicates that the θpm is of great significance for the performance of two-throat nozzle ejectors. In addition, the figure also depicts that the optimal θpm hardly changes in response to variations of Pp.
The phenomenon of Er variations with θpm in Figure 13 can be further explained with the contribution of axial static pressures distribution and distribution of velocity fields at Pp of 9.3 MPa illustrated in Figure 14 and Figure 15, respectively. It is noticed from Figure 14 that the curves representing axial static pressures over the total scope of cases are of no dissimilarity before nozzle outlet. The axial static pressures observed from pre-mixing chamber decrease with the increment of θpm. It means that the two-throat nozzle ejectors with relatively bigger θpm own better entrainment performance, which is demonstrated by Figure 15 that the SF velocity inside the pre-mixing chamber is accelerated better. From Figure 14 it can be also noticed that the fall rate of Er in pre-mixing chamber slows down with the reduction of θpm. The possible reason for this can be that the decrease of θpm results in the increase of flow area, leading to more SF floods into the pre-mixing chamber to transform energy with PF. Figure 15 presents that the PF of larger θpm keeps high velocity for longer distance at mixing chamber, which is confirmed with the aid of pressure differences inside the mixing chamber in Figure 14. And the possible reason behind this variation could be that the different levels of mixing occur between PF and SF. The relationship between Er and θpm should be taken into account to design high-performance two-throat nozzle ejectors.

4. Conclusions

In this study, 3D models for two-throat nozzle ejectors were developed and numerical simulations were implemented. The performance of a two-throat nozzle ejector changed with different dimensional parameters, including pre-mixing chamber converence angle, mixing chamber diameter and length, was studied under different working conditions. The simulated method was initially validated by actual experimental data from relevant literature, and then 3D models with different mixing chamber structural parameters were generated to investigate the entrainment ratio of the two-throat nozzle ejector. The influences of different mixing chamber geometries on entrainment performance were analyzed by axial static pressures and velocity fields. Based on above simulated results, the following conclusions can be drawn:
(1)
The parameter Dm is of great significance in a two-throat nozzle ejector. When the simulated Pp ranges from 8.9 MPa to 9.5 MPa, Er reaches its peak value from 0.85 to 0.97 corresponding to the Dt from 8 mm to 9 mm and there exists a tendency that the optimum Dm increases as Pp rises.
(2)
The relationship between Er and Lm should be taken into account when designing a two-throat nozzle ejector. When the maximum Er is obtained, the optimal Lm is in the scope of 64 mm-82 mm with the Pp from 8.9 MPa to 9.5 MPa. The optimal Lm of each curve representing Er increases for the rise of Pp.
(3)
The two-throat nozzle ejector performance is sensitive to the changes of θpm. The percentages of increment in Er by changing θpm from 49.7° to 69.0° for 4 sets of Pp are 9.2%, 7.9%, 6.3% and 6.0% respectively, and the optimal θpm hardly changes in response to the variations of Pp.
The results may contribute to the prospective design work of two-throat nozzle ejector mixing chambers in refrigeration systems. The relationship between mixing chamber geometries and performance was analyzed for a certain set of working conditions. There exists an optimum Dm, Lm and θpm for which a two-throat nozzle ejector achieves its maximum entrainment performance. The following research can be developed around the design of two-throat nozzle geometries to strengthen the ejector performance.

Author Contributions

Writing—original draft preparation, F.J.; writing—review and editing, D.Y. and J.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by National 13th Five-Year Plan (2019YFD0901604), Science and Technology Innovation Action Plan of Shanghai Science and Technology Commission (19DZ1207503), Public Service Platform Project of Shanghai Science and Technology Commission (20DZ2292200).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Nomenclature

PStatic pressure, MPa
TStatic temperature, °C
ωMass flow rate, kg·s−1
DDiameter, mm
LLength, mm
θangel, °
δError, %
ρDensity, kg·m−3
uVector velocity, m s−1
ETotal energy
τviscous stress, Pa
tTime, s
Abbreviations
CFDComputational fluid dynamics
ErEntrainment ratio
PFPrimary flow
SFSecondary flow
Subscripts
pPrimary flow inlet
sSecondary flow inlet
cBack flow
mMixing chamber
pmPre-mixing chamber

References

  1. Fu, W.; Liu, Z.; Li, Y.; Wu, H.; Tang, Y. Numerical study for the influences of primary steam nozzle distance and mixing chamber throat diameter on steam ejector performance. Int. J. Therm. Sci. 2018, 132, 509–516. [Google Scholar] [CrossRef]
  2. Bolaji, B.O.; Huan, Z. Ozone depletion and global warming: Case for the use of natural Refrigerant—A review. Renew. Sustain. Energy Rev. 2013, 18, 49–54. [Google Scholar] [CrossRef]
  3. Bodys, J.; Smolka, J.; Palacz, M.; Haida, M.; Banasiak, K.; Nowak, A.J.; Hafner, A. Performance of fixed geometry ejectors with a swirl motion installed in a multi-ejector module of a CO2 refrigeration system. Energy 2016, 117, 620–631. [Google Scholar] [CrossRef]
  4. Gullo, P.; Hafner, A.; Banasiak, K. Transcritical R744 refrigeration systems for supermarket applications: Current status and future perspectives. Int. J. Refrig.-Rev. Int. Du Froid 2018, 93, 269–310. [Google Scholar] [CrossRef]
  5. Karampour, M.; Sawalha, S. Energy efficiency evaluation of integrated CO2 trans-critical system in supermarkets: A field measurements and modelling analysis. Int. J. Refrig.-Rev. Int. Du Froid 2017, 82, 470–486. [Google Scholar] [CrossRef] [Green Version]
  6. Rony, R.U.; Yang, H.J.; Krishnan, S.; Song, J. Recent Advances in Transcritical CO2 (R744) Heat Pump System: A Review. Energies 2019, 12, 457. [Google Scholar] [CrossRef] [Green Version]
  7. Austin, B.T.; Sumathy, K. Transcritical carbon dioxide heat pump systems: A review. Renew. Sustain. Energy Rev. 2011, 15, 4013–4029. [Google Scholar] [CrossRef]
  8. Jin, Z.Q.; Eikevik, T.M.; Neksa, P.; Hafner, A. Investigation on CO2 hybrid ground-coupled heat pumping system under warm climate. Int. J. Refrig.-Rev. Int. Du Froid 2016, 62, 145–152. [Google Scholar] [CrossRef]
  9. Yang, J.Y.; Yu, B.B.; Chen, J.P. Improved genetic algorithm-based prediction of a CO2 micro-channel gas-cooler against experimental data in automobile air conditioning system. Int. J. Refrig.-Rev. Int. Du Froid 2019, 106, 517–525. [Google Scholar] [CrossRef]
  10. Luger, C.; Rieberer, R. Multi-objective design optimization of a rail HVAC CO2 cycle. Int. J. Refrig. 2018, 92, 133–142. [Google Scholar] [CrossRef]
  11. Kimura de Carvalho, B.Y.; Melo, C.; Pereira, R.H. An experimental study on the use of variable capacity two-stage compressors in transcritical carbon dioxide light commercial refrigerating systems. Int. J. Refrig. 2019, 106, 604–615. [Google Scholar] [CrossRef]
  12. Mastrullo, R.; Mauro, A.W.; Perrone, A. A model and simulations to investigate the effects of compressor and fans speeds on the performance of CO2 light commercial refrigerators. Appl. Therm. Eng. 2015, 84, 158–169. [Google Scholar] [CrossRef]
  13. Sian, R.A.; Wang, C.-C. Comparative study for CO2 and R-134a heat pump tumble Dryer—A rational approach. Int. J. Refrig. 2019, 106, 474–491. [Google Scholar] [CrossRef]
  14. Mancini, F.; Minetto, S.; Fornasieri, E. Thermodynamic analysis and experimental investigation of a CO2 household heat pump dryer. Int. J. Refrig. 2011, 34, 851–858. [Google Scholar] [CrossRef]
  15. Purohit, N.; Sharma, V.; Fricke, B.; Gupta, D.K.; Dasgupta, M.S. Parametric analysis and optimization of CO2 trans-critical cycle for chiller application in a warm climate. Appl. Therm. Eng. 2019, 150, 706–719. [Google Scholar] [CrossRef]
  16. Li, Y.; Deng, J.; Ma, L. Experimental study on the primary flow expansion characteristics in transcritical CO2 two-phase ejectors with different primary nozzle diverging angles. Energy 2019, 186, 115839. [Google Scholar] [CrossRef]
  17. Elbel, S.; Lawrence, N. Review of recent developments in advanced ejector technology. Int. J. Refrig. 2016, 62, 1–18. [Google Scholar] [CrossRef]
  18. Yadav, S.K.; Murari Pandey, K.; Gupta, R. Recent advances on principles of working of ejectors: A review. Mater. Today Proc. 2021, 45, 6298–6305. [Google Scholar] [CrossRef]
  19. Li, S.Z.; Li, W.; Liu, Y.J.; Ji, C.; Zhang, J.Z. Experimental Investigation of the Performance and Spray Characteristics of a Supersonic Two-Phase Flow Ejector with Different Structures. Energies 2020, 13, 1166. [Google Scholar] [CrossRef] [Green Version]
  20. Huang, K.; Guo, X.M.; Zhang, P.L. Influence of Structural Parameters of Two-Throat Nozzle Ejector on Performance of Two-phase Flow Ejector Refrigeration System. In 8th International Conference on Applied Energy; Yan, J., Sun, F., Chou, S.K., Desideri, U., Li, H., Campana, P., Xiong, R., Eds.; Elsevier Science Bv: Amsterdam, The Netherlands, 2017; Volume 105, pp. 5091–5097. [Google Scholar]
  21. Wang, X.; Wang, L.; Song, Y.; Deng, J.; Zhan, Y. Optimal design of two-stage ejector for subzero refrigeration system on fishing vessel. Appl. Therm. Eng. 2021, 187, 116565. [Google Scholar] [CrossRef]
  22. Chen, G.; Zhang, R.; Zhu, D.; Chen, S.; Fang, L.; Hao, X. Experimental study on two-stage ejector refrigeration system driven by two heat sources. Int. J. Refrig. 2017, 74, 295–303. [Google Scholar] [CrossRef]
  23. Chang, Y.J.; Chen, Y.M. Enhancement of a steam-jet refrigerator using a novel application of the petal nozzle. Exp. Therm. Fluid Sci. 2000, 22, 203–211. [Google Scholar] [CrossRef]
  24. Yang, X.; Long, X.; Yao, X. Numerical investigation on the mixing process in a steam ejector with different nozzle structures. Int. J. Therm. Sci. 2012, 56, 95–106. [Google Scholar] [CrossRef]
  25. Wang, X.; Yu, J. Experimental investigation on two-phase driven ejector performance in a novel ejector enhanced refrigeration system. Energy Convers. Manag. 2016, 111, 391–400. [Google Scholar] [CrossRef]
  26. Banu, J.P.; Mani, A. Numerical studies on ejector with swirl generator. Int. J. Therm. Sci. 2019, 137, 589–600. [Google Scholar] [CrossRef]
  27. Zhou, M.; Wang, X.; Yu, J. Theoretical study on a novel dual-nozzle ejector enhanced refrigeration cycle for household refrigerator-freezers. Energy Convers. Manag. 2013, 73, 278–284. [Google Scholar] [CrossRef]
  28. Rao, S.M.V.; Jagadeesh, G. Novel supersonic nozzles for mixing enhancement in supersonic ejectors. Appl. Therm. Eng. 2014, 71, 62–71. [Google Scholar] [CrossRef]
  29. Chen, X.; Omer, S.; Worall, M.; Riffat, S. Recent developments in ejector refrigeration technologies. Renew. Sustain. Energy Rev. 2013, 19, 629–651. [Google Scholar] [CrossRef]
  30. He, S.; Li, Y.; Wang, R.Z. Progress of mathematical modeling on ejectors. Renew. Sustain. Energy Rev. 2009, 13, 1760–1780. [Google Scholar] [CrossRef]
  31. Mohamed, S.; Shatilla, Y.; Zhang, T. CFD-based design and simulation of hydrocarbon ejector for cooling. Energy 2019, 167, 346–358. [Google Scholar] [CrossRef]
  32. Kim, S.; Jeon, Y.; Chung, H.J.; Kim, Y.J.A.T.E. Performance optimization of an R410A air-conditioner with a dual evaporator ejector cycle based on cooling seasonal performance factor. Appl. Therm. Eng. 2018, 131, 988–997. [Google Scholar] [CrossRef]
  33. Wang, L.; Yan, J.; Wang, C.; Li, X.B. Numerical study on optimization of ejector primary nozzle geometries. Int. J. Refrig.-Rev. Int. Du Froid 2017, 76, 219–229. [Google Scholar] [CrossRef]
  34. Hou, W.; Wang, L.; Yan, J.; Li, X.; Wang, L. Simulation on the performance of ejector in a parallel hybrid ejector-based refrigerator-freezer cooling cycle. Energy Convers. Manag. 2017, 143, 440–447. [Google Scholar] [CrossRef]
  35. Nakagawa, M.; Marasigan, A.R.; Matsukawa, T.; Kurashina, A. Experimental investigation on the effect of mixing length on the performance of two-phase ejector for CO2 refrigeration cycle with and without heat exchanger. Int. J. Refrig. 2011, 34, 1604–1613. [Google Scholar] [CrossRef]
  36. Zhang, H.; Wang, L.; Yan, J.; Li, X.; Wang, L. Performance investigation of a novel EEV-based ejector for refrigerator-freezers. Appl. Therm. Eng. 2017, 121, 336–343. [Google Scholar] [CrossRef]
  37. Wu, H.; Liu, Z.; Han, B.; Li, Y. Numerical investigation of the influences of mixing chamber geometries on steam ejector performance. Desalination 2014, 353, 15–20. [Google Scholar] [CrossRef]
  38. Dong, J.; Chen, X.; Wang, W.; Kang, C.; Ma, H. An experimental investigation of steam ejector refrigeration system powered by extra low temperature heat source. Int. Commun. Heat Mass Transf. 2017, 81, 250–256. [Google Scholar] [CrossRef] [Green Version]
  39. Dong, J.; Hu, Q.; Yu, M.; Han, Z.; Cui, W.; Liang, D.; Ma, H.; Pan, X. Numerical investigation on the influence of mixing chamber length on steam ejector performance. Appl. Therm. Eng. 2020, 174, 115204. [Google Scholar] [CrossRef]
  40. Guangming, C.; Xiaoxiao, X.; Shuang, L.; Lixia, L.; Liming, T. An experimental and theoretical study of a CO2 ejector. Int. J. Refrig. 2010, 33, 915–921. [Google Scholar] [CrossRef]
  41. Palacz, M.; Haida, M.; Smolka, J.; Plis, M.; Nowak, A.J.; Banasiak, K. A gas ejector for CO2 supercritical cycles. Energy 2018, 163, 1207–1216. [Google Scholar] [CrossRef]
  42. Ramesh, A.S.; Sekhar, S.J. Experimental and numerical investigations on the effect of suction chamber angle and nozzle exit position of a steam-jet ejector. Energy 2018, 164, 1097–1113. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a two-throat nozzle and a Laval nozzle.
Figure 1. Schematic diagram of a two-throat nozzle and a Laval nozzle.
Energies 14 06900 g001
Figure 2. Structural diagram of the two-throat nozzle ejector.
Figure 2. Structural diagram of the two-throat nozzle ejector.
Energies 14 06900 g002
Figure 3. Grid of the two-throat nozzle ejector 3D model.
Figure 3. Grid of the two-throat nozzle ejector 3D model.
Energies 14 06900 g003
Figure 4. Variation of Er with the number of grids.
Figure 4. Variation of Er with the number of grids.
Energies 14 06900 g004
Figure 5. Structural sizes of ejector from the published literature [40].
Figure 5. Structural sizes of ejector from the published literature [40].
Energies 14 06900 g005
Figure 6. The comparison of Er between CFD and experimental data from the published literature [40].
Figure 6. The comparison of Er between CFD and experimental data from the published literature [40].
Energies 14 06900 g006
Figure 7. Variations of Er with different Dm and Pp (Lm of 64 mm, θpm of 60.4°).
Figure 7. Variations of Er with different Dm and Pp (Lm of 64 mm, θpm of 60.4°).
Energies 14 06900 g007
Figure 8. Variations of axial static pressures under different Dm (Pp of 9.3 MPa).
Figure 8. Variations of axial static pressures under different Dm (Pp of 9.3 MPa).
Energies 14 06900 g008
Figure 9. Fields of velocity under different Dm (Pp of 9.3 MPa).
Figure 9. Fields of velocity under different Dm (Pp of 9.3 MPa).
Energies 14 06900 g009
Figure 10. Variations of Er with different Lm and Pp (Dm of 8 mm, θpm of 60.4°).
Figure 10. Variations of Er with different Lm and Pp (Dm of 8 mm, θpm of 60.4°).
Energies 14 06900 g010
Figure 11. Variations of axial static pressures under different Lm (Pp of 9.3 MPa).
Figure 11. Variations of axial static pressures under different Lm (Pp of 9.3 MPa).
Energies 14 06900 g011
Figure 12. Fields of velocity under different Lm (Pp of 9.3 MPa).
Figure 12. Fields of velocity under different Lm (Pp of 9.3 MPa).
Energies 14 06900 g012
Figure 13. Variations of Er with different θpm and Pp (Dm of 8 mm, Lm of 64 mm).
Figure 13. Variations of Er with different θpm and Pp (Dm of 8 mm, Lm of 64 mm).
Energies 14 06900 g013
Figure 14. Variations of axial static pressure under different θpm (Pp of 9.3 MPa).
Figure 14. Variations of axial static pressure under different θpm (Pp of 9.3 MPa).
Energies 14 06900 g014
Figure 15. Fields of velocity under different θpm (Pp of 9.3 MPa).
Figure 15. Fields of velocity under different θpm (Pp of 9.3 MPa).
Energies 14 06900 g015
Table 1. Key structural sizes for the two-throat nozzle ejector.
Table 1. Key structural sizes for the two-throat nozzle ejector.
Geometric ObjectsValue/mmGeometric ObjectsValue/mm
Nozzle inlet diameter/Dp17Diameter of the second throat/Dst3.4
Suction nozzle diameter/Ds34Second nozzle outlet diameter/Dso3.6
Diameter of the first throat/Dft3.6Diffuser outlet diameter/Dd28
First nozzle outlet diameter/Dfo4.8Length of diffuser/Ld72
Table 2. Simulation parameters.
Table 2. Simulation parameters.
Primary flow inletPressure inlet
Secondary flow inletPressure inlet
OutletPressure outlet
WallNo-slip
SolverPressure-based
TimeSteady
Turbulence modelStandard k-ε
Near-wall treatmentStandard wall functions
MaterialsCO2
Pressure-velocity couplingSIMPLEC
Convective termsSecond-order upwind
Convergence criteria1 × 10−6
Table 3. Measurement uncertainties of instrument from the published literature [40].
Table 3. Measurement uncertainties of instrument from the published literature [40].
ParametersUncertaintyUnit
Temperature±0.5°C
Pressure (refrigerant)±0.032MPa
Mass flow rate (refrigerant)±0.001Kg·s−1
Table 4. Size changes for ejector.
Table 4. Size changes for ejector.
PressureStructural Parameter
Pp (MPa)Ps (MPa)Pc (MPa)Dm (mm)Lm (mm)θpm (°)
8.9~9.54.14.76.0~10.052~9449.7~69.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jia, F.; Yang, D.; Xie, J. Numerical Investigation on the Performance of Two-Throat Nozzle Ejectors with Different Mixing Chamber Structural Parameters. Energies 2021, 14, 6900. https://doi.org/10.3390/en14216900

AMA Style

Jia F, Yang D, Xie J. Numerical Investigation on the Performance of Two-Throat Nozzle Ejectors with Different Mixing Chamber Structural Parameters. Energies. 2021; 14(21):6900. https://doi.org/10.3390/en14216900

Chicago/Turabian Style

Jia, Fatong, Dazhang Yang, and Jing Xie. 2021. "Numerical Investigation on the Performance of Two-Throat Nozzle Ejectors with Different Mixing Chamber Structural Parameters" Energies 14, no. 21: 6900. https://doi.org/10.3390/en14216900

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