Accepted Manuscript Computational Fluid-Dynamics modeling of supersonic ejectors: screening of turbulence modeling approaches Giorgio Besagni, Fabio Inzoli PII: DOI: Reference:
S1359-4311(16)32829-0 http://dx.doi.org/10.1016/j.applthermaleng.2017.02.011 ATE 9891
To appear in:
Applied Thermal Engineering
Received Date: Revised Date: Accepted Date:
28 October 2016 22 January 2017 4 February 2017
Please cite this article as: G. Besagni, F. Inzoli, Computational Fluid-Dynamics modeling of supersonic ejectors: screening of turbulence modeling approaches, Applied Thermal Engineering (2017), doi: http://dx.doi.org/10.1016/ j.applthermaleng.2017.02.011
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Computational Fluid-Dynamics modeling of supersonic ejectors: screening of turbulence modeling approaches Giorgio Besagni*, Fabio Inzoli Politecnico di Milano, Department of Energy, Via Lambruschini 4, 20156 *Corresponding author: Giorgio Besagni, +39-02-2399-3826;
[email protected]; Address: Politecnico di Milano, Department of Energy Via Lambruschini 4a, 21056 Milan, Italy.
ABSTRACT Ejector refrigeration has not been able to penetrate the market because of the low coefficient of performance and because of the high influence of ejector performance on the efficiency of the refrigeration system. Improving the performance of ejector refrigeration systems relies on the understanding of the fluid dynamic phenomena, which can be obtained through computational fluid-dynamics approaches. Unfortunately, no agreement has been reached yet on the closures for the turbulence modeling in the Reynolds Averaged Navier-Stokes (RANS) approach. This paper contributes to the existing discussion by presenting a numerical study of the turbulent compressible fluid in a supersonic ejector. Seven RANS turbulence closures have been compared and, in addition, the turbulence models were tested under different near-wall modelling options in order to investigate the wall treatment effect on the numerical results. The numerical results have been validated by literature data consisting in entrainment ratio and wall static pressure measurements, for different ejector geometries and operating conditions. As a result, the k– omega model shows better performance in terms of global and local flow phenomena predictions for the different ejector geometries and operating conditions: (a) on the global point of view, the entrainment ratio is well predicted with a maximum relative error equal to about 10%, (b) on the local point of view, the shock wave position, the pressure recovery and the wall static pressure values are well predicted. The results, taking into account previous numerical studies, suggest the use of the k–omega model to simulate ejector fluid dynamics.
Keywords: CFD; Ejector refrigeration; Computational Fluid Dynamics; Turbulence models; Modeling
1
NOMENCLATURE Acronyms CFD
Computational Fluid-dynamics
CFL
Courant–Friedrichs–Lewy
COP
Coefficient of performance
EWT
Enhanced-Wall-Treatment
FMG
Full multi-grid initialization
RANS
Reynolds Averaged Navier-Stokes
RMSE
Root mean square error, Eq. (2)
SWF
Standard-Wall-Function
Symbols ER
Relative error [-]
[-]
m
Mass flow rate in Figure 1a
[kg/s]
I
Turbulence intensity
[-]
p
Pressure
[Pa]
Q
Rate of heat in Figure 1b
[W]
T
Temperature
[°C]
Wp
Pump power in Figure 1b
[W]
X
Generic variable in Eq. (1) and Eq. (2)
Subscripts c
Condenser
CFD
Calculated value from the CFD model
e
Evaporator
exp
Experimental measurement
g
Generator
p
Primary flow
s
Secondary flow
w
Wall static measurement
2
1
INTRODUCTION
Ejector technology is receiving growing attention in thermal engineering and in refrigeration industries (e.g., building, industrial and commercial refrigeration—see, for example, refs. [1-3]) thanks to the many advantages: an ejector (Figure 1a) provides a combined effect of compression, mixing and entrainment, with no-moving parts and without limitations concerning working fluids. On the practical point of view, the simplicity of construction, the lack of any mechanically operated parts, the reliability, the little maintenance, the low cost and the long lifespan are some of the practical advantages in ejector technology. Ejector technology can be applied in building, industrial and commercial refrigeration: (a) in buildings refrigeration, heat-driven ejector refrigeration would allow a significant reduction of energy consumption (cooling energy and electricity demand); (b) in industrial refrigeration, heat-driven ejector refrigeration may contribute to the energy efficiency, by exploiting the low grade energy sources [4, 5]; (c) in commercial refrigeration, owing to the recent regulations concerning refrigerants (see ref. [6]), ejectors are receiving growing attention, to overcome the low efficiency of the trans-critical R744 systems in warm climates [7-9]. The simplest ejector refrigeration system is shown in Figure 1b and consists of a generator, a condenser, an evaporator, an ejector, a circulation pump and a throttle valve. The heat energy is delivered to the generator for the working fluid vaporization; subsequently, the high-pressure vapor (“primary flow”) flows out from the generator, enters inside the motive nozzle, and draws low-pressure vapor (“secondary flow”) from the evaporator. The primary and the secondary flows mix, the pressure of the mixed stream is raised in the diffuser, and the flow reaches the condenser where it changes phases from vapor to liquid rejecting heat. Once condensation takes place, the flow is splitted in two parts, one heading to the generator and the ejector and another to the evaporator. For further details, the reader may refer to the most recent reviews (see refs. [3, 10-12]) for a comprehensive literature survey concerning the different ejector refrigeration systems. Despite the simple component design, ejector refrigeration technologies were not been able to penetrate the market because of (a) the low coefficient of performance (COP) and (b) the high influence of the ejector performance on the efficiency of the whole system, especially in off-design operating conditions [12]. Indeed, the COP of the ejector-based system (the “refrigeration-system-scale”) relies, mainly, on the ejector performance (the “component-scale”, i.e., the primary and secondary mass flow rates, the pressure recovery,, the operation modes, …). Unfortunately, ejectors are characterized by extremely complex fluid dynamic interactions and, for this reasons, their correct design, operation, and scale-up (towards large-scale industrial applications) rely on the knowledge of the fluid dynamics at the “local-
3
scale” (i.e., boundary layers subject to adverse pressure gradients, shock waves, under-expanded jets, flow reparation, recirculation, turbulence mixing phenomena bounded by near-wall regions,…) and at the “component-scale” (i.e., the entrainment ratio—defined as the ratio between the secondary and the primary mass flow rates; it measures the recirculation ratio and it is a measure of the performance of the ejector). Indeed, the global behavior, at the “component-scale, is the result of the flow features inside the ejector at the “local-scale”. The many relationships between the different scales makes the estimation of ejector fluid dynamics a very challenging task; this task is even more complex owing to the many relationships between the ejector fluid dynamics and the various variables characterizing the system (i.e., geometrical parameters, operating temperature and pressures, refrigerant properties, …). In this respect, the prediction of ejector local flow phenomena is actually matter of intensive research and is usually approached by using Computational Fluid-Dynamics (CFD) method, within the Reynolds Averaged NavierStokes (RANS) approach. Unfortunately, RANS approaches are generally validated with global fluid dynamic properties (i.e. the entrainment ratio) [13, 14] in a limited range of operating conditions. However, local parameters are needed to verify the correspondence between the experiments and the numerical simulations: indeed validated CFD model may be able to correctly determine the global parameters, but it might not accurately estimate the local flow phenomena (i.e. mixing losses, friction losses, …) [15]. In particular, this issues was demonstrated by Hemidi et al. [16]: the k-ε Standard and k-ω SST predicted similar entrainment ratios (global flow features), but different local flow phenomena. Indeed, the ejector flow phenomena concerns shock waves, under-expanded jet and flow reparation requiring the validation performed with local data and the evaluation of different RANS models before studying a broader range of operating conditions [17-21]. In this respect, different studies were proposed in the last decades and a brief summary is proposed in the following. Bartosiewicz et al. [22] compared the performance of six RANS turbulence models (k-ε Standard, k-ε RNG, k-ε Realizable, k-ω Standard , k-ω SST, RSM). The k-ε RNG and k-ω Standard best predict the shock phase, strength, and the mean line of pressure recovery. The validated model has been used to reproduce different operation modes of a supersonic ejector [18]. Bouhanguel et al. [23] compared the performance of several RANS turbulence models in a supersonic air ejector. The results strongly depended on turbulence modeling and the authors observed that none of the RANS turbulence models tested was able to capture the shock reflection pattern in the nozzle exit region. ElBehery and Hammed [24] studied different turbulence models for the case of axi-symmetric diffuser. They have shown that the k–ω Standard, k–ω SST and v2-f models clearly performed better than other models when an adverse
4
pressure gradient was present. The RSM model shows an acceptable agreement with the velocity and turbulent kinetic energy profiles but it fails to predict the location of separation and attachment points. The k-ε Standard and the low-Re k-ε models give very poor results. Dvorak and Vit [25] compared different turbulence models with hot wire anemometry experimental measurements in terms of the static pressure at wall, velocity profile and turbulence intensity. They suggested the k-ε Realizable as the most promising turbulence model. Kolar and Dvorak [26] verified the k–ω SST turbulence model by comparison with experimental Schlieren picture: they found good agreement of the shock wave prediction and boundary layer separation, but the shear stress between streams was over-predicted. Bouhanguel et al. [27] used the laser sheet method to investigate shock structure, flow instabilities, and mixing process. The results of this visualization were used for the validation of the turbulence models by the same authors in ref. [23]. Hemidi et al. [28] compared the k-ε Standard and k-ω SST models for supersonic ejector working with air. The k-ε model provided better results for on-design conditions, with errors mostly less than 10%. The k-ω SST model, instead, yields errors often more than 20%, but with better prediction at off-design operation conditions. Thus, for the analysis of ejector performance in a wide range of operating conditions the k-ω SST or other models should be considered. In addition, Hemidi et al. [16], in the second part of their study, found that the k-ε Standard and k-ω SST predicted similar entrainment ratios (global flow features), but different local flow phenomena. C. Li and Y.Z. Li [29] investigated the behavior and performance of single-phase and two-phase ejectors with different working fluids. In their analysis, the k-ε Standard, the k-ω Standard and the k-ω SST models were considered. The k-ε model provides better results in terms of entrainment ratio prediction: the relative errors were less then 25% in almost all cases. Moreover, it appears that the k-ω SST model significantly over-predicts the entrainment ratio for small pressure difference between the entrained pressure and the discharge pressure. Ruangtrakoon, Thongtip et al. [30] investigated the effect of the primary nozzle geometries on the performance of an ejector. In this study, the k-ε Realizable and k-ω SST models were used and compared. It was found that the simulated results based on the k-ω SST model more closely corresponded to the experimental values than those based on the k-ε Realizable model. According to the authors, a possible reason is that the k-ε Realizable model is unable to accurately predict the performance of the ejector under a strong adverse pressure gradient. Gagan et al. [19] used the PIV technique for a supersonic ejector flow visualization and compared it with numerical results obtained for various turbulence models. They recommended the k-ε Standard model. The k-ε Standard, the k-ε Realizable and the RSM models predict accurately the entrainment ratio. The k-ε RNG, the RSM as well as the k–ω Standard do not predict vortex a region
5
downstream nozzle. The k-ε Realizable and the k–ω SST model under predict the scale of the vortex area. Zhu and Jiang [31] investigated the entrainment performance and the shockwave structures in a 3D ejector. Four turbulence models are used: k-ε Standard, k-ε RNG, k-ε Realizable and k-ω SST. The results show that the k-ε RNG and k-ω SST models agree best with the experimental entrainment ratio and shockwave structures. The k-ε Standard model fails in predicting shock location after the second shock and, in the critical operating conditions, the k-ε Standard and k-ε Realizable models over-predict the first shock wavelength and fail in predicting the reflected shocks. Croquer et al. [32, 33] investigated a R134a ejector and compared four different turbulence models (k–ε Standard, k–ε Realizable, k– ω SST, and stress–ω RSM): they concluded that k–ω SST showed the best performance. The authors also tested different approaches to model equations of state of refrigerants. Mazzelli et al. [34] investigated a rectangular crosssection ejector and compared four different turbulence models (k–ε Standard, k–ε Realizable, k–ω SST, and stress–ω RSM): they found that the entrainment ratio was reasonably predicted at on-design operating conditions, whereas the numerical model poorly predicts the off-design operation (i.e., the entrainment ratio and the critical back pressure). Among the different turbulence models, the k–ω SST showed the best performance. At present, there are no clear recommendations on the selection of the numerical approaches to simulate the ejector fluid dynamics. In our previous paper we proposed a comparison of several turbulence closures (under different nearwall modelling options) for a subsonic ejector [35]. This paper further contributes to the present discussion and extends the previous numerical results to the case of a supersonic ejector by considering both global and local flow features. This work provides guidelines and a rational basis to simulate supersonic ejectors to be applied in refrigeration systems. In this respect, it is worth noting that future studies will consider benchmarks with other refrigerants. The paper is structured as follows. In Section 2 the numerical benchmark is presented. In Section 3 the CFD approach is outlined, and in Section 4 the numerical results are presented and discussed. First, the sensitivity analysis to set up the numerical model are presented (grid independency and near wall modeling). Secondly, the numerical approach is tested considering all the operating conditions of the benchmark used. Finally, the main conclusions are outlined.
2
THE NUMERICAL BENCHMARK
The validation of the numerical model has been performed by using the experimental dataset provided by Sriveerakul et al. [36]. In this study, both global (the entrainment ratio, representing the “component-scale”) and local (the wall
6
static pressure along the ejector, representing the “local-scale”) measurements are available for a complete validation of the approach (as discussed in the introduction). It is worth noting that wall static pressure measurements are considered local measurements as they highly depends on the local flow properties, i.e., boundary layers subject to adverse pressure gradients, shock waves, under-expanded jets, flow reparation, recirculation, turbulence mixing phenomena bounded by near-wall regions. For this reason, the proposed benchmark allows a complete validation of the CFD approach, owing to the multi-scale dataset employed. Unfortunately, the proposed benchmark does not consider local velocity profiles (as the data employed in our previous study, ref. [35]): future studies will be devoted to extend the screening of turbulence approaches proposed in the following to other numerical benchmarks. The experimental investigation of Sriveerakul et al. [36] considered a steam ejector that works under different working conditions (Tg = 120÷130°C, Te = 5÷15°C and Tc = 24÷40°C; where the subscripts stand for generator, evaporator and condensed respectively – See Figure 1b) and with different geometries. The main geometrical characteristics of the ejector and the operating conditions are listed in Table 1 and Table 2, respectively. The reader should refer to the original reference [36] for a complete description of the cases tested.
3
NUMERICAL MODEL
3.1 Mesh generation The mesh generations process has been handled with GAMBIT 2.4.6. A 2D axi-symmetric domain has been applied for modeling the ejector, based on its nature. Concerning the 2D axi-symmetric, the reader may also refer to the discussion proposed by Croquer et al. [32]. The three ejector geometries considered (Table 1) are modeled using three meshes and, after a grid independency study, the final meshes are composed of approximately 70000 quadrilateral elements and were refined on proximity of the wall. The independence of the results from the grid density has been analyzed and the results are presented in Section 4.1 and in Table 3 the mesh quality parameters are summarized. Please note that the dynamic solution-adaptive mesh refinement, performed during the simulations, ensures a good representation of the local flow phenomena (oblique shock-waves, boundary layer, mixing process). This point was also discussed by Besagni et al. [35].
7
3.2 Numerical setting 3.2.1 Solver The fluid flow in the ejector is compressible and turbulent and, in the present case, it is considered as steady state. The relationship between density, velocity and temperature is given by the conservation of continuity, momentum, energy equations, in the form of a set of partial differential equations. In this paper, the finite volume commercial code ANSYS Fluent Release 15.0 has been used to solve the steady state 2D axisymmetric Reynolds Averaged NavierStokes (RANS) equations for the turbulent compressible Newtonian fluid flow. The governing equations have been discretized as follows. Second order numerical schemes have been used for the spatial discretization, in order to limit the numerical diffusion. Second order schemes also for the turbulence quantities have been used. Gradients are evaluated by a least-squares approach. After the discretization process of the governing equations, a system of algebraic equations is obtained and it has been solved: due to the strong coupling between mass, momentum and energy equations, determined by the presence of a high speed compressible flow, a density-based coupled solver has been chosen and applied. IN ANSYS FLUENT, both density-based and pressure-based algorithms are available: a comparative discussion concerning density-based and pressure-based algorithms to solve ejector fluid dynamics was provided by Croquer et al. [32]. The Coupled algorithm applied solves the continuity, momentum, and energy (density is computed through the equation of state selected (See Section 3.2.3)) equations simultaneously. The formulation of the coupled algorithm requires setting up a Courant–Friedrichs–Lewy (CFL) condition; it was set CFL = 0.1 in order to run a stable simulation. Because of to the complex fluid dynamics and in order to achieve fast convergence, a full multi-grid (FMG) initialization scheme has been adopted.
3.2.2 Turbulence modeling and near wall treatment The turbulent behavior has been treated using the RANS approach. Seven RANS turbulence models have been tested and compared: Spalart-Allmaras, k-ε Standard, k-ε Realizable, k-ε RNG, k-ω Standard, k-ω Standard and RSM. These models are well known in the literature and a detailed description of their mathematical structure is beyond the scope of the present paper. For further information about their mathematical formulation the reader should refer to the discussions proposed by Wilcox [37]; for further information about their mathematical formulation and their application in ejector modeling the reader may refer to Mazzelli et al. [34] and Croquer et al. [32] proposed a brief and complete summary on the main turbulence models; For further information about their implementation in the
8
commercial code ANSYS-Fluent the reader should refer to ref. [38]. All the considered models have been used in their original implementation without further modifications. Beside turbulence models, the near wall formulation should be considered. The Spalart-Allmaras, k-ω Standard, k-ω Standard models do not need a near wall treatment because their mathematical structure already emphasizes on the flow close to the wall. However, some aspects, in their ANSYS FLUENT implementations, have to be considered. Concerning the Spalart-Allmaras model, its boundary conditions have been implemented so that the model may work on coarse mesh (such as would be appropriate for the wall function approach) and on a fine mesh in its low-Reynolds formulation. Concerning the k-ω Standard and the k-ω SST, a low-Reynolds implementation may be enabled in ANSYS FLUENT for fully resolver near wall grids. Conversely, the k-ε Standard, k-ε Realizable, k-ε RNG and RSM models require a near wall modeling. In the present study, a sensitivity analysis on the following wall treatments has been performed: •
Standard-Wall-Function (SWF). The viscous sub-layer is not resolved and the first grid cell need to be in the region 30 < y+ < 300. It becomes less reliable when the local equilibrium assumption is not valid (i.e. strong pressure gradient, large curvature, highly 3D flow).
•
Non-Equilibrium-Wall-Function. The Non-Equilibrium-Wall-Function has the same validity range of the SWF approach, but relaxes the local equilibrium assumption in the turbulent region of the wall-neighboring cells.
•
Enhanced-Wall-Treatment (EWT). The Enhanced-Wall-Treatment requires a very fine mesh near the wall, to +
resolve the viscous sub-layer (y < 1). This approach is indicated for low-Reynolds flows or flows with complex near-wall phenomena. The detailed investigation and the results of the sensitivity analysis on the near wall approaches can be found in section 4.1.2. As a result of the sensitivity analysis, the Standard Wall Functions approach is suggested and applied in the comparison of the turbulence models in Section 4.2. For the sake of clearness, here we can just state that the use of wall functions coupled with high-Reynolds turbulence models is based on the absence of detachment, reattachment of the flow, adverse pressure gradient and general non-equilibrium conditions. The absence of these phenomena allows using a y+ value higher than unity reducing the computational costs and high volume aspect ratio close to the walls. This observation agrees with the results discussed in ref. [35] .
9
3.2.3 Working fluid The working fluid employed in the benchmark is water vapour and, because of the low operating pressure, the ideal gas assumption has been considered. Furthermore, no phase transition/condensation effect has been considered. The thermophysical properties of water vapour, provided by the Fluent database, have been used and have been considered as constant (see Error! Reference source not found.Table 4); the density has been evaluated using the ideal gas law.
3.2.4 Boundary conditions The boundary conditions at the inlets and at the outlet (See the notations in Figure 1a and Figure 1b) have been assigned as the saturation temperatures and saturation pressures, accordingly to the benchmark operating conditions (as displayed in Table 2). In particular, pressure-based boundary conditions (in the ANSYS FLUENT implementation) have been applied at the primary flow inlet, at the secondary flow inlet, as well as at the outlet section (see the notation in Figure 1b). In order to ensure a stable solution, the value of the pressure-outlet boundary condition was discretely increased until convergence during the initialization process. The boundary conditions at the walls have been set as adiabatic and no-slip. The turbulence intensity and hydraulic diameter have been chosen as turbulence boundary conditions; unfortunately, no turbulence measurements have been performed and the same approach as the one described in our previous paper [35] has been applied: (i) the hydraulic diameter and (ii) a turbulence intensity equal to I = 5% for the primary flow and equal to I = 2% for the secondary flow. Future studies will be devoted to study the effects of uncertainties on turbulence boundary conditions on the prediction of the numerical model (See, for example, ref. [39]).
3.3 Convergence criteria The numerical solution is considered as converged when all the following converging criteria were satisfied simultaneously: (a) a decrease in numerical residuals by six orders of magnitude; (b) the normalized difference of mass -7
flow rates at the inlet and at the outlet passing through the modelled ejector had to be less than 10 ; (c) the areaweighed-averaged value for inlet velocity of primary and secondary flow is constant. Please refer also to the discussion concerning physical and numerical convergence in ref. [35].
4
NUMERICAL RESULTS
Herein the numerical results are presented and discussed considering both the global (i.e., the entrainment ratio – the “component-scale”) and the local (i.e., wall static pressure along the ejector - the “local-scale”) quantities. First, the
10
sensitivity analysis to set up the numerical model are presented. Secondly, the numerical approach is tested considering all the operating conditions of the benchmark (Table 2). Finally, some considerations on the convergence capability and the computational effort are proposed, for the sake of clarity. The effectiveness of the numerical results (compared with the corresponding experimental data) has been evaluated in terms of the relative error ER defines as follows:
ER ( X ) =
Xexp − XCFD Xexp
⋅100
(1)
Where X is the considered variable, Xexp and XCFD are the experimental measurement and the CFD model estimates, respectively. Moreover, to better asses the validity of the turbulence models, the mean squared deviation has been computed as follows:
RMSE( X) = ∑( XCFD − Xexp )
2
(2)
n
4.1 Sensitivity anlaysis In this section, the sensitivity analyses are presented and discussed: first, (a) the mesh independency analysis and, second, (b) the near-wall treatment sensitivity analysis. These sensitivity analyses are needed to set up the numerical model that has been used for a broader range of operating conditions in Section 4.2. The results are discussed considering the global (i.e., entrainment ratio - Table 5 and Table 6) and the local (i.e., wall static pressure profiles Figure 2 and Figure 3) measurements.
4.1.1 Grid independency analysis In this study, mesh grids having approximately 70000 quadrilateral elements have been used for the detailed analyses performed in Section 4.2. The applied mesh density is the result of a grid convergence analysis, carried out to obtain a numerical solution not affected by discretization errors. To this end, simulations have been performed on three different meshes (see the details in Table 3): (a) “coarse” mesh (approximately 40000 quadrilateral elements); (b) “medium” mesh (approximately 70000 quadrilateral elements); (c) “fine” mesh (approximately 280000 quadrilateral elements). Please notice that the above-mentioned mesh densities do not consider the mesh refinement procedure considered during the simulations. In order to assess the grid sensitivity of the results, the simulations corresponding to the case “RUN 1” have been performed with the three different meshes (the “coarse”, the “medium” and the “fine”
11
mesh) and considering all seven turbulence models. The results of the grid convergence analysis are summarized in Table 5 (global quantities: the entrainment ratio – the “component-scale”) and in Figure 2 (local quantities: wall static pressure along the ejector - the “local-scale”). Considering the entrainment ratios (Table 5), the results enlighten that an asymptote has been reached with ”medium” and the ”fine” meshes for all turbulence models, expect for the Spalart-Allmaras model. However, in the case of the Spalart-Allmaras model, the variation of the predicted entrainment ratio between the “medium” mesh and the “fine” mesh is almost negligible and within the expected range of the experimental uncertainty. Similarly, an asymptotic convergence is obtained by considering the wall static pressure profiles for all turbulence models, except for the Spalart-Allmaras model. In particular, agreement between the “medium” mesh and the “fine” mesh is very good (the results for the “medium” and the “fine” meshes are almost overlapping) in the case of: (a) k-ε Standard (b) k-ε Realizable, (c) k-ω Standard, (d) k-ω Standard, (e) k-ω SST and (f) RSM. Conversely, a fair agreement between the “medium” mesh and the “fine” mesh is observed for the k-ε RNG model and slightly worst agreement is observed in the case of Spalart-Allmaras model. To explain these performances of the Spalart-Allmaras model and, in general, when discussing about non-asymptotic convergence, it is worth noting that the high degree of coupling between the equations and closure relations together with the inherent non-linearity of the equations are such that the results hardly display a monotonic convergence of the resolved variables, regardless of the mesh resolutions used. In conclusion, the grid independency has been verified for the 70000 “medium” grid for all the turbulence models: therefore, the ”medium” mesh represents the right trade-off between accuracy and number of cells.
4.1.2 Wall treatment analysis To investigate the effects of the near-wall treatments on the numerical results, a near-wall treatment sensitivity analysis has been performed. In particular, the near-wall treatment sensitivity analysis has been performed on two meshes: (a) the “medium” mesh (described above and in Table 3) with the grid point spacing at approximately y+ ≈ 40, and the fine mesh (y+ < 1 ≈ 0.25 to 0.75), derived from the “medium” one by refining the region near the wall boundaries. The first mesh has been used to study the wall-function approaches: (a) the Standard-Wall-Function (SWF), and (ii) the Non-Equilibrium-Wall-Function. The second mesh has been used to study the influence of Enhance Wall Treatment (EWT) approaches. It is worth noting that the k–ε Standard, k–ε RNG, k–ε Realizable, and the RSM models require a wall-function approaches (SWF), Non-Equilibrium-Wall-Function, (EWF). Conversely, the k–ω Standard, k–ω SST, and Spalart–Allmaras models do not consider a wall function approach because of their
12
mathematical formulation: (a) the k–ω Standard and the k–ω SST have been implemented with the low-Reynolds correction; (b) the Spalart-Allmaras model is a low Reynolds model and do not need any further specification for the wall treatment. When considering these results, it should be considered that the refinement of the grid near the wall involves an increase of the computational cost due to the growth of the cell number. The results of the near-wall treatment sensitivity analysis are summarized in Table 6 (global quantities: the entrainment ratio – the “component-scale”) and in Figure 3 (local quantities: wall static pressure along the ejector the “local-scale”). As expected, a change in the wall treatment method resulted in a modification of the numerical predictions, especially for the wall static pressure profiles along the ejector (which is expected, as these data are strictly related to the near wall modeling). Comparing the experimental and the numerical entrainment ratios (Table 6), it is found that the use of Non-Equilibrium-Wall-Function instead of SWF had little effects on the numerical results: (a) the prediction of the entrainment ratio by applying the k-ε Standard model with Non-Equilibrium-Wall-Function improved compared with the use of SWF (the relative error decreased from +13.59% to 3.56%); (b) the prediction of the entrainment ratio by applying the k-ε Realizable model with Non-Equilibrium-Wall-Function is similar to the results obtained by using SWF (the relative error for both near wall treatments is +2.27%); (c) the use of NonEquilibrium-Wall-Function in the k-ε RNG and RSM models results in a worsens of the convergence capability -3
(residuals for the continuity equation do not fall below 10 ). These issues in the convergence behavior are in agreement with the results reported by Besagni et al. [35]. As consequence, Non-Equilibrium-Wall-Function were not considered in the followings. Conversely, EWT approaches allow achieving generally better results (in particular, when considering the RSM and the k-ε models that involve wall functions). However, the differences between the SWF approach and the EWT approach were negligible in most of the cases. The Spalart-Allmaras model provides very similar results between the different meshes, due to its low-Reynolds formulation: it does not need wall treatment because already implemented in its mathematical structure and, therefore, it has a slight benefit from the near-wall grid refinement. Similarly, the Low-Reynolds correction of the k-ω SST model does not improve the results between the different meshes. Moreover, the independence of the k-ω SST model from the wall treatment, in this range of y+ (y+ < 40), is a great advantage compared with the other turbulence models and, for these reasons (also considering the outcomes of Section 4.1.1), is applied in the following. Similar conclusions were drawn by Besagni et al. [35] when considering a convergent-nozzle ejector.
13
4.2 Analysis of the different operating conditions: comparison of the turbulence models 4.2.1 Global quantities: the entrainment ratio Table 7 compares the experimental entrainment ratios with the entrainment ratios computed from the CFD model, for the different RUNs. Table 7 also summarizes the relative errors and the RMSE (Eq. (2)) committed in the prediction of the entrainment ratios. In addition, Figure 3 displays the frequency distribution of the relative errors for the seven turbulence models employed. According to the results, all the turbulence models are able to predict the global quantities (the entrainment ratio) with acceptable errors and in line with the literature [28, 29]. The maximum relative error has been obtained with the k-ε Standard model (equal to 24.5%) and, in general, the k-ε models have lower performance than the other turbulence models. The k-ω Standard model has achieved a maximum error equal to 16.3%, but the half of the relative errors is less than 5%. The k-ω SST and RSM models, instead, attain errors less than 12%. However, the k-ω SST model is the best turbulence model in terms of prediction ability of the entrainment ratio of the ejector. This result is supported by the comparison between the RMSE: the minimum RMSE value is achieved by the k-ω Standard (0.064) and the k-ω SST (0.092) models. The RSM model showed low RMSE value but, in the RUN 5 it could con convergence. The other turbulence models showed higher value of the RMSE, ranging from 0.147 (k-ε Realizable) to 0.201 (k-ε Standard).
4.2.2 Local quantities: static wall pressure profiles The validation of CFD models with global parameters (Section 4.2.1) provides preliminary assessment of the RANS turbulence models, but it is not enough: the CFD model may be able to correctly determine the entrainment ratios (global parameters), but it might not accurately estimate the local flow phenomena (i.e. mixing losses and friction losses), as demonstrated by Hemidi et al. [16]. Indeed, every RANS model was developed for particular test cases, flow phenomena and the prediction of the local flow features should be always verified [37]. In this respect, we have performed a comparison between the local measurements (wall static pressure along the ejector) and the local numerical results. For each simulation, we have summarized the results in a table, compared the pressure profiles by means of a graph and reported the flow field, in terms of the Mach number, for all the turbulence models. The results are displayed in Tables 8-15 and Figures 5-20. Tables 8-15 also summarize the relative errors and the RMSE (Eq. (2)) committed in the prediction of the entrainment ratios.
14
It is observed that the overall trend of the wall static pressure distribution is quite well predicted by all the turbulence models and the predictions are similar to those presented by Sriveerakul et al. (2007) [36] and other literature studies (see ref. [40, 41]). This observation suggests that the overall CFD modeling approach is suitable to predict the ejector fluid dynamics. Considering the influence of the operating conditions, the prediction of the CFD approach increases as the outlet pressure decreases; conversely, for higher outlet pressures (i.e., RUN 5, Table 12, Figure 13 and Figure 15) the results are slightly worse for all the turbulence models and, in particular, the RSM has convergence issues in this case. In this case, having high outlet pressure, it is possible to observe that the ‘shock starts’, corresponding to the rapid increase in pressure, is predicted in a downstream position respect with the experimental measurements. In the previous paragraph, some general comments on the CFD model predictions are provided. Now, we better compare the predictions of the different turbulence approaches. The Spalart-Allmaras model, characterized by low computational costs, allows achieving satisfactory performance, especially in cases, RUN 2 (Table 9, Figure 7 and Figure 8), RUN 3 (Table 10, Figure 9 and Figure 10) and RUN 8 (Table 15, Figure 19 and Figure 20) with a low outlet pressure (pc = 3000 Pa). In the case of RUN 1 (Table 8, Figure 5 and Figure 6), it over predicts the pressure before the shock wave, but it is able to correctly predict the pressure recovery. In the other cases, it largely underestimates the pressure recovery. The k-ε Standard model predicts a normal shock wave at the end of the mixing chamber wave and an increase in the static pressure is observed at the inlet of the diffuser for the RUN 1 (Table 8, Figure 5 and Figure 6), RUN 2 (Table 9, Figure 7 and Figure 8), RUN 3 (Table 10, Figure 9 and Figure 10), and RUN 6 (Table 13, Figure 16 and Figure 17). A similar observation was made by Croquer et al. [32] . These cases correspond to the lower outlet pressures on geometry G1 (Table 3). Conversely, the predictions of the k-ε Standard model are similar to the k-ω Standard model and the k- ω SST model for geometry G1 (RUN 4— Table 11, Figure 11 and Figure 12— and RUN 5— Table 12, Figure 13 and Figure 14) and to the k- ε RNG model, k- ε Realizable model, and the RSM model for the geometries G2 and G3 (RUN 7— Table 14, Figure 17 and Figure 18— and RUN 8— Table 15, Figure 19 and Figure 20). The k- ε RNG model, k- ε Realizable model, and the RSM model showed similar performances: expect for the case RUN 5 (discussed above), they generally predicts the shock wave in a upstream position compared with the experimental measurements and they overpredicts the pressure recovery. Conversely, the k-ω Standard model and the k- ω SST model better fit the experimental results than the other turbulence models (in terms of shock wave position, pressure recovery and fitting of the experimental data before the shock wave). These results are supported by the comparison between the RMSE (see the values in Tables 8-15): the minimum RMSE value is achieved by the k-ω Standard and the
15
k-ω SST models. The other turbulence models showed higher value of the RMSE. The overall better performance of the k-ω SST model also emerges from the literature [30, 32-34, 42, 43] and from our previous study concerning a subsonic ejector [35]. A more detailed understanding of the CFD model results is provided considering the Mach contours. Generally, the flow patterns show a supersonic jet exited from the nozzle outlet and extended into the mixing chamber of the ejector. Thus, shockwaves occur in a determined position in the mixing chamber. According to the value of the discharge pressure, the shock is more or less close to the diffuser. The flow downstream of the shock wave is subsonic. The under-expanded wave at the nozzle exit is well described by the turbulence models for all the simulated geometries and operating conditions. The comparison among the different cases shows that the expansion angle depends also on the secondary flow conditions. For example, Figure 21 shows the Mach contour lines for the cases RUN 1 and RUN 2 computed with the k-ω SST model. It can be observed that an increase of the secondary flow pressure determines a smaller shocks region (jet-flow core) and thus a greater secondary mass flow rate can be entrained in the mixing chamber. The major difference among the turbulence models consists in the prediction of the flow behavior in correspondence of the shocking position. It can be observed that the k-ε models predict the shock in advance position respect to the other turbulence models for all the simulated cases. The k-ω SST model seems to have a greater relevance to the actual physical behavior and to several literature analyses [44-46]. According to ref. [47], concerning an investigation about the effect of the downstream pressure, shows that the shock will not affect the mixing behavior of the two streams because the discharge pressure does not exceed the critical backpressure (outlet pressure). Indeed, the flow structures in front of the shocking position are unchanged and the size of the primary jet core remained constant and independent from downstream conditions. However, the shocking position changes with the value of the backpressure: an increase of the discharge pressure moves the shock position to the upstream of the ejector (Figure 22). In conclusion, according to the results, the best agreement with both global and local parameters is achieved with the k-ω SST model. To the author’s opinion, the better performance of the k–ω SST model can be explained by its inherent structure, designed to be accurate for both near-wall and free-stream regions, and calibrated to yield better results for transonic to moderate supersonic regimes (as also remarked by Mazzelli et al. [34] and Croquer et al. [32]) .The interested reader may refer to the discussion proposed by Croquer et al. [32] concerning the turbulence quantitates and the local flow phenomena. In future studies, the predictions of the RANS models will be improved by modification
16
of the model constants (i.e., the production and dissipation terms), a correction in the model diffusion coefficients and/or the inclusion of cross-diffusion contributions, as also suggested in our previous paper [35]. It is worth noting that the proposed CFD approach has been also applied to study the local and the global fluid dynamics in R134a ejector [48].
4.3 Convergence capability and computational effort According to the convergence criteria reported in Section 3.3, we have performed a convergence analysis in order to compare the models from this aspect. Considering fixed the discretization scheme, the grid density, and the numerical methods the computing time mainly depends upon the turbulence model used [35]. For this purpose, we have selected for all the turbulence models the fine mesh (280000 elements), used in the grid sensitivity analysis. For a rigorous analysis, the main factors that influence the convergence capability and the computing time (i.e. grid resolution, discretization scheme, numerical methods, CFL and under-relaxation factor) must be the same for all the simulations. Thus, we have changed these parameters (i.e. increase of the CFL number, first/second order method switch) in the same way for every turbulence model. In all the simulated cases, the reduction of the mass and energy residuals has been the most difficult to achieve. In particular, these residuals have not fallen below 10-4 with the RSM. The convergence problems of this model are mainly due to the high degree of non-linearity. Conversely, the SpalartAllmaras, k-ω Standard and k-ω SST models easily reach convergence. From a computational point of view, the Spalart-Allmaras is the fastest model to converge, while the most onerous model is the RSM. This is due to the number of equations that must be resolved. The results of the analysis are reported in the Table 16. The computational effort is calculated based on the number of iterations needed to achieve convergence (which is related to the interaction between the turbulence model and the flow field, see Appendix A in ref. [35]), in relative terms, taking as reference the Spalart-Allmaras model. These results can be also compared with the convergence study proposed in our previous paper [35] and the one presented by El-Behery and Hamed [24].
5
CONCLUSIONS
In this paper, we have presented numerical results concerning a supersonic ejector. The numerical approach has been validated against global and local measurements. In particular, we have compared different turbulence models (Spalart-Allmaras, k-ε Standard RNG k- ε, Realizable k- ε, k-ω Standard , SST k- ω, RSM) and the different turbulence models have been tested under different near-wall modelling options in order to investigate the wall treatment effect on the numerical results. The present paper is intended as a complete evaluation of turbulence approaches in order to
17
provide guidelines and a rational basis to simulate supersonic ejectors to be applied in refrigeration systems. As a result, the k- ω SST has showed the best agreement with the experimental measurements concerning both global and local flow quantities. The results show that the k-ω SST model performs better than the other employed models both in global and local parameters prediction. Indeed, the entrainment ratio is well predicted under different geometries and operating conditions with a maximum relative error equal to about 10%. Moreover, the predicted wall pressure distribution is quite well fitted with the experimental data. Considering the results in our previous paper (a complete evaluation of turbulence approaches for a subsonic ejector), as well as the other studies from the literature, we suggest the k- ω SST model for future numerical investigations of supersonic ejectors (i.e., ejectors applied in refrigeration systems). Future studies should be devoted to extend the screening of turbulence approaches presented in this paper and in ref. [35], to take into account the (a) effects of other refrigerants (i.e., the behavior of real gases, see refs. [32, 48]) and (b) phase change inside ejector (in steam ejectors [49, 50] and in ejectors using other refrigerants [51]).
ACKNOWLEDGMENTS The authors thank Giuseppe Di Leo for performing and post-processing the simulations, during its master of science thesis. The authors would like to thank the anonymous reviewers for their valuable comments.
18
REFERENCES [1] J. Godefroy, R. Boukhanouf, S. Riffat, Design, testing and mathematical modelling of a small-scale CHP and cooling system (small CHP-ejector trigeneration), Applied Thermal Engineering, 27 (2007) 68-77. [2] R. Ben Mansour, M. Ouzzane, Z. Aidoun, Numerical evaluation of ejector-assisted mechanical compression systems for refrigeration applications, International Journal of Refrigeration, (2014). [3] G. Besagni, R. Mereu, F. Inzoli, Ejector refrigeration: a comprehensive review, Renewable and Sustainable Energy Reviews, 53 (2016) 373-407. [4] A.B. Little, S. Garimella, Comparative assessment of alternative cycles for waste heat recovery and upgrade, Energy, 36 (2011) 4492-4504. [5] G. Besagni, R. Mereu, G. Di Leo, F. Inzoli, A study of working fluids for heat driven ejector refrigeration using lumped parameter models, International Journal of Refrigeration, 58 (2015) 154-171. [6] EU 517/2014, REGULATION (EU) No 517/2014 OF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 16 April 2014 on fluorinated greenhouse gases and repealing Regulation (EC) No 842/2006 (Text with EEA relevance), 2014., in, 2014. [7] F. Liu, E.A. Groll, J. Ren, Comprehensive experimental performance analyses of an ejector expansion transcritical CO2 system, Applied Thermal Engineering, 98 (2016) 1061-1069. [8] M. Palacz, J. Smolka, W. Kus, A. Fic, Z. Bulinski, A.J. Nowak, K. Banasiak, A. Hafner, CFD-based shape optimisation of a CO2 two-phase ejector mixing section, Applied Thermal Engineering, 95 (2016) 62-69. [9] P. Gullo, A. Hafner, G. Cortella, Multi-ejector R744 booster refrigerating plant and air conditioning system integration – A theoretical evaluation of energy benefits for supermarket applications, International Journal of Refrigeration, (2017). [10] J. Chen, S. Jarall, H. Havtun, B. Palm, A review on versatile ejector applications in refrigeration systems, Renewable and Sustainable Energy Reviews, 49 (2015) 67-90. [11] S. Elbel, N. Lawrence, Review of recent developments in advanced ejector technology, International Journal of Refrigeration, 62 (2016) 1-18. [12] A.B. Little, S. Garimella, A critical review linking ejector flow phenomena with component- and system-level performance, International Journal of Refrigeration, 70 (2016) 243-268. [13] S. He, Y. Li, R.Z. Wang, Progress of mathematical modeling on ejectors, Renewable and Sustainable Energy Reviews, 13 (2009) 1760–1780. [14] D.A. Arias, T.A. Shedd, CFD analysis of compressible flow across a complex geometry venturi, Journal of Fluids Engineering, 129 (2007) 1193-1202. [15] G. Besagni, R. Mereu, E. Colombo, CFD study of ejector efficiencies, in: ASME 2014 12th Biennal Conference on Engineering Systems Design and Analysis, Vol. Dynamics, Vibration and Control; Energy; Fluid Engineering; MIcro and Nano Manufacturing, ESDA2014-20053, Copenhagen, Denmark, July 25-27, 2014, 2014, pp. V02T11A004. [16] A. Hemidi, F. Henry, S. Leclaire, J.-M. Seynhaeve, Y. Bartosiewicz, CFD analysis of a supersonic air ejector. Part II: Relation between global operation and local flow features, Applied Thermal Engineering, 29 (2009) 2990-2998. [17] S. Ghahremanian, B. Moshfegh, Evaluation of RANS models in predicting low reynolds, free, turbulent round jet, Journal of Fluids Engineering, 136 (2013) 011201-011201.
19
[18] Y. Bartosiewicz, Z. Aidoun, Y. Mercadier, Numerical assessment of ejector operation for refrigeration applications based on CFD, Applied Thermal Engineering, 26 (2006) 604-612. [19] J. Gagan, K. Smierciew, D. Butrymowicz, J. Karwacki, Comparative study of turbulence models in application to gas ejectors, International Journal of Thermal Sciences, 78 (2014) 9-15. [20] M. Yazdani, A.A. Alahyari, T.D. Radcliff, Numerical modeling and validation of supersonic two-phase flow of CO2 in converging-diverging nozzles, Journal of Fluids Engineering, 136 (2013) 014503-014503. [21] J. García del Valle, J. Sierra-Pallares, P. Garcia Carrascal, F. Castro Ruiz, An experimental and computational study of the flow pattern in a refrigerant ejector. Validation of turbulence models and real-gas effects, Applied Thermal Engineering, 89 (2015) 795-811. [22] Y. Bartosiewicz, Z. Aidoun, P. Desevaux, Y. Mercadier, CFD-experiments integration in the evaluation of six turbulence models for supersonic ejectors modeling, in: Proceedings of Integrating CFD and Experiments, Vol. 26, Glasgow, 2003, pp. 71-78. [23] A. Bouhanguel, P. Desevaux, E. Gavignet, 3D CFD simulation of supersonic ejector, in: International Seminar on Ejector/jet-pump Technology and Application, Vol. CD, Paper No. 15, Louvain-La-Neuve, Belgium 2009. [24] S.M. El-Behery, M.H. Hamed, A comparative study of turbulence models performance for turbulent flow in a planar asymmetric diffuser, World Academy of Science, Engineering and Technology, 53 (2009) 769-780. [25] V. Dvorak, T. Vit, Experimental and numerical study of constant area mixing, in: 16th International Symposium on Transport Phenomena, Prague, 2006. [26] J. Kolář, V. Dvořák, Verification of K-ω SST turbulence model for supersonic internal flows, World Academy of Science, Engineering and Technology, 81 (2011) 262-266. [27] A. Bouhanguel, P. Desevaux, E. Gavignet, Flow visualisation in supersonic ejectors using laser tomography techniques, in: International Seminar on Ejector/jet-pump Technology and Application, Vol. CD, Paper No. 16, LouvainLa-Neuve, Belgium 2009. [28] A. Hemidi, F. Henry, S. Leclaire, J.-M. Seynhaeve, Y. Bartosiewicz, CFD analysis of a supersonic air ejector. Part I: experimental validation of single-phase and two-phase operation, Applied Thermal Engineering, 29 (2009) 1523-1531. [29] C. Li, Y.Z. Li, Investigation of entrainment behavior and characteristics of gas–liquid ejectors based on CFD simulation, Chemical Engineering Science, 66 (2011) 405-416. [30] N. Ruangtrakoon, T. Thongtip, S. Aphornratana, T. Sriveerakul, CFD simulation on the effect of primary nozzle geometries for a steam ejector in refrigeration cycle, International Journal of Thermal Sciences, 63 (2013) 133-145. [31] Y. Zhu, P. Jiang, Experimental and numerical investigation of the effect of shock wave characteristics on the ejector performance, International Journal of Refrigeration, 40 (2014) 31-42. [32] S. Croquer, S. Poncet, Z. Aidoun, Turbulence modeling of a single-phase R134a supersonic ejector. Part 1: Numerical benchmark, International Journal of Refrigeration, 61 (2016) 140-152. [33] S. Croquer, S. Poncet, Z. Aidoun, Turbulence modeling of a single-phase R134a supersonic ejector. Part 2: Local flow structure and exergy analysis, International Journal of Refrigeration, 61 (2016) 153-165. [34] F. Mazzelli, A.B. Little, S. Garimella, Y. Bartosiewicz, Computational and experimental analysis of supersonic air ejector: Turbulence modeling and assessment of 3D effects, International Journal of Heat and Fluid Flow, 56 (2015) 305-316. [35] G. Besagni, R. Mereu, P. Chiesa, F. Inzoli, An Integrated Lumped Parameter-CFD approach for off-design ejector performance evaluation, Energy Conversion and Management, 105 (2015) 697-715.
20
[36] T. Sriveerakul, S. Aphornratana, K. Chunnanond, Performance prediction of steam ejector using computational fluid dynamics: Part 1. Validation of the CFD results, International Journal of Thermal Sciences, 46 (2007) 812-822. [37] D.C. Wilcox, Turbulence modeling for CFD, DCW industries La Cañada, CA, 2006. [38] Ansys FLUENT 13 - Theory guide, Ansys FLUENT, 2010. [39] X. Han, P. Sagaut, D. Lucor, On sensitivity of RANS simulations to uncertain turbulent inflow conditions, Computers & Fluids, 61 (2012) 2-5. [40] K. Chunnanond, S. Aphornratana, An experimental investigation of a steam ejector refrigerator: the analysis of the pressure profile along the ejector, Applied Thermal Engineering, 24 (2004) 311-322. [41] I.W. Eames, S. Wu, M. Worall, S. Aphornratana, An experimental investigation of steam ejectors for applications in jet-pump refrigerators powered by low-grade heat, Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy, 213 (1999) 351-361. [42] Y. Bartosiewicz, Z. Aidoun, P. Desevaux, Y. Mercadier, Cfd-experiments integration in the evaluation of six turbulence models for supersonic ejector modeling, in, 2003. [43] Y. Bartosiewicz, Z. Aidoun, P. Desevaux, Y. Mercadier, Numerical and experimental investigations on supersonic ejectors, International Journal of Heat and Fluid Flow, 26 (2005) 56-70. [44] K. Pianthong, W. Seehanam, M. Behnia, T. Sriveerakul, S. Aphornratana, Investigation and improvement of ejector refrigeration system using computational fluid dynamics technique, Energy Conversion and Management, 48 (2007) 2556-2564. [45] N. Sharifi, M. Sharifi, Reducing energy consumption of a steam ejector through experimental optimization of the nozzle geometry, Energy, 66 (2014) 860-867. [46] S. Alimohammadi, T. Persoons, D.B. Murray, M.S. Tehrani, B. Farhanieh, J. Koehler, A validated numericalexperimental design methodology for a movable supersonic ejector compressor for waste-heat recovery, Journal of Thermal Science and Engineering Applications, 6 (2014) 021001. [47] T. Sriveerakul, S. Aphornratana, K. Chunnanond, Performance prediction of steam ejector using computational fluid dynamics: Part 2. Flow structure of a steam ejector influenced by operating pressures and geometries, International Journal of Thermal Sciences, 46 (2007) 823-833. [48] G. Besagni, R. Mereu, F. Inzoli, Numerical investigation of R-134a ejector, in: R. B. (ed.) st International Conference IIR of Cryogenics and Refrigeration Technology, ICCRT 2016, Vol. 22-25-June-2016, Bucharest; Romania, 2016, pp. 124-133. [49] Q.L. Dang, R. Mereu, G. Besagni, V. Dossena, F. Inzoli, Simulation of R718 flash boiling flow inside motive nozzle of ejector, in: R. B. (ed.) 1st International Conference IIR of Cryogenics and Refrigeration Technology, ICCRT 2016;, Vol. 22-25-June-2016, Bucharest; Romania, 2016, pp. 116-123. [50] F. Giacomelli, G. Biferi, F. Mazzelli, A. Milazzo, CFD Modeling of the Supersonic Condensation Inside a Steam Ejector, Energy Procedia, 101 (2016) 1224-1231. [51] G. Biferi, F. Giacomelli, F. Mazzelli, A. Milazzo, CFD Modelling of the Condensation Inside a Supersonic Ejector Working with R134a, Energy Procedia, 101 (2016) 1232-1239.
21
FIGURE CAPTION Figure 1
Ejector refrigeration system
Figure 2
Wall static pressure distribution along the ejector – Grid sensitivity analysis.
Figure 3
Wall static pressure distribution along the ejector – Enhanced wall treatment performance.
Figure 4
Absolute frequency distribution of the errors.
Figure 5
Wall static pressure distribution along the ejector (RUN 1).
Figure 6
Mach contours of the ejector flow field (RUN 1).
Figure 7
Wall static pressure distribution along the ejector (RUN 2).
Figure 8
Mach contours of the ejector flow field (RUN 2).
Figure 9
Wall static pressure distribution along the ejector (RUN 3).
Figure 10
Mach contours of the ejector flow field (RUN 3).
Figure 11
Wall static pressure distribution along the ejector (RUN 4).
Figure 12
Mach contours of the ejector flow field (RUN 4).
Figure 13
Wall static pressure distribution along the ejector (RUN 5).
Figure 14
Mach contours of the ejector flow field (RUN 5).
Figure 15
Wall static pressure distribution along the ejector (RUN 6).
Figure 16
Mach contours of the ejector flow field (RUN 6).
Figure 17
Wall static pressure distribution along the ejector (RUN 7).
Figure 18
Mach contours of the ejector flow field (RUN 7).
Figure 19
Wall static pressure distribution along the ejector (RUN 8).
Figure 20
Mach contours of the ejector flow field (RUN 8).
Figure 21
Mach contour lines comparison.
Figure 22
Mach contours of the ejector flow field – Effect of the discharge pressure (SST k-ω ω).
22
(a) Ejector layout
Ejector [Tg,pg]
Qc
[pc] Condenser [Te,pe]
Boundary conditions in the CFD approach
Primary flow
Throttle valve Evaporator Pump
Secondary flow
Generator Wp
Qg (b) Ejector refrigeration system
23
Spalart-Allmaras model
3000 Exp Spalart-Allmaras (coarse mesh)
2500
Spalart-Allmaras (medium mesh) Spalart-Allmaras (fine mesh)
2000
1500
3000 Exp
Standard k-e (medium mesh)
1500
500
500
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0
0.4
Standard k-e (fine mesh)
2000
1000
0
Standard k-e (coarse mesh)
2500
1000
0
Standard k-e model
3500
Static pressure [Pa]
Static pressure [Pa]
3500
0
0.05
0.1
0.15
Distance along ejector [m]
RNG k-e model
3000 Exp RNG k-e (coarse mesh) 2500
RNG k-e (medium mesh) RNG k-e (fine mesh)
2000
1500
0.15
0.2
0.25
0.3
0.35
0
0.05
0.1
0.15
Static pressure [Pa]
Standard k-w (coarse mesh) Standard k-w (medium mesh) Standard k-w (fine mesh)
1500
0.15
0.2
0.25
0.3
0.35
1500
0
0.4
SST k-w (fine mesh)
0
0.05
0.1
0.15
Distance along ejector [m]
0.2
RSM
3000 Exp RSM (coarse mesh) RSM (medium mesh) RSM (fine mesh)
2000
1500
1000
500
0
0
0.05
0.25
0.3
Distance along ejector [m]
3500
2500
0.4
SST k-w (medium mesh)
500
0.1
SST k-w (coarse mesh)
2000
1000
0.05
0.35
Exp 2500
500
0
0.3
3000
1000
Static pressure [Pa]
Static pressure [Pa]
Exp
0
0.25
SST k-w model
3500
3000
2000
0.2
Distance along ejector [m]
Standard k-w model
2500
0.4
1500
Distance along ejector [m]
3500
0.35
Realizable k-e (fine mesh)
0
0.4
Realizable k-e (medium mesh)
2000
500
0.1
0.4
Realizable k-e (coarse mesh)
500
0.05
0.35
Exp 2500
1000
0
0.3
3000
1000
0
0.25
Realizable k-e model
3500
Static pressure [Pa]
Static pressure [Pa]
3500
0.2
Distance along ejector [m]
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
24
0.35
0.4
k-e model
3000
Exp k-e (SWF)
2500 k-e (EWT) 2000
1500
3000 Exp
RNG k-e (EWT) 2000
1500
1000
500
500
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0
0.4
RNG k-e (SWF)
2500
1000
0
RNG k-e model
3500
Static pressure [Pa]
Static pressure [Pa]
3500
0
0.05
0.1
Distance along ejector [m]
Static pressure [Pa]
Static pressure [Pa]
Exp Realizable k-e (SWF) Realizable k-e (EWT) 2000
1500
0.15
0.2
0.25
0.3
0.35
0
0.05
0.1
Distance along ejector [m]
Static pressure [Pa]
Static pressure [Pa]
Exp Spalart-Allmaras Spalart-Allmaras (EWT) 2000
1500
0.15
0.2
0.25
0.3
0.35
0.4
Distance along ejector [m]
0.35
0.4
SST k-w SST k-w (EWT)
1500
500
0.1
0.4
2000
1000
0.05
0.3
Exp 2500
500
0
0.25
3000
1000
0
0.2
SST k-w model
3500
3000
2500
0.15
Distance along ejector [m]
Spalart-Allmaras model
3500
0.35
1500
0
0.4
RSM (SWF) RSM (EWT)
500
0.1
0.4
2000
500
0.05
0.35
Exp 2500
1000
0
0.3
3000
1000
0
0.25
RSM
3500
3000
2500
0.2
Distance along ejector [m]
Realizable k-e model
3500
0.15
0
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
25
RUN 1
Static pressure [Pa]
3500
3000
Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
2500
2000
1500
1000
500
0
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
26
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
27
RUN 2
Static pressure [Pa]
3500
Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
3000
2500
2000
1500
1000
500
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
28
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
29
RUN 3
Static pressure [Pa]
4000
3500
Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
3000
2500
2000
1500
1000
500
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
30
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
31
RUN 4
4500
Static pressure [Pa]
4000 Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
3500 3000 2500 2000 1500 1000 500
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
32
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
33
RUN 5
5000
Static pressure [Pa]
4500 Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w
4000 3500 3000 2500 2000 1500 1000 500
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
34
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
Mach Number
35
RUN 6
Static pressure [Pa]
3500
3000
Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
2500
2000
1500
1000
500
0
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
36
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
37
Static pressure [Pa]
3500
3000
2500
2000
Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
1500
1000
500
0 0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
38
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
39
RUN 8
Static pressure [Pa]
3500
3000 Exp Spalart-Allmaras k-e RNG k-e Realizable k-e k-w SST k-w RSM
2500
2000
1500
1000
500
0
0
0.05
0.1
0.15
0.2
0.25
0.3
Distance along ejector [m]
40
0.35
0.4
Spalart-Allmaras
Standard k-ε
RNG k-ε
Realizable k-ε
Standard k-ω
SST k-ω
RSM
Mach Number
41
Run 1: pg = 270280 Pa, pe = 872.5 Pa, pc = 3000 Pa
Run 2: pg = 270280 Pa, pe = 1228.1 Pa, pc = 3000 Pa
42
Run 2: pg = 270280 Pa, pe = 1228.1 Pa, pc = 3000 Pa
Run 3: pg = 270280 Pa, pe = 1228.1 Pa, pc = 3500 Pa
Run 4: pg = 270280 Pa, pe = 1228.1 Pa, pc = 4000 Pa
Run 5: pg = 270280 Pa, pe = 1228.1 Pa, pc = 4500 Pa
Mach Number
43
Geometry code name Nozzle throat diameter [mm] Nozzle exit diameter [mm] Mixing chamber inlet diameter [mm] Throat length [mm]
Run Geometry Tg [°C] pg [Pa] Te [°C] pe [Pa] Tc [°C] pc [Pa]
1 G1 130 270280 5 872.5 24.08 3000
G1 2 8 24 95
2 G1 130 270280 10 1228.1 24.08 3000
3 G1 130 270280 10 1228.1 26.67 3500
Geometry Number of elements Aspect ratio worst value >8 Size-change worst value > 1.1 Cell surface worst value > 0.2 Equi-angle skew worst value > 0.65
G2 2 8 24 57
4 G1 130 270280 10 1228.1 28.96 4000
5 G1 130 270280 10 1228.1 31.01 4500
G1 71289 10.1 0.06 % 1.27 0.37 % 0.22 1.91 % 0.86 0.13 %
G3 2 8 19 95
6 G1 120 198670 10 1228.1 24.08 3000
7 G2 130 270280 5 872.5 24.08 3000
G2 68129 10.1 0.06 % 1.27 0.39 % 0.22 2.00 % 0.86 0.14 %
8 G3 130 270280 5 872.5 24.08 3000
G3 71289 10.1 0.06 % 1.19 0.27 % 0.22 1.91 % 0.86 0.13 %
Density
Molecular mass
Cp
Termal conductivity
Vicosity
[kg/m3]
[kg/kmol]
[J/(kg⋅⋅K)] 2014.0
[W/(m⋅⋅K)] 0.0261
[kg/(m⋅⋅s)]
Ideal gas law
Coarse mesh Medium mesh Fine mesh Experimental data
18.01534
Spalart-Allmaras 0.298 0.303 0.302 0.309
k-ε Standard 0.351 0.349 0.304 0.309
k-ε RNG 0.313 0.312 0.309 0.309
k-ε Realizable 0.302 0.302 0.306 0.309
1.34 ⋅ 10-5
k-ω Standard 0.262 0.273 0.292 0.309
k-ω SST 0.298 0.303 0.303 0.309
RSM 0.273 0.273 0.284 0.309
Experimental data = 0.309 [-] Mesh
Medium mesh y+ ≈ 40
Wall treatment Standard Wall function (SWF)
Non Equilibrium Wall Function
Spalart-Allmaras
0.303 (ER = 1.94%) (ii)
k-ε Standard k-ε RNG k-ε Realizable k-ω Standard k-ω SST 0.349 (12.94%)
0.320 (3.56%)
44
0.312 (0.97%) (i)
0.302 (2.27%)
0.302 (2.27%)
0.273(ii) (11.65%)
RSM
0.273 0.303(ii) (11.65%) (1.94%) (i)
Refined mesh Enhanced wall treatment (EWT) +
y <1 (i) (ii) (iii) (iv)
0.301 (ER = 2.59%) (iv)
0.322 (4.21%)
0.314 (1.62%)
0.301(iii) (2.59%)
0.319 (3.24%)
0.302(iii) (2.27%)
0.318 (2.91%)
-3
Difficulties in convergence: residuals for the continuity equation do not fall below 10 Near wall treatment not needed because already implemented in the mathematical structure of the model Low-Re correction for the k-ω models activated in the viscous panel [31] The Spalart-Allmaras model, in its mathematical formulation, is a low Reynolds model and do not need any further specification for the wall treatment
Experimental data Spalart-Allmaras
Entrainment ratio [] Entrainment ratio [] Relative error ER [%] Entrainment ratio [k-ε Standard ] Relative error ER [%] Entrainment ratio [k-ε RNG ] Relative error ER [%] Entrainment ratio [k-ε Realizable ] Relative error ER [%] Entrainment ratio [k-ω Standard ] Relative error ER [%] Entrainment ratio [k-ω SST ] Relative error ER [%] RSM Entrainment ratio [] Relative error ER [%] *Convergence was not reached
Run 1 0.309
Run 2 0.397
Run 3 0.400
Run 4 0.400
Run 5 0.403
Run 6 0.527
Run 7 0.301
Run 8 0.172
0.303
0.477
0.477
0.477
0.477
0.656
0.304
0.169
1.94 0.349
20.15 0.338
19.25 0.302
19.25 0.390
18.36 0.387
24.48 0.622
1.00 0.307
1.74 0.177
12.94 0.312
14.86 0.485
24.50 0.487
2.50 0.487
3.97 0.487
18.03 0.577
1.99 0.312
2.91 0.176
0.97 0.302
22.17 0.470
21.75 0.469
21.75 0.469
20.84 0.469
9.49 0.574
3.65 0.304
2.33 0.172
2.27 0.273
18.39 0.394
17.25 0.394
17.25 0.394
16.38 0.394
8.92 0.531
1.00 0.252
0.00 0.158
11.65 0.303
0.76 0.438
1.50 0.438
1.50 0.438
2.23 0.438
0.76 0.476
16.28 0.302
8.14 0.169
1.94 0.273
10.33 0.413
9.50 0.413
9.50 0.414
8.68 n.c.*
9.68 0.550
0.33 0.274
1.74 0.158
11.65
4.03
3.25
3.50
n.c.*
4.36
8.97
8.14
RMSE, Eq. (2) -
0.201
0.155
0.181
0.147
0.064
0.092
0.058
** Value obtained excluding Run 5
45
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
30
70
90
110
150
177.5
205
245
280
315
350
RMSE, Eq. (2)
661
488
564
752
755
854
653
1130
2076
2532
2892
-
1011
901
941
902
979
1019
1068
798
1979
2535
2846
52.95
84.63
66.84
19.95
29.67
19.32
-63.55
29.38
4.67
-0.12
1.59
718
537
507
756
870
937
1001
673
2530
2828
2942
-8.62
10.04
10.11
-0.53
15.23
-9.72
-53.29
40.44
21.87
11.69
-1.73
738
552
550
857
1055
1209
1437
1899
2654
2889
2963
11.65
13.11
2.48
13.96
39.74
41.57
120.06
68.05
27.84
14.10
-2.46
753
911
809
1382
533
626
811
965
1062
1211
1806
2655
2910
2965
13.92
-9.22
10.99
-7.85
27.81
24.36
-85.45
59.82
27.89
14.93
-2.52
791
724
731
723
734
618
716
1334
1946
2446
2807
19.67
48.36
29.61
3.86
2.78
27.63
-9.65
18.05
6.26
3.40
2.94
751
531
580
729
813
761
805
1381
1998
2468
2836
3.76
2.53
1.94
1165
484
350
13.62
-8.81
-2.84
3.06
-7.68
10.89
-23.28
22.21
768
656
834
901
971
1061
1192
1889
2568
2858
2960
16.19
34.43
47.87
19.81
28.61
24.24
-82.54
67.17
23.70
12.88
-2.35
46
1202
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
30
70
90
110
150
177.5
205
245
280
315
350
884
877
864
792
849
993
840
1332
1867
2466
2840
1020
911
946
899
977
1020
1059
796
1989
2548
2852
15.08
-2.72
-26.07
40.24
-6.53
-3.33
-0.42
1321
831
2414
2786
2920
12.98
-2.82
642
15.38
-3.88
-9.49
13.51
957
692
917
956
1177
1232
38.63
24.07
-57.26
37.61
29.30
1060
-8.26
21.09
-6.13
20.71
738
552
550
857
1055
1209
1439
1865
2644
2889
2963
21.75
-71.31
40.02
41.62
17.15
-4.33
16.52
37.06
36.34
-8.21
24.26
1015
1326
842
963
982
1228
1455
1704
1424
2395
2837
2932
14.82
3.99
11.46
23.99
44.64
46.53
102.86
-6.91
28.28
15.04
-3.24
1089
1034
973
861
882
889
867
1244
1851
2377
2767
23.19
17.90
12.62
-8.71
-3.89
10.47
-3.21
6.61
0.86
3.61
2.57
1032
953
913
867
968
1024
1077
1306
1953
2431
2793
-3.12
-28.21
1.95
-4.61
1.42
1.65
1265
342
344
16.74
-8.67
-5.67
-9.47
14.02
1064
993
998
1170
1170
1328
1470
1714
2452
2796
2943
20.36
13.23
15.51
47.73
37.81
33.74
-75.00
28.68
31.33
13.38
-3.63
30
70
90
110
150
177.5
205
245
280
315
350
820
869
811
781
846
959
720
1877
2545
3072
3375
1011
902
941
901
979
1019
1068
1720
2723
3199
3412
23.29 780
-3.80
15.36 918
15.72 1030
-6.26
-48.33
8.36
-6.99
-4.13
-1.10
749
16.03 841
1059
1087
1307
3225
3413
3471
4.88
13.81
-3.70
2397
2619
26.72 3186
11.10 3377
-2.84
955
10.43 1604
30.37
697
21.75 1371
-50.97
996
17.54 1072
3455
21.46 1030
19.79
17.76 975
37.26 986
62.06 1282
67.26 1399
232.92 1582
39.53 2564
25.19 3220
-9.93
-2.37
3397
3460
25.61 1089
-3.57
26.25 861
51.54 909
45.88 890
119.72 879
36.60 1764
26.52 2494
10.58 3347
-2.52
1034
20.22 972
3375
32.80 1032
18.99 953
19.85 914
10.24 868
-7.45
7.19
-22.08
6.02
2.00
-8.95
0.00
968
1023
1077
1879
2599
3097
3397
25.85 1064
-9.67
11.14 1184
14.42 1175
-6.67
-49.58
-0.11
-2.12
-0.81
-0.65
993
12.70 997
1328
1470
2540
3149
3378
3461
29.76
14.27
22.93
51.60
38.89
38.48
104.17
35.32
23.73
-9.96
-2.55
900
RMSE, Eq. (2) -
47
1194
RMSE, Eq. (2) -
534
1062
2175
1510
507
470
1409
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
[mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
Spalart-Allmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
30
70
90
110
150
177.5
205
245
280
315
350
RMSE, Eq. (2)
870
871
866
942
1221
1664
2127
3192
3582
3828
4096
-
1011
901
941
901
980
1020
1068
2521
3410
3803
3940
16.21
-3.44
-8.66
4.35
19.74
38.70
49.79
21.02
4.80
0.65
3.81
1090
1074
1033
897
921
932
1265
2476
3297
3767
3942
25.29
23.31
19.28
4.78
24.57
43.99
40.53
22.43
7.96
1.59
3.76
1224
1083
1203
1257
1403
1514
1914
3578
3849
3940
3977
40.69
24.34
38.91
33.44
14.91
9.01
10.01
12.09
-7.45
-2.93
2.91
1029
1458
1452
855
902
983
1003
1229
1348
2198
3504
3808
3914
3966
18.28
-3.56
13.51
-6.48
-0.66
18.99
-3.34
-9.77
-6.31
-2.25
3.17
1089
1034
972
861
883
889
1283
2417
3235
3749
3938
25.17
18.71
12.24
8.60
27.68
46.57
39.68
24.28
9.69
2.06
3.86
1032
953
913
867
968
1024
1156
2588
3322
3789
3950
18.62
-9.41
-5.43
7.96
20.72
38.46
45.65
18.92
7.26
1.02
3.56
1064
993
997
1181
1174
1377
2461
3419
3782
3913
3968
22.30
14.01
15.13
25.37
3.85
17.25
15.70
-7.11
-5.58
-2.22
3.13
30
70
90
110
150
177.5
205
245
280
315
350
RMSE, Eq. (2)
879
1044
953
696
1689
2506
3231
4217
4655
4727
4860
-
1010
901
941
902
980
1020
1166
3397
4116
4369
4454
41.98
59.30
63.91
19.45
11.58
7.57
8.35
567
1507
1383
662
2883
14.90
13.70
1.26
29.60
1092
1046
973
874
892
1166
1964
3242
4051
4373
4462
47.19
53.47
39.21
23.12
12.98
7.49
8.19
24.23
-0.19
-2.10
25.57
1224
2390
1083
1204
1257
1403
1514
1913
4078
4349
4439
4477
39.25
-3.74
26.34
80.60
16.93
39.58
40.79
3.30
6.57
6.09
7.88
1311
1244
1193
1130
1215
1389
1782
2917
3735
4249
4438
49.15
19.16
25.18
62.36
28.06
44.57
44.85
30.83
19.76
10.11
8.68
1090
1034
973
860
882
1169
1951
3228
4045
4371
4461
47.78
53.35
39.62
23.45
13.10
7.53
8.21
968
1026
2054
3393
4118
4394
4464
42.69
59.06
36.43
19.54
11.54
7.04
8.15
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
n.c*
24.00
0.96
-2.10
23.56
1032
953
914
867
17.41 n.c*
n.c*
n.c*
24.57 n.c*
n.c*
n.c*
n.c*
n.c*
8.72
4.09
*Convergence was not reached
48
1909
2643
2405
2322
n.c*
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
30
70
90
110
150
177.5
205
245
280
315
350
RMSE, Eq. (2)
981
921
942
948
931
925
879
1666
2302
2688
2893
-
1011
901
941
902
980
1020
1068
798
1978
2535
2846
21.50
52.10
14.07
5.69
1.62
967
-3.06
2.17
0.11
4.85
-5.26
10.27
719
536
508
756
871
939
1003
670
2529
2832
2947
59.78
-9.86
-5.36
-1.87
1236
26.71
41.80
46.07
20.25
6.44
-1.51
14.11
738
552
550
857
1055
1209
1437
1900
2654
2889
2963
30.70
63.48
14.05
15.29
-7.48
-2.42
994
24.77
40.07
41.61
9.60
13.32
750
534
631
809
967
1061
1215
1803
2659
2921
2967
38.23
-8.22
15.51
-8.67
-2.56
811
23.55
42.02
33.01
14.66
-3.87
14.70
1090
1044
998
965
829
801
933
1758
2360
2761
2937
11.11
13.36
-5.94
-1.79
10.96
13.41
-6.14
-5.52
-2.52
-2.72
-1.52
751
533
580
728
813
761
808
1381
1997
2468
2836
23.45
42.13
38.43
23.21
12.67
17.73
8.08
17.11
13.25
8.18
1.97
764
659
832
909
969
1066
1195
1893
2572
2852
2963
-4.08
15.24
35.95
13.63
11.73
-6.10
-2.42
280
809
22.12
28.45
11.68
4.11
49
638
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
x [mm]
Axial position
Experimental data
Wall static pressure [Pa] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%] Wall static pressure [Pa] Relative error ER [%]
SpalartAllmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
70
90
110
150
177.5
205
245
280
315
350
385
RMSE, Eq. (2)
452
488
639
683
824
739
1679
2269
2602
2714
2893
-
744
647
509
696
682
712
1078
1962
2544
2879
2984
64.60
32.58
20.34
-1.90
17.23
3.65
35.80
13.53
2.23
-6.08
-3.15
742
632
536
859
1053
1156
1662
2622
2891
2964
2991
64.16
29.51
16.12
25.77
27.79
56.43
1.01
15.56
11.11
-9.21
-3.39
733
624
542
859
1065
1158
1157
2587
2893
2965
2993
62.17
27.87
15.18
25.77
29.25
56.70
31.09
14.01
11.18
-9.25
-3.46
741
643
516
788
931
1035
1413
2627
2884
2941
2987
63.94
31.76
19.25
15.37
12.99
40.05
15.84
15.78
10.84
-8.36
-3.25
789
741
725
713
603
676
1347
1929
2405
2777
2967
74.56
51.84
13.46
-4.39
26.82
8.53
19.77
14.98
7.57
-2.32
-2.56
746
653
518
731
707
748
1315
1919
2417
2829
2983
65.04
33.81
18.94
-7.03
14.20
-1.22
21.68
15.43
7.11
-4.24
-3.11
765
694
663
891
1040
1064
1722
2457
2839
2958
2997
69.25
42.21
-3.76
30.45
26.21
43.98
-2.56
-8.29
-9.11
-8.99
-3.59
802
808
950
756
716
674
707
30
70
90
110
150
177.5
205
245
280
315
350
RMSE, Eq. (2)
461
251
169
340
392
519
534
1222
1983
2273
2518
-
746
489
436
517
540
1237
1987
2575
2904 15.33
462
352
61.82
-84.06
108.28
-43.82
-11.22
0.39
-1.12
-1.23
-0.20
13.29
729
547
463
681
784
905
1027
2015
2731
2917
2967
58.13
117.93
173.96
100.29
100.00
74.37
92.32
64.89
37.72
28.33
17.83
729
546
472
675
782
910
1022
1909
2700
2922
2971
58.13
117.53
179.29
-98.53
-99.49
75.34
91.39
56.22
36.16
28.55
17.99
735
525
454
653
755
797
879
1796
2724
2936
2972
59.44
109.16
168.64
-92.06
-92.60
53.56
64.61
46.97
37.37
29.17
18.03
763
481
428
497
627
1347
1947
2441
2814
10.23
1.82
-7.39
11.76
465
428
65.51
-85.26
153.25
-41.47
-9.18
4.24
17.42
745
512
479
539
620
1394
1952
2466
2879
14.08
1.56
-8.49
14.34
468
363
61.61
-86.45
114.79
-50.59
-22.19
-3.85
16.10
752
498
449
619
711
729
815
1819
2525
2868
2966
-98.41
165.68
-81.38
40.46
52.62
48.85
27.33
26.18
17.79
63.12
-82.06
50
651
1647
1587
1477
606
639
1316
Computational effort
Spalart-Allmaras
k-ε Standard
k-ε RNG
k-ε Realizable
k-ω Standard
k-ω SST
RSM
1
1.33
1.53
1.45
1.42
1.84
3.73
TABLE CAPTION Table 1
Main geometrical parameters of the ejectors.
Table 2
Operating conditions of numerical simulations.
Table 3
Mesh quality parameters.
Table 4
Water vapour properties.
Table 5
Grid sensitivity analysis – Entrainment ratio prediction.
Table 6
Run 1 – Turbulence models and wall treatments.
Table 7
CFD models prediction of the entrainment ratio.
Table 8
Run 1 - Experimental and numerical results of the wall static pressure.
Table 9
Run 2 – Experimental and numerical results of the wall static pressure.
Table 10
Run 3 – Experimental and numerical results of the wall static pressure.
Table 11
Run 4 – Experimental and numerical results of the wall static pressure.
Table 12
Run 5 – Experimental and numerical results of the wall static pressure.
Table 13
Run 6 – Experimental and numerical results of the wall static pressure.
Table 14
Run 7 – Experimental and numerical results of the wall static pressure.
Table 15
Run 8 – Experimental and numerical results of the wall static pressure.
Table 16
Computational effort of CFD simulations with different turbulence models.
51
Highlights • The fluid dynamics in a supersonic ejector for refrigeration systems was studied • Different turbulence modeling approaches were compared • Different mesh sizes and near wall modeling approaches were compared • A computational fluid dynamics model was validated
52