Next Article in Journal
Forest Fuel Drying, Pyrolysis and Ignition Processes during Forest Fire: A Review
Next Article in Special Issue
Simulative Investigation of Different DLD Microsystem Designs with Increased Reynolds Numbers Using a Two-Way Coupled IBM-CFD/6-DOF Approach
Previous Article in Journal
Effects of Alkaline-Reduced Water on Gastrointestinal Diseases
Previous Article in Special Issue
Study on a New Transient Productivity Model of Horizontal Well Coupled with Seepage and Wellbore Flow
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mathematical Modelling of the Entrainment Ratio of High Performance Supersonic Industrial Ejectors

Department of Land, Environment, Agriculture and Forestry, University of Padova, Agripolis, Viale dell’Università 16, 35020 Legnaro, Italy
Processes 2022, 10(1), 88; https://doi.org/10.3390/pr10010088
Submission received: 14 November 2021 / Revised: 29 December 2021 / Accepted: 29 December 2021 / Published: 2 January 2022
(This article belongs to the Special Issue Computational Modeling of Multiphase Flow (II))

Abstract

:
For many years now, manufacturers have been producing supersonic ejectors with a high entrainment ratio for the chemical, oil, and food industries. In the present work, mathematical modelling of the entrainment ratio of such industrial ejectors is carried out, in which a variation of the diffuser efficiency is also assumed to be a function of the Mach number of the motive gas. To determine this unknown relationship, the mathematical modelling was overturned by inserting the entrainment ratios of ten different high-performance industrial ejectors, as identified through an experimental investigation. The mathematical modelling, completed through the use of the relationship between the diffuser efficiency and the Mach number of the motive gas, was applied to sixty-eight ejectors, built and tested experimentally over the last twenty years as part of research aimed at the development of thermal ejector refrigeration systems (ERSs), to obtain the entrainment ratios proposed by the manufacturers (industrial entrainment ratios). A comparison of the experimental entrainment ratios with respect to the industrial ones demonstrated that the former were always lower, ranging from a minimum of −17% to a maximum of −82%. These results indicate that the lab-built ejectors for ERS prototypes can be improved. Therefore, in the future, researchers should apply numerical analysis iteratively, starting from a given geometry of the ejector, and modifying it until the numerical analysis provides the industrial value of the entrainment ratio.

1. Introduction

An ejector, or jet pump, is an apparatus (Figure 1) of which the task is the compression of a fluid—called the induced fluid—by molecular entrainment and diffusion, through a jet produced by the expansion of another fluid—called the motive fluid [1,2]. Fluids can be liquid or gas and, in this case, are in supersonic motion. In this work, supersonic ejectors are discussed, as they are widespread in the chemical, oil, and food industries, where they are used in the following:
-
Vacuum pumps to eliminate incondensable gases (sucked fluid) in various treatment plants, where the motive fluid is steam from a boiler; or
-
Steam compressors in concentration and distillation plants. The vapor resulting from boiling liquid is thermodynamically requalified and then reused to heat the boiling liquid itself. This ejector–heat exchanger system is, therefore, a heat pump, and the ejector is also called a thermo-compressor. Furthermore, in this case, the motive fluid is steam from a boiler.
Various studies focused on the use of supersonic ejectors as thermo-compressors in heat-driven ejector refrigeration systems (ERSs), where both fluids are refrigerating gases, have been underway for many years. This type of system, which has been known for more than a century [3], is now being re-proposed for environmental reasons, as they can operate on solar energy [4].
The peculiarities of these ejectors are as follows: The absence of moving parts and, therefore, reliability and durability; simplicity of construction and, therefore, relatively low investment cost; operational rigidity, as they operate correctly only in design conditions; and the complexity of the phenomena of mixing and diffusion of the two fluids operating in transonic and supersonic conditions, therefore leading to difficulty in the mathematical modelling and design of high-performance ejectors. In confirmation of this last point, it should be noted that there are no more than a dozen manufacturers in the world capable of producing high-performance supersonic ejectors.
The performance of an ejector can be identified by the entrainment ratio ω, defined as the ratio between the mass flow rate of the induced gas Gi to that of the motive gas Gm. Its calculation—and, therefore, prediction of the performance of an ejector—can be carried out with the aid of a one-dimensional gas-dynamics theory. Keenan and Neumann [5] used such a one-dimensional approach with the hypothesis of ideal gas behavior considering the equations of conservation of mass, momentum, and energy; however, they excluded frictional losses and diffusion modelling. In a subsequent work, Keenan, Neumann, and Lustwerk [6] introduced these two phenomena; however, the mathematical description of the dissipative phenomenon of shock waves occurring in the transition from supersonic to subsonic motion of the gas mixture was lacking.
A later work by Lukasiewicz [7] on diffuser ducts highlighted the presence of several curved or oblique shock waves. Furthermore, Shapiro [8], in his 1953 textbook, referred to a series of bifurcated normal shocks in diffuser ducts. Matsuo [2] has called this phenomenon a shock train, but other expressions are also in use, besides those mentioned by Shapiro and Lukasiewicz: a series of shocks [9], Lambda foot shock system [10], Bifurcated normal shock waves [11], a series of oblique shocks [12], multiple normal shocks [13], and multiple shocks [14].
Munday and Bagster [15] introduced a mathematical formulation of the phenomenon of diffusion in the ejector from supersonic to subsonic stream using the Prandtl–Meyer equation, which predicts the presence of a frontal shock wave. Furthermore, they proposed that the induced gas, in the first part of the mixing chamber (Figure 1), remained separate from the motive one and was forced to expand up to the sonic conditions. This was due to widening of the motive gas jet, leading to a reduction in the section available for the induced gas. From the sonic condition of the induced gas, the real mixing hypothesized at constant pressure then started.
With the renewed interest in ejectors used as thermo-compressors in refrigeration cycles, Eames et al. [16] verified this one-dimensional analytical approach for various ejectors inserted in a heat-driven ejector refrigeration system (ERS) operated with steam, noting that the experimental entrainment ratio was always less than that predicted theoretically. Similar differences have also been identified by Huang et al. [17], considering different ejectors inserted in an ERS with refrigerant fluid R141b. Rogdakis and Alexis [18] also experimentally verified the theory with an ejector operating in an ERS with refrigerant fluid R717, always identifying a more optimistic theoretical entrainment ratio.
Therefore, in order of a better comprehension of the phenomena of mixing and diffusion, the computer fluid dynamics CFD modelling is useful. Consequently, for about 20 years, computational studies on the operation and design of ejectors have continued to be produced, focused on both on-design and off-design conditions [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46]. Most of the results of the numerical analyses presented in these works have been experimentally validated by the authors: in many cases, the entrainment ratio in the on-design condition predicted by the numerical analysis has presented a negligible difference, compared with the experimental one.
Parallel to this incessant scientific commitment—which began over 80 years ago and has accelerated in the last twenty years due to the availability of numerical analysis software—manufacturers of supersonic ejectors for various industrial sectors (i.e., chemical, oil, and food) have been present on the market for over a century.
For example, in Germany, Koerting Gebrueder began manufacturing supersonic ejectors in 1920 and, in 1964, began using computers for the fluid dynamic design of supersonic jet pumps (as ejectors are also called). Therefore, it is very likely that the design guidelines adopted by Koerting AG were not and (even more so today) are not purely empirical. Moreover, in Germany, another company was founded in the 1940s, Wiegand (which has recently been absorbed by GEA AG), by the researcher Joachim Wiegand, who in 1940, published a 24-page booklet full of theory and experimental data on the design of ejectors [47]. In Europe, other relevant manufacturers, such as the French SCAM, Kinetic-therm, and LVI and the Spanish Equirepsa, subsequently established themselves, while in the United States [1], Croll-Reynolds, Fox-Valve, Graham Manufacturing, Jet-vac technologies, and Nash-kinema have long been known. It is likely that some names have been unintentionally forgotten, but it should be noted that there are only a dozen manufacturers internationally, and not all of them produce high-performance ejectors (i.e., with high entrainment ratio = industrial entrainment ratio).
In the first part of this work, we detail an experimental investigation carried out to determine the entrainment ratio values obtainable from high-performance supersonic ejectors designed under different boundary conditions (i.e., pressure and temperature of the motive gas, the induced gas, and the mixed gases at the outlet). These ejectors offer high-performance (optimal entrainment ratio) because they are built by a big European manufacturer that has been present on the international market for 80 years. In the second part, mathematical modelling is carried out, which, starting from the Wiegand equations [47], correlates the entrainment ratio of the aforementioned ejectors of the European manufacturer to the gas properties, the boundary conditions (pressures and temperatures), and the isentropic efficiency of the supersonic nozzle (motive gas), suction chamber (induced gas), and diffuser (mixed gases). In the third part of the work, the mathematical model is used to obtain the values of the entrainment ratio as a function of the gas properties and the boundary conditions (pressures and temperatures) of ejectors built and tested in the laboratory. These ejectors are considered benchmarks, as they are used to test the results of the numerical analyses proposed in many papers.
A comparison between the industrial entrainment ratios predicted from the mathematical model elaborated in the second part and the entrainment ratios obtained experimentally from the benchmark ejectors highlights important differences; namely, the experimental values are always lower than the industrial ones.
This is due to the fact that all of the numerical analyses conducted on the ejectors are characterized by computational thermo-fluid dynamics investigations, which provide entrainment ratio values for given geometries of the elements of the ejector, and all the validations made in the laboratory were used to experimentally check the entrainment ratio predicted by the numerical analysis, but not to improve the geometry in order to maximize the entrainment ratio. This is because numerical analysis does not allow for derivation of the geometry by imposing the value of the entrainment ratio; that is, it does not allow for a direct design [48]. In other words, numerical analyses can estimate, with great precision, the entrainment ratio for a given ejector component geometry without having to build it; however, if one wishes to obtain a high-performance ejector, it is necessary to analyze various geometries, until one having a suitably high entrainment ratio is obtained. On the other hand, maximizing the entrainment ratio involves maximizing the coefficient of performance COP [19] of the ejector refrigeration system ERS. Therefore, it is necessary to know the maximum values of the entrainment ratio that international manufacturers—including that of the ejectors investigated in this work—can obtain and which is the result of thousands of industrial tests and millions of ejectors that have been built and sold.

2. Materials and Methods

2.1. High Performance of Industrial Ejectors

At an Italian company for the construction of food evaporators equipped with ejectors/thermo-compressors purchased from a European company that has manufactured ejectors for 80 years, a survey was conducted to determine the performance, represented by the entrainment ratio, of various industrial ejectors installed in the evaporators. Eight ejectors were identified with a wide variability of boundary conditions—that is, in terms of the pressure of the motive gas pm, pressure of the induced gas pi, and pressure of the mixed gases p4 at the outlet (Figure 1)—to constitute a heterogeneous population.
In Table 1, the operating data of these eight industrial ejectors (ejectors No. 1–8) are presented, including the pressures and stagnation temperatures of the motive gas (pm and Tm, respectively), those of the induced gas (pi and Ti, respectively), the outlet pressure (p4), the mass flow rates of the motive gas (Gm) and the induced gas (Gi), and the diameters of the intake pipes (Di) and outlet (D4). All of the ejectors had motive gas consisting of dry saturated steam and induced gas also consisting of dry saturated steam. Finally, Table 1 shows the industrial entrainment ratios ωind declared by the European manufacturer. They are to be considered optimal, as these ejectors have been improved to have the maximum value of the entrainment ratio ω. The European manufacturer has also declared that their ejectors are sized for a critical pressure pc that is 5% higher than the outlet pressure p4 (Figure 2). This is a safety margin to avoid the risk of operating in off-design mode, in which the entrainment ratio ω would drop drastically, stopping the operation of the evaporator in which the ejector is installed.
The survey was conducted with an inspection of the evaporators installed in Italy. This survey allowed for verification of the regular operation of the evaporators and, therefore, of the installed ejectors. It could not be otherwise, as this European manufacturer is a company that is known internationally and has operated for eighty years in the construction of ejectors.
In addition, ejectors No. 9 and 10 are two ejectors from the same European company, subjected to tests at the Food Processes Engineering Laboratory at the TESAF Dept. of University of Padova. The two ejectors had motive gas consisting of dry saturated steam produced by a boiler operating at pm = 5 bar and Gm = 128 kg h−1 (Figure 3a), and the induced gas consisting of air at pressure of 0.3 and 0.5 bar, respectively, due to the use of a multiple-orifice device (flute) built according to the guidelines of Power [1]. The tests (Figure 3b) confirmed the performances—represented by the entrainment ratio ωind—declared by the manufacturer, as shown in Table 1.

2.2. Mathematical Modelling

2.2.1. Calculation of the Pressure at the Outlet of the Suction Chamber and the Diffuser Efficiency of High-Performance Industrial Ejectors

Below, a mathematical model is developed to determine the entrainment ratio, with respect to the pressure and temperature of the motive gas (pm and Tm, respectively), the pressure and temperature of the induced gas (pi and Ti, respectively), the pressure of the mixed gases at the outlet (p4), the physical properties of the gases themselves, and the isentropic efficiencies, which take into account the frictional effects during expansion and diffusion.
Figure 1 presents the schematic of an ejector, in which an appropriate acceleration of the fluid induced in the suction chamber is provided to reduce the speed difference between the two gases in the mixing chamber [49], therefore reducing irreversibility. In this way, an increase in the entrainment ratio is achieved.
The momentum theorem applied to the mixing chamber [47,50,51], neglecting pressure variations and friction with the mixing chamber wall, gives the following:
w 1 · G m + w 2 · G i = w 3 · ( G m + G i ) ,
where Gm is the mass flow rate of the motive gas (kg s−1), w1 is the supersonic velocity of the motive stream at the inlet of the mixing chamber (m s−1), Gi is the mass flow rate of the induced gas (kg s−1), and w2 is the subsonic velocity of the induced stream at the inlet of the mixing chamber (m s−1). It should be noted that, unlike Eames [3] and other subsequent researchers, the mixing efficiency ηm is not included in Equation (1); that is, following the old German school, ηm is considered unitary.
By defining the entrainment ratio as ω = G i / G m , from Equation (1), we obtain the following:
ω = w 1 w 3 w 3 w 2 .
The three speeds inside the ejector (w1, w2, and w3) can be obtained from the energy equation (first law of thermodynamics) [8,50] applied to the irreversible adiabatic steady flow, as follows:
1.
Motive stream in the nozzle:
w 1 = 2 · η E 1 · Δ h 1 ,
2.
Induced stream in the suction chamber:
w 2 = 2 · η E 2 · Δ h 2 ,
3.
Mixed stream in the mixing chamber:
w 3 = 2 · Δ h 3 η D ,
where Δh1, Δh2, and Δh3, are the isentropic difference in enthalpies of the motive stream in the nozzle, induced stream in the suction chamber, and mixed stream in the diffuser, respectively; ηE1 is the isentropic efficiency in the nozzle; ηE2 is the isentropic efficiency in the suction chamber; and ηD is the isentropic efficiency of the diffuser, which accounts for the whole loss during the pressure gain process due to the shock train and subsonic diffuser section.
In Equations (3)–(5), the velocities of the gases in the m, i, and 4 sections of the ejector (i.e., wm, wi, and w4; see Figure 1) have been omitted, as the corresponding kinetic energies are smaller than 2% of those corresponding to the velocities inside the ejector (i.e., w1, w2, and w3).
By introducing Equations (3)–(5) into Equation (2), we obtain the following:
ω = η E 1 · η D · Δ h 1 Δ h 3 Δ h 3 η E 2 · η D · Δ h 2 .
The isentropic difference in enthalpies for the ideal gas—a valid model for the gases operating in the ejector—are as follows:
4.
Motive stream in the nozzle:
Δ h 1 = k m k m 1 R m · T m [ 1 ( p 1 p m ) k m 1 k m ] ,
5.
Induced stream in the suction chamber:
Δ h 2 = k i k i 1 R i · T i [ 1 ( p 2 p i ) k i 1 k i ] ,
6.
Mixed stream in the diffuser:
Δ h 3 = k 4 k 4 1 R 4 · T 4 [ 1 ( p 3 p 4 ) k 4 1 k 4 ] ,
where km is the ratio of specific heats cp/cv of the motive gas; ki is the ratio of specific heats cp/cv of the induced gas; k4 is the ratio of specific heats cp/cv of the mixed gases; Rm is the specific constant of the motive gas; Ri is the specific constant of the induced gas; R4 is the specific constant of the mixed gases; p1 is the pressure of the motive gas at the exit of the nozzle, which is equal to p2; p2 is the pressure of the induced gas at the exit of the suction chamber, which is equal to the pressure at the inlet of mixing chamber; pm is the pressure of the motive gas at the inlet of the nozzle; pi is the pressure of the induced gas at the inlet of the suction chamber; p3 is the pressure of the mixed gas at the end of mixing chamber, which is supposed to be equal to the pressure at the inlet of mixing chamber p2 (isobaric process of mixing); p4 is the pressure of the mixed gases at the exit of the diffuser; and Tm is the absolute stagnation temperature (K) of the motive gas, which is supposed to be equal to the absolute temperature at the inlet of nozzle (tm is the same temperature in °C); Ti is the absolute stagnation temperature of the induced gas, which is supposed to be equal to the absolute temperature at the inlet of the suction chamber; and T4 is the absolute stagnation temperature of the mixed gases, which is supposed to be equal to the absolute temperature at the exit of the diffuser.
The equations for the calculation of the specific heat at constant pressure cp4 and that at constant volume cv4, the ratio of specific heats k4, the absolute stagnation temperature T4, and the specific constant of the mixed gases R4 are provided in Appendix A.
By combining Equations (6)–(9), we obtain the following:
ω = η E 1   ·   η D   ·   k m k m     1 R m   ·   T m [ 1 ( p 2 p m ) k m 1 k m ] k 4 k 4 1 R 4   ·   T 4 [ 1 ( p 2 p 4 ) k 4 1 k 4 ] k 4 k 4 1 R 4   ·   T 4 [ 1 ( p 2 p 4 ) k 4 1 k 4 ] η E 2   ·   η D   ·   k i k i 1 R i   ·   T i [ 1 ( p 2 p i ) k i 1 k i ]
In this equation, considering the appropriate design of the nozzle, the high Reynolds number, and the accelerating flow with the consequence of a very thin boundary layer compared to the nozzle section, the isentropic efficiency of the nozzle ηE1 has limited variation, ranging from 0.95 to 0.99 [8,19]. Furthermore, the isentropic efficiency of the suction chamber ηE2, before the mixing chamber, also has limited variation (from 0.92 to 0.99) [8,19]. For caution, here, ηE1 is assumed to have a constant value of 0.95 and ηE2 has a constant value of 0.92, as the wall of the suction chamber is considered to be rough.
In the diffusion process from the last part of the mixing chamber to the ejector outlet, the isentropic efficiency ηD is not constant, due the presence of a shock train. In a similar case considering a cylindrical diffuser with initial supersonic motion, Shapiro [8] indicated that ηD is dependent of the inlet Mach number. In this work, a dependence of ηD on the Mach number of the motive stream at the nozzle exit, Mm−2, was found; this Mach number is as follows [52]:
M m 2 = 2 · η E 1 k m 1 [ ( p m p 2 ) k m 1 k m 1 ] .
Equation (10) shows that the entrainment ratio, ω, depends on two important variables: the efficiency of the diffuser ηD and the pressure of induced gas at the exit of suction chamber p2, which is equal to the pressure at the nozzle exit p1 (Figure 1).
The diagram shown in Figure 4 represents the function of the entrainment ratio ω with respect to the pressure p2 and the isentropic efficiency of the diffuser ηD; that is, ω = f ( p 2 , η D ) .
The diagram shows that, for each diffusion efficiency value ηD, the function ω = f ( p 2 ) has a maximum (Figure 5). Therefore, the entrainment ratio ω has a maximum value corresponding to a precise value of the pressure p2, which accelerates the induced gas such that the lower speed difference between the motive gas and the induced gas reduces the frictional effects in the mixing chamber, while avoiding an excessive compression ratio in the diffuser which, instead, negatively affects the entrainment ratio.
The values of the entrainment ratio ωind presented in Table 1 concern high-performance industrial ejectors, in which the manufacturer has also foreseen the correct acceleration of the induced stream in the suction chamber to obtain the high entrainment ratio (i.e., ωind). If these ωind values are introduced into (10), the equation is left with two unknowns: The pressure p2 and the diffuser efficiency ηD. It seems to be a problem without solution, but studying the function ω = f ( p 2 ) , which is parametric in ηD (shown in Figure 5), as well as the function η D = f ( p 2 ) , which is parametric in ω (shown in Figure 6) we find that when the former, ω = f ( p 2 ) , presents its maximum, the latter, η D = f ( p 2 ) , presents its minimum.
Therefore, it is possible to determine the value of p2 that minimizes the isentropic efficiency of the diffuser, by zeroing the derivative of the function η D = f ( p 2 , ω ) , providing the parameter ω with the values ωind in Table 1 for the high-performance industrial ejectors. The solution provides p2 values that minimize the diffuser efficiency while also maximizing the entrainment ratio. The function η D = f ( p 2 , ω i n d ) is obtained from (10):
η D = k 4 k 4     1 R 4   ·   T 4 [ 1     ( p 2 p 4 ) k 4     1 k 4 ]   ·   ( 1   +   ω i n d ) 2 { η E 1   ·   k m k m     1 R m   ·   T m [ 1     ( p 2 p m ) k m     1 k m ] + ω i n d 2   ·   η E 2   ·   k i k i     1 R i   ·   T i [ 1     ( p 2 p i ) k i     1 k i ] } 2
Taking the derivative of (12) and equating it to zero, we have the following:
η D p 2 = R 4   ·   T 4   ·   ( p 2 ) 1 k 4 ( p 4 ) k 4 1 k 4   ·   ( 1   +   ω i n d ) 2 [ F m + ω i n d 2   ·   F i ] 2 + 2   ·   F 4   ·   ( 1   +   ω i n d ) 2 [ η E 1   ·   R m   ·   T m   ·   ( p 2 ) 1 k m ( p m ) k m 1 k m F m   +   ω i n d 2   ·   η E 2   ·   R i   ·   T i   ·   ( p 2 ) 1 k i ( p i ) k i 1 k i ω i n d 2   ·   F i ] [ F m   +   ω i n d 2   ·   F i ] 3 = 0 ,
where Fm, Fi, and F4 are defined as follows:
F m = η E 1 · k m k m 1 R m · T m [ 1 ( p 2 p m ) k m 1 k m ] ,
F i = η E 2 · k i k i 1 R i · T i [ 1 ( p 2 p i ) k i 1 k i ] ,
F 4 = k 4 k 4 1 R 4 · T 4 [ 1 ( p 2 p 4 ) k 4 1 k 4 ] .
Equation (13), which can be solved an iterative method, provides the value of p2 that improves the ejector, as it is also the one that maximizes the entrainment ratio ω.
The value of p2, inserted into Equation (12), allows us to obtain the efficiency of the diffuser ηD in high-performance industrial ejectors.
As mentioned above, it is to be expected that ηD depends on the Mach number of the motive gas in the second section, Mm−2, at the inlet of the mixing chamber. As already said, the calculation method of p2 described by Equation (13) and that of ηD through Equation (12) can be applied to the high-performance industrial ejectors detailed in Table 1. The obtained p2 values also allow for obtaining the Mach number, Mm−2, through Equation (11). Therefore, the relation η D = f ( M m 2 ) can be easily found, as discussed in the results section.

2.2.2. Calculation of the Entrainment Ratio for Ejectors to Be Improved

The calculation of the Mach number Mm−2 of the motive stream passing from the pressure pm to the best pressure p2 requires determination of p2, which is still unknown. This means that an iterative calculation procedure must be implemented; however, this can be avoided if the Mach number Mm−2 is replaced by the Mach number Mm−i of the motive stream at the final pressure equal to the suction pressure pi (Figure 1). In fact, pi is known as it is a boundary condition of the ejector. As detailed in the results, the equation η D = f ( M m 2 ) obtained by regression (Figure 7) by the procedure of Section 2.2.1, has a good coefficient of determination R2, but the function η D = f ( M m i ) (Figure 8) has a higher R2. This is a good reason to always use the relationship η D = f ( M m i ) . Therefore, the procedure for calculating the entrainment ratio ω, as outlined below, begins with the calculation of Mm−i:
M m i = 2 · η E 1 k m 1 [ ( p m p i ) k m 1 k m 1 ] .
The next calculation concerns the efficiency of the diffuser ηD through the function η D = f ( M m i ) , as presented in the results (Section 3.1).
Setting the derivative of Equation (10) equal to zero allows us to obtain the value of the best pressure p2 which maximizes the entrainment ratio ω:
ω p 2 = [ R 4 · T 4 · ( p 2 ) 1 k 4 ( p 4 ) k 4 1 k 4 F 4 2 η D · η E 1 · R m · T m · ( p 2 ) 1 k m ( p m ) k m 1 k m η D F m 2 ] ( F 4 η D F i ) 1 { [ η D F m F 4 η D F i 2 ] ( η D · η E 2 · R i · T i · ( p 2 ) 1 k i ( p i ) k i 1 k i ) [ R 4   ·   T 4   ·   ( p 2 ) 1 k 4 ( p 4 ) k 4 1 k 4 F 4 2 ] } ( F 4 η D F i ) 2 = 0 ,
where Fm, Fi, and F4 are given by Equations (14)–(16), respectively.
Equation (18), similar to Equation (13), also requires solution through an iterative method, for example, by a spreadsheet.
Finally, the pressure value p2 obtained from Equation (18) allows for obtaining the entrainment ratio ωind maximized through Equation (10).
The solution of Equations (18) and (10) requires knowledge of the values of k4, R4, and T4 which, according to Equations (A1)–(A5), presented in Appendix A, depend on the unknown value of the entrainment ratio ω.
If the induced gas is equal to the motive gas, then km = ki = k4 and Rm = Ri = R4, but the need remains to determine the stagnation temperature T4 through Equation (A5), starting from the stagnation temperatures Tm and Ti and from the entrainment ratio ω. The absolute stagnation temperatures T4, Tm, and Ti practically coincide with the absolute temperatures as (see Figure 1), in Section 4, m, and i, the gases generally have velocities lower than 50 m s−1, often around 30 m s−1. The problem arises of needing to know ω a priori to calculate T4 which, in turn, serves to calculate the same ω by Equation (18). The problem can be overcome by inserting into the spreadsheet, before Equations (18) and (10), a cell containing an initial value of ω (experience suggests ω = 1) and cells containing Equations (A1)–(A5), which provide the values of cp4, cv4, R4, k4, and T4 resulting from this initial ω. Additionally, it is necessary to enter an initial value for p2 (experience suggests p 2 = 0.99 · p i ).
Therefore, the procedure includes the following:
(1)
Execution of the iterative method (e.g., by spreadsheet) to search for the ω value associated to p 2 = 0.99 · p i , using Equation (10);
(2)
Execution of the iterative method (e.g., by spreadsheet) to search for the best value of p2, using Equation (18);
(3)
Final execution of the iterative method (e.g., by spreadsheet) to recalculate the final industrial ω value, now for the best p2 obtained in step 2, by Equation (10); and
(4)
Final execution of the iterative method (e.g., by spreadsheet) to recalculate the final best p2 value, using Equation (18).
Only two iterations of steps (1) and (3) are sufficient to maximize ω and only two for steps (2) and (4) are required to improve p2, as the solution method is strongly convergent.

2.3. Experimental Entrainment Ratio of Benchmark Ejectors Built and Tested to Use in ERSs

Benchmark ejectors built and tested by various researchers in their laboratories over the past twenty years, for use in ERS prototypes, have been considered. Details on an initial series of such benchmark ejectors, in terms of their geometric characteristics, boundary conditions, and performance (represented by the entrainment ratio measured in the laboratory) have been presented in the work of Besagni [19] exactly in its Tables 6, 7 and 11.
To check whether these ejectors were improved (i.e., if they were high-performance ejectors), the mathematical model for calculating the industrial high entrainment ratio ωind proposed in this work was applied.
For some of these ejectors, only the experimental values of the boundary conditions and the entrainment ratio corresponding to the critical point were available while, for others, these values were available both in on-design and off-design contexts (Figure 2). For the latter, using the results reported by [19], the value of the entrainment ratio ω and the corresponding outlet pressure p4 of the critical point were identified. The critical point was not always uniquely identifiable. Therefore, in these cases, two pairs of ω–p4 values were taken into consideration, that is, those that could be closest (one to the right and one to the left) to the ω–p4 pair of the critical point.
Therefore, Table 2 summarizes the values of the boundary conditions and the experimental entrainment ratio ωexp corresponding to the critical point for each ejector, except in some cases, in which there were two critical pseudo-points, marked with the symbol *. The alphanumeric codes reported, which distinguish each ejector (e.g., A1f), are the same as those used in [19].
A second series of benchmark ejectors, with their geometric characteristics, boundary conditions, and performance represented by the entrainment ratio measured in the laboratory, was found in the works [62,63,64,65,66], as summarized in Table 3.

3. Results

3.1. High-Performance Industrial Ejectors

For the ten high-performance industrial ejectors produced by the European manufacturer, Equation (13) was applied, together with Equations (A1)–(A5), in order to determine the best values of the pressure p2 (Table 4) of the induced gas at the exit of the suction chamber. It should be noted that the value of the specific heat ratio of motive steam, km, used in the calculations, was not equal to 1.3 (typical of superheated steam), but was equal to an average empirical value of 1.14 (which corresponds to wet steam) [67]. In fact, the motive steam, during the expansion in the nozzle starting from the dry saturated condition, assumes the wet condition. The ki value was also equal to 1.14 for the first eight ejectors, which intake dry saturated steam which, when expanded in the suction chamber, became wet steam. Meanwhile, for ejectors 9 and 10, which intake air, ki was set to 1.4.
Equation (12) was applied to the same ten high-performance industrial ejectors, in order to find the corresponding values of the diffuser efficiency ηD. Table 4 also reports the values of the Mach number Mm−2 calculated using Equation (11), as well as the values of the Mach number Mm−i calculated using Equation (17). The values of ηD vs. Mm−2 are plotted in Figure 7, which also shows the regression line (R2 = 0.9725).
As mentioned in Section 2.2.2, determining the Mach number Mm−2 requires prior knowledge of the pressure p2. To overcome this problem, Equation (17) was used to calculate the Mach number Mm−i reached by the motive gas at the outlet of the nozzle with the pressure pi of the induced gas, as if this were not accelerated in the suction chamber by wi to w2 (Figure 1). The resulting diagram (Figure 8) shows that the correlation between the efficiency of the diffuser ηD and the Mach number Mm−i was better than the previous one, with a coefficient of determination R2 equal to 0.9741. Therefore, the relationship found η D = f ( M m i ) , as a result of linear regression, as follows:
η D = 0.0609 · M m i + 0.932

3.2. Validation of the Mathematical Modelling of the Entrainment Ratio vs. the Values of Industrial Ejectors

To validate the results of the mathematical modelling of the industrial entrainment ratio, represented by Equations (18) and (10), it is necessary to compare the calculated entrainment ratios from Equation (10) with those declared by the manufacturer of the ten high performance ejectors (Table 1). This comparison produced an average error only of 1.7%, due to the linear relationship presented in Equation (19), characterized by R2 = 0.9741 < 1.

3.3. Calculation of the Industrial Entrainment Ratio on Benchmark Ejectors Built and Tested to Use in ERSs

Following the method indicated in Section 2.2.2, Equation (17) allowed for calculation of the Mach Number Mm−i. Then, Equation (19) made it possible to calculate the efficiency of the diffuser ηD. Consequently, Equation (18) allowed us to calculate the best pressure p2, and finally, Equation (10) allowed us to calculate the industrial entrainment ratio ωind.
This method was applied to each of the benchmark ejectors presented in Table 2 and Table 3. The calculation results are shown in Table 5 and Table 6, respectively, together with the experimental entrainment ratio value ωexp and the relative percentage difference: δ = ω e x p ω i n d ω i n d · 100 .
Equations (10), (17) and (18), as well as the auxiliary Equations (14)–(16), contain the quantities km, ki, k4, Rm, Ri, and R4. In all ejectors presented in Table 2 and Table 3, the induced gas was the same as the motive gas. Therefore, km = ki = k4 = k and Rm = Ri = R4 = R. Table 5 and Table 6 shows the values of the gas constant R calculated for each gas by the following formula:
R = R ¯ m m ,
where mm is the molecular mass found in [19] and R ¯ is the universal gas constant equal to 8314 (J kmole−1K−1).
Table 5 and Table 6 also show the values of the ratio of specific heat k. As the bibliography provided disparate values, they were calculated as follows: For each group of ejectors characterized by the same gas, the average of the pressures pm and temperatures Tm (Figure 1 and Table 2 and Table 3) and the average of the pressures pi were calculated, noting that the S.D. was limited. The average value of the pressure pm and temperature Tm of each gas made it possible to trace the point of the boundary condition m (Figure 1) on the p–h diagram of the gas. Considering the isentropic expansion of the motive gas in the nozzle, the mean value of pi made it possible to trace, on the p–h diagram, the point of isentropic conditions at the outlet of the nozzle. In correspondence to the two points (m and i) the p–h diagram provided the enthalpy values hm and hi-is and, therefore, the isentropic difference in enthalpies Δ h m - i s = ( h m h i - i s ) . As the temperature Tm-is can be read from the p–h diagram, and as Δ h m - i s = c p · ( T m T i - i s ) , the specific heat at constant pressure cp can therefore be obtained as follows:
c p = Δ h m - i s ( T m T i - i s ) .
Meyer’s relationship provides the specific heat at constant volume cv:
c v = c p R .
Finally, k, the ratio of the two specific heats, is obtained as follows:
k = c p c v .
After having obtained the best pressure p2 with the method in Section 2.2.2, Equation (8), with ki = k and Ri = R, provides the value of the isentropic difference in the enthalpies Δh2. Plotting of the points i and 2is on the p–h diagram made it possible to measure the exact value of Δh2, which was found to be superimposable to the value provided by Equation (8), with an error of less than 2%.
Therefore, the ratio of specific heats k obtained from the expansion of the motive gas was also usable for the expansion of the induced gas and, according to Equations (A1)–(A3), also for the diffusion of the mixed gases.
On the other hand, this method for calculating the specific heat ratio k is the same as that adopted previously [67] to quantify the k of the wet steam result equal to 1.14. It is very different from the value of 1.3 for dry steam, as the specific heat cp of the wet steam calculated by Equation (21) considers the latent heat released during the partial condensation and which is present in the quantification of Δ h m - i s = ( h m h i - i s ) by the h–s Mollier diagram.
Differences between the experimental and industrial entrainment ratios presented in Table 5 and Table 6 are also depicted in the histograms of Figure 9 and Figure 10, respectively.

4. Discussion and Conclusions

The long experience, extending for over a century, of the manufacturers of ejectors for the chemical, oil, and food industries, has led some of them to improve their products, where improving an ejector means maximizing the entrainment ratio ω. For this reason, these high entrainment ratios are called here: industrial entrainment ratios.
However, Chen, in his 1997 work [63], wrote: “Although ejector manufacturers possess much valuable experimental data, they are understandably reluctant to publish this data”. Almost a quarter of a century has passed since then, and Chen’s statement is still valid. It is even impossible to know the entrainment ratio values with respect to the boundary conditions unless the ejector is purchased.
The first studies conducted in Germany in the 1930s led to a mathematical model for calculating the entrainment ratio ω based on the momentum theorem and on the identification of isentropic efficiencies of expansion and diffusion. The same German studies showed that it was necessary to accelerate the induced gas before it entered the mixing chamber (Figure 1). The acceleration in the suction chamber—and, therefore, the reduction in the pressure p2 at the inlet of the mixing chamber—cannot be indiscriminate but should be designed to reduce the speed difference between the motive gas and the induced gas in the mixing chamber, while avoiding an excessive compression ratio in the diffuser which, instead, negatively affects the entrainment ratio ω.
The best value of this acceleration—and, therefore, the best value of the discharge pressure from the suction chamber of the induced gas p2—is usually determined [68] by searching for the maximum of the function represented by Equation (10), assuming the value of the diffuser efficiency ηD to be constant.
In this work, the diffuser efficiency ηD was instead considered to be variable, depending on the Mach number of the motive gas at the inlet in the mixing chamber.
Therefore, Equation (10) was studied analytically as a function of two variables: p2 and ηD (Figure 4, Figure 5 and Figure 6). Two equations were produced as a result: Equation (13), of which the solution provides the best values of p2 with respect to the industrial entrainment ratio ωind, and Equation (12), which provides the efficiency of the diffuser ηD always with respect to the industrial entrainment ratio ωind.
If the entrainment ratio ωind inserted into Equations (13) and (12) is that of a high-performance ejector, such as those produced by the most competitive manufacturers mentioned at the beginning of this discussion, then Equation (13) gives the relative value of p2 and Equation (12) provides the efficiency of the diffuser ηD of this ejector.
This calculation method was applied to ten high-performance industrial ejectors from a major European company, which has operated in the global ejector market for eighty years. This choice was made in order to have a heterogeneous population of boundary conditions (i.e., pm, pi, and p4). To circumvent the problem of scarce data availability for industrial ejectors, as stated by Chen [63], a manufacturer of evaporators for the food industry was asked to make available their data archive of the ejectors purchased over the course of twenty years from a big European company.
For each ejector, in addition to p2 and ηD, the Mach number of the motive gas Mm−i was also calculated, using Equation (17). Comparison between the efficiency of the diffuser ηD and the Mach number Mm−i produced Equation (19), which is a linear regression function characterized by a high coefficient of determination R2, confirming that the hypothesis η D = f ( M m i ) is consistent.
In the second part of the present work, comprising mathematical modelling, from Equation (10), the function (18) was obtained, of which the solution gives the value of the pressure p2 which maximizes the entrainment ratio. Equation (18) contains the quantity ηD which depends, through Equation (19), on the Mach Number Mm−i which, in turn, according to Equation (17), is only dependent on the known boundary conditions pm and pi. Finally, Equation (10) can provide the maximized value of entrainment ratio, that is industrial entrainment ratio ωind.
This procedure was applied to a series of 68 ejectors built and tested experimentally over the last twenty years or so (Table 2 and Table 3), as a part of research activities aimed at developing heat-driven ejector refrigeration systems (ERSs); for this reason, they were called benchmark ejectors [19]. The renewed interest in this type of refrigeration system, mainly due to environmental concerns (as they can operate on solar energy), has activated the interest of many researchers towards the study of the functioning of supersonic ejectors through computational methods. Such numerical investigations have produced more and more refined results, are now able to simulate the functioning of the ejectors and to predict, with relative precision, the experimental entrainment ratios ωexp [19] of the benchmark ejectors.
Again, with reference to the second part of this work, the results of the procedure for calculating the industrial entrainment ratio ωind, applied to the series of 68 benchmark ejectors, are presented in Table 5 and Table 6 (as well as in Figure 9 and Figure 10), where they were compared with the experimental entrainment ratio values ωexp.
The experimental entrainment ratios ωexp were always lower than the industrial ones ωind, with a minimum difference of −17% for the ejector of [55] and a maximum difference of about −80% for the ejector of [60]. Limited differences of about −30% were also found for the ejectors of [56,58,59] and only −25% for the ejector of [64]. For many other ejectors, the differences were consistent. This means that, while numerical analysis in this field has shown promise, it must be extended to the search for the best geometry of the ejector elements, first ensuring that the suction chamber has an outlet section which provides the best pressure p2, to be obtained a priori, for example, using the analytical method proposed herein. Second, it is necessary to pay close attention to the shape and size of the mixing chamber, wherein the shock train phenomena are very complex. Therefore, researchers who have the objective of designing a high-performance ejector should repeat the numerical analysis, which has only a checking character, with simulated variations in the shape and size of the mixing chamber, in an attempt to reach the high value of the entrainment ratio of industrial ejector. The industrial value can be calculated, for example, using the method proposed in this work, and corresponds to those of high-performance ejectors already on the market.
Finally, Equation (19) proposed in this work, which indicates the influence of the Mach number of motive gas at the inlet of mixing chamber Mm−i on the diffuser efficiency ηD with good precision, allows for calculation of the latter in the extreme situations encountered during the survey on 68 laboratory ejectors (Table 5 and Table 6). These are the ejectors [61] with a minimum Mach number Mm−i of about 1, for which Equation (19) proposes a diffuser efficiency ηD of 0.870 and, excluding the ejector of [54], the ejector of [62] with the relative maximum Mm−i of almost four, for which the diffuser efficiency ηD is 0.710. It must be remembered that, in this work, the diffuser efficiency ηD takes into account the whole loss during the pressure gain process due to shock train and subsonic diffuser section and, as a unitary mixing efficiency was assumed in Equation (1), any energy losses during mixing are also represented by ηD, due to the adopted mode of calculation of ηD.
Furthermore, study of the partial derivative of Equation (10) shows an average value of ω η D = 2.1 ± 0.1 . If a constant diffuser efficiency ηD had been chosen, for example, equal to 0.770, as indicated by [68] (or 0.807, as indicated by [16,23]), in the ejector of [61], the absolute error in the maximum entrainment ratio prediction ωind would have been Δ ω = ω Δ η D · Δ η D = 2.1 · ( 0.770 0.870 ) = 0.21 , corresponding to a relative error of Δ ω % = Δ ω ω t o p · 100 = 0.21 1.02 · 100 = 20.5 % . Similarly, the case of the ejector of [62], the absolute error is Δ ω = ω η D · Δ η D = 2.1 · ( 0.770 0.710 ) = 0.126 , corresponding to a relative error of Δ ω % = Δ ω ω t o p · 100 = 0.126 0.588 · 100 = 21.4 % . Therefore, Equation (19) allows the elimination of relative errors in the ωind prediction between about ±20%. The next step will be to expand the sample of industrial ejectors, especially towards the high values of the Mach number, in order to verify Equation (19) even in cases such as the (rare) one of [54], with Mm−i equal to about 7.
In conclusion, it is useful to say that in this work, the objective was not for the design method of ejectors to have the optimal entrainment ratio, but rather for an algorithm to determine the maximum value of the entrainment ratio, built through the application of Calculus [69,70,71]. This algorithm can be used to determine the maximum (therefore optimal) value of the entrainment ratio ω of the ejector operating at the expected boundary conditions represented by the pressures and temperatures of the motive gas (pm, Tm), of the induced gas (pi, Ti), and of the pressure of the mixed gases (p4). The maximum values, therefore optimal, of entrainment ratios obtained with the algorithm are those (ωind) corresponding to the industrial ejectors of the big European manufacturer chosen because it is the one with the longest experimental research history and with the greatest diffusion in the world market of ejectors.
Starting from the best value of entrainment ratio ω = ωind, obtainable with the algorithm of this work, the researchers with experience in CFD applied to ejectors, will have to use the computational methods repeatedly, varying the dimensions especially of the mixing chamber, until the CFD provides the value of the entrainment ratio equal to ωind. This is a similar method adopted in machine design with the FEA [72].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are contained within the article.

Acknowledgments

Roberto Stivanello, technician of Food Processes Engineering Laboratory, for his contribution during the tests on the industrial ejectors.

Conflicts of Interest

The author declares no conflict of interest.

Nomenclature

SymbolQuantity
cpspecific heat at constant pressure
cvspecific heat at constant volume
Diintake pipe diameter of the induced gas
D4exit pipe diameter of the mixed gases
Gmmass flow rate of the motive gas
Gimass flow rate of the induced gas
kmratio of specific heats cp/cv of the motive gas
kiratio of specific heats cp/cv of the induced gas
k4ratio of specific heats cp/cv of the mixed gases
Mm−2Mach number of the motive gas at the nozzle exit when pressure is p1 = p2
Mm−iMach number of the motive gas at the nozzle exit when pressure is p1 = pi
pmpressure of the motive gas at the inlet of the nozzle
pipressure of the induced gas at the inlet of the suction chamber
p1pressure of the motive gas at the nozzle exit
p2pressure of the induced gas at the exit of the suction chamber, equal to p1
p3pressure of the mixed gas at the end of mixing chamber, equal to p2
p4pressure of the mixed gases at the exit of the diffuser
Rmspecific constant of the motive gas
Ri specific constant of the induced gas
R4specific constant of the mixed gases
Tmabsolute stagnation temperature (K) of the motive gas
tmstagnation temperature (°C) of the motive gas
Tiabsolute stagnation temperature (K) of the induced gas
T4absolute stagnation temperature (K) of the mixed gases
w1supersonic velocity of the motive gas at the inlet of mixing chamber
w2subsonic velocity of the induced gas at the inlet of mixing chamber
w3velocity of the mixed gases in the mixing chamber
wmvelocity of the motive gas at the inlet of the nozzle
wivelocity of the induced gas at the inlet of the suction chamber
w4velocity of the mixed gases at the exit of the diffusor
Greek symbolQuantity
Δh1isentropic difference in enthalpies of the motive stream in the nozzle
Δh2isentropic difference in enthalpies of the induced stream in the suction chamber
Δh3isentropic difference in enthalpies of the mixed stream in the diffusion
ηE1isentropic efficiency in the nozzle
ηE2isentropic efficiency in the suction chamber
ηDisentropic efficiency of the diffusion
ωEntrainment ratio (Gi/Gm)
ωindEntrainment ratio of the industrial ejectors (therefore optimal)
ωexpEntrainment ratio of the experimental laboratory ejectors designed by CFD

Appendix A

The specific heat at constant pressure of the mixed gases, cp4, is dependent on the specific heat of motive gas, cpm, of the specific heat of induced gas, cpi, and of the entrainment ratio ω:
c p 4 = c p m + ω · c p i 1 + ω
Similarly, the specific heat at constant volume of the mixed gases, cv4, is as follows:
c v 4 = c v m + ω · c v i 1 + ω
Hence, the specific heat ratio of the mixed gases, k4, is as follows:
k 4 = c v 4 c p 4 .
The specific gas constant of mixed gases, R4, is given by the following:
R 4 = c p 4 c v 4 .
Finally, due to enthalpy balance, the absolute stagnation temperature of the mixed fluid, T4, is as follows:
T 4 = c p m · T m + ω · c p i · T i c p m + ω · c p i .

References

  1. Power, R.B. Steam Jet Ejector for the Process Industries, 1st ed.; Mac Graw-Hill: New York, NY, USA, 1994; pp. 7–18. [Google Scholar]
  2. Matsuo, K.; Miyazato, Y.; Kim, H.D. Shock train and pseudo-shock phenomena in internal gas flows. Prog. Aerosp. Sci. 1999, 35, 33–100. [Google Scholar] [CrossRef]
  3. Thevenot, R. Histoire du Froid Artificiel, 1st ed.; Institut International du Froid: Paris, France, 1978; pp. 63–64. [Google Scholar]
  4. Pollerberg, C.; Heinzel, A.; Weidner, E. Model of a solar driven steam jet ejector chiller and investigation of its dynamic operational behaviour. Sol. Energy 2009, 83, 732–742. [Google Scholar] [CrossRef]
  5. Keenan, J.H.; Neumann, E.P. A simple air ejector. ASME J. Appl. Mech. 1942, 9, A75–A81. [Google Scholar] [CrossRef]
  6. Keenan, J.H.; Neumann, E.P.; Lustwerk, F. An investigation of ejector design by analysis and experiment. ASME J. Appl. Mech. 1950, 17, 299–309. [Google Scholar] [CrossRef]
  7. Lukasiewicz, J. Diffusers for supersonic wind tunnels. J. Aeronaut. Sci. 1953, 20, 617–626. [Google Scholar] [CrossRef]
  8. Shapiro, A.H. The Dynamics and Thermodynamics of Compressible Fluid Flow, 1st ed.; Ronald Press: New York, NY, USA, 1953; Volume 1, pp. 99, 135–139, 152. [Google Scholar]
  9. Nill, L.D.; Mattick, A.T. An experimental study of shock structure in a normal shock train. In Proceedings of the AIAA 34th Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 15–18 January 1996; p. 96-0799. [Google Scholar] [CrossRef]
  10. McCormick, D.C. Shock boundary-layer interaction control with vortex generators and passive cavity. AIAA J. 1993, 31, 91–96. [Google Scholar] [CrossRef]
  11. Sullins, G.; McLafferty, G. Experimental results of shock trains in rectangular ducts. In Proceedings of the AIAA 4th International Aerospace Planes Conference, Orlando, FL, USA, 1–4 December 1992; p. 92-5103. [Google Scholar] [CrossRef]
  12. Carroll, B.F.; Dutton, J.C. Multiple normal shock wave/turbulent boundary-layer interactions. J. Propuls. Power 1992, 8, 441–448. [Google Scholar] [CrossRef]
  13. Lin, P.; Rao, G.V.R.; O’Connor, G.M. Numerical analysis of normal shock train in a constant area isolator. In Proceedings of the AIAA 27th Joint Propulsion Conference, Sacramento, CA, USA, 24–26 June 1991; p. 91-2162. [Google Scholar] [CrossRef]
  14. Sajben, M.; Donovan, J.F.; Morris, M.J. Experimental investigation of terminal shock sensors for mixed-compression inlets. J. Propuls. Power 1992, 8, 168–174. [Google Scholar] [CrossRef]
  15. Munday, J.T.; Bagster, D.F. A new ejector theory applied to steam jet refrigeration. Ind. Eng. Chem. Process Des. Dev. 1977, 16, 442–449. [Google Scholar] [CrossRef]
  16. Eames, I.W.; Aphornaratana, S.; Haider, H. A theoretical and experimental study of a smallscale steam jet refrigerator. Int. J. Refrig. 1995, 18, 378–385. [Google Scholar] [CrossRef]
  17. Huang, B.J.; Chang, J.M.; Wang, C.P.; Petrenko, V.A. A 1-D analysis of ejector performance. Int. J. Refrig. 1999, 18, 354–364. [Google Scholar] [CrossRef]
  18. Rogdakis, E.D.; Alexis, G.K. Investigation of ejector design at optimum operating condition. Energy Convers. Manag. 2000, 41, 1841–1849. [Google Scholar] [CrossRef]
  19. Besagni, G.; Cristiani, N.; Croci, L.; Guedon, G.R.; Inzoli, F. Computational fluid-dynamics modelling of supersonic ejectors: Screening of modelling approaches, comprehensive validation and assessment of ejector component efficiencies. Appl. Therm. Eng. 2021, 186, 116431. [Google Scholar] [CrossRef]
  20. Rusly, E.; Aye, L.; Charters, W.W.S.; Ooi, A. CFD analysis of ejector in a combined ejector cooling system. Int. J. Refrig. 2005, 28, 1092–1101. [Google Scholar] [CrossRef]
  21. Pianthong, K.; Seehanam, W.; Behnia, M.; Sriveerakul, T.; Aphornratana, S. Investigation and improvement of ejector refrigeration system using computational fluid dynamics technique. Energy Convers. Manag. 2007, 48, 2556–2564. [Google Scholar] [CrossRef]
  22. Zhu, Y.; Cai, W.; Wen, C.; Li, Y. Numerical investigation of geometry parameters for design of high-performance ejectors. Appl. Therm. Eng. 2009, 29, 898–905. [Google Scholar] [CrossRef]
  23. Varga, S.; Oliveira, A.C.; Diaconu, B. Numerical assessment of steam ejector efficiencies using CFD. Int. J. Refrig. 2009, 32, 1203–1211. [Google Scholar] [CrossRef]
  24. Varga, S.; Oliveira, A.C.; Ma, X.; Omer, S.A.; Zhang, W.; Riffat, S.B. Comparison of CFD and experimental performance results of a variable area ratio steam ejector. Int. J. Low-Carbon Technol. 2010, 6, 119–124. [Google Scholar] [CrossRef]
  25. Ji, M.; Utomo, T.; Woo, J.; Lee, Y.; Jeong, H.; Chung, H. CFD investigation on the flow structure inside thermo vapor compressor. Energy 2010, 35, 2694–2702. [Google Scholar] [CrossRef]
  26. Wang, X.D.; Dong, J.L. Numerical study on the performances of steam-jet vacuum pump at different operating conditions. Vacuum 2010, 84, 1341–1346. [Google Scholar] [CrossRef]
  27. 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]
  28. Yazdani, M.; Alahyari, A.A.; Radcliff, T.D. Numerical modeling of two-phase supersonic ejectors for work-recovery applications. Int. J. Heat Mass Transf. 2012, 55, 5744–5753. [Google Scholar] [CrossRef]
  29. Sharifi, N.; Boroomand, M. An investigation of thermo-compressor design by analysis and experiment: Part 1. Validation of the numerical method. Energy Convers. Manag. 2013, 69, 217–227. [Google Scholar] [CrossRef]
  30. Wang, X.; Dong, J.; Li, A.; Lei, H.; Tu, J. Numerical study of primary steam superheating effects on steam ejector flow and its pumping performance. Energy 2014, 78, 205–211. [Google Scholar] [CrossRef]
  31. Gagan, J.; Smierciew, K.; Butrymowicz, D.; Karwacki, J. Comparative study of turbulence models in application to gas ejectors. Int. J. Therm. Sci. 2014, 78, 9–15. [Google Scholar] [CrossRef]
  32. Zhu, Y.; Jiang, P. Experimental and numerical investigation of the effect of shock wave characteristics on the ejector performance. Int. J. Refrig. 2014, 40, 31–42. [Google Scholar] [CrossRef]
  33. Chandra, V.V.; Ahmed, M.R. Experimental and computational studies on a steam jet refrigeration system with constant area and variable area ejectors. Energy Convers. Manag. 2014, 79, 377–386. [Google Scholar] [CrossRef]
  34. Banasiak, K.; Palacz, M.; Hafner, A.; Bulinski, Z.; Smołka, J.; Nowak, A.J.; Fic, A. A CFD-based investigation of the energy performance of two-phase R744 ejectors to recover the expansion work in refrigeration systems: An irreversibility analysis. Int. J. Refrig. 2014, 40, 328–337. [Google Scholar] [CrossRef]
  35. Rao, S.; Gopalan, J. Observations on the non-mixed length and unsteady shock motion in a two dimensional supersonic ejector. Phys. Fluids 2014, 26, 036103. [Google Scholar] [CrossRef]
  36. Arun, K.M.; Tiwari, S.; Mani, A. Three-dimensional numerical investigations on rectangular cross-section ejector. Int. J. Therm. Sci. 2017, 122, 257–265. [Google Scholar] [CrossRef]
  37. Kong, F.; Kim, H.D. Analytical and computational studies on the performance of a two-stage ejector–diffuser system. Int. J. Heat Mass Transf. 2015, 85, 71–87. [Google Scholar] [CrossRef]
  38. Besagni, G.; Inzoli, F. Computational fluid-dynamics modeling of supersonic ejectors: Screening of turbulence modeling approaches. Appl. Therm. Eng. 2017, 117, 122–144. [Google Scholar] [CrossRef]
  39. Varga, S.; Soares, J.; Lima, R.; Oliveira, A.C. On the selection of a turbulence model for the simulation of steam ejectors using CFD. Int. J. Low-Carbon Technol. 2017, 12, 233–243. [Google Scholar] [CrossRef]
  40. Mazzelli, F.; Giacomelli, F.; Milazzo, A. CFD modeling of condensing steam ejectors: Comparison with an experimental test-case. Int. J. Therm. Sci. 2018, 127, 7–18. [Google Scholar] [CrossRef]
  41. Haghparast, P.; Sorin, M.V.; Nesreddine, H. The impact of internal ejector working characteristics and geometry on the performance of a refrigeration cycle. Energy 2018, 162, 728–743. [Google Scholar] [CrossRef]
  42. Thongtip, T.; Aphornratana, S. Development and performance of a heat driven R141b ejector air conditioner: Application in hot climate country. Energy 2018, 160, 556–572. [Google Scholar] [CrossRef]
  43. 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]
  44. Smierciew, K.; Gagan, J.; Butrymowicz, D. Application of numerical modelling for design and improvement of performance of gas ejector. Appl. Therm. Eng. 2019, 149, 85–93. [Google Scholar] [CrossRef]
  45. Mohamed, S.; Shatilla, Y.; Zhang, T. CFD-based design and simulation of hydrocarbon ejector for cooling. Energy 2019, 167, 346–358. [Google Scholar] [CrossRef]
  46. Atmaca, M.; Ezgi, C. Three-dimensional CFD modeling of a steam ejector, Energy Sources, Part A: Recovery. Utilizat. Environ. Effects 2019, 201223538. [Google Scholar] [CrossRef]
  47. Wiegand, J. Bemessung von Dampfstrahl-Verdichtern und ihr Verhalten bei Wechselnden Betriebsbedingungen; VDI-Forschungsheft, VDI-Verlag: Berlin, Germany, 1940; Volume 401, pp. 1–24. [Google Scholar]
  48. Grazzini, G.; Milazzo, A.; Mazzelli, F. Ejectors for Efficient Refrigeration: Design. Applications and Computational Fluid Dynamics, 1st ed.; Springer: Cham, Switzerland, 2018. [Google Scholar] [CrossRef]
  49. Bosnjakovic, F. Dampfbefeuchtung in Strahlgebläsen; VDI-Forschungsheft, VDI-Verlag: Berlin, Germany, 1940; Volume 11, pp. 210–215. [Google Scholar]
  50. Bosnjakovic, F. Über dampfstrahlgebläse. Zeitschr. Ges. Kälteindustrie 1936, 43, 229–233. [Google Scholar]
  51. Weydanz, W. Die vorgänge in strahlapparaten. Beiheft Z. Zeitschr. Ges. Kälteindustrie 1939, Reihe 2, H.8. [Google Scholar]
  52. Grazzini, G.; Mariani, A. A simple program to design a multi-stage jet-pump for refrigeration cycles. Energy Convers. Manag. 1998, 39, 1827–1834. [Google Scholar] [CrossRef]
  53. Han, Y.; Wang, X.; Sun, H.; Zhang, G.; Guo, L.; Tu, J. CFD simulation on the boundary layer separation in the steam ejector and its influence on the pumping performance. Energy 2019, 167, 469–483. [Google Scholar] [CrossRef]
  54. Sriveerakul, T.; Aphornratana, S.; Chunnanond, K. Performance prediction of steam ejector using computational fluid dynamics: Part 1. Validation of the CFD results. Int. J. Therm. Sci. 2007, 46, 812–822. [Google Scholar] [CrossRef]
  55. Gagan, J.; Smierciew, K.; Butrymowicz, D. Performance of ejection refrigeration system operating with R-1234ze(E) driven by ultra-low grade heat source. Int. J. Refrig. 2018, 8, 458–471. [Google Scholar] [CrossRef]
  56. Poirier, M.; Giguère, D.; Sapoundjiev, H. Experimental parametric investigation of vapor ejector for refrigeration applications. Energy 2018, 162, 1287–1300. [Google Scholar] [CrossRef]
  57. García Del Valle, J.; Sierra-Pallares, J.; Garcia Carrascal, P.; Castro Ruiz, F. An experimental and computational study of the flow pattern in a refrigerant ejector. Validation of turbulence models and real-gas effects. Appl. Therm. Eng. 2015, 89, 795–811. [Google Scholar] [CrossRef]
  58. García Del Valle, J.; Saíz Jabardo, J.M.; Castro Ruiz, F.; San José Alonso, J.F. An experimental investigation of a R-134a ejector refrigeration system. Int. J. Refrig. 2014, 46, 105–113. [Google Scholar] [CrossRef]
  59. Shestopalov, K.O.; Huang, B.J.; Petrenko, V.O.; Volovyk, O.S. Investigation of an experimental ejector refrigeration machine operating with refrigerant R245fa at design and off-design working conditions. Part 2. Theoretical and experimental results. Int. J. Refrig. 2015, 55, 212–223. [Google Scholar] [CrossRef]
  60. Thongtip, T.; Aphornratana, S. An experimental analysis of the impact of primary nozzle geometries on the ejector performance used in R141b ejector refrigerator. Appl. Therm. Eng. 2017, 110, 89–101. [Google Scholar] [CrossRef]
  61. Chong, D.; Hu, M.; Chen, W.; Wang, J.; Liu, J.; Yan, J. Experimental and numerical analysis of supersonic air ejector. Appl. Energy 2014, 130, 679–684. [Google Scholar] [CrossRef]
  62. Kitrattana, B.; Aphornratana, S.; Thongtip, T.; Ruangtrakoon, N. Comparison of traditional and CRMC ejector performance used in a steam ejector refrigeration. Energy Procedia 2017, 138, 476–481. [Google Scholar] [CrossRef]
  63. Chen, Y.-M.; Sun, C.-Y. Experimental study of the performance characteristics of a steam-ejector refrigeration system. Exp. Therm. Fluid Sci. 1997, 15, 384–394. [Google Scholar] [CrossRef]
  64. Reddick, C.; Sorin, M.; Sapoundjiev, H.; Aidoun, Z. Effect of a mixture of carbon dioxide and steam on ejector performance: An experimental parametric investigation. Exp. Therm. Fluid Sci. 2018, 92, 353–365. [Google Scholar] [CrossRef]
  65. Hamzaoui, M.; Nesreddine, H.; Aidoun, Z.; Balistrou, M. Experimental study of a low grade heat driven ejector cooling system using the working fluid R245fa. Int. J. Refrig. 2018, 86, 388–400. [Google Scholar] [CrossRef]
  66. Yan, J.; Chen, G.; Liu, C.; Tang, L.; Chen, Q. Experimental investigations on a R134a ejector applied in a refrigeration system. Appl. Therm. Eng. 2017, 110, 1061–1065. [Google Scholar] [CrossRef]
  67. Chaussin, C.; Hilly, G. Chaleur et Thermodynamique, 4th ed.; Dunod: Paris, France, 1962; pp. 135–136. [Google Scholar]
  68. Paliwoda, A. Design of supersonic ejectors operating as booster compressor in refrigerating systems. In Proceedings of the XIth International Congress of Refrigeration Science and Technology, Munich, Germany, 31 August 1963; 1965; pp. 589–598. [Google Scholar] [CrossRef]
  69. Friso, D. An approximate analytic solution to a non-Linear ODE for air jet velocity decay through tree crops using piecewise linear emulations and rectangle functions. Appl. Sci. 2019, 9, 5440. [Google Scholar] [CrossRef] [Green Version]
  70. Friso, D. Conveyor-belt dryers with tangential flow for food drying: Development of drying ODEs useful to design and process adjustment. Inventions 2021, 6, 6. [Google Scholar] [CrossRef]
  71. Friso, D. Mathematical modelling of conveyor-belt dryers with tangential flow for food drying up to final moisture content below the critical value. Inventions 2021, 6, 43. [Google Scholar] [CrossRef]
  72. Gu, Z.; Chen, M.; Wang, C.; Zhuang, W. Static and dynamic analysis of a 6300 KN cold orbital forging machine. Processes 2021, 9, 7. [Google Scholar] [CrossRef]
Figure 1. General schematic of an ejector: The motive gas enters on the left with mass flow rate Gm; the induced gas enters from below with mass flow rate Gi.
Figure 1. General schematic of an ejector: The motive gas enters on the left with mass flow rate Gm; the induced gas enters from below with mass flow rate Gi.
Processes 10 00088 g001
Figure 2. Ejector operating curve: The pressure at the outlet of the diffuser is p4. To operate correctly—that is, keeping the entrainment ratio constant and equal to the design value (on-design)—the pressure p4 must be equal to or lower than the critical pressure pc.
Figure 2. Ejector operating curve: The pressure at the outlet of the diffuser is p4. To operate correctly—that is, keeping the entrainment ratio constant and equal to the design value (on-design)—the pressure p4 must be equal to or lower than the critical pressure pc.
Processes 10 00088 g002
Figure 3. Bench tests of the ejectors 9 and 10: (a) Motive steam generator; (b) Ejectors during the tests: the induced gas was air, which passes through the calibrated orifices of the flute at sonic speed. The number and diameter of the orifices must be calibrated appropriately, in order to produce the suction mass flow rate Gi and the pressure pi at the inlet of the suction chamber.
Figure 3. Bench tests of the ejectors 9 and 10: (a) Motive steam generator; (b) Ejectors during the tests: the induced gas was air, which passes through the calibrated orifices of the flute at sonic speed. The number and diameter of the orifices must be calibrated appropriately, in order to produce the suction mass flow rate Gi and the pressure pi at the inlet of the suction chamber.
Processes 10 00088 g003
Figure 4. Diagram (derived from Equation (10)) of the entrainment ratio of an ejector (N. 1, Table 1) vs. pressure at the exit of the suction chamber p2 and the diffuser efficiency ηD: ω = f ( p 2 , η D ) . The red curve represents the function: ω = f ( p 2 ) . The yellow curve represents the function: η D = f ( p 2 ) . The p2 pressure value that maximizes the entrainment ratio ω is also the one that minimizes the diffuser efficiency ηD.
Figure 4. Diagram (derived from Equation (10)) of the entrainment ratio of an ejector (N. 1, Table 1) vs. pressure at the exit of the suction chamber p2 and the diffuser efficiency ηD: ω = f ( p 2 , η D ) . The red curve represents the function: ω = f ( p 2 ) . The yellow curve represents the function: η D = f ( p 2 ) . The p2 pressure value that maximizes the entrainment ratio ω is also the one that minimizes the diffuser efficiency ηD.
Processes 10 00088 g004
Figure 5. Function: ω = f ( p 2 ) parametric in ηD.
Figure 5. Function: ω = f ( p 2 ) parametric in ηD.
Processes 10 00088 g005
Figure 6. Function: η D = f ( p 2 ) parametric in ω.
Figure 6. Function: η D = f ( p 2 ) parametric in ω.
Processes 10 00088 g006
Figure 7. Experimental relationship between the diffuser efficiency ηD and the Mach number Mm−2 of the motive gas at the final pressure p2.
Figure 7. Experimental relationship between the diffuser efficiency ηD and the Mach number Mm−2 of the motive gas at the final pressure p2.
Processes 10 00088 g007
Figure 8. Experimental relationship between the diffuser efficiency ηD and the Mach number Mm−i of the motive gas at the final pressure pi.
Figure 8. Experimental relationship between the diffuser efficiency ηD and the Mach number Mm−i of the motive gas at the final pressure pi.
Processes 10 00088 g008
Figure 9. Industrial entrainment ratios ωind vs. experimental ones ωexp for the first series of benchmark ejectors (Table 5).
Figure 9. Industrial entrainment ratios ωind vs. experimental ones ωexp for the first series of benchmark ejectors (Table 5).
Processes 10 00088 g009
Figure 10. Industrial entrainment ratios ωind vs. experimental ones ωexp for the second series of benchmark ejectors (Table 6).
Figure 10. Industrial entrainment ratios ωind vs. experimental ones ωexp for the second series of benchmark ejectors (Table 6).
Processes 10 00088 g010
Table 1. Operating data of the high-performance industrial ejectors.
Table 1. Operating data of the high-performance industrial ejectors.
Ejector N.12345678910
Motive gasSteamSteamSteamSteamSteamSteamSteamSteamSteamSteam
Motive gas pressure pm (bar)6999101210855
Motive gas temperature tm (°C)159175175175180188180170152152
Induced gasSteamSteamSteamSteamSteamSteamSteamSteamAirAir
Induced gas pressure pi (bar)0.5200.3800.09580.03160.04240.5330.01310.1230.50.3
Ind. gas temperature ti (°C)82754525308311502020
Aspiration diameter Di (mm)1252003005003501252002505050
Discharge pressure p4 (bar)1.0130.80.2650.07320.0961.7200.0960.311.0131.013
Discharge diameter D4 (mm)1502003005003501502002505050
Critical pressure pc = 1.05·p4 (bar)1.0530.8420.2790.07710.1011.8110.1010.3261.0531.053
Motive mass flow rate Gm (kg h−1)11601270157511005501900235842128128
Induced mass flow rate Gi (kg h−1)8501130129513417006809575010449
Industrial entrainment ratio ωind0.7350.8900.8221.2201.2700.3580.4040.8900.8100.380
Table 2. Operating data of the first series of benchmark ejectors built and tested experimentally by the authors indicated in the first column.
Table 2. Operating data of the first series of benchmark ejectors built and tested experimentally by the authors indicated in the first column.
Ref.Geom.
[19]
Case
[19]
Run
[19]
Motive GasInduced Gaspm
(bar)
Tm
(°C)
pi
(bar)
Ti
(°C)
p4 = pc
(bar)
ωexp
[53]A1fSteamSteam3.6139.80.032250.0531.2
[54]M1iSteamSteam2.7031300.012100.050.4
[54]M2eSteamSteam1.9871200.012100.0380.513
[54]M3gSteamSteam2.7031300.00950.0480.31
[55]B1dR1234zeR1234ze11.7963.72.188.45.090.24
[55]B2cR1234zeR1234ze11.6064.72.72114.920.408
[56]C1h*R134aR134a26.331003.5157.20.329
[56]C1p*R134aR134a26.331003.5158.00.252
[56]C2d*R134aR134a26.331003.8217.67.50.393
[56]C2k*R134aR134a26.331003.8217.68.20.293
[56]C3d*R134aR134a26.331004.2520.77.70.489
[56]C3h*R134aR134a26.331004.2520.78.30.390
[56]C4e*R134aR134a26.331004.6523.57.80.554
[56]C4i*R134aR134a26.331004.6523.58.50.438
[57]F1 R134aR134a23.3384.24.15206.50.592
[57]F2 R134aR134a31.901003.50156.90.361
[57]F3 R134aR134a31.901004.15206.90.433
[58]I1c*R134aR134a28.8994.44.146207.320.459
[58]I1d*R134aR134a28.8994.44.146207.540.416
[58]J1dR134aR134a28.8994.44.146207.50.476
[58]K1eR134aR134a28.8994.44.146207.660.504
[59]G1cR245faR245fa11.298950.75881.9790.31
[59]G2cR245faR245fa11.298950.895122.0830.4
[59]G3b*R245faR245fa11.298951.052162.1200.55
[59]G3c*R245faR245fa11.298951.052162.2020.48
[59]H1b*R245faR245fa11.298950.75881.8450.41
[59]H1c*R245faR245fa11.298950.75881.9010.34
[59]H2b*R245faR245fa11.298950.895121.9110.54
[59]H2c*R245faR245fa11.298950.895122.0080.48
[59]H3b*R245faR245fa11.298951.052161.9790.68
[59]H3c*R245faR245fa11.298951.052162.1160.59
[60]D1f*R141bR141b6.7731000.435101.0500.247
[60]D1g*R141bR141b6.7731000.435101.0700.239
[60]D2f*R141bR141b6.7731000.33641.0370.123
[60]D2g*R141bR141b6.7731000.33641.0590.116
[60]D3fR141bR141b6.7731000.256−21.0270.063
[61]L1fAirAir10.0255.0256.00.452
Table 3. Operating data of the second series of benchmark ejectors built and tested experimentally by the authors indicated in the first column.
Table 3. Operating data of the second series of benchmark ejectors built and tested experimentally by the authors indicated in the first column.
Ref.Figure
of Ref.
CaseRunMotive GasInduced Gaspm
(bar)
Tm
(°C)
pi
(bar)
Ti
(°C)
p4 = pc
(bar)
ωexp
[62]4aCON13.4-D1.4SteamSteam2.7011300.010367.50.045750.367
[62]4aCON19-D2.0SteamSteam2.7011300.010367.50.045750.367
[62]4aCON13.4-D1.7SteamSteam2.7011300.010367.50.05800.188
[62]4aCON19-D2.4SteamSteam2.7011300.010367.50.06550.167
[62]4bCOND1.4SteamSteam2.7011300.010367.50.04560.276
[62]4bCRMCD1.4SteamSteam2.7011300.010367.50.04440.390
[62]4bCOND1.7SteamSteam2.7011300.010367.50.05820.190
[62]4bCRMCD1.7SteamSteam2.7011300.010367.50.05620.267
[63]5pm116SteamSteam1.1601040.01306110.02370.617
[63]5pm153SteamSteam1.5301120.01306110.02940.486
[63]5pm198SteamSteam1.9801200.01306110.03920.389
[63]5pm270SteamSteam2.7001300.01306110.04750.343
[63]6pi1.306SteamSteam1.9801200.01306110.03920.389
[63]6pi1.933SteamSteam1.9801200.01933170.04090.643
[63]6pi2.346SteamSteam1.9801200.02346200.04180.769
[64]9pm350SteamSteam3.501390.70901.2310.430
[64]9pm450SteamSteam4.501480.70901.5170.324
[64]9pm550SteamSteam5.50155.50.70901.8040.222
[64]10pi90SteamSteam4.501480.9096.71.5600.464
[64]10pi70SteamSteam4.501480.70901.5170.324
[64]10pi50SteamSteam4.501480.5081.31.4770.181
[65]3pi90R245faR245fa5.57840.90122.1710.132
[65]3pi95R245faR245fa5.57840.95142.1820.152
[65]3pi100R245faR245fa5.57841.00152.1930.172
[65]3pi105R245faR245fa5.57841.05162.2050.193
[65]3pi110R245faR245fa5.57841.10172.2160.213
[65]3pi115R245faR245fa5.57841.15182.2270.233
[65]3pi120R245faR245fa5.57841.20192.2400.251
[66]3pm2148R134aR134a21.480713.124.6800.626
[66]3pm2340R134aR134a23.400753.125.1200.528
[66]3pm2495R134aR134a24.950773.125.3300.481
Table 4. Results of mathematical modelling applied to high-performance industrial ejectors to calculate diffuser efficiency.
Table 4. Results of mathematical modelling applied to high-performance industrial ejectors to calculate diffuser efficiency.
Ejector N.12345678910
Industrial entrainment ratio ωind0.7350.8900.8221.2201.2700.3580.4040.8900.8100.380
Diffuser efficiency ηD0.7880.7790.7340.7000.7170.7440.6820.7560.8150.785
Pressure of induced gas pi (bar)
at the inlet of suction chamber
0.5200.3800.09580.03160.04240.5330.01310.1230.50.3
Pressure of induced gas p2 (bar)
at the exit of suction chamber
0.41360.28880.07250.02390.03150.45490.01090.09130.40770.2562
Mach number Mi−2 of induced
gas at the pressure p2
0.6120.6720.6760.6760.6990.5080.5530.7000.5260.461
Mach number Mm−2 of motive
gas at the final pressure p2
2.2972.6713.3113.8133.7372.5914.2203.1522.2122.445
Mach number Mm−i of motive
gas at the final pressure pi
2.1802.5393.1843.6873.6022.5144.1343.0152.1062.367
Table 5. Results of mathematical modelling applied to the first series of benchmark ejectors (Table 2), in order to determine the industrial entrainment ratio ωind vs. the experimental one ωexp.
Table 5. Results of mathematical modelling applied to the first series of benchmark ejectors (Table 2), in order to determine the industrial entrainment ratio ωind vs. the experimental one ωexp.
Test
n.
Ref.Geom.
[19]
Case
[19]
Run
[19]
Rkcp
(J kg−1K−1)
Mm−i Mach N.ηDp2
(bar)
ωindωexpδ
(%)
1[53]A1f461.81.1437573.2660.7330.02432.0111.200−40.3%
2[54]M1i461.81.1437576.9340.5100.01030.7080.400−43.5%
3[54]M2e461.81.1437576.7670.5200.01010.9400.513−45.4%
4[54]M3g461.81.1437577.0920.5000.00790.5510.310−43.8%
5[55]B1d72.931.1256561.7710.8241.8950.2890.240−17.0%
6[55]B2c72.931.1256561.6300.8332.2920.4490.408−18.2%
7[56]C1h*81.51.1655751.9520.8132.8550.5860.329−43.9%
8[56]C1p*81.51.1655751.9520.8132.9170.4470.252−43.6%
9[56]C2d*81.51.1655751.9030.8163.1180.6180.393−36.4%
10[56]C2k*81.51.1655751.9030.8163.1700.4880.293−39.9%
11[56]C3d*81.51.1655751.8420.8203.4640.7070.489−30.8%
12[56]C3h*81.51.1655751.8420.8203.4990.5740.390−32.1%
13[56]C4e*81.51.1655751.7900.8233.7860.8220.554−32.6%
14[56]C4i*81.51.1655751.7900.8233.8240.6410.438−31.6%
15[57]F1 81.51.1655751.7860.8233.3610.9900.592−40.2%
16[57]F2 81.51.1655752.0570.8072.7940.7330.361−50.8%
17[57]F3 81.51.1655751.9640.8123.3081.0160.433−57.4%
18[58]I1c*81.51.1655751.9080.8163.3370.8160.459−43.8%
19[58]I1d*81.51.1655751.9080.8163.3460.7550.416−44.9%
20[58]J1d81.51.1655751.9080.8163.3450.7650.476−37.8%
21[58]K1e81.51.1655751.9080.8163.3520.7250.504−30.5%
22[59]G1c621.1006822.3000.7920.5940.5710.310−45.7%
23[59]G2c621.1006822.2190.7970.6980.6510.400−38.6%
24[59]G3b*621.1006822.1390.8020.8170.8050.550−31.7%
25[59]G3c*621.1006822.1390.8020.8200.7410.480−35.2%
26[59]H1b*621.1006822.3000.7920.5880.6540.410−37.3%
27[59]H1c*621.1006822.3000.7920.5900.6170.340−44.9%
28[59]H2b*621.1006822.2190.7970.6910.7770.540−30.5%
29[59]H2c*621.1006822.2190.7970.6950.7020.480−31.6%
30[59]H3b*621.1006822.1390.8020.8130.9370.680−27.4%
31[59]H3c*621.1006822.1390.8020.8170.8080.590−27.0%
32[60]D1f*71.121.1156902.3250.7900.3380.6660.247−62.9%
33[60]D1g*71.121.1156902.3250.7900.3390.6430.239−62.8%
34[60]D2f*71.121.1156902.4490.7830.2670.4850.123−74.6%
35[60]D2g*71.121.1156902.4490.7830.2680.4670.116−75.1%
36[60]D3f71.121.1156902.5770.7750.2100.3540.063−82.2%
37[61]L1f2871.410051.0200.8704.4561.0230.452−55.8%
Table 6. Results of mathematical modelling applied to the second series of benchmark ejectors (Table 3), in order to determine the industrial entrainment ratio ωind vs. the experimental one ωexp.
Table 6. Results of mathematical modelling applied to the second series of benchmark ejectors (Table 3), in order to determine the industrial entrainment ratio ωind vs. the experimental one ωexp.
Test
n.
Ref.Figure
of Ref.
CaseRunRkcp
(J kg−1K−1)
Mm−i Mach N.ηDp2
(bar)
ωindωexpδ
(%)
38[62]4aCON13.4-D1.4461.81.1437573.6470.7100.008030.5880.367−37.6%
39[62]4aCON19-D2.0461.81.1437573.6470.7100.008030.5880.367−37.6%
40[62]4aCON13.4-D1.7461.81.1437573.6470.7100.008290.4490.188−58.1%
41[62]4aCON19-D2.4461.81.1437573.6470.7100.008450.3880.167−57.1%
42[62]4bCOND1.4461.81.1437573.6470.7100.008030.5910.276−53.3%
43[62]4bCRMCD1.4461.81.1437573.6470.7100.008000.6100.390−36.1%
44[62]4bCOND1.7461.81.1437573.6470.7100.008290.4470.190−57.5%
45[62]4bCRMCD1.7461.81.1437573.6470.7100.008250.4650.267−42.6%
46[63]5pm116461.81.1437573.1580.7400.009841.6270.617−62.1%
47[63]5pm153461.81.1437573.2850.7320.009721.2050.486−59.7%
48[63]5pm198461.81.1437573.4020.7250.009790.8510.389−54.5%
49[63]5pm270461.81.1437573.5420.7160.009930.7090.343−51.6%
50[63]6pi1.306461.81.1437573.4020.7250.009790.8480.389−54.0%
51[63]6pi1.933461.81.1437573.2230.7360.014431.3130.643−51.0%
52[63]6pi2.346461.81.1437573.1350.7410.017741.7040.769−54.9%
53[64]9pm350461.81.1437571.7220.8270.57610.6250.430−31.2%
54[64]9pm450461.81.1437571.8670.8180.58260.4350.324−25.5%
55[64]9pm550461.81.1437571.9770.8120.59290.3330.222−33.3%
56[64]10pi90461.81.1437571.7220.8270.73910.6550.464−29.2%
57[64]10pi70461.81.1437571.8670.8180.58260.4350.324−25.5%
58[64]10pi50461.81.1437572.0500.8070.43190.2650.181−31.8%
59[65]3pi90621.1006821.8500.8190.76980.3280.132−59.8%
60[65]3pi95621.1006821.8210.8210.80820.3550.152−57.2%
61[65]3pi100621.1006821.7920.8230.84890.3840.172−55.2%
62[65]3pi105621.1006821.7640.8250.88790.4130.193−53.4%
63[65]3pi110621.1006821.7370.8260.92700.4450.213−52.1%
64[65]3pi115621.1006821.7120.8280.96380.4780.233−51.2%
65[65]3pi120621.1006821.6870.8291.00350.5110.251−50.9%
66[66]3pm214881.51.1655751.9060.8162.48761.2550.626−50.1%
67[66]3pm234081.51.1655751.9540.8132.47231.0230.528−48.4%
68[66]3pm249581.51.1655751.9890.8112.46630.9530.481−49.5%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Friso, D. Mathematical Modelling of the Entrainment Ratio of High Performance Supersonic Industrial Ejectors. Processes 2022, 10, 88. https://doi.org/10.3390/pr10010088

AMA Style

Friso D. Mathematical Modelling of the Entrainment Ratio of High Performance Supersonic Industrial Ejectors. Processes. 2022; 10(1):88. https://doi.org/10.3390/pr10010088

Chicago/Turabian Style

Friso, Dario. 2022. "Mathematical Modelling of the Entrainment Ratio of High Performance Supersonic Industrial Ejectors" Processes 10, no. 1: 88. https://doi.org/10.3390/pr10010088

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