Next Article in Journal
Gaussian Amplitude Amplification for Quantum Pathfinding
Previous Article in Journal
Optimal Information Update for Energy Harvesting Sensor with Reliable Backup Energy
Previous Article in Special Issue
Numerical Study of Fin-and-Tube Heat Exchanger in Low-Pressure Environment: Air-Side Heat Transfer and Frictional Performance, Entropy Generation Analysis, and Model Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Energy Characteristics of a Bidirectional Axial-Flow Pump with Two Impeller Airfoils Based on Entropy Production Analysis

1
Research Center of Fluid Machinery Engineering and Technology, Jiangsu University, Zhenjiang 212013, China
2
Wenling Fluid Machinery Technology Institute of Jiangsu University, Wenling 317525, China
*
Author to whom correspondence should be addressed.
Entropy 2022, 24(7), 962; https://doi.org/10.3390/e24070962
Submission received: 7 June 2022 / Revised: 2 July 2022 / Accepted: 4 July 2022 / Published: 11 July 2022
(This article belongs to the Special Issue Entropy in Computational Fluid Dynamics III)

Abstract

:
This research sought to determine the spatial distribution of hydraulic losses for a bidirectional axial-flow pump with arc- and S-shaped impellers. The unsteady Reynolds time-averaged Stokes (URANS) approach with the SST k-omega model was used to predict the internal flow field. The total entropy production (TEP) and total entropy production rate (TEPR) were used to evaluate the overall and local hydraulic losses. The results show that the distribution of TEP and TEPR was similar for both impeller cases. Under a forward condition, TEP mainly comes from the impeller and elbow pipe. The high TEPR inside the impeller can be found near the shroud, and it shifts from the leading edge to the trailing edge with an increase in the flow rate due to the decline in the attack angle. The high TEPR inside the elbow pipe can be seen near the inlet, and the area shrinks with an increase in the flow rate caused by a reduction in the velocity circulation. Under the reverse condition, TEP mainly comes from the impeller and the straight pipe. The TEPR of the region near the shroud is obviously higher than for other regions, and the area of high TEPR near the suction side shrinks with an increase in the flow rate. The high TEPR of the straight pipe can be found near the inlet, and declines in the flow direction. These results provide a theoretical reference for future work to optimize the design of the bidirectional axial-flow pump.

1. Introduction

The bidirectional axial-flow pump can achieve two-way pumping by directly changing the rotation direction of the motor, which has the advantages of low civil engineering costs, stable operation, and easy maintenance. Therefore, it is becoming more widely used in agricultural engineering for both drainage and irrigation requirements [1,2]. When the conventional one-way axial-flow impeller rotates in the opposite direction, the camber of the airfoil is negative, which leads to strong flow separation near the blade surface [3,4]. The large-scale flow separation vortices near the blade surface collide with the main stream, which results in serious cavitation, vibration, and an obvious decline in hydraulic performance [5]. To balance the forward and reverse pumps’ performance, the airfoils of the impeller in the bidirectional axial-flow pump were designed to be arc- and S-shaped [6]. Due to the special airfoils, the hydraulic design theory of the traditional axial-flow impeller is not applicable. The current lack of mature theoretical guidance on this airfoil design hinders the further improvement of the hydraulic performance of the bidirectional axial-flow pump.
In benefiting from the development of computational fluid dynamics (CFD) technology [7], the hydraulic performance [8,9] and internal flow structure [10,11] of an axial-flow pump can be predicted accurately, which can provide a reference for hydraulic design. For instance, Shi [12,13] explored the effect of backflow clearance on the internal unstable flow structure and pressure pulsation characteristics of an axial-flow pump. The researcher asserted that leakage flow caused by backflow clearance reduced the axial velocity at the impeller inlet, resulting in a reduction in the hydraulic performance and high-efficiency operating range. Meanwhile, Feng [14] investigated the influence of the tip clearance radius on the internal pressure pulsation characteristics of an axial-flow pump. The results showed that the intensity of the tip clearance leakage vortex increases with an increase in the tip clearance radius, enhancing the pressure fluctuation intensity on the blade. Elsewhere, Fei [15] studied the effect of the blade angle on the hydraulic performance and tip leakage vortex (TLV). The results showed that the pump head, best efficiency point, and TLV intensity increase with an increase in the blade angle. Beyond this, Yang [16] analyzed the influence of the deflection angle on the internal flow characteristics in the inlet passage of an axial-flow pump. He found that the inflow pattern gradually worsened with an increase in the deflection angle, and the dominant frequency amplitude of the monitoring point under the deflection inflow was significantly higher than that under the vertical inflow. However, analysis results for the velocity and pressure fields in the axial-flow pump alone cannot directly determine the hydraulic losses. As the impeller airfoil is affected by numerous control parameters, it is necessary to determine the spatial distribution of hydraulic losses in a bidirectional axial-flow pump to optimize the control parameters efficiently.
Thanks to the continuous improvement of entropy production theory by Herwig and Knock [17,18,19], the hydraulic losses in rotating machinery [20,21] can be quantitatively evaluated based on the entropy production, which can be calculated in the CFD post-processing. Many scholars have used entropy production theory to visually analyze the distribution of hydraulic losses in pumps. Guan [22], for instance, studied the effects of flow rates on the distribution of entropy production in a double-suction centrifugal pump. The results showed that the distribution of entropy production was mainly affected by the main flow characteristic. The vortex size decreased with an increase in the flow rate, which led to a decline in the entropy production rate. Meanwhile, Zhang [23] compared the distributions of entropy production in a side-channel pump, and found that the entropy production in the inner radius of the impeller passage was higher than for the outer radius. Elsewhere, Ji [24] and Shen [25] evaluated the hydraulic losses in a mixed-flow pump and an axial-flow pump, respectively, based on entropy production. They proposed that the intensity of the tip leakage flow rose with an increase in the tip clearance radius under the design condition, which led to an increase in the entropy production in the impeller. When compared with the traditional axial-flow pump, the hydraulic design of the bidirectional axial-flow pump differs in that it needs to balance the forward and reverse hydraulic performance. Under the reverse condition, the inlet anti-arch guide vanes distort the impeller inflow and produce great hydraulic losses. Accordingly, the hydraulic design of the bidirectional axial-flow pump is complex, and the high hydraulic losses need to be located to provide a theoretical reference. No existing public literature could be found that analyzed the distribution of the entropy generation in the bidirectional axial-flow pump.
In this study, the unsteady Reynolds time-averaged Stokes (URANS) approach was used to predict the internal flow structures of arc- and S-shaped bidirectional axial-flow pumps, and the calculated results were validated by external characteristic test data. Based on entropy production theory, the total entropy production in each hydraulic component was determined under different flow rates. The spatial distributions of entropy production rates in the impellers and guide vanes were analyzed, and combined with the velocity field. The results provide useful suggestions for how to optimize the design of arc- and S-shaped bidirectional axial-flow pumps.

2. Numerical Simulation

2.1. Three Models and Meshes

In this study, a bidirectional axial-flow pump with an arc-shaped impeller was used, as shown in Figure 1. Except for the impeller, the two kinds of axial-flow pumps were the same in that they contained the same straight pipe, guide vanes, and elbow pipe. Figure 2 shows a comparison of the arc- and S-shaped impellers both based on five airfoils. The blade number, diameter, and tip clearance for the two impellers were 3, 300, and 0.2 mm, respectively. The blade number and hub diameter of the guide vanes were 5 and 120 mm, respectively. Further main design parameters for the two bidirectional axial-flow pumps are listed in Table 1.
The computational domains of the two bidirectional axial-flow pumps were discretized by a hexahedral mesh. As shown in Figure 3, ICEM CFD was used to generate meshes for the straight and elbow pipes. TurboGrid was applied to create meshes for the arc-shaped impeller, S-shaped impeller and guide vanes. The average Y+ values of the straight pipe, arc-shaped impeller, S-shaped impeller, guide vanes, and elbow pipe were 27.9, 23.3, 24.6, 11.2, and 33.3, respectively, which met the requirements of the k ω SST turbulence model [26]. The grid numbers of the two bidirectional axial-flow pumps were determined by grid independence analysis, as shown in Figure 4. Since the total grid nodes exceeded 5.41 million, the forward and reverse design heads’ relative growth rates for the arc-shaped impeller case were less than 0.02% and 0.17%, respectively. Accordingly, the number of grid nodes of the straight pipe, arc-shaped impeller, guide vanes, and elbow pipe were determined to be 909,706, 2,079,720, 1,478,235, and 950,172, respectively. To ensure that the number of nodes would not affect the comparison of the two impeller cases, the number of grid nodes for the S-shaped impeller was calculated as 2,279,160.

2.2. Boundary Condition

In this study, the governing equation of the computational domain was an RANS equation, as follows [27]:
v ¯ j x j = 0
( ρ v ¯ i ) t + ( ρ v ¯ i v ¯ j ) x j = p ¯ x i + x j ( μ v ¯ i x j ρ v i v j ¯ ) + ρ f i
where ρ , p ¯ , and μ are the water density, time-averaged pressure, and dynamic viscosity, respectively, and v ¯ is the time-averaged velocity. The subscripts i and j represent the x, y, and z directions in the Cartesian coordinate system.
The steady calculation results were used as the initial values of the unsteady calculations, which were completed to predict the internal flow fields of the two bidirectional axial-flow pumps. The inlet and outlet conditions were set as “Mass Flow Rate” [28] and “Opening Pres. And Dirn.” The inlet flow rate was adjusted according to the operating conditions, and the outlet relative pressure was set to 0 Pa. The wall surface roughness scores of the inlet section, impeller, guide vanes, and outlet section were 0.05, 0.0125, 0.0125, and 0.05 mm, respectively. The interface between the stators was set as “None.” The interfaces between the rotor and stator for steady and unsteady calculations were set as “Stage” [29] and “Transient Rotor Stator” [30], respectively. In addition, the time step was 0.00037037 s, i.e., a rotation of 3° per time step [31]. The total computation time was 0.533333 s.

3. Entropy Production Theory

Entropy production is a state quantity that can parameterize the transformation of energy within the computational domain. Based on the second law of thermodynamics, the time-averaged transport equation for entropy production can be obtained as follows [18,19]:
ρ ( S t + v 1 S x + v 2 S y + v 3 S z ) = d i v ( q T ) + Φ T + Φ θ T 2
where S is the specific entropy, T and q represent the temperature and the heat flux density vector, respectively, Φ T and Φ θ T 2 represent the entropy production by dissipation and heat transfer, respectively, and v 1 , v 2 , and v 3 represent the x, y, and z directions of the Cartesian coordinate system, respectively.
During the operation of a bidirectional axial-flow pump, part of the mechanical energy of the motor is irreversibly converted into internal energy due to the existence of working medium viscosity and Reynolds stress, resulting in irreversible hydraulic losses. The working medium is pure water, and there are no chemical reactions leading to significant temperature fluctuations in the bidirectional axial-flow pump. The internal flow field is thus considered to be at a constant temperature and incompressible, and the local hydraulic losses can be evaluated quantitatively by Φ T .
Since the governing equation of the computational domain is an RANS equation, the transport equation for the entropy production shroud can be time-averaged, so Φ T can be calculated in the CFD post-processing. After time-averaging, ( Φ T ) ¯ can be divided into the entropy production rate by direct dissipation Φ D ¯ T ¯ and the entropy production rate by indirect dissipation Φ D T ¯ , as follows [17,18]:
( Φ T ) ¯ = Φ D ¯ T ¯ + Φ D T ¯
Φ D ¯ T ¯ = μ T ¯ { 2 × [ ( v 1 ¯ x ) 2 + ( v 2 ¯ y ) 2 + ( v 3 ¯ z ) 2 ] + ( v 1 ¯ y + v 2 ¯ x ) 2 + ( v 3 ¯ x + v 1 ¯ z ) 2 + ( v 2 ¯ z + v 3 ¯ y ) 2 }
Φ D T ¯ = μ T ¯ { 2 × [ ( v 1 x ) 2 ¯ + ( v 2 y ) 2 ¯ + ( v 3 z ) 2 ¯ ] + ( v 1 y + v 2 x ) 2 ¯ + ( v 3 x + v 1 z ) 2 ¯ + ( v 2 z + v 3 y ) 2 ¯ }
where the superscripts ¯ and represent the time-averaged component and the fluctuation component, respectively. Since the velocity fluctuation component cannot be obtained by solving the URANS equation, Knock [18,19] proposed a model equation to obtain the approximate solution of Φ D T ¯ , as follows:
Φ D T ¯ = ρ ε T ¯
where ε is the dissipation rate of turbulent kinetic energy. In summary, the overall hydraulic losses of each hydraulic component of a bidirectional axial-flow pump can be characterized by the overall entropy production, which can be obtained by integrating Φ D ¯ T ¯ and Φ D T ¯ .
S P R O , D ¯ = V Φ D ¯ T ¯ d V
S P R O , D = V Φ D T ¯ d V
S P R O , D = S P R O , D ¯ + S P R O , D

4. Results and Discussion

4.1. Test Validation

To verify the accuracy of the numerical simulation, the external characteristics testing of an arc-shaped bidirectional axial-flow pump was completed. The test bench was a double-layered vertical structure, and the upper and lower heights were 4.6 m and −2.6 m, respectively, as shown in Figure 5. Low- and high-pressure tanks were located up- and downstream of the bidirectional axial-flow pump, respectively, to stabilize the water pressure. The electromagnetic flowmeter was located on the bottom layer to prevent interference from the bidirectional axial-flow pump. The specific parameters of the test instrument are shown in Table 2, and the measurement uncertainty of the system was calculated as less than 0.3% by E S = E Q 2 + E H 2 + E T 2 + E n 2 .
Figure 6 compares the hydraulic performance of the arc-shaped bidirectional pump based on the CFD results and test data. The head coefficient C h and flow coefficient C q can be calculated based on Equations (11)–(13). The calculated efficiency is slightly lower than the test efficiency under part-loaded conditions, but the calculated efficiency is slightly higher than the test efficiency under overloaded conditions. Under forward design flow rate, the relative error between the calculations and experiments of the efficiency and the head coefficient is 2.0% and −6.2%, respectively; under the reverse design point, those values are −4.3% and −8.5%, respectively.
u 2 = π d 2 n 60
C q = Q π d 2 ( r s r h ) u 2
C h = 2 g H u 2 2
where u 2 and n are the circumferential velocity and rotation speed, respectively, d 2 is the diameter of the impeller outlet, and r s and r h are the radii of the hub and the shroud, respectively.

4.2. Energy Characteristics of the Two Bidirectional Axial-Flow Pumps

Figure 7 compares the hydraulic performance of the arc- and S-shaped cases. Under the forward condition, the efficiency and head of the arc-shaped case are higher. The highest efficiencies of the arc- and S-shaped cases are 79.9% and 77.8% respectively, and the best efficiency point for both cases can be found at Cq = 0.21. The efficiency of the two cases is close—within Cq = 0.15–0.21—but the efficiency of the S-shaped case decreases sharply from Cq = 0.21 to Cq = 0.24, which indicates that the low-camber trailing edge of the S-shaped impeller leads to strong flow separation under a large flow coefficient. Under the reverse condition, the efficiency and head of the arc-shaped case are lower. The highest efficiencies of the arc- and S-shaped cases are 62.2% and 65.5%, respectively. The best efficiency points for the arc- and S-shaped cases can be seen at Cq = 0.19 and Cq = 0.17, respectively. The results indicate that the matching degree between the arc-shaped impeller and the inlet guide vanes is poor. The high-camber leading edge of the arc-shaped impeller decreases the attack angle of the inflow and, thus, increases the best efficiency point when compared with the S-shaped impeller.
In this study, TEP is taken as the quantitative evaluation index of hydraulic loss, which is the key factor to determine the efficiency and head of an axial-flow pump. Figure 8 shows the TEP distribution with different flow coefficients for arc- and S-shaped cases. Under the forward condition, TEP first decreases and then increases with an increase in Cq, and the minimum can be found under design condition Cq = 0.21, due to good matching between the impeller and the guide vanes. Except for Cq = 0.19, the TEP of S-shaped cases is higher than that of arc-shaped cases under all flow coefficients. The relative deviation between arc- and S-shaped cases is −1.12% under Cq = 0.21. The maximum relative deviation is −5.09%, which can be found under Cq = 0.24. Under the reverse condition, TEP decreases with a decline in Cq, which shows that an excessive attack angle of the impeller inflow due to inverted arch guide vanes is the main cause of hydraulic loss. The TEP of arc-shaped cases is higher than that of S-shaped cases under all flow coefficients, and the relative deviation between arc- and S-shaped cases is 13.95% under Cq = 0.17 (design condition). The maximum relative deviation is 15.04%, which can be seen under Cq = 0.14. Since the head and efficiency of axial-flow pumps decrease with an increase in hydraulic loss, the TEP curve can effectively express the influence of the impeller airfoil on the internal hydraulic loss, as shown in Figure 7 and Figure 8.
Figure 9 shows the TEP proportion of each hydraulic component for arc- and S-shaped cases under three forward flow coefficients. The straight pipe is located upstream of the impeller, so there is no significant difference in TEP proportions between the two impeller cases. The TEP proportion of the impeller is obviously higher than that of the rest of the hydraulic components, showing that unstable flow inside the impeller is the main source of TEP. Under Cq = 0.15 and 0.21, the TEP proportion of the arc-shaped impeller is lower than that for the S-shaped impeller, showing that the unstable flow losses inside the S-shaped impeller lead to lower hydraulic performance. Under Cq = 0.24, the TEP proportion of each hydraulic component in the arc-shaped case is similar to that for the S-shaped case. The internal flow losses for the two impeller cases are approximately the same, but the lower blade outlet angle of the S-shaped impeller reduces the airfoil lift coefficient, and results in lower efficiency of the S-shaped impeller case. Figure 10 shows the TEP proportion of each hydraulic component for the arc- and S-shaped cases under three reverse flow coefficients. The TEP proportion of the straight pipe and impeller is obviously higher than that of the rest of the hydraulic components due to the distorted impeller inflow caused by an inverted arch guide vane. The TEP proportion of the arc-shaped impeller is higher than that of the S-shaped impeller, indicating that a higher blade inlet angle of the arc-shaped impeller leads to a strong impeller–guide vanes interference effect and, thus, hampers the pump efficiency.

4.3. Distribution of Local Entropy Production Rates under the Forward Condition

The radial coefficient R* of the turbo surface was defined as r i r h r s r h , where r i is the radius of the calculated turbo surface. To quantitatively analyze the TEPR distribution in different sections of the impeller passage, the turbo surface i (i = 1, 2, 3…11) was used to divide the impeller passage evenly into 10 approximate hollow cylinders, as shown in Figure 11. The radial coefficient R* of the turbo surface i could be calculated by 0.1 × i . The inner and outer walls of the hollow cylinder i are turbo surface i and turbo surface i + 1.
Figure 12 shows the volume-averaged TEPR of 10 hollow cylinders in arc- and S-shaped impellers under the forward condition. The average TEPR of each volume is the highest and lowest under Cq = 0.15 and 0.21, respectively. Due to the wall effect flow, the average TEPR of the volumes first increased and then declined from hub to shroud. Compared with the arc-shaped impeller, the average TEPR for the S-shaped impeller was higher from volumes 5 to 9, which led to lower efficiency. In addition, the average TEPR of volumes 9 and 10 was obviously higher than those of the rest of the volumes, due to the tip clearance leakage flow. Accordingly, the spatial distribution of TEPR in the turbo surface near the impeller shroud (R* = 0.95) was obtained, as shown in Figure 13. Under Cq = 0.15, there was a large area of high TEPR near the leading edge due to strong flow separation caused by a large attack angle, and the TEPR gradually decreased from inlet to outlet. Under Cq = 0.21 and 0.24, the TEPR distributions of the two cases were similar, and there was no obvious high TEPR near the leading edge, since the attack angle decreased with an increase in the flow rate. However, mixing of the wake vortex and the main stream led to a small area of high TEPR near the trailing edge. In addition, the area of high TEPR near suction and the trailing edge of S-shaped blade case was slightly larger than that of the arc-shaped blade case, due to the anti-arch trailing edge of the S-shaped suction side.
To quantitatively analyze the TEPR of different sections under the forward condition, the passage of the guide vanes was also divided into 10 volumes. Figure 14 compares the volume-averaged TEPR of 10 hollow cylinders in guide vanes between the arc- and S-shaped impeller cases. The average TEPR decreased from volumes 1 to 9 due to the wall effect. Under Cq = 0.15, the average TEPR was extremely high, and the volume-averaged TEPR of the S-shaped impeller case was lower than that of the arc-shaped impeller case. In addition, the volume-averaged TEPR of the arc-shaped impeller case decreased with an increase in the flow rate, but the minimum for the S-shaped impeller case could be found under the design flow coefficient Cq = 0.21.
Since the average TEPR of volume 1 was significantly higher than those of the rest of the volumes, this indicated that unstable flow structures were found near the hub. Thus, we determined the distribution of TEPR in the turbo surface with R* = 0.05, as shown in Figure 15. Due to interference between the impeller and the guide vanes, the TEPR near the inlet was high under all flow coefficients. However, the position of high TEPR near the surface of the guide vanes differed under the three flow coefficients. Under Cq = 0.15, the secondary vortex boundary near the leading edge was mixed with the main stream, resulting in high TEPR away from the suction side. Meanwhile, under Cq = 0.21 and 0.24, the internal fluid flowed along the surface of the guide vanes, meaning that the high TEPR could be found close to the surface of the guide vanes due to the wall effect.
Figure 16 shows the distribution of TEPR in the elbow pipe under three forward flow coefficients. The TEPR distribution in the elbow pipe of S- and arc-shaped impeller cases was similar. There was high TEPR near the inlet, since the circulation of the internal flow gradually decreased in the flow direction, and the area of high TEPR decreased with an increase in the flow coefficients. Under Cq = 0.15, the backflow was upstream of the bend corner due to the high circumferential velocity, leading to high TEPR. Under Cq = 0.24, there was flow separation downstream of the bend corner due to the high axial velocity, which also resulted in high TEPR.

4.4. Distribution of Local Entropy Production Rate under the Reverse Condition

Figure 17 shows the volume-averaged TEPR of 10 hollow cylinders in arc- and S-shaped impellers under the reverse condition, which is obviously higher than that under the forward condition, owing to the distorted impeller inflow. The average TEPR first decreases and then increases with an increase in the volume number due to the wall effect. Except for volume 1, the average TEPR of the rest of the volumes declines with an increase in the flow rate resulting from an improvement to the inflow pattern of the impeller. The average TEPR of each volume in the arc-shaped impeller is higher than that in the S-shaped impeller, which indicates that the large inlet angle of the arc-shaped blade leads to large, unstable flow structures. In addition, the average TEPR in volume 10 is obviously higher than those in the rest of the volumes, caused by tip clearance leakage flow. Given that, we determined the TEPR distribution in the turbo surface with R* = 0.95 of the arc- and S-shaped impellers under the reverse condition, as shown in Figure 18. Since the circumferential velocity decreases with an increase in the flow rate, the position of high TEPR caused by flow separation differs under the three flow coefficients. Under Cq= 0.15, the high TEPR can be found near the impeller inlet and the leading edge, due to backflow caused by the large circumferential velocity. Under Cq= 0.21, the high-TEPR area in both of the impellers shrinks significantly due to a reduction in the circumferential velocity. However, strong flow separation still exists at the leading edge of the arc-shaped impeller, resulting in a large area of high TEPR. Under Cq= 0.24, the flow pattern in both impellers is stable, and only a small area of high TEPR can be found near the trailing edge, caused by mixing of the wake vortices with the mainstream. Due to the larger inlet angle of the arc-shaped blade, the area of high TEPR near the suction and trailing edge is higher, compared with S-shaped blade.
Figure 19 shows the TEPR distribution in the straight pipe under the reverse condition. Since the internal TEPR of the straight pipe is mainly affected by the velocity circulation of the inflow, the internal TEPR decreases with an increase in the flow rate. High TEPR is located near the inlet, and decreases in the flow direction. According to Figure 7 and Figure 18, the velocity circulation of the inflow and the intensity of the wake vortex of the S-shaped impeller are higher and stronger than those of the arc-shaped impeller under the same flow coefficient. Thus, the area of high TEPR is larger in the S-shaped impeller case.

5. Conclusions

In this study, the URANS approach was used to predict the hydraulic performance and inner flow structure of bidirectional axial-flow pumps with arc- and S-shaped impellers. Entropy production theory was applied to evaluate the spatial distribution of internal hydraulic losses. The distribution of TEP and TEPR for the two impeller cases was similar, and the following conclusions can be drawn:
(1)
Under the forward condition, the optimal operation point is not affected by the impeller airfoil, and the hydraulic performance of the arc-shaped case is higher than that of the S-shaped case. Under the reverse condition, the hydraulic performance is higher, and the optimal operation point shifts to a small flow rate of the S-shaped case, compared with the arc-shaped case.
(2)
Under the forward condition, the TEP of the impeller and the elbow pipe is dominant. The anti-arch form of the trailing edge of the S-shaped suction side creates a barrier effect, which results in higher average TEPR near the region of the hub and the rim compared with the arc-shaped blade. In addition, the TEPR distribution in the elbow pipe is similar for the arc- and S-shaped cases, and the high TEPR can be found near inlet due to a decrease in the rotational kinetic energy along the flow direction.
(3)
Under the reverse condition, the TEP of the impeller and the straight pipe is dominant. Due to the large inlet angle of the arc-shaped blade, the flow separation and TEPR near the suction side are stronger and higher, compared with the S-shaped case. Due to the stronger wake vortex, the area of high TEPR in the straight pipe of the arc-shaped case is higher than that of the S-shaped case.
These results provide useful suggestions for how to optimize the design of bidirectional axial-flow pumps.

Author Contributions

Conceptualization, F.M.; data curation, F.M.; methodology, F.M.; software, F.M.; formal analysis, F.M.; resources, F.M. and Y.L.; visualization; F.M.; methodology, F.M.; project administration, Y.L.; supervision, Y.L.; validation, F.M.; writing—original draft, F.M.; writing—review and editing, F.M. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Latin Letters
p ¯ Time-averaged pressure S Specific entropy
v ¯ Time-averaged velocity T Temperature
q Heat flux density vector C h Head coefficient
C q Flow coefficient u 2 Circumferential velocity
n Rotation speedd2Diameter of impeller outlet
r s Hub radius r h Shroud radius
EQMeasurement uncertainty of flow rateEHMeasurement uncertainty of head
ETMeasurement uncertainty of torqueEnMeasurement uncertainty of rotation speed
R*Radial coefficient S P R O , D ¯ Entropy production by direct dissipation
S P R O , D Entropy production by indirect dissipation S P R O , D Total entropy production
Greek Letters
ρ Water density μ Dynamic viscosity
Φ T Entropy production by dissipation Φ θ T 2 Entropy production by heat transfer
Φ D ¯ T ¯ Entropy production rate by direct dissipation Φ D T ¯ Entropy production rate by indirect dissipation
Abbreviations
URANSUnsteady Reynolds time-averaged Navier–StokesCFDComputational fluid dynamics
EXPExperimentTLVTip leakage vortex
TEPTotal entropy productionTEPRTotal entropy production rate

References

  1. Xia, Y.; Tang, F.; Shi, L.; Xie, C.; Zhang, W. Numerical simulation and experimental analysis of bidirectional shaft tubular pump device. China Rural Water Hydropower 2017, 7, 149–153. (In Chinese) [Google Scholar]
  2. Shi, L.; Liu, X.; Tang, F.; Yao, Y.; Xie, R.; Zhang, W. Numerical simulation and experimental analysis of bidirectional shaft tubular pump device. Trans. Chin. Soc. Agric. Mach. 2016, 47, 85–91. (In Chinese) [Google Scholar]
  3. Ma, P.; Wang, J.; Wang, H. Investigation of performances and flow characteristics of two bi-directional pumps with different airfoil blades. Sci. China Technol. Sci. 2018, 61, 1588–1599. [Google Scholar] [CrossRef]
  4. Ma, P.; Wang, J. An analysis on the flow characteristics of bi-directional axial-flow pump under reverse operation. Proc. Inst. Mech. Eng. Part A J. Power Energy 2017, 231, 239–249. [Google Scholar] [CrossRef]
  5. Liu, H.; Chen, X.; Wang, K.; Tan, M.; Zhou, X. Multi-condition optimization and experimental study of impeller blades in a mixed-flow pump. Adv. Mech. Eng. 2016, 8, 1687814016651817. [Google Scholar] [CrossRef] [Green Version]
  6. Heo, M.W.; Kim, K.Y.; Kim, J.H.; Choi, Y.S. High-efficiency design of a mixed-flow pump using a surrogate model. J. Mech. Sci. Technol. 2016, 30, 541–547. [Google Scholar] [CrossRef]
  7. Wang, W.; Tai, G.; Pei, J.; Pavesi, G.; Yuan, S. Numerical investigation of the effect of the closure law of wicket gates on the transient characteristics of pump-turbine in pump mode. Renew. Energy 2022, 194, 719–733. [Google Scholar] [CrossRef]
  8. Shi, L.; Zhu, J.; Tang, F.; Wang, C. Multi-disciplinary optimization design of axial-flow pump impellers based on the approximation model. Energies 2020, 13, 779. [Google Scholar] [CrossRef] [Green Version]
  9. Shi, L.; Fu, L.; Xia, Y.; Tang, F.; Sun, D.; Zhai, L. Influence of Maximum Airfoil Camber Position on Hydraulic Performance of Axial-flow Pump. Trans. Chin. Soc. Agric. Mach. 2018, 49, 148–154. (In Chinese) [Google Scholar]
  10. Kim, S.; Kim, Y.I.; Kim, J.H.; Choi, Y.S. Three-objective optimization of a mixed-flow pump impeller for improved suction performance and efficiency. Adv. Mech. Eng. 2019, 11, 1687814019898969. [Google Scholar] [CrossRef]
  11. Suh, J.W.; Yang, H.M.; Kim, Y.I.; Lee, K.Y.; Kim, J.H.; Joo, W.G.; Choi, Y.S. Multi-objective optimization of a high efficiency and suction performance for mixed-flow pump impeller. Eng. Appl. Comput. Fluid Mech. 2019, 13, 744–762. [Google Scholar] [CrossRef] [Green Version]
  12. Shi, L.; Yuan, Y.; Jiao, H.; Tang, F.; Cheng, L.; Yang, F.; Jin, Y.; Zhu, J. Numerical investigation and experiment on pressure pulsation characteristics in a full tubular pump. Renew. Energy 2021, 163, 987–1000. [Google Scholar] [CrossRef]
  13. Shi, L.; Zhang, W.; Jiao, H.; Tang, F.; Wang, L.; Sun, D.; Shi, W. Numerical simulation and experimental study on the comparison of the hydraulic characteristics of an axial-flow pump and a full tubular pump. Renew. Energy 2020, 153, 1455–1464. [Google Scholar] [CrossRef]
  14. Feng, J.; Luo, X.; Guo, P.; Wu, G. Influence of tip clearance on pressure fluctuations in an axial flow pump. J. Mech. Sci. Technol. 2016, 30, 1603–1610. [Google Scholar] [CrossRef]
  15. Fei, Z.; Xu, H.; Zhang, R.; Zheng, Y.; Mu, T.; Chen, Y. Numerical simulation on hydraulic performance and tip leakage vortex of a slanted axial-flow pump with different blade angles. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2021, 236, 09544062211032989. [Google Scholar] [CrossRef]
  16. Yang, F.; Li, Z.; Yuan, Y.; Liu, C.; Zhang, Y.; Jin, Y. Numerical and Experimental Investigation of Internal Flow Characteristics and Pressure Fluctuation in Inlet Passage of Axial Flow Pump under Deflection Flow Conditions. Energies 2021, 14, 5245. [Google Scholar] [CrossRef]
  17. Herwig, H.; Kock, F. Direct and indirect methods of calculating entropy generation rates in turbulent convective heat transfer problems. Heat Mass Transf. 2007, 43, 207–215. [Google Scholar] [CrossRef]
  18. Kock, F.; Herwig, H. Local entropy production in turbulent shear flows: A high-Reynolds number model with wall functions. Int. J. Heat Mass Transf. 2004, 47, 2205–2215. [Google Scholar] [CrossRef]
  19. Kock, F.; Herwig, H. Entropy production calculation for turbulent shear flows and their implementation in CFD codes. Int. J. Heat Fluid Flow 2005, 26, 672–680. [Google Scholar] [CrossRef]
  20. Tao, R.; Zhao, X.; Wang, Z. Evaluating the transient energy dissipation in a centrifugal impeller under rotor-stator interaction. Entropy 2019, 21, 271. [Google Scholar] [CrossRef] [Green Version]
  21. Deng, Q.; Pei, J.; Wang, W.; Lin, B.; Zhang, C.; Zhao, J. Energy loss and radial force variation caused by impeller trimming in a double-suction centrifugal pump. Entropy 2021, 23, 1228. [Google Scholar] [CrossRef] [PubMed]
  22. Guan, H.; Jiang, W.; Yang, J.; Wang, Y.; Zhao, X.; Wang, J. Energy loss analysis of the double-suction centrifugal pump under different flow rates based on entropy production theory. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2020, 234, 4009–4023. [Google Scholar] [CrossRef]
  23. Zhang, F.; Appiah, D.; Hong, F.; Zhang, J.; Yuan, S.; Adu-Poku, K.A.; Wei, X. Energy loss evaluation in a side channel pump under different wrapping angles using entropy production method. Int. Commun. Heat Mass Transf. 2020, 113, 104526. [Google Scholar] [CrossRef]
  24. Ji, L.; Li, W.; Shi, W.; Chang, H.; Yang, Z. Energy characteristics of mixed-flow pump under different tip clearances based on entropy production analysis. Energy 2020, 199, 117447. [Google Scholar] [CrossRef]
  25. Shen, S.; Qian, Z.; Ji, B. Numerical analysis of mechanical energy dissipation for an axial-flow pump based on entropy generation theory. Energies 2019, 12, 4162. [Google Scholar] [CrossRef] [Green Version]
  26. Jia, X.; Yu, J.; Li, B.; Zhang, L.; Zhu, Z. Effect of Incident Angle of Wear-ring Clearance on Pressure Pulsation and Vibration Performance of Centrifugal Pump. Front. Energy Res. 2022, 250. [Google Scholar] [CrossRef]
  27. Huang, X.; Li, Y.; Liu, Z.; Yang, W. Numerical investigation of flow features in the vaneless region of a centrifugal pump by large eddy simulation. Eng. Comput. 2018, 35, 395–410. [Google Scholar] [CrossRef]
  28. Lin, P.; Yang, T.; Xu, W.; Zhu, Z. Influence of Different Offset Angles of Inlet Guide Vanes on Flow Characteristics of Centrifugal Pump. Front. Energy Res. 2022, 10. [Google Scholar] [CrossRef]
  29. Yang, F.; Li, Z.; Hu, W.; Liu, C.; Jiang, D.; Liu, D.; Nasr, A. Analysis of flow loss characteristics of slanted axial-flow pump device based on entropy production theory. Royal Soc. Open Sci. 2022, 9, 211208. [Google Scholar] [CrossRef]
  30. Yang, J.; Liu, J.; Liu, X.; Xie, T. Numerical study of pressure pulsation of centrifugal pumps with the compressible mode. J. Therm. Sci. 2019, 28, 106–114. [Google Scholar] [CrossRef]
  31. Yang, F.; Chang, P.; Hu, W.; Mao, B.; Liu, C.; Li, Z. Numerical Study on Pressure Pulsation in a Slanted Axial-Flow Pump Device under Partial Loads. Processes 2021, 9, 1404. [Google Scholar] [CrossRef]
Figure 1. Three models of the bidirectional axial-flow pump.
Figure 1. Three models of the bidirectional axial-flow pump.
Entropy 24 00962 g001
Figure 2. Comparison of arc- and S-shaped impeller airfoils.
Figure 2. Comparison of arc- and S-shaped impeller airfoils.
Entropy 24 00962 g002
Figure 3. Mesh of each hydraulic component of the bidirectional axial-flow pump.
Figure 3. Mesh of each hydraulic component of the bidirectional axial-flow pump.
Entropy 24 00962 g003
Figure 4. Grid independence analysis of the arc-shaped impeller case under (a) forward and (b) reverse conditions.
Figure 4. Grid independence analysis of the arc-shaped impeller case under (a) forward and (b) reverse conditions.
Entropy 24 00962 g004
Figure 5. Structure of the test bench (1: butterfly damper, 2: electromagnetic flowmeter, 3: circulating pump, 4: high-pressure tank, 5: bidirectional axial-flow pump, 6: low-pressure tank).
Figure 5. Structure of the test bench (1: butterfly damper, 2: electromagnetic flowmeter, 3: circulating pump, 4: high-pressure tank, 5: bidirectional axial-flow pump, 6: low-pressure tank).
Entropy 24 00962 g005
Figure 6. Comparison of hydraulic performance between the CFD results and the test data of the arc-shaped bidirectional axial-flow pump under (a) forward and (b) reverse conditions.
Figure 6. Comparison of hydraulic performance between the CFD results and the test data of the arc-shaped bidirectional axial-flow pump under (a) forward and (b) reverse conditions.
Entropy 24 00962 g006
Figure 7. Comparison of pump performance between arc- and S-shaped cases under (a) forward and (b) reverse conditions.
Figure 7. Comparison of pump performance between arc- and S-shaped cases under (a) forward and (b) reverse conditions.
Entropy 24 00962 g007
Figure 8. TEP distribution for arc- and S-shaped cases under (a) forward and (b) reverse conditions.
Figure 8. TEP distribution for arc- and S-shaped cases under (a) forward and (b) reverse conditions.
Entropy 24 00962 g008
Figure 9. Proportion of entropy production of each hydraulic component for (a) arc-shaped and (b) S-shaped cases under the forward condition.
Figure 9. Proportion of entropy production of each hydraulic component for (a) arc-shaped and (b) S-shaped cases under the forward condition.
Entropy 24 00962 g009
Figure 10. Proportion of entropy production of each hydraulic component for (a) arc-shaped and (b) S-shaped cases under the reverse condition.
Figure 10. Proportion of entropy production of each hydraulic component for (a) arc-shaped and (b) S-shaped cases under the reverse condition.
Entropy 24 00962 g010
Figure 11. Position of the turbo surface and the volume.
Figure 11. Position of the turbo surface and the volume.
Entropy 24 00962 g011
Figure 12. Volume average TEPR distribution in (a) arc-shaped and (b) S-shaped impellers under three forward flow coefficients.
Figure 12. Volume average TEPR distribution in (a) arc-shaped and (b) S-shaped impellers under three forward flow coefficients.
Entropy 24 00962 g012
Figure 13. TEPR in the sections of the impeller passage under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Figure 13. TEPR in the sections of the impeller passage under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Entropy 24 00962 g013aEntropy 24 00962 g013b
Figure 14. Volume-averaged TEPR distribution in guide vanes for (a) arc-shaped and (b) S-shaped cases under three forward flow coefficients.
Figure 14. Volume-averaged TEPR distribution in guide vanes for (a) arc-shaped and (b) S-shaped cases under three forward flow coefficients.
Entropy 24 00962 g014
Figure 15. TEPR in the turbo surface of guide vanes under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Figure 15. TEPR in the turbo surface of guide vanes under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Entropy 24 00962 g015aEntropy 24 00962 g015b
Figure 16. TEPR and velocity vector in the elbow pipe under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Figure 16. TEPR and velocity vector in the elbow pipe under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Entropy 24 00962 g016aEntropy 24 00962 g016b
Figure 17. Volume-averaged TEPR distribution in (a) arc-shaped and (b) S-shaped impellers under three reverse flow coefficients.
Figure 17. Volume-averaged TEPR distribution in (a) arc-shaped and (b) S-shaped impellers under three reverse flow coefficients.
Entropy 24 00962 g017
Figure 18. TEPR in the turbo surfaces of the impeller under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Figure 18. TEPR in the turbo surfaces of the impeller under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24. (R* = 0.95).
Entropy 24 00962 g018
Figure 19. TEPR and velocity vector in the straight pipe under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24.
Figure 19. TEPR and velocity vector in the straight pipe under (a) Cq = 0.15, (b) Cq = 0.21, and (c) Cq = 0.24.
Entropy 24 00962 g019
Table 1. Rated operational parameters of two bidirectional axial-flow pumps (calculated data).
Table 1. Rated operational parameters of two bidirectional axial-flow pumps (calculated data).
ParameterUnitArc-ShapedS-Shaped
Forward-design flow ratem/s0.340.34
Forward-design headm3.32.9
Forward rotation speedr/min13501350
Forward specific speed 1173.51292.9
Reverse-design flow ratem/s0.280.28
Reverse-design headm2.93.1
Reverse rotation speedr/min13501350
Reverse specific speed 1626.01116.1
Table 2. Specific parameters of the test instruments.
Table 2. Specific parameters of the test instruments.
MeasurementItemsTest InstrumentMeasurement
Range
Measurement
Uncertainty
Flow rateElectromagnetic flowmeter OPTIFLUX2000F0~1800 m3/h E Q = 0.2%
HeadIntelligent differential pressure transmitter EJA0~10 m E H = 0.1%
TorqueIntelligent torque speed sensor JCL10~200 N·m E T = 0.1%
Rotation speed E n = 0.1%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Meng, F.; Li, Y. Energy Characteristics of a Bidirectional Axial-Flow Pump with Two Impeller Airfoils Based on Entropy Production Analysis. Entropy 2022, 24, 962. https://doi.org/10.3390/e24070962

AMA Style

Meng F, Li Y. Energy Characteristics of a Bidirectional Axial-Flow Pump with Two Impeller Airfoils Based on Entropy Production Analysis. Entropy. 2022; 24(7):962. https://doi.org/10.3390/e24070962

Chicago/Turabian Style

Meng, Fan, and Yanjun Li. 2022. "Energy Characteristics of a Bidirectional Axial-Flow Pump with Two Impeller Airfoils Based on Entropy Production Analysis" Entropy 24, no. 7: 962. https://doi.org/10.3390/e24070962

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