Next Article in Journal
A DPSIR-Driven Agent-Based Model for Residential Choices and Mobility in an Urban Setting
Previous Article in Journal
The Performance and Qualitative Evaluation of Scientific Work at Research Universities: A Focus on the Types of University and Research
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of the Sustainable Cooperation between a Multi-Piped Impeller and a Concentric Casing Using Experimental Planning

by
Bartłomiej Chomiuk
,
Janusz Skrzypacz
* and
Marcin Bieganowski
Department of Energy Conversion Engineering, Faculty of Mechanical and Power Engineering, Wroclaw University of Science and Technology, 50-370 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(18), 8179; https://doi.org/10.3390/su16188179
Submission received: 18 August 2024 / Revised: 15 September 2024 / Accepted: 18 September 2024 / Published: 19 September 2024
(This article belongs to the Section Sustainable Engineering and Science)

Abstract

:
Centrifugal pumps are one of the most commonly used devices in all branches of industry. Their application area is very broad and practically related to all sectors of the economy. Such pumps are widely applied in the chemicals industry, liquid gas technology, and machine lubrication systems, among others. On the basis of available data, it can be assumed that pumps consume approximately 27% of the world’s electricity. For this reason, their efficiency has a significant impact on the energy consumption of many technological processes. This article presents a summary of the research concerning the cooperation between a multi-piped impeller and a concentric casing. A multi-piped impeller is a completely new concept of a centrifugal impeller, which is characterized by the fact that the energy of the liquid is transferred as a result of the flow of media through the internal flow channels, and also the flow of media around it. The type and design of the casing, with regard to the flow in it and around the impeller’s pipes, plays an extremely important role. Numerical analysis was used as the main research method, which was based on a rotatable experimental plan. The CFD simulation results were validated on a test stand. Finally, the mathematical model and the rules for selecting the geometric features of such a stator were proposed in order to achieve the highest possible efficiency when cooperating with the multi-piped impeller, which constitutes a scientific novelty.

1. Introduction

Centrifugal pumps are one of the most frequently used devices in all branches of industry. It is estimated that the transport of liquids in the economy consumes nearly 27% of the electricity produced in the world [1]. It is extremely difficult to design centrifugal pumps with low specific speed and at an acceptable level of efficiency due to increasing losses. For this reason, their efficiency has an important impact on energy consumption in the global economy. Sustainability of centrifugal pump components and increasing the efficiency of all pump units by only 1% would result in annual savings of EUR 27 billion (electricity consumption in the global economy is estimated at 2700 TWh per year [1]). To illustrate the scale of the phenomenon, we will use an example from the Polish energy sector. In 2017, electricity production in Poland amounted to slightly more than 165 TWh (according to the Polish Power Grid Operator), with the majority generated from coal-fired power plants: approximately 50% from hard coal and about 32% from lignite. Assuming that the electricity consumption for the self-operation of thermal power plants related to pump operation is approximately 6.2%, this translates to an energy utilization figure of around 8.1 TWh, which corresponds to CO2 emissions at a level of 5.3 Tg (assuming 630 kg CO2 per 1 MWh). The savings achieved by optimizing the operation of pumps in professional power plants lead to a significant reduction in the carbon footprint as well as considerable economic and environmental benefits.
Although centrifugal pumps have been well known and used for many years, there is still room to develop innovative solutions for their flow elements. An example of such an element is a multi-piped impeller, the concept and name of which were developed and patented at Wroclaw University of Science and Technology [2]. The idea of such a solution uses flow-through internal flow channels, as in a classic centrifugal impeller. Additionally, it uses the external flow around the channels, which in turn generates an additional 30% of lifting height [3,4,5].
Information regarding such impellers can be found in the publication [3], which presents the results of energy tests of such an impeller in comparison with hole impellers and also presents the principles of flow modeling in such structures. Publication [4] presents the results of comprehensive research on the influence of the geometric features of a multi-piped impeller on operating parameters.
An extremely important factor affecting the total efficiency of a pump is the cooperation between the impeller and the casing. This aspect becomes particularly important in the case of a multi-piped impeller due to the phenomena of flow around the impeller’s channels. For this reason, the casing must be selected and designed in a completely different way than for a classic centrifugal or hole impeller, as illustrated in Figure 1.
Concentric casings (Figure 2) are one of the possible types of elements for removing liquid from an impeller—mainly in single-stage pumps. They are used, for example, for the hydrotransport of liquids containing solids. The principles of constructing such elements are limited to calculating the cross-sectional area of the casing, while at the same time maintaining a constant average velocity of the liquid in the cu2 channel (in the case of a channel without a bladeless casing). They are widely described in the literature [6,7]. Additionally, it is assumed that the ratio of the casing’s width to the impeller’s outlet width should be approximately b3/b2 = 1.2 ÷ 1.4, and the ratio b3 to the casing’s height h will probably be optimal within the range b3/h = 1.5 ÷ 2.5.
A concentric casing is less hydraulically advantageous than a collecting spiral due to the fact that it mixes liquid streams with different velocities. Paper [8] states that the efficiency of extremely low-specific speed centrifugal pumps with a concentric casing is sometimes higher than the efficiency of a unit with a spiral volute.
The disadvantage of this type of stator is the fact that liquid streams with different velocities mix with each other, causing an increase in hydraulic losses. Preliminary research by the authors of [8] also shows that a concentric casing can be a highly efficient structure that competes with a spiral liquid drainage element. Papers [9,10] present the results of experimental research concerning the geometry of a concentric casing that cooperates with semi-open and open blade impellers. The optimized geometry of the stator was characterized by having higher efficiency than spiral channels that were designed for the same discriminants of specific speed.
There is a lack of data and knowledge of how a concentric casing will cooperate with a multi-piped impeller, as well as what flow phenomena will occur (and what their intensity will be), and also how to optimally determine the dimensions of such a stator in order to ensure the highest possible efficiency of the entire pump.
The need to know and understand the working process that takes place in a concentric casing, as well as the answer to the question of how to optimally design the flow geometry of this element, constitute an important problem of pump technology. In this paper, numerical calculations were used as the main research method, and they were based on a rotatable experimental plan. The CFD simulation results were validated on a test stand. Statistical analysis of the results allowed a mathematical model that describes the energy equation of the pump in question to be proposed.

2. Research Object and Test Rig

The object of the numerical and experimental research is a model pump (operating parameters: capacity Q = 4.8 m3/h, lifting height Hu = 24.43 m, and kinematic-specific speed factor nq = 9.54), which consists of a multi-piped impeller (Figure 3a) and a casing with a radial diffuser (Figure 3b). The liquid drainage element was designed in accordance with guidelines in the literature [6,7,8]. Experimental tests were performed in accordance with the PN—EN ISO 9906:2012 standard (Grade 1) [11]. The measurement is made from the pump’s maximum capacity to its minimum capacity, and then from its minimum capacity to its maximum capacity. Each measured quantity is sampled 10 times in order to eliminate random errors. The experimental tests were carried out on a fully automated test stand, which is shown schematically in Figure 3c.
The geometry of the impeller and its dimensions were unchanged during all the performed tests. The geometric parameters of the multi-piped impeller and the model concentric casing, which were tested on the measurement station, are presented in Table 1 and Table 2.

3. Numerical Modeling

In order to understand the flow phenomena that occur in a pump with a multi-piped impeller and a concentric casing, and to obtain a reliable research tool for further research work, numerical simulations of fluid dynamics were used. They were performed as non-stationary calculations using the commercial software ANSYS FLUENT 16.2 [12]. The domain moving at a given rotational speed of n = 2870 rpm was a multi-piped impeller. The Sliding Mesh Model (SMM) [12,13] was used while taking into account the following settings:
  • Pressure–velocity model according to the SIMPLE scheme;
  • Isothermal calculations;
  • Discretization of all equations using the second-order upwind scheme;
  • The criterion of convergence of continuity equations was set at 1∙10−5 (for RMS values);
  • K-ω SST turbulence model (the boundary conditions of the model for clear water);
  • The model made one rotation of the impeller, which was divided into 120 time steps, in turn giving 3° of rotation per time step—its value amounted to t1s = 1.74·10−4 s, with one rotation of the impeller lasting t1obr = 2.09·10−2 s;
  • A maximum number of 1000 iterations was assumed for one time step;
  • Fluid: pure water with a density of ρ = 998.2 kg/m3, dynamic viscosity of μ = 1003 ·10−3 Pa·s, and at a temperature of t = 20 °C;
  • The initial parameters of the numerical model were determined based on the stationary flow solution.
The discrete model of the base pump consisted of five water solids (Figure 4a). Unstructured, tetragonal meshes were used for discretization, with a densification of elements near the walls (the mapping of the boundary layer)—Figure 4b,c. The criterion of y+ < 1 was used as the criterion for evaluating the computational mesh, which is consistent with the assumptions for the k-ω SST turbulence model [13,14] for non-stationary calculations. The thickness of the first layer was Δy1 = 0.02 mm. Detailed information regarding modeling has been included in the publications [3,4].
In order to determine the optimal mesh size in terms of accuracy and speed of calculations, GIT (Grid Independence Test) [15,16] was used, the results of which are presented in Table 3. On its basis, it can be concluded that mesh no. 4 (marked in the table in bold) is the smallest possible discrete mesh size, the number of elements of which does not affect the solution. The difference in values for the effective lifting height of the pump Hu between mesh variants no. 3 and no. 4 is δHu_3/4 = 1.70%, and for mesh no. 4 and mesh no. 5—δHu_4/5 = 0.46%. In the case of torque Mwir, the difference in values is δMwir_3/4 = 3.63% and δMwir_4/5 = 0.65%.
To obtain a reliable research tool in the form of a numerical model of a base pump with a multi-piped impeller, the discretization error was analyzed using the GCI (Gird Convergence Index) test [17,18]. The obtained values of GCIfine21 = 1.39% and GCIfine32 = 1.42% for the considered model parameters Mwir and Hu were at a very low level.
The pump’s operating parameters, obtained from the numerical calculations, were determined on the basis of the following relationships:
  • Pump lifting height:
H u = p o u t p i n ρ g
  • where pout is the total pressure at the outlet of the model (Figure 4—OUTLET) and pin is the total pressure at the inlet to the model (Figure 4—INLET);
  • Power on the pump’s shaft:
P w = M w i r ω
  • where Mwir is the total torque on the moving outer and inner walls of the multi-piped impeller;
  • Computational (CFD) hydraulic efficiency:
η h = ρ g Q H u M w i r ω
  • Total unit efficiency:
η c = η h η v η m 1 ζ t
  • where ηv is the volumetric efficiency of the pump (ηv = 0.92), ηm is the mechanical efficiency of the unit (ηm = 0.96), and ζt is the friction loss coefficient of the rotating discs assumed based on the diagram ζt = f(Re) [7] (ζt = 0.0047).
In the presented considerations, volumetric and mechanical losses were not modeled. The assumed efficiency values of ηv and ηm in Equation (4) result from the authors’ previously conducted research [2,3,4]. They had a constant value over the entire efficiency range, which may contribute to the errors in the numerical modeling, especially in the case of the extreme values of the pump’s efficiency.
In order to determine the accuracy of the obtained numerical simulation results for the model pump with regard to the experimental test results, the numerical characteristics were compared with the experimental characteristics, as shown in Figure 5.
After analyzing the data presented in Figure 5, the following conclusions can be drawn:
  • In the pump’s efficiency range of 2.2 m3/h < Qn < 5.2 m3/h, the maximum discrepancy between the numerical and experimental results does not exceed 1.4%;
  • In the case of lifting height, the difference between the numerical and experimental values at the operating point is equal to δHu_BEP = 0.36%;
  • For the total efficiency, the difference between the results at the operating point is equal to δηc._BEP = 0.64%—assuming a volumetric and mechanical efficiency at the level of ηv = 0.92 and ηm = 0.96;
  • The accuracy of matching the results of the numerical simulations of the curve of the power on the pump’s shaft to the real results, for the efficiency of Q = Qn, is at the level of δPw_BEP = 0.71%.

4. Testing the Cooperation between the Multi-Piped Impeller and the Concentric Casing

CFD numerical tests were planned in order to analyze the flow phenomena that occur during cooperation between the multi-piped impeller and the concentric casing, as well as to determine the impact of the geometric features of the stator on the pump’s operating parameters. In order to limit the number of the performed numerical simulations, experiment planning techniques were used. From the dimensional analysis [19,20,21], the relationship between the unit energy and geometric parameters of the concentric casing was obtained. To limit the number of parameters that influence the energy conversion process in a centrifugal pump with a multi-piped impeller, the following assumptions were made:
  • A constant flow geometry of the multi-piped impeller;
  • A constant value of the opening angle of the outlet diffuser (δmax = 8°);
  • The liquid discharge element is symmetrical in relation to the axis of the impeller’s flow channels (pipes);
  • A rectangular shape of the cross-section of the stator [8].
Based on the theory of dimensional analysis [19,20,21], and by taking into account the above assumptions, a function expressing the influence of the geometric dimensions of the concentric casing on the amount of unit energy related to the outer diameter of the impeller d2 was obtained and written in the form:
g H u n 2 d 2 2 = f π 8 , π 9 = f d 4 d 2 , b 3 d 2 = > d 4 = Z 8 ,   b 3 = Z 9
Taking into account that the concentric casing is an extremely simple element, and that only two parameters determine its geometry—width and diameter—(Figure 6), variability in these parameters was assumed within the following limits: width b3 = 18 ÷ 22 mm; external diameter of the casing d4 = 160 ÷ 170 mm. In order to design the research, a rotatable experimental plan was used, which is widely applied in hydraulic machines [22,23]. Nine stator geometries were prepared and then numerically tested. The values of the geometric parameters, which were determined according to the experimental plan, are presented in Table 4.

4.1. Results of Numerical Calculations

Figure 7 shows the results of the numerical calculations for the nine tested pump geometries, which are presented with regard to the lifting height in the function of the pump’s efficiency.
The highest value of the lifting height, which is a measure of the energy transported to the liquid, was obtained by the kk7 casing geometry with dimensions b3 = 17.2 mm and d4 = 165 mm (model pump b3 = 22.5 mm and d4 = 180 mm). As a result of reducing the flow cross-section of the concentric casing by more than 30%, the effective lifting height of the pump increased by almost 17%.

4.2. Analysis of the Results of the Experiment Plan

In order to check the influence of the pump capacity Q on the change in the unit energy Y, a statistical analysis [24,25] of the obtained results of the numerical simulations for the pump efficiency Q (shown in Figure 7) was carried out. Based on the statistical analysis of the regression function, a mathematical model of the pump with the multi-piped impeller and concentric casing was determined, and written as
Y = g H u = Q 2 d 2 4 173.64 ζ d 4 994.66 ζ b 3 + 274.74 + n Q d 2 27.87 ζ d 4 + 126.31 ζ b 3 178.847 + n 2 d 2 2 5.199 ζ d 4 7.962 ζ b 3 + 13.202
where
ζd4—discriminant of the external diameter of the stator d4/d2 = π8
ζb3—discriminant of the cross-sectional width of the stator b3/d2 = π9.
For the calculated regression coefficients of the polynomial (6), the obtained coefficients of determination R2 ∈ (0.802; 0.941) prove that the model fits very well. Moreover, it was shown, based on the F-Snedecor test, F ∈ (5,33, 39,43) > Fkr = 5.14), and t-student test, t ∈ (17.79; 37.01) > tkr = 2.45), that all the variables have a significant impact on the polynomial. In engineering practice, the analytical model from Equation (6) can facilitate the design process of the concentric casings of pumps with multi-piped impellers.

4.3. Verification of the Test Results

In order to verify the results of the tests of the derived mathematical model and the results of the numerical calculations, the flow geometry of the optimal solution of the concentric casing was designed. This stator was then printed (PET-G) using FDM technology (Figure 8a,b) and analyzed on a modernized test stand (Figure 8c,d).
The results of the conducted research include the energy characteristics of the pump, which are shown in Figure 9. These characteristics were then compared with the initial characteristics of the base model.
As can be seen in Figure 9, the increase in the operating parameters of the optimized geometry is identical to the results obtained within the experimental design. The experimental studies confirm that the numerical modeling that was used to conduct the research as part of the experimental design can be treated as a reliable research tool for assessing the impact of the geometric parameters of the casing of a pump on the energy characteristics of a pump with a multi-piped impeller. Validation of the discrete model showed that the maximum error between the results of the numerical simulations and the results of the experimental measurements did not exceed 3% (in BEP the discrepancy in the results was below 1%). The total efficiency of the unit increased by over 16 percentage points (for Q = 4.8 m3/h). Power demand dropped by nearly 24%.
The accuracy of the mathematical model of the centrifugal pump with the multi-piped impeller and the optimized concentric casing in relation to the numerical and experimental calculations is shown in Figure 10.
The analysis of the lifting height characteristics (Figure 10) allows for the conclusion that the mathematical model and the performed numerical calculations allow (with satisfactory accuracy) the actual operating parameters of the pump to be approximated and the following observations to be formulated:
  • The maximum discrepancy (in the case of lifting height) between the experimental results of the pump with the modernized kk7 concentric casing and the numerically modeled pump with the same casing geometry amounts to ∆H ≈ 4.5 m for Q = 6.2 m3/h. In the optimal point Qn_opt = 4.8 m3/h, this deviation amounted to ∆H ≈ 0.9 m (approximately 3.2% more than for the pump from the experimental stand);
  • When comparing the mathematical model of the pump with the kk7 stator and the results of the numerical CFD analysis for the same geometry of the stator, ∆H = 0.76 m (2.91% when compared to the experimental tests) for Q = 6.2 m3/h. In the optimal point Qn_opt = 4.8 m3/h, this deviation is ∆H = 0.132 m (approximately 0.46% more than for the discrete pump with the kk7 stator). Within the range of 2.2 m3/h < Qn < 5.2 m3/h, the maximum error between the results obtained from the mathematical model of the pump and the results of the pump with the kk7 concentric casing (numerically simulated) does not exceed 0.82%;
  • For Qn_opt = 4.8 m3/h, the difference in the lifting height between the base pump and the unit with the optimized flow geometry of the kk7 stator is ∆H ≈ 3.3 m;
  • The obtained results prove that the mathematical model fits well with the results of the numerical simulations of the pump with the stator kk7, as well as the results of real tests of the modernized model pump.
From the analysis of the partial characteristics of the influence of the geometric features of the stator on the increase in unit energy Y, the following dimensional relationships can be determined:
  • In the case of diameter d4:
d 4 = ( 1.053 ÷ 1.1 ) d 2
  • In the case of width b3:
b 3 = ( 0.115 ÷ 0.133 ) d 2
The obtained Equations (7) and (8) can be used to correctly design stators that cooperate with the multi-piped impeller in slow-running centrifugal pumps.

4.4. Analysis of the Graphical Numerical Results

The performed non-stationary numerical calculations allow for a quantitative and qualitative assessment of the energy conversion process that takes place in the pump with the multi-piped impeller and the concentric casing. The assessment of liquid flow is best done by analyzing the vorticity distributions (Figure 11), thanks to which the mechanism of energy transfer can be assessed. Vorticity is a measure of a fluid’s tendency to rotate. Together with the Shear Strain Ratio, it can be used to assess losses that occur in the flow [10]. The vorticity distributions are shown in the cross-section of the casing on two planes: P4 passing through the axis of the tube and P5 in the plane halfway between the impeller’s tubes (Figure 1a).
Figure 12 shows a comparison of the total pressure distributions in the pump with the multi-piped impeller and the concentric casing (on three control planes).
The reduction in the flow geometry of the concentric casing results in a reduction in the value and area of vorticity in the space between the impeller’s pipes (Figure 11c,d) when compared to the base pump model. In the case of the basic variant, the liquid flowing around the tubes generates an increase in vorticity, which in turn may result in an increase in losses. In the optimal variant, the intensity of vorticity around the edges of the pipes is reduced. The change in vorticity translates into the pump’s efficiency. The reduction of the cross-sectional area of the concentric casing—in particular, its width b3—causes a decrease in the intensity of liquid turbulence. This is especially the case in the area between the impeller’s pipes, which in turn results in an increase in the hydraulic efficiency of the pump.
Similar conclusions can be drawn from the analysis of the total pressure distribution shown in Figure 12. The reduction in the cross-section of the concentric casing influenced the equalization of pressure within the impeller’s pipes (Figure 12e). As the rotor radius increases, an increase in pressure in the concentric casing space around the tubes can be observed (Figure 12b,e). In the gaps between the outer walls of the impeller’s flow channels, and the walls of the kk7 pump’s casing, the pressure fields within the impeller were equalized in the case of the basic geometry of the concentric stator (Figure 12d,f).

5. Conclusions

A single-stage centrifugal pump with a multi-piped impeller that works with a properly selected and designed concentric casing is an interesting alternative to classic centrifugal pumps, both in terms of its operating parameters and implementation costs. Stators should be designed in order to minimize hydraulic losses, and at the same time to maximize the effect of liquid circulation. This phenomenon can also be used to process the intensification of the simultaneous mixing and pumping process of mixtures or substances.
The developed calculation formulas allow the geometric features of the concentric casing to be linked with the diameter d2 and the achieved lifting height Hu, thanks to which they can be used in the design procedures of slow-running pumps with multi-piped impellers. The obtained mathematical model may complement the design algorithm for centrifugal pumps with multi-piped impellers. It also provides a foundation for further research for the development of hybrid centrifugal-side channel pumps.
According to the results of the numerical and experimental studies, modification of the flow geometry of the casing reduces liquid recirculation in the cross-section of the pump and significantly improves its operating parameters.
The performed numerical simulations, laboratory measurements, and statistical analyses allow the following conclusions to be formulated:
  • It was confirmed that numerical modeling can be treated as a reliable research tool for assessing the impact of the geometric parameters of a casing on the energy characteristics of a pump with a multi-piped impeller. Validation of the discrete model showed that the maximum error between the results of the numerical simulations and the results of the experimental measurements did not exceed 3% (in BEP the discrepancy in the results was below 1%);
  • The highest value of the lifting height of the pump with the concentric casing, with an optimal efficiency of Qn = 4.8 m3/h, was obtained for the kk7 casing. At the nominal point, Hu_exp_opt = 27.65 m in relation to Hu_exp_base = 24.43 m for the basic concentric casing;
  • The highest increase in efficiency Δηc was also obtained for the kk7 casing; Δηc = 37.63%—this translates into over 17 percentage points of efficiency increase;
  • The ranges of the geometric features of the concentric casing in relation to the outlet diameter of the multi-piped impeller d2, for which the best operating parameters are achieved, were determined;
  • Taking into account the production advantages, a rational element for draining the liquid of a single-stage centrifugal pump with a multi-piped impeller (which operates in the range of low and extremely low values of the kinematic-specific speed factor) seems to be a concentric casing with the relative dimensions d4 = 1.1d2 and b3 = 0.115d2.

Author Contributions

Conceptualization, B.C.; methodology, B.C.; software, M.B.; validation, B.C. and M.B.; formal analysis, B.C.; investigation, B.C. and J.S.; resources, J.S.; data curation, B.C., J.S. and M.B.; writing—original draft preparation, B.C.; writing—review and editing, B.C., J.S. and M.B.; visualization, B.C. and J.S.; supervision, J.S. All authors have read and agreed to the published version of the manuscript.

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 available on request from the corresponding author.

Acknowledgments

Calculations were carried out using resources provided by the Wroclaw Centre for Networking and Supercomputing (http://wcss.pl, accessed on 18 August 2024), grant No. 444/2017.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

b2impeller widthmm
b3concentric casing widthmm
d2impeller outlet diametermm
d3concentric casing inlet diametermm
d4concentric casing external diametermm
ggravity accelerationm/s2
Hupump lifting heightm
inext value-
Mwirtotal torque on the moving walls of the impellerNm
nrotational speeds−1
nqKinematic-specific speed factor-
Nnumber of experiments in the experiment plan-
pintotal pressure at the inlet of the modelPa
pouttotal pressure at the outlet of the modelPa
Qcapacitym3/h
R2determination coefficient-
ReReynolds number-
ttime step, time, temperature-, s, °C
y+Reynolds number in cell-
Yunit energyJ
Δy1thickness of the boundary first layermm
znumber of impeller flow channels-
Greek Symbols
βinflow angle, offset angle°
δdiffuser opening angle, relative error°
Δvariability, difference-
ζtfriction loss coefficient-
ηefficiency-
μdynamic viscosity coefficientPa∙s
νkinematic viscosity coefficientm2/s
πnumber, dimensionless variable-
ρfluid densitykg/m3
ωangular velocityrad/s
Subscripts
baseapplies to casing of model pump-
ctotal-
CFDapplies to numerical simulations-
expexperimental-
hhydraulic-
kkapplies to concentric casing in a rotatable plan
mmechanical-
nnominal-
optoptimal-
vvolumetric-

References

  1. Market Analysis Raport, Global Industrial Pump Market Size Report 2020–2027, Grand View Research, Report ID: GVR-1-68038-325-6, USA. 2020. Available online: https://www.grandviewresearch.com/industry-analysis/industrial-pumps-market (accessed on 18 August 2024).
  2. Skrzypacz, J. Wirnik Pompy Wirowej (Impeller Pump Rotor). PL Patent 386135A1, 23 September 2008. [Google Scholar]
  3. Skrzypacz, J. Numerical modelling of flow phenomena in a pump with a multi-piped impeller. Chem. Eng. Process. 2014, 75, 58–66. [Google Scholar] [CrossRef]
  4. Skrzypacz, J. Investigating the impact of multi-piped impellers design on the efficiency of rotodynamic pumps operating at ultra-low specific speed. Chem. Eng. Process. 2014, 86, 145–152. [Google Scholar] [CrossRef]
  5. Varchola, M. Špeciálne Hydrodynamické Čerpadlá (Special Hydrodynamic Pumps); Slovenská Technická Univerzita: Bratislava, Slovakia, 2017. [Google Scholar]
  6. Stępniewski, M. Pompy (Pumps); WNT: Warsaw, Poland, 1985. [Google Scholar]
  7. Karassik, J.I. Centrifugal Pumps, 2nd ed.; International Thomson Publishing: New York, NY, USA, 1998. [Google Scholar]
  8. Gulich, J.F. Centrifugal Pumps, 4th ed.; Springer Nature: Cham, Switzerland, 2020. [Google Scholar]
  9. Matsui, J.; Kurokawa, J.; Choi, Y.; Nishino, K. Flow in the low specific speed Centrifugal pump with circular casing. In Proceedings of the XXIIIrd IAHR Symposium on Hydraulic Machinery and Systems, Yokohama, Japan, 17–21 October 2006. [Google Scholar]
  10. Kagawa, S.; Kurokawa, J.; Matsui, J.; Choi, Y. Performance of very low specific speed centrifugal pumps with circular casing. J. Fluid Sci. Technol. 2007, 2, 130–138. [Google Scholar] [CrossRef]
  11. EN ISO 9906:2012; Rotodynamic Pumps. Hydraulic Performance, Acceptance Tests., Grades 1, 2 and 3. BSI: London, UK, 2012.
  12. ANSYS Inc. ANSYS FLUENT Theory Guide. Release 16.2; SAS IP, Inc.: Canonsburg, PA, USA, 2016. [Google Scholar]
  13. Salim, M.S.; Cheah, C.S. Wall y+ Strategy for Dealing with Wall-bounded Turbulent Flows. In Proceedings of the International Multi Conference of Engineers and Computer Scientists, Hong Kong, China, 18–20 March 2009; Volume II. [Google Scholar]
  14. Wileox, D.C. Turbulence Modelling for CFD, 3rd ed.; DCW Industries: San Diego, CA, USA, 2006. [Google Scholar]
  15. Lee, M.; Park, G.; Park, C.; Kim, C. Improvement of Grid Independence Test for Computational Fluid Dynamics Model of Building Based on Grid Resolution. Adv. Civ. Eng. 2020, 2020, 11. [Google Scholar] [CrossRef]
  16. Wang, H.; Zhai, Z. Analyzing grid independency and numerical viscosity of computational fluid dynamics for indoor environment applications. Build. Environ. 2012, 52, 107–118. [Google Scholar] [CrossRef]
  17. Roache, P.J. Verification of codes and calculations. AIAA J. 1998, 36, 696–702. [Google Scholar] [CrossRef]
  18. Celik, I.; Ghia, U.; Roache, P.J. Procedure for Estimation and Reporting of Uncertainty Due to Discretization in CFD Applications. ASME J. Fluid Eng. 2008, 130, 078001. [Google Scholar]
  19. Roslanowski, J. Valuation of rotodynamic pumps operation, by means of, dimensional analysis. J. Pol. CIMAC 2014, 2, 1–8. [Google Scholar]
  20. Szirtes, T. Applied Dimensional Analysis and Modeling; McGraw-Hill: New York, NY, USA, 1997. [Google Scholar]
  21. Barenblatt, G.I. Dimensional Analysis and Intermediate Asymptotics; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar]
  22. Dean, A.; Voss, D.; Dragulić, D. Design and Analysis of Expermients, 2nd ed.; Springer: Genewa, Switzerland, 1999. [Google Scholar]
  23. Si, Q.; Yuan, S.; Yuan, J.; Wang, C.; Lu, W. Multiobjective Optimization of Low-Specific-Speed Multistage Pumps by Using Matrix Analysis and CFD Method. J. Appl. Math. 2013, 10, 10. [Google Scholar] [CrossRef]
  24. Freedman, A.D. Statistical Models: Theory and Practice; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  25. Katalazhnova, I. Correlation and Regression Method of Centrifugal Pump Geometry Optimization. In Proceedings of the 4th International Conference on Industrial Engineering, ICIE 2018, Moscow, Russia, 15–18 May 2018; Springer: Cham, Switzerland, 2018; pp. 1839–1845. [Google Scholar]
Figure 1. List of structures and the identification of flow phenomena during the flow around (red arrows identify the liquid flow around the impeller): (a) a multi-piped impeller; (b) a hole impeller.
Figure 1. List of structures and the identification of flow phenomena during the flow around (red arrows identify the liquid flow around the impeller): (a) a multi-piped impeller; (b) a hole impeller.
Sustainability 16 08179 g001
Figure 2. Concentric casings (a) with radial liquid discharge; (b) with an axial discharge diffuser.
Figure 2. Concentric casings (a) with radial liquid discharge; (b) with an axial discharge diffuser.
Sustainability 16 08179 g002
Figure 3. Flow elements of the model pump analyzed on the test stand: (a) multi-piped impeller printed using SLS technology; (b) concentric casing; (c) scheme of the test stand: 1—pressure tank; 2—electromagnetic flowmeter; 3—automatic ball valve on the discharge pipe; 4—pressure transducer on the suction port; 5—pressure transducer on the discharge port; 6—electric motor; 7—flexible coupling; 8—pump’s body; 9—automatic ball valve on the suction pipe; 10—manual ball valves; 11—temperature sensor.
Figure 3. Flow elements of the model pump analyzed on the test stand: (a) multi-piped impeller printed using SLS technology; (b) concentric casing; (c) scheme of the test stand: 1—pressure tank; 2—electromagnetic flowmeter; 3—automatic ball valve on the discharge pipe; 4—pressure transducer on the suction port; 5—pressure transducer on the discharge port; 6—electric motor; 7—flexible coupling; 8—pump’s body; 9—automatic ball valve on the suction pipe; 10—manual ball valves; 11—temperature sensor.
Sustainability 16 08179 g003
Figure 4. Discrete model of the pump with a computational mesh: (a) computational domains; (b) boundary layer on the inner and outer walls of the multi-piped impeller; (c) boundary layer on the inner walls of the stator.
Figure 4. Discrete model of the pump with a computational mesh: (a) computational domains; (b) boundary layer on the inner and outer walls of the multi-piped impeller; (c) boundary layer on the inner walls of the stator.
Sustainability 16 08179 g004
Figure 5. Validation of the numerical characteristics and the experimental results. Solid curves and solid markers are the experimental results and dashed curves and cross markers are the numerical analysis results.
Figure 5. Validation of the numerical characteristics and the experimental results. Solid curves and solid markers are the experimental results and dashed curves and cross markers are the numerical analysis results.
Sustainability 16 08179 g005
Figure 6. Schematic presentation of the analyzed parameters of the concentric casing.
Figure 6. Schematic presentation of the analyzed parameters of the concentric casing.
Sustainability 16 08179 g006
Figure 7. Lifting height characteristics Hu = f(Q) for the tested flow geometries of the stator from the experimental plan.
Figure 7. Lifting height characteristics Hu = f(Q) for the tested flow geometries of the stator from the experimental plan.
Sustainability 16 08179 g007
Figure 8. Components of the optimal concentric casing: (a) ring body with rear disc; (b) front disc; (c) open-stator model; (d) complete casing.
Figure 8. Components of the optimal concentric casing: (a) ring body with rear disc; (b) front disc; (c) open-stator model; (d) complete casing.
Sustainability 16 08179 g008
Figure 9. Comparison of the experimental energy characteristics of the model pump (solid curves and intense markers), the modernized pump with an optimal geometry of the concentric casing (solid curves and brightened markers), and the numerical energy characteristics of the modernized pump (dashed curves and grey markers).
Figure 9. Comparison of the experimental energy characteristics of the model pump (solid curves and intense markers), the modernized pump with an optimal geometry of the concentric casing (solid curves and brightened markers), and the numerical energy characteristics of the modernized pump (dashed curves and grey markers).
Sustainability 16 08179 g009
Figure 10. Comparison of the energy characteristics of the model pump and the pump with the optimal geometry of the kk7 stator. The numerical results are represented by dashed curves and the experimental results by solid lines.
Figure 10. Comparison of the energy characteristics of the model pump and the pump with the optimal geometry of the kk7 stator. The numerical results are represented by dashed curves and the experimental results by solid lines.
Sustainability 16 08179 g010
Figure 11. Comparison of the vorticity distributions of the liquid circulating in the cross-section of the pump with the multi-piped impeller and the concentric casing (on control planes—scheme): (a) base variant P4; (b) base variant P5; (c) optimal variant P4; (d) optimal variant P5.
Figure 11. Comparison of the vorticity distributions of the liquid circulating in the cross-section of the pump with the multi-piped impeller and the concentric casing (on control planes—scheme): (a) base variant P4; (b) base variant P5; (c) optimal variant P4; (d) optimal variant P5.
Sustainability 16 08179 g011
Figure 12. Comparison of the total pressure distributions in the pump with the multi-piped impeller and the concentric casing (on control planes—scheme): basic—(a) P1; (b) P2; (c) P3—and optimal—(d) P1; (e) P2; (f) P3.
Figure 12. Comparison of the total pressure distributions in the pump with the multi-piped impeller and the concentric casing (on control planes—scheme): basic—(a) P1; (b) P2; (c) P3—and optimal—(d) P1; (e) P2; (f) P3.
Sustainability 16 08179 g012
Table 1. Summary of the main geometric parameters of the multi-piped impeller.
Table 1. Summary of the main geometric parameters of the multi-piped impeller.
No.NameSymbolValueUnit
1Hub diameterdp20mm
2Inlet diameterd140mm
3Outer diameterd2150mm
4Pipe diameterdk = b26mm
5Inlet angleβ150°
6Outlet angleβ280°
7Number of pipesz5
Table 2. Summary of the main geometric parameters of the concentric casing.
Table 2. Summary of the main geometric parameters of the concentric casing.
No.NameSymbolValueUnit
1Channel widthb322.5mm
2Inlet diameterd3155mm
3Outer diameterd4180mm
4Inlet diameter (into the diffuser)dt22mm
5Diffuser opening angleδmax2°
Table 3. Main parameters of the numerical meshes and the pump parameters for the GIT test.
Table 3. Main parameters of the numerical meshes and the pump parameters for the GIT test.
No.Number of Mesh ElementsAverage SkewnessAverage Quality of ElementsAverage Aspect RatioQMwirHu
112,956,5660.260.722.854.82.8620.82
216,843,5360.240.772.454.82.7823.53
322,065,0330.230.782.224.82.7325.34
428,684,5430.200.802.084.82.6325.78
537,003,0600.200.812.044.82.6125.90
647,363,9170.190.822.024.82.5925.84
Table 4. Values of the geometric parameters according to the numerical simulation plan (abbreviations: kk1–kk9 denote subsequent geometries of the concentric casing).
Table 4. Values of the geometric parameters according to the numerical simulation plan (abbreviations: kk1–kk9 denote subsequent geometries of the concentric casing).
NZ8π8Z9π9
d4 (mm)b3 (mm)
kk11601.067180.120
kk21701.133180.120
kk31601.067220.147
kk41701.133220.147
kk51581.053200.133
kk61721.147200.133
kk71651.10017.200.115
kk81651.10022.900.153
kk91651.100200.133
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chomiuk, B.; Skrzypacz, J.; Bieganowski, M. Analysis of the Sustainable Cooperation between a Multi-Piped Impeller and a Concentric Casing Using Experimental Planning. Sustainability 2024, 16, 8179. https://doi.org/10.3390/su16188179

AMA Style

Chomiuk B, Skrzypacz J, Bieganowski M. Analysis of the Sustainable Cooperation between a Multi-Piped Impeller and a Concentric Casing Using Experimental Planning. Sustainability. 2024; 16(18):8179. https://doi.org/10.3390/su16188179

Chicago/Turabian Style

Chomiuk, Bartłomiej, Janusz Skrzypacz, and Marcin Bieganowski. 2024. "Analysis of the Sustainable Cooperation between a Multi-Piped Impeller and a Concentric Casing Using Experimental Planning" Sustainability 16, no. 18: 8179. https://doi.org/10.3390/su16188179

APA Style

Chomiuk, B., Skrzypacz, J., & Bieganowski, M. (2024). Analysis of the Sustainable Cooperation between a Multi-Piped Impeller and a Concentric Casing Using Experimental Planning. Sustainability, 16(18), 8179. https://doi.org/10.3390/su16188179

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